MIC CHI GANSTATE ll lllllllllzllll llllllllllll‘llllllllllllllll 93 01026 2073 LIBRARY Michigan State University This is to certify that the dissertation entitled Investigation of the Ion Formation and Unimoleoular Decomposition Mechanisms in the Desorption/Ionization Technique Fast Atom Bombardment presented by Jason C. Rouse has been accepted towards fulfillment of the requirements for PhD. Chemistry degree in /%fl% Major professor Date 4 Jeff 73 MS U is an Affirmative Action/Equal Opportunity Institution 0- 12771 mouthi- checkmnflom your rooofd. PLACE iN RETURN 80X to data duo. 1’0 AVOID FINES roturn on or baton MSU is An Affirmative Action/Equal Opportunity institution Want ____————’ INVESTIGATION OF THE ION FORMATION AND UNIMOLECULAR DECOMPOSITION MECHANISMS FOR THE DESORPTION/IONIZATION TECHNIQUE FAST-ATOM BOMBARDMENT (FAB) MASS SPECTROMETRY By Jason Craig Rouse A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1993 ABSTRACT INVESTIGATION OF THE ION FORMATION AND UNIMOLECULAR DECOMPOSITION MECHANISMS FOR THE DESORPTION/IONIZATION TECHNIQUE FAST-ATOM BOMBARDMENT (FAB) MASS SPECTROMETRY By Jason Craig Rouse A divided probe that incorporates a potassium aluminosilicate glass target and an analyte/glycerol matrix target, spatially-separated, was used to inject potassium ions (K+) into the high pressure "selvedge" region formed above the analyte/glycerol matrix target during fast-atom bombardment (FAB). [M+K]+ adduct ions that represent the types of gas-phase neutral molecules (M) present in the selvedge region are observed. Computer modeling assisted in designing the divided target and an additional ion optical element for the FAB ion source to optimize interactions between K+ ions and the desorbed neutral molecules. The capability of injecting K+ ions into the FAB experiment has utility in both mechanistic studies and analysis. Experimental results here are consistent with a' model for the desorption/ionization processes in FAB in which some types of neutral analyte molecules, M, are desorbed intact and are subsequently protonated by glycerol- chemical ionization. Unstable protonated molecules undergo unimolecular decompositions to yield observed fragment ions. The use of K+ cationization of analytes for molecular weight confirmation is demonstrated, as well as its utility in FAB experiments in which mixtures are encountered. Peptide sequence ion fragmentation mechanisms that define the relationship between the relative intensity (RI) information in liquid secondary-ion mass spectrometry (LSIMS), metastable ion (MI) decomposition, and single-collision, high-energy collisional activation dissociation (CAD) mass spectra, and the protonated structure of pentaglycine (no side chains) and methionine enkephalin (side chains) were evaluated by using the RI distributions in the respective spectra and estimated reaction enthalpies for the formation of each sequence fragment ion, an, bn, on, x", yn and z". The most likely protonation sites in these pentapeptides are the amide nitrogens based on their high estimated proton affinities and by the straightforward fragmentation mechanisms that can be written from this site to explain the formation of each sequence ion. The RI distributions and reaction enthalpies suggest that both simple cleavages and secondary interactions (i.e., when a remote site of the molecule interacts with the protonation site to induce chemistry) are competitive processes from this site in LSIMS and CAD whereas only secondary interactions are prominent for M1. The ratios of fragment ions from each protonated amide nitrogen were determined not to be equal and it appears secondary interactions either redistribute the initial protonation sites and/or induce specific bn, y" and on fragmentation. Although the most favorable pathways are not always followed, the most stable fragment ions and neutrals are formed in every peptide fragmentation process, depending on the amount of the internal energy in the [M+H]+, whether simple cleavages or secondary interactions are required to propagate the chemistry. iv This dissertation is dedicated to my Father and in memory of my loving Mother, Lee and Phyllis, for their endless love, guidance, encouragement, and support which was essential in my pursuit of this advanced degree. I want to also thank my late Grandfather, Andrew T. F enrich, who referred to me as "his grandson, the doctor", for initially introducing me to science and providing me with love, encouragement, and support needed to pursue this profession. ACKNOWLEDGMENTS This advanced degree would have been impossible for me to complete alone, and I am eternally grateful to everyone who provided me with assistance, guidance, and moral, technical and financial support in my five years of graduate school at MSU. I am very grateful to Jeff Gilbert because he made my move from The College of Wooster, a small liberal arts college, to MSU a smooth transition both socially and academically. Jeff and I were both members of the Beta Kappa Phi Fraternity at Wooster, and, here in graduate school at MSU, we became best friends, roommates, and members of the Allison research group. Karen Light-Wahl was another College of Wooster graduate and a former Allison research group member who initially taught me how to operate a mass spectrometer. I appreciate all her efforts to help me begin my research project and I hope we can "spike" a beer (and break glass!) when I graduate. The first year of graduate school is very intellectually challenging and I am indebted to my fellow colleagues Kurt Kneen, Ed Townsend, Mike Waldo, and Julie Homer for their genuine friendships and their helpful contributions in our first-year study groups. I was glad to become good friends with fellow MSU students Kris Kurtz, Kimberly Gilbert, Gary Schultz, Tracy Kneen, Man'a Townsend, Laura Pence, Dave Gale, Jim Ridge, Jon Wahl, Dave Wagner, Art Harms, and Mary Puzycki. Amazingly, everyone mentioned above would get together quite often during the week to celebrate events at the local bars, go out to dinner, see a movie, play volleyball, etc., which was a needed break from the stresses of graduate school. In the NIH supported Mass Spectrometry Facility (MSF) in Biochemistry at MSU, I am appreciative of all the technical assistance from J. Throck Watson, Doug vi Gage, Mike Davenport, Mel Micke, Bev Chamberlin, and Melina Beming. The MSF has provided me a great working atmosphere, state-of-the-art instrumentation, and the opportunity to attend the numerous mass spectrometry meetings, both local and national. A most rewarding experience for me was performing maintenance with Mike Davenport on the JEOL HX-l 10 because I learned much about the electronics, nuts and bolts, and workings of a double-focusing mass spectrometer. I am obliged to Kermit Johnson and Evy Jackson in the Max T. Rodgers NMR Facility in Chemistry for their run-time assistance and technical support for the "bradykinin" project. Although the project is not finished, I still learned much about the cutting-edge techniques in NMR spectroscopy. I want to acknowledge Joe Leykam, the manager of the Structure Facility in Biochemistry, for teaching me the principles of peptide synthesis and purification. Last, I want to acknowledge NIH, the MSU Research Excellence Fund, and the Analytical Sciences Group at Dow Chemical for moneys supporting my research assistantships and supplies. I met my loving wife, Sue, at the College of Wooster my sophomore year. After I started at MSU, we endured a difficult, long-distance relationship until my third year of graduate school when we were finally married on August 17, 1991. We have been very happy ever since. Sue has provided me with love, encouragement, feedback, and moral support during my entire graduate school endeavor, and with her by my side, the difficult times and tasks became much easier knowing she was there rooting for me. Last, but by far the most important person during my graduate career at MSU was my advisor John Allison. I am very grateful to receive from JA a wealth of graduate career guidance, encouragement, and technical, moral and financial support over the years. The most rewarding times in graduate school were one-on-one meetings with JA either interpreting data, discussing possible research approaches, learning the correct way to give a seminar, or writing (or downsizing) a paper together. I learned the most about real science during those times. JA, I thank you for being a challenging but fair advisor, and I wish the best for you in the future. TABLE OF CONTENTS LIST OF TABLE xi LIST OF FIGURES x LIST OF ABBREVIATIONS xvii CHAPTER ONE. INTRODUCTION 1 1. Mass Spectrometry l 2. Overview of Ionization Techniques in Mass Spectrometry 3 3. The Focus, Goals, and Organization of the Dissertation 8 CHAPTER TWO. THEORY AND METHODS 13 1. The Ion Formation and Dissociation Processes in the Traditional Ionization Techniques Electron Ionization (E1) and Chemical Ionization (CI) - 13 2. Potassium Ion Ionization of Desorbed Species (K+IDS) 29 3. Fast-Atom Bombardment (FAB) Mass Spectrometry and Related Particle Bombardment D/I Techniques 34 4. Principles and Utility of Collisionally-Activated Dissociation (CAD) 58 5. The Mass Analyzers Used for D/I Experiments 62 CHAPTER THREE. K+IDS-BY-FAB TARGET DEVELOPMENT AND APPLICATION TO MECHANISTIC STUDIES OF ION FORMATION IN FAST-ATOM BOMBARDMENT MASS SPECTROMETRY 77 1. Introduction 77 2. Development and Characterization of the K+IDS-by-FAB Target 81 3. Mechanistic Studies of Ion Formation in FAB using KflDS-by-FAB 126 4. Application of NMR for Determining the Charge State of Peptides in Solution 142 5. Conclusions 145 vii viii CHAPTER FOUR. THE EVOLUTION OF FRAGMENT IONS FROM PROTON ATED PEPTIDES IN LIQUID SECONDARY-ION MASS SPECTROMETRY: MECHANISTIC AND THERMOCHEMICAL CONSIDERATIONS 1. Introduction 2. Experimental Details 3. Results and Discussion 4. Conclusions CHAPTER FIVE. FUNDAMENTAL AND ANALYTICAL APPLICATIONS OF GAS-PHASE METAL ION/BIOMOLECULE REACTIONS 1. Introduction 2. Reactivities between Metal Ions and Analyte Molecules in M+IDS 3. Gas-Phase Metal Cationization in FAB Using M+IDS-by-FAB APPENDD( ONE. INJECTION OF REAGENT IONS INTO THE SELVEDGE REGION IN FAST—ATOM BOMBARDMENT MASS SPECTROMETRY APPENDIX TWO. DATA FOR ESTIMATION OF REACTION ENTHALPIES FOR PEPTIDE SKELETAL BOND CLEAVAGES LIST OF REFERENCES 147 147 150 152 204 207 207 210 221 244 255 261 LIST OF TABLES Table 1. PAs and IE5 for small molecules with unique functional groups. All values were obtained from Lias et a1. [40] except when noted. ix Figure 1. Figure 2. Figure 3. Figure 4. Figure 5. Figure 6. Figure 7. Figure 8. Figure 9. Figure 10. Figure 1 1. Figure 12. Figure 13. Figure 14. Figure 15. Figure 16. Figure 17. LIST OF FIGURES Schematic diagram of a mass spectrometer. Schematic diagram of an E1 ion source. An EI clastogram for methanol (Adapted from [39]). Thermochemical considerations for unimolecular decomposition of M“. Estimated Wahrhaftig diagram for ionized CH3CH2CH2CH20CH3. Schematic diagram of the K+IDS experiment. An Arrhenius plot that demonstrates the competition between desorption and degradation processes in thermal heating. Schematic diagram of the fast atom gun (actually a SIMION model drawn to scale) and the FAB experimental configuration on the JEOL HX—l 10. Schematic diagram of the fast ion gun (actually a SIMION model drawn to scale) and the LSIMS experimental configuration on the JEOL HX-l 10. LSIMS mass spectrum of lug/11L of triglycine (MW 189) dissolved in the glycerol matrix. Energy deposition by a keV-Xe beam induces desorption of atoms, fragment ions, preionized molecules, and intact molecules (not drawn to scale). The neutral and ionic species desorbed into the selvedge region above the FAB target during Xe bombardment. The extraction and acceleration of M” a) from an ion source and b) the representative potential energy surface for this ion source. Schematic diagram of a double-focusing mass spectrometer. Schematic diagram of a quadrupole mass analyzer. A plot of the B, E plane for linked scans at constant B/E and B2/E on double-focusing instruments. A KHDS-by-FAB divided target under Xe bombardment demonstrating the injection of desorbed K+ ions into the selvedge region for the Figure 18. Figure 19. Figure 20. Figure 21. Figure 22. Figure 23. Figure 24. Figure 25. Figure 26. xi identification of desorbed neutral molecules. Figure 18. a) FAB mass spectra of K+ glass emitter and of b) KCl solid film on a divided target half without matrix or analyte present on the opposite half. c) KHDS-by-FAB mass spectrum of a KCl emitter and glycerol on either half of a dual-surface space-divided target. The latter spectrum is presented to show K0 is just as effective as K+ glass emitter in a K+IDS- by-FAB experiment, but the KCl emitter is more difficult to apply cleanly to only one target surface plus it needs to be regenerated for each run. a) The first K+IDS-by-FAB type target employing a K+ glass emitter reported by Akerrnann et al. [164], b) the wall and slit targets reported by Munster et al. [166, 168], c) the dual-surface target reported by Miller et al. [169], d) the KHDS-by-FAB "set screw" target reported by Schultz et al. [172], e) the K+IDS-by-FAB "wall-divided" target reported by Kassel [171], f) the K+IDS-by-FAB "K+ bead pop-up" target, g) the K+IDS-by- FAB "vertical arrangement" FD/FAB type divided target, and h) the vertical, dual—surface space-divided target, all which provided unsatisfactory gas-phase K+ adduct ion formation for mechanistic studies. a) FAB mass spectrum of glycerol (MW 92) matrix, b) typical K+IDS-by- FAB mass spectrum when the K+ glass emitter on any divided target is contaminated with glycerol/matrix, and c) K+IDS-by-FAB mass spectrum of the glycerol matrix obtained with the KflDS-by-FAB space-divided target and LO lens element installed in the ion source. Different views of the a) K+IDS-by-FAB space-divided target with the K+ glass emitter bead and analyte/matrix sample holder, b) its orientation with the impinging FAB beam and exiting secondary ions, and its typical c) side and d) top dimensions. a) FAB mass spectrum and b) early K+IDS-by-FAB mass spectrum of H- LGG-OH (MW 245) in glycerol on the standard JEOL target and K+IDS— by-FAB space-divided target, respectively. Early K+IDS-by-FAB mass spectrum of sucrose (MW 342) in glycerol using the K+IDS-by-FAB space-divided target. Schematic diagram for the implementation of thermally-assisted K+IDS-by- FAB on the JEOL HX-l 10. Figure 25. Thermally-assisted K+IDS-by-FAB mass spectra of a) thermally- heated K+ glass emitter, FAB off, and no analyte on FD/FAB sample target; b) thermally-heated K+ glass emitter, FAB off, and glycerol on FD/FAB target; and c) same as b except FAB is on. SIMION models of the FD/FAB ion source coupled with the a) standard FAB target, b) wall-divided target, c) dual-surface space-divided target, and K+IDS-by-FAB space-divided target showing (I) analyte/matrix ion and e) K+ ion trajectories. xii Figure 27. SIMION models of the FD/FAB ion source and L0 ion optical element Figure 28. coupled with the a) standard FAB target, b) wall-divided target, c) dual- surface space-divided target, and KflDS-by-FAB space-divided target showing (I) analyte/matrix ion and e) K+ ion trajectories. Estimated extraction times for typical ions from the a) FAB target, b) dual- surface space-divided target, and c) KflDS-by-FAB space-divided target SIMION models, with and without L0 present, shown in Figures 26 and 27a, c, d, and e, respectively. The KflDS-by-FAB space-divided target combined with L0 is best technique for injecting low energy K+ ions, with longer residence times, into the selvedge region. Figure 29. SIMION model of dual-surface space-divided target and full FD/FAB ion Figure 30. Figure 31. Figure 32. Figure 33. source a) shows K“ and glycerol ions, extracted from each respective target half, leave the ion source in two separate beams, b) mass spectrum of K+ beam from K+ glass emitter target half, c) mass spectrum of the glycerol beam from glycerol target half, (I) "rare" mass spectrum of both K+ and glycerol ion beams, e) typical mass spectrum of both K+ and glycerol ion beams showing glycerol contamination of K+ glass emitter, and 1') same SIMION model as in 3, except with L0 installed, shows a single well-defined beam leaving the ion source which eliminates ion beam focusing problems. Sequential K+IDS-by-FAB mass spectra of glycerol using the K+IDS-by- FAB space-divided target a) with L0 installed initially and b) then with L0 removed which demonstrates the effectiveness of L0 for enhanced [M+K]+ adduct formation. The KfiDS-by-FAB mass spectra of glycerol from sequential experiments in which the KflDS-by-FAB space-divided target was positioned a) distant from, and b) then close to the L0 lens element resulting in an increase in the intensity of the peak representing [G+K]+ in the latter experiment owing to the possible creation of a higher pressure region between the target and L0. The K+ cationization efficiency of desorbed neutral molecules and the effect of L0 slit width on K+ adduct formation is demonstrated in sequential experiments using thevetin A (MW 872) and B (MW 858). A comparison between the a) FAB mass spectrum, and the K+IDS-by-FAB mass spectra b) with a narrow L0 slit width (out of ion optical tune) and c) with a wide L0 slit width reveals 30% of total desorbed neutral molecules are K+ cationized, if [M+H]+ and [M+K]+ are summed from b and c, averaged, and compared to [M+H]+ in a. Slightly more K+ adducts are obtained. from a narrow slit width. a) A FAB mass spectrum, b) KflDS-by-FAB mass spectrum, using with the K+IDS-by-FAB space-divided target with L0 installed, and c) same as b, except L0 was not installed, were obtained for Kassinin (MW 1334) at a resolution of 1000. A comparison between b and c demonstrates that L0 enhances resolution of the peaks. Figure 34. Figure 35. Figure 36. Figure 37. Figure 38. Figure 39. Figure 40. Figure 41. xiii The utility of the LO lens element in front of the FAB target for delaying ion extraction and possibly providing a higher pressure region above the FAB target is demonstrated, in sequential scans, for the analysis of methionine enkephalin (MW 573) in glycerol. With L0 present, a) an intense peak representing [M+H]+ was observed in the initial FAB mass spectrum, b) a successful KflDS-by-FAB mass spectrum validated a, and last, c) L0 was removed resulting in a decrease in the intensity of [M+H]+. Sequential experiments to prove K+ forms adduct ions in the gas phase and K+ is not readily deposited into the upper layers of glycerol in K+IDS-by- FAB. First, a) K+IDS-by-FAB mass spectrum of glycerol on K+IDS-by- FAB space-divided target which yields 6 counts for [G+K]+, then b) the K+ glass emitter is detached from this target and the residual glycerol is reanalyzed by FAB to probe K+ contamination which yields 0.6 counts, and last c) the residual K+ in new glycerol placed on the same target in b is probed. Cholic acid (MW 408), in glycerol, a) FAB mass spectrum, b) K+IDS-by— FAB mass spectrum, c) FAB mass spectrum with 0.01 M KCl salts, d) FAB mass spectrum with 0.1 M KCl salts, e) FAB mass spectrum of analyte/matrix contaminated K+ glass emitter. Two possible mechanisms for the formation of the [M+H-3H20]+ fragment ion in FAB of cholic acid (MW 408). A comparison between the molecular weight region peak intensities for the analytes: 3) glycerol, b) cholic acid, c) digoxin, d) stachyose, e) kassinin, and f) bradykinin dissolved in glycerol and analyzed by FAB, by K+IDS- by-FAB, by FAB in 0.01 M KCl also, by FAB in 0.1 M KCl also, and on the surface of the K+ glass emitter by FAB. By applying an extremely small amount of glycerol on the target to reduce the glycerol-related reagent ions, K+ became the dominant reagent ion as shown in this K+IDS-by-FAB mass spectrum of digoxin (MW 780) since the peak representing the [M+K]+ adduct ion is more intense in comparison to [M+H]+ and there is no fragmentation observed from either species. a) FAB mass spectrum and b) KHDS-by-FAB mass spectrum of benzyltriethylammonium chloride (MW 227) in glycerol, with L0 present, obtained from a regular FD/FAB target and KflDS-by-FAB space-divided target, respectively. a) FAB mass spectrum and b) K+IDS-by-FAB mass spectrum, obtained with a regular FD/FAB target and the KflDS-by-FAB space-divided target, respectively, with L0 present, of a glycerophospholipid mixture containing l-stearoyl-2-palmitoleoyl-3-phosphatidylcholine and/or 1- palmitoyl-2-oleoyl-3-phosphatidylcholine as M] (MW 759), and l- stearoyl-2-oleoyl-3-phosphatidylcholine and/or 1-palmitoyl-2- arachidonoyl-3-phosphatidylcholine as M2 (MW 787), in glycerol. Figure 42. Figure 43. Figure 44. Figure 45. Figure 46. Figure 47. Figure 48. Figure 49. Figure 50. Figure 51. Figure 52. Figure 53. Figure 54. xiv Fractional composition diagrams for polyprotic peptides a) H-VGVAPG- 0H, b) kassinin, c) bradykinin, and d) sleep-inducing peptide in which the prominent forms at pH=7 and other peptides exhibiting similar behavior are listed on left hand column. a) FAB mass spectrum of H-VGVAPG-OH (MW 498), b) KflDS-by-FAB mass spectrum of H-VGVAPG-OH, and c) KHDS-by-FAB mass spectrum of H-VGVAUDPG-OH (MW 499) in glycerol using either the FD/FAB or K+IDS-by-FAB space-divided target, respectively, with L0 installed. a) FAB mass spectrum and b) KflDS-by-FAB mass spectrum of the sleep- inducing peptide, H-WAGGDASGE-OH (MW 848), in glycerol, with L0 installed, using the FD/FAB and K+IDS-by—FAB space—divided targets, respectively. Representative fragment ion structures for the peptide methionine enkephalin (MW 573). a) LSIMS, b) MI, c) MI/CAD, and (1) "true" CAD mass spectra for methionine enkephalin (MW 573). See text for description of spectra and peptide sequence ion nomenclature. a) LSIMS, b) MI, c) MI/CAD, and d) "true" CAD mass spectra for pentaglycine (MW 303). See text for description of spectra and peptide sequence ion nomenclature. Estimated PAs for various basic sites on methionine enkephalin (H- YGGFM-OH) are compared to PA of glycerol. Bar graphs showing the percent of skeletal bond cleavage products versus skeletal bond positions from a) LSIMS, b) MI, c) "true", single collision CAD mass spectra of the peptide pentaglycine (H -GGGGG-0H ). Structure and skeletal bonds are shown at the top. Bar graphs showing the percent of skeletal bond cleavage products versus skeletal bond positions from a) LSIMS, b) MI, c) "true", single collision CAD mass spectra of the peptide methionine enkephalin (H-YGGFM- 0H). Structure and skeletal bonds are shown at the top. Estimated reaction enthalpies for peptide skeletal bond cleavages in kcal/mol. Unimolecular decomposition pathways and their estimated reaction enthalpies (in kcal/mol) for excited peptide [M+H]+ ions. Bar graphs comparing normalized RIs of bu and y" peaks in the LSIMS and MI spectra of a) triglycine, b) tetraglycine, c) pentaglycine, and d) hexaglycine. Fragmentation barrier profiles based on estimated reaction enthalpies for an and b" ions from unimolecular decompositions of a) methionine Figure 55. Figure 56. Figure 57. Figure 58. Figure 59. Figure 60. Figure 61. Figure 62. Figure 63. Figure 64. XV enkephalin and b) pentaglycine protonated molecules. Bar graphs comparing normalized RIs of 1),, and yn peaks in the MI spectra of underivatized and acetylated forms of a) triglycine, b) tetraglycine, and c) pentaglycine. Fragmentation map showing the metastable decomposition pathways for [M+H]+ of methionine enkephalin obtained from linked scans at constant B/E and B2/E. The dotted arrows represent missing data, but the pathways are presumed to occur based on sequence ions in same series. a) Schematic diagram of the vacuum chamber, K+IDS power supply unit, and picoammeter used for testing M+ thermionic emission ion currents from aluminosilicate emitters doped with various metal substances, and b) a plot of In+ emission ion current and emitter temperature versus heating current for the In2032Ale3z2Si02 emitter. a) K+IDS and b) In+IDS mass spectra of acetone (MW 58), and c) Al+IDS mass spectrum of CCl4 (MW 152) leaked into HP ion source to 1x10'5 torr base pressure. a) K+IDS, b) Ga+IDS, and c) Al+IDS mass spectra of palmitic acid (MW 256), and d) Al+IDS mass spectrum of tridecanoic acid (MW 214) on the HP mass spectrometer. K+IDS mass spectra of a) glycerol (MW 92), b) triethanolamine (MW 149), c) thioglycerol (MW 108), d) nitrobenzyl alcohol (MW 153), and e) cholic acid (MW 408) mixed in glycerol. FAB mass spectra of a) Co(NO3)3, b) Cu(NO3)3, c) Ni(NO3)3, d) K+ glass emitter bead on wall-divided target in Figure 19c, e) Na+ glass emitter bead on "pop-up" divided target in Figure 19f, D NaflDS-by-FAB mass spectrum of glycerol, g) FAB mass spectrum of Li+ glass emitter bead on K+IDS-by-FAB space-divided target, h) LifiDS-by-FAB mass spectrum of glycerol. 3) CoflDS-by-FAB mass spectrum of glycerol and Co(NO3)2 solid separated on a wall-divided target in Figure 1%, b) CoflDS-by-FAB mass spectrum of H-LGG-OH (MW 245) in glycerol and Co(NO3)2 separated on a wall—divided target, c) FAB mass spectrum of digoxin (MW 780) and Co(NO3)2 mixed in glycerol, and d) [digoxin-H+Co]+ metastable decomposition mass spectrum obtained by a linked scan at constant B/E. The [M+K]+ adduct ions of H—VGVAPG-OH (MW 498) in glycerol formed by a) KfiDS-by-FAB and c) FAB of 0.1 M KI salts in glycerol were selected as the precursor ion to undergo metastable decomposition. In b), the structure of H -VGVAPG-0H is shown with the prominent fragmentation sites. The * represents a glycerol contamination peak. The [M+K]+ adduct ions of H-VGVAPG-OH (MW 498) in glycerol formed xvi by a) K+IDS-by-FAB and c) FAB of 0.1 M KI salts in glycerol were selected as the precursor ion for CAD at 50% beam attenuation. In b), the structure of H—VGVAPG-OH is shown with the prominent fragmentation sites. Figure 65. Metastable decomposition mass spectra of the a) [M+H]+, b) [M+Li]+, and c) [M+Na]+ precursor ions of the peptide H-VGVAPG-OH (MW 498) in glycerol, 0.1 M Lil in glycerol, and 0.1 M NaI in glycerol, respectively. A * is a glycerol contamination peak. Figure 66. CAD mass spectra of the a) [M+H]+, b) [M+Li]+, and c) [M+Na]+ precursor ions of the peptide H-VGVAPG-OH (MW 498) in glycerol, 0.1 M LiI in glycerol, and 0.1 M NaI in glycerol, respectively, at 50% beam attenuation. Figure 67. a) Metastable decomposition mass spectrum and c) single-collision, high- energy CAD mass spectrum of the [M+K]+ of methionine enkephalin (MW 573) formed by FAB of 0.1 M KI salts in glycerol. The methionine enkephalin structure and prominent fragmentation sites are shown in b). Both [M-i-K]+ precursor peaks are on scale. LIST OF ABBREVIATIONS ! l l v' . I l . 3N-6 ........................... Degrees of freedom in M (N is # of atoms) AID ........................... Analog-to-digital conversion ac ........................... Alternating current AE ........................... Appearance energy A ........................... Alanine APCI ........................... Atmospheric pressure CI API ........................... Atmospheric pressure ionization Art ........................... Argon ion B ........................... Magnetic sector, Magnetic field strength BAT ........................... Best anode temperature BET ........................... Best emitter temperature CAD ........................... Collisionally-activated dissociation CE ........................... Charge exchange / Capillary electrophoresis C ........................... Cysteine CF-FAB ........................... Continuous-flow FAB CI ........................... Chemical ionization Cs+ ........................... Cesium ion D ........................... Distance, Aspartic acid D/I ........................... Desorption/ionization Da ........................... Dalton dc ........................... Direct current DCI ........................... Direct CI / Desorption CI DEI ........................... Desorption EI DIP ........................... Direct insertion probe E ........................... Electric sector, Electric field strength E ........................... Internal energy, Glutamic acid e' ........................... Electron E0 ........................... Activation energy E0r ........................... Reverse activation energy Ecom ........................... Center-of—mass energy in CAD EE ........................... Even-electron species EI ........................... Electron ionization Eiab ........................... Translational energy of projectile ion ESI ........................... Electrospray ionization eV ........................... Electron volt E2t ........................... Energy of the kinetic shift F ........................... Phenylalanine F+ ........................... Fragment ion FAB ........................... Fast-atom bombardment FD ........................... Field desorption xvii xviii FI ........................... Field ionization G ........................... Glycerol molecule, Glycine GB ........................... Gas phase basicity GC/MS ........................... Gas chromatography/mass spectrometry Gp+ ........................... Glycerol fragment ion H ........................... Histidine H+ ........................... Proton He ........................... Helium I ........................... Isoleucine ICR ........................... Ion Cyclotron Resonance IE ........................... Ionization energy k(E) ........................... Unimolecular decomposition rate function ........................... Lysine K+ ........................... Potassium ion K+IDS ........................... ‘ Potassium ion ionization of desorbed species KE ........................... Kinetic energy keV ........................... Kiloelectron volts LC ........................... Liquid chromatography L ........................... Leucine LD ........................... Laser desorption Li+ ........................... Lithium ion LSIMS ........................... Liquid secondary-ion mass spectrometry M ........................... Analyte molecule, Methionine m ........................... Mass or Minute M+ ........................... Metal cation or Molecular cation M+‘ ........................... Molecular ion or Radical cation M+'* ........................... Excited molecular ion M' ........................... Radical anion m/z ........................... Mass-to-charge ratio MALDI ........................... Matrix-assisted laser desorption ionization MeV ........................... Megaelectron volts MS ........................... Mass spectrometry MS/MS ........................... Tandem mass spectrometry mt ........................... Mass of target gas MW ........................... Molecular weight or Megawatt MI ........................... M” critical configuration in transition state N ........................... Neutral molecule in fragmentation Na+ ........................... Sodium ion OE ........................... Old-electron species 0 ........................... Omithine Q ........................... Glutarnine P(E) ........................... Probability distribution function P ........................... Proline P+ ........................... Product ion formed in CAD PA ........................... Proton affinity PD ........................... Plasma desorption PE ........................... Potential energy PID ........................... Photo-induced dissociation QET ........................... Quasi-equilibrium theory R, r ........................... Radius of E and B, respectively; Arginine RA ........................... Relative abundance xix Rb+ ........................... Rubidium ion Re ........................... Rhenium RF ........................... Radio frequency RI ........................... Relative intensity RRKM ........................... Rice, Ramsperger, Kassel, Marcus theory 8 ........................... Second or degrees of freedom, 3N-6 S ........................... Serine SID ........................... Surface—induced dissociation SIMS ........................... Secondary-ion mass spectrometry t ........................... Time TIC ........................... Total ion current T ........................... Threonine TOF ........................... Time-of-flight UV ........................... Ultraviolet radiation V ........................... Volt, Accelerating voltage, Valine W ........................... Tryptophan v ........................... Velocity v ........................... Vibrational frequency MI ........................... Metastable ion MI/CAD ........................... CAD mass spectrum containing MI Xe+ ........................... Xenon ion Y ........................... Tyrosine ze ........................... Charge [G+H]+ ........................... Protonated glycerol [M+H]+ ........................... Protonated molecule (M) [M+H]+* ........................... Activated/excited protonated molecule [M+H]+* ........................... Transition state, Critical configuration [M+K]+ .................. _. ........ Potassium adduct ion of M [M-H]' ........................... Deprotonated molecule (M) AGrxn ........................... Gas-phase basicity AHf ........................... Heat of formation Dn ........................... Thermal degradation neutral fragment AH,“ ........................... Reaction enthalpy CHAPTER ONE. INTRODUCTION 1. Mass Spectrometry Mass spectrometry is an extremely powerful analytical technique that provides molecular weight (MW) and/or structural information for nanomole quantities or less of analyte [1]. A mass spectrometer is an analytical instrument, comprised of a sample inlet system, ion source, mass analyzer, ion detection system, and computer, connected as illustrated in Figure 1. The ion source, mass analyzer and ion detector regions are held under high vacuum (<10"6 torr) in order to prevent ion scattering and additional chemistry from ion/molecule collisions. An elaborate sample-inlet system uses a direct insertion probe (DIP) to transfer the analyte from 760 torr into the evacuated ion source. In the ion source, a user-selected ionization technique converts gas-phase analyte molecules (M) into ionized intact analyte molecules, representative of the MW. Positively-charged intact analyte molecules, M4”, [M+H]+, and [M~i-K]+ are formed by Computer Scan Signal Sample ‘ Ion Mass Ion Inlet ' Source ' Analyzer Detector Vacuum System l——— Figure 1. Schematic diagram of a mass spectrometer. the energetic removal of an electron (e') from, proton (H+) attachment to, or K+ attachment to an analyte molecule (M), respectively, and negatively-charged intact analyte ions, M' and [M-H]', are formed from the attachment of an e' or removal of a H+, respectively. The ionization process leaves some intact analyte ions with excess vibrational internal energy. For ions containing sufficient internal energy, subsequent unimolecular decomposition reactions will occur to form unique fragment ion and neutral molecule pairs, representative of analyte structure. The various types of fragment ions formed following ionization have unimolecular decomposition mechanisms with favorable thermodynamical properties such as activation energy, reaction enthalpy, reaction entropy, etc. However, the abundances of these fragment ions depend on the unimolecular decomposition kinetics for each mechanism, which depends on the internal energy content and structural features of the ionized intact analyte molecule such as charge-site location, gas-phase secondary structure, etc. All analyte ions formed in the ion source volume are extracted by the ion source optics system, accelerated into a well- defined beam, and directed towards the mass analyzer such that each ion has the same final kinetic energy (KB). The mass analyzer is designed to separate the ion beam that emerges from the ion source into groups of ions according to mass-to—charge (m/z) ratios. Under computer control, the mass analyzer is scanned so each group of ions with identical m/z values is sequentially separated from the ion beam and focused into the ion detector (e.g., electron multiplier, etc.) for generation of a signal proportional to the ion current. The computer acquires the ion abundance signal for each m/z value following A/D conversion of a voltage proportional to ion current and then generates a mass spectrum for the range of m/z values scanned. The mass spectrum is a plot of relative abundances for each ion, normalized to the most abundant ion, versus the respective m/z values. A certain peak at high mass represents the stable ionized intact analyte molecules which provides the analyte MW information. The unique "fingerprint" region in the mass spectrum, or the peaks representing the fragment ions from the ionized analyte molecule, provide the structural information. 2. Overview of Ionization Techniques in Mass Spectrometry Compatibility between the analyte of interest and the available ionization techniques on a mass spectrometer centers around how well the ionization technique interfaces with the physical properties of the analyte and whether the ionization technique can provide the desired MW and/or structural information. Today, a wide array of ionization techniques is available in response to the wide variety of organic, inorganic, biological or polymeric analytes that benefit from analysis by mass spectrometry. Analytes can be successfully analyzed whether they are nonpolar or highly polar, low or high MW, and gaseous, liquid or solid. A brief survey of the popular ionization techniques is presented below. Historically, W (El) [2] was the first successful routine ionization technique in mass spectrometry. Analytes analyzed by El usually show extensive fragmentation which is desirable for obtaining structural information. However, often times a peak representing the intact ionized analyte molecules, which provides the analyte MW information in the mass spectrum, is missing or has low abundance. This led researchers to develop "softer" (i.e., less fragmentation) alternative ionization methods, such as finlsLinnizntinn (F1) [3. 4]. nhntninnLinminn (CI) [5], and W (CE) [6] to complement EI and provide the missing MW information. All of these traditional ionization techniques require gaseous analytes or condensed-phase analyte molecules that are highly volatile at 10'6 torr so an ample number of analyte molecules reach the gas phase for ionization because the ionization technique does not assist in the phase transition. In these traditional ionization techniques, ionization of the analyte to form positive ions is accomplished by removal of an e' by a collision with a 70-eV e' in El, exothermic transfer of a H+ from a reagent ion in CI, exothermic transfer of an e' from a reagent ion in CB, and removal of an e' in a high electric field by tunneling in F1. MW (APCI) [7] is essentially CI of volatile compounds at 760 torr. Thus, volatility limits these traditional ionization techniques to the analysis of the less polar, lower MW analytes. The advent of W (D/I) [8] techniques, beginning with field desorption (FD) [9] in 1969, has advanced mass spectrometry from a routine, small molecule, organic analysis tool to a cutting-edge analytical technique for protein sequencing in biological research. D/I techniques provide MW and/or structural information for biomolecules and polymers that are fragile, high MW, highly functionalized, highly polar and thermally labile. In contrast to the traditional ionization techniques, D/I techniques provide the means for producing analyte-related ions directly from a condensed phase sample, which eliminates the need for derivatization methods. Historically, derivatization methods were used in E1 and CI to reduce the analyte polarity and increase volatility, but the intact derivatized analyte ions usually fragmented differently from the underivatized analyte ions making it difficult to study the underivatized analyte, specifically. Although the exact ion formation mechanisms are unknown, evidence and theory predict condensed-phase analyte molecules are first desorbed intact into the gas-phase and, subsequently, ionized in a manner specific to the DI] technique used. Alternatively, preionized analyte molecules may be directly desorbed into the gas phase eliminating the need for further ionization. mm (FD) [10, 11] was the first D/I technique amenable for the analysis of polar, high MW, thermally labile analytes. From the end of a DIP, intact ionized analyte molecules and a few fragment ions are generated from thermally heating a microneedle-containing pyrolytic carbon emitter, on which a solid or liquid analyte resides, in the presence of a very high electric field. FD is based on F1 principles, specifically the field image theory [12]. Nonpolar analyte molecules are thermally desorbed into the gas phase and, subsequently, near the microneedle tips, the high electric field removes an electron by tunneling from these thermally activated molecules to form a M“ ion. For polar molecules, the combination of thermal energy and high field strengths induce the desorption of preionized intact analyte molecules such as [M+H]+ or [M+Na]+ from the condensed phase [13, 14, 15]. Although FD works well to provide MW information exclusively, the technique is difficult and non-routine because the carbon emitter is fragile and consumable after a few experiments, exact sample placement is critical, the "best anode temperature" (BAT) needs to be optimized for maximum intact analyte desorption for each analyte, and the types of ions generated are transient [16]. Another early D/I technique, first introduced in 1973 by Baldwin and McLafferty [17], was 51mg (DCI). DCI involved placing a condensed-phase analyte on a target, attached to the end of a DIP, and inserting it into a heated EI/CI ion source volume near the electron beam. By increasing the ion source temperature at some rate, the analyte on the target, initially at room temperature, was rapidly heated which promoted thermal desorption of analyte molecules into the gas phase for ionization by the traditional techniques E1 or C1 [18]. Similar techniques, W (also DCI) [19] and W (DEI) [20], introduced by Hunt et al. and Holland et al. in 1977, respectively, involved rapidly heating an analyte on a pyrolytic carbon emitter to induce thermal desorption for gas-phase ionization by C1 or E1. Daves [21] later performed DCI by heating the analyte "dried" on a bare rhenium wire attached to a DIP, which is the preferred DCI method used today. DCI and DEI using either an emitter or a wire have problems similar to those in FD because the best emitter temperature (BET), similar to the BAT, needs to be optimized for each analyte, and the analyte ions generated are transient and last for only several seconds making accurate spectral acquisition difficult. In addition, polar, highly functionalized analytes are thermally labile which means the analyte is highly unstable during heating, and, as a result, neutral thermal degradation products are desorbed into the gas phase instead of intact analyte molecules. In the late 19708, particle bombardment D/I techniques, such as W (PD) [22], lnnerdnsntntinn (LD) [23], and "513119" 0r"mnlnnnl_ar"§9£9m1a§tifln_mm W (SIMS) [24], were actively pursued as alternatives to FD and the thermal desorption techniques since the initial analyses of fragile, highly polar, high MW, thermally labile biomolecules and polymers were successful and held much promise. Instead of thermally heating the solid analyte as required by FD and DCI for desorbing ions and/or molecules, the solid analyte was "dried" on a target, meaning solvents were evaporated, and the target, attached to a DIP, was bombarded with an energetic particle beam to induce desorption and ionization of the analyte. The respective bombardment particles for PD, LD and "static" SIMS are MeV-fission products of radioactive califomium—252, photons generated by a UV laser with power densities of l MW/cm2 or greater, and low beam densities (1 nA/cmz) of keV-ions such as Ar+ or Xe+. Unfortunately, the analyte ions were usually short-lived and transient similar to those from FD and DCI. Bombardment of the solid analyte sample is simply another method to energize the analyte molecules to effect ionization. In early 1981, the advent of the DI] technique, W (FAB) [25, 26], changed mass spectrometry forever because, for the first time, these fragile, high MW, highly polar, thermally labile biomolecule analytes could be analyzed routinely. In FAB, on the end of a DIP, a low density beam of keV-xenon atoms impinges on a target on which the analyte resides similar to that in "static" SIMS. However, FAB differed from SIMS, PD, and LD because the analyte is dissolved in a polar, viscous matrix, typically glycerol (MW 92), instead of being presented as a "dried" layer of the analyte on the target. The matrix was the key to success in FAB because it had a relatively low vapor pressure, and thus persisted as a solvent for a solution of analyte under vacuum. This provided strong, long-lived, stable ion currents without degradation as opposed to the transient signals previously obtained. However, there are drawbacks to using a matrix because it may interfere with the analyte spectrally if a matrix peak overlaps with an analyte peak and/or chemically if the matrix undergoes condensed-phase reactions with the analyte during bombardment. Two years later, W (LSIMS) [27] was introduced which combined the "static" SIMS technique with the liquid matrix identical to that in FAB. LSIMS is generally more sensitive than FAB, having detection limits extending down into the low picomole region, since the keV-Cs+ beam can be highly focused at high energies which results in higher ion yields. PD has also enjoyed renewed success since recently the analyte is dissolved in a solution of reduced glutathion (1:1 mole ratio) and electrosprayed on a nitrocellulose substrate (i.e., the matrix) target instead of onto a bare thin metal target as before [28]. WW (MALDI) combines LD with a solid matrix [29]. Here, the solid analyte/matrix is routinely prepared by dissolving picomole amounts of analyte in a solution containing excess low MW, highly UV-absorbing organic matrix molecules, and subsequently drying a portion on the DIP target. With the addition of the matrix, LSIMS, PD and MALDI are capable of desorbing and ionizing analytes having MW 5 of 10, 35, and 300 kDa, respectively. Most recently, the ionization techniques, continuous-flow FAB (CF-FAB) [30], and W (ESI) [31], have enabled the direct introduction of trace amounts of analytes in their native aqueous media by flow injection, or from a liquid chromatography (LC) or capillary electrophoresis (CE) experiment for mass spectral analysis. In CF-FAB, the aqueous solution containing the analyte is pumped through a capillary in the DIP to the FAB target surface which is simultaneously bombarded with Xe atoms. This analyte-aqueous solution contains only 5% glycerol and is continuously regenerated on the target surface, resulting in both higher signal-to-background noise and lower detection limits. In ESI, small highly charged droplets of the analyte solution are sprayed into a high electric field and a high pressure N2 bath gas similar to those used in FD and APCI, respectively. As the droplets evaporate, their surface charge density increases until the Rayleigh limit is reached, and, at this point, a coulomb explosion occurs to produce an array of smaller charged dr0plets. Eventually, when droplets are fully evaporated, a distribution of multiply protonated analyte molecules emerges. Mass analysis of the array of differently charged analytes provides a database from which an average analyte MW is calculated. ESI allows detection of analytes with MW 5 up to 100 kDa. 3. The Focus, Goals, and Organization of the Dissertation The power of mass spectrometry for analytical applications is the ability to identify the molecular weight and structure of the analyte under consideration through interpretation of the respective mass spectrum. The ability of the analyst to interpret a mass spectrum with confidence requires complete knowledge of the ion formation and fragmentation mechanisms specific to the analyte under consideration. General interpretation rules and fragmentation mechanisms for mass spectrometry were successfully developed by analyzing small volatile molecules, containing no more than two functional groups, with E1 and CI. This feat was possible because the ion formation mechanisms for both EI and CI were well understood and these small molecules gave very straightforward results. However, using the DH techniques, FAB and LSIMS, to elucidate structures of polar, highly functionalized, higher MW analytes are not always straightforward because the fragmentation mechanisms are not known with the same degree of certainty as in E1 and CI. Unfortunately, in FAB and LSIMS, the ion formation mechanisms are highly complex and not well understood, and the analytes commonly analyzed, usually biomolecules, are structurally much more complex than the small. molecules analyzed in El and CI because they have many ionizable functional groups and three dimensional structures that are capable of folding which makes the fragmentation mechanisms more complex to elucidate. FAB provides an abundance of very useful structural information for biomolecules which is not readily obtainable by other analytical or mass spectrometric techniques. Often times, the m/z and the relative intensity information of most peaks in a typical mass spectrum cannot be fully utilized because the fragmentation chemistry is not well understood. This dissertation will focus on the fundamental aspects of the DH techniques, FAB and LSIMS, with the goal of trying to better understand the ion formation and dissociation mechanisms for typical biomolecules in an attempt to further aid the interpretation process. In the traditional ionization techniques, EI and C1, the mechanisms by which gas- phase analyte molecules are ionized are based on well-known gas-phase chemistry principles such as ionization energies (IE) and proton affinities (PA), respectively. The formation of a particular fragment ion from the ionized analyte molecule has a high probability of occurring if the overall reaction enthalpy (Aern) for the reaction mechanism requires little energy. More importantly, whether this particular fragment ion is formed to a great extent depends on whether the average amount of internal energy in the ionized analyte molecules is sufficiently high to initiate the unimolecular decomposition processes. However, if the rate of another fragmentation reaction mechanism is more competitive in the same range of internal energies and/or if there is steric hindrance or interactions between two distant groups in the rate determining step for rearrangements, then a thermodynamically favorable reaction mechanism may not occur to a great extent. Most EI and CI fragmentation mechanisms have been elucidated previously because the e' beam in E1 or the CI reagent ion deposits a controllable, average amount of internal energy into the gas-phase analyte so the appearance and disappearance of fragment ions can be controlled by adjusting the internal energies. As a result, the changes in relative abundances of the fragment ions show the kinetic relationships between the ionized analyte molecule, the first generation of fragment ions, the second generation, and so on. Thermochemical data such as IE3 and appearance energies of fragment ions can be obtained from these studies for estimating the AHrxn of 10 fragmentation mechanisms. Fragmentation mechanisms may involve rearrangements or simple bond cleavages, and isotopic labeling can be used to differentiate between these two processes. Isotopic labels placed in strategic locations on the analyte molecules may or may not be shifted depending on the exact mechanism as indicated by unique shifts in fragment ion m/z values. The standard set of fragmentation mechanisms for El and C1 are summarized in references 2 and 5, respectively. Thus, in El and CI, the fragmentation mechanisms are well understood for most small analytes, and, as a result, unknown analyte structures may be elucidated because these standard mechanisms represent the links between the fragment ions in the unknown mass spectrum and the unknown analyte structure. In contrast to the development of El and C1, the ability to perform the particle bombardment experiments, FAB and LSIMS, has preceded our understanding of the underlying mechanism(s) by which the condensed-phase analyte molecules are transformed into ionized, gas-phase analyte molecules and fragment ions. Currently, there are many excellent theoretical studies and/or indirect studies which rely on the condensed- and gas-phase chemistry of a select group of analytes to reveal the mechanism(s) responsible for the production of ionized analyte molecules in FAB, LSIMS and other D/I techniques. While much of what is known to date about the ion formation mechanisms in FAB and LSIMS was obtained from these indirect studies, the fact is, different analytes have different condensed- and gas-phase properties depending on their structural features which ultimately determines the manner by which they are ionized in FAB and LSIMS. As a result, there is an over-abundance of mechanisms and multiple interpretations which makes the fundamental mechanism(s) that much more difficult to identify. A better approach would be to conduct direct experimental studies to probe for the ion formation mechanism(s). However, the particle beam energy is not directly related to the ionization of the condensed-phase analyte molecules as in E1 or C1, and these direct experiments need to be performed in or near the micro-volume (i.e., 11 where gas-phase chemistry is believed to occur) above the sample probe tip, under high vacuum, and in a high-voltage ion source. Even though direct experimentation is extremely difficult, it has the potential to yield a plethora of new mechanistic information. The elucidation of fragmentation mechanisms is significantly more complicated in FAB than in E1 because the analytes are higher MW and more highly functionalized, and the ion formation and dissociation mechanisms are not well understood. For biomolecules, there are more polar sites on the molecule that have the potential to be ionized and each of these ionized sites may unimolecularly decompose with competitive rates to give rise to an array of fragment ions. In contrast, in E1, only one or two sites are ionized. In many cases, the exact sites of protonation in FAB are difficult to locate in biopolymers which makes exact mechanisms more difficult to determine. Another inability to elucidate fragmentation mechanisms with certainty, similar to that for E1, is not knowing the average internal energy imparted into the ionized intact analyte molecules by the FAB ion formation mechanism. This internal energy is not directly related to the energy of the keV-particle beam as in El. In addition, some fragment ions in the mass spectrum may result from the specific conformations of the ionized analyte molecules in the gas phase. Determining the ionized analyte molecular structure/confirmation in the gas phase is beyond our current analytical technologies and is a question mass spectrometry will answer someday. While FAB expertise does exist in some laboratories, a complete understanding of all the underlying mechanisms is still some years away even with the research effort currently underway. Without understanding the actual ion formation and dissociation mechanisms in a FAB D/I experiment which link the mass spectrum to the condensed-phase analyte, it is unlikely that ( 1) FAB and other D/I techniques will be advanced to assume new levels of performance, (2) applications to new chemical systems or studies will occur and, (3) new ionization techniques will evolve from these existing techniques in a timely manner. 12 Unfortunately, many D/I parameters are interlinked and conclusions are usually limited. Chapter two will discuss FAB and LSIMS; other research related ionization techniques such as EI, CI, and K+IDS; unimolecular decomposition theory; fundamental mass spectrometry principles and MS/MS instrumental scanning techniques and CAD in an attempt to bring the reader up-to-speed before discussing specific, novel research. Both Chapter three and the published paper in Appendix one identify and evaluate the D/I mechanisms operative in FAB using a novel, direct experimental method referred to as KflDS-by-FAB in an attempt to better understand the ion formation mechanism in FAB. Chapter four evaluates or elucidates the mechanisms for the formation of peptide fragment ions under FAB and LSIMS using either available literature information, estimated reaction enthalpies and other novel experimental methods in an attempt to better understand the fragmentation mechanisms in FAB and LSIMS. Chapter five describes some preliminary work that explores the metal ion/biomolecule reactions in the gas phase with the ionization techniques, M+IDS and M+IDS-by-FAB. CHAPTER TWO. THEORY AND METHODS I. The Ion Formation and Dissociation Processes in the Traditional Ionization Techniques Electron Ionization (E1) and Chemical Ionization (CI) The theories and mechanisms surrounding ion formation and ion dissociation are well-characterized experimentally for the ionization techniques, EI and CI, as will be described below. This knowledge base for E1 and CI may serve as a starting point for understanding the theories and mechanisms for ion formation and ion dissociation in FAB and the other related D/I techniques. ELECTRON IONIZA TION (El) Historically, E1 was first introduced by Dempster in 1921 [32] and it evolved into the first routine ionization technique in mass spectrometry. Improvements were made to Dempster's design in 1929 by Bleakly [33] who collimated the e' beam with a parallel external magnetic field, and in 1940 by Nier [34, 35] who placed a field—free region between the El ion source and the mass analyzer to eliminate any influences from the magnetic and electric fields on production and acceleration of ions. Before the 19505, EI mass spectrometry (El-MS) was primarily used for accurately determining the elemental isotopic masses and abundances [36]. Then, beginning in the 19505, the power of EI-MS for the analysis of small organic molecules was first realized [37] because, in addition to. molecular weight information, researchers noted that EI-MS also produced an unique "fingerprint" of abundant fragment ions which provided specific structural information for each analyte structure. Over the years, most fragmentation mechanisms have been elucidated for EI-MS so unknown ionized analyte structures can be derived from the 13 14 various types of fragment ions present in the unknown mass spectrum. Presently, E1 is a routine, well characterized ionization technique that gives reproducible, well understood spectra, and is primarily utilized as the ionization technique in the mass spectrometer detector of a gas chromatograph (GC/MS). EI Instrumentation. A schematic for an El ion source is shown in Figure 2. In El, a monoenergetic beam of electrons (e') is the ionizing reagent for analyte molecules. Gas-phase electrons are generated by thermionic emission from an incandescent rhenium filament. The gas- phase electrons are accelerated into a well-defined beam with a final energy of 70 eV since the potential difference between the rhenium filament held at 9930 V, for example, and ion source volume held at 10000 V is 70 V. The beam of electrons traveling across the ion source is well collimated owing to an external magnetic field (not shown in Figure 2) parallel to the electron beam which forces the electrons to travel in small helices about the field. The electron trap is held at a higher potential than the ion source volume in order to collect the remaining electron beam so electronic feedback can be provided to the filament power supply to ensure that the electron emission current is kept constant. The ’ \ 2'! £5. E’, Repeller 1 G'OOOGOOOOOO'GO + I H O D 3 fl. 0 VJ - 'N1. + ZZ 3 lo .31 ZZz Z ZZZ Z e' Ion Source Y Volume Electron Trap Figure 2. Schematic diagram of an El ion source. 15 repeller in the rear of the ion source "pushes" any newly formed ions in the ion volume towards the front of the ion source for extraction and acceleration by the ion optics. These ions are subsequently analyzed by the mass analyzer. Analyte Volatility in El. In El, either a gaseous analyte is leaked into the ion source volume or condensed- phase analyte molecules are thermally desorbed by heating a sample holder on the end of a direct insertion probe. Thus, volatility limits the analytes amenable for analysis by E1 to low molecular weight and low polarity (i.e., frequently less than two polar functional groups). For a successful analysis by El, the analyte pressure needs to be at least lxlO‘5 torr. If it is greater, then ion/molecule chemistry readily occurs. Unfortunately, it is well known that less than 1% of the total analyte is ionized while the remaining 99%+ is pumped away by the vacuum system. The current used to heat the rhenium filament and ion source heaters (not shown in Figure 2) raise the ion source temperature to approximately 250° C which prevents the analyte molecules from condensing on the ion source walls. Formation of Ions in E1. Ionization of a gas-phase analyte molecule (M) can occur if a 70-eV 6 passes sufficiently close to M and upsets its electronic equilibrium. The activated M* may either eject an e‘ to rid energy equal to its ionization energy (IE) or remain in an excited neutral state. As a result, a positively-charged analyte radical cation, M+', also known as the molecular ion, is formed initially by the ionization process as illustrated in Equation 1 M (g) + e' (70 eV) —-) M'” -i- 2 e' (1) and Figure 2. For electron ionization to occur, the energy of the colliding electron must be greater than the IE of the gas-phase M. The amount of energy imparted into M is dependent on the closeness of the interaction between 6 and M and the structural features 16 of M. If the M” has low internal energy after ionization, it will probably be detected intact and molecular weight information will be obtained. If M“ contains excess internal energy after ionization, it has a high probability of undergoing unimolecular ion decomposition processes [38] to form a fragment ion and neutral radical pair as shown in Equation 2 and Figure 2. The fragment ion, Fn+, can undergo further unimolecular M+‘* —> E,“ + Nno (2) decomposition to form a secondary fragment ion if ample internal energy remains. Since these fragmentation mechanisms in El are well characterized, many Fn+ reveal structural features of M‘”. Utility of Clastograrns in E1. The clastogram (or breakdown curve) in Figure 3 is a plot of the relative abundances of M‘” and the Fn+ ions for methanol versus energy of the e' beam. At 70 1007 r: 90f [:1 30% '3 70.. U 5 .5 60., fl 3 50.. <1 .5. 40“ E 30-- % 20L a: lO-- 0 j 0 Electmn Beam Energy (eV) Figure 3. An EI clastogram for methanol (Adapted from [39]). 17 eV, the rates of unimolecular decomposition are independent of e' energy because only a small fraction of the 70 eV can be deposited into M. Since the relative abundances (RA) of ions formed in E1 are highly reproducible and uniform at 70 eV, EI mass spectra are obtained at this e‘ energy. The efficiency of the e' collision and the structural features of M determine how much energy is absorbed, which is summarized by the probability distribution function, P(E). For a 10-kV ion source, if the voltage potential on the filament is increased from 9930 V to 9980 V, then the e' beam energy will be decreased from 70 eV to 20 eV. At 20 eV, the rates of ion formation are highly dependent on the energy of e‘ beam, which is not the case at 70 eV. The IE of methanol is approximately 11 eV so the CH3OH+', or M‘”, curve starts at 11 eV and 100% RA in the clastogram. Therefore, at low e' energies near the IE, M“ will be the most intense ion in the mass spectrum which is useful for molecular ion identification purposes. As the e' energy is increased, the RA of M‘” decreases while the RA of the CH20H+ fragment ion increases because, at higher e' energies, more M“ ions have more excess internal energy after ionization which results in a higher probability of unimolecular decomposition occurring. The shapes of the curves in Figure 3 indicate that the formation of CO+ and CHO+ fragment ions are secondary processes which originate from CH20H+ precursor. Thus, a clastogram is useful for fragmentation studies in B] because it "maps" out the kinetics of the fragmentation events, contained in the RAs, over an entire range of average internal energies imparted into M+' by the e' beam energy. This aids in identifying the precursor ions of each fragment ion. ii" A clastogram is also useful for identifying the thermodynamics and possibly the mechanism of each fragmentation process. In Figure 3, the appearance energy (AE) of CH2=OH+ fragment ion is approximately 12 eV. The AB is equal to the sum of the IE of CH3OH and the activation energy, E0, of the fragmentation pathway. At the e“ beam energy of 12 eV in Figure 3, a certain population of M” ions have approximately 1 eV of excess internal energy since ionization of CH30H requires 11 eV. As a plausible 18 mechanism, CH3OH+' may decomposes into CH2=OH+ if a hydrogen loss occurs by a radical-initiated alpha-cleavage as shown in Equation 3. The estimated reaction enthalpy I[I/\ [1+ 18 kcal/mol + CH2—OH -——> CH2=OH + H- 3 202 kcal/mol 168 kcal/mol 52 kcal/mol ( ) (AHrm) for the unimolecular decomposition mechanism of CH3OH+' shown in Equation 3 is 18 kcal/mol or 0.8 eV, which corresponds well to the experimental AB in Figure 3. The heats of formation (AHf) in Equation 3 for small ions and neutral radicals are available and listed in reference 40. Ionization Theory for M in E]. A physical description of the ionization and fragmentation processes in E1 is provided by the quasi-equilibrium theory (QET), developed by Rosenstock et al. [41] in 1952. In El, if M* chooses, an e' is removed in approximately 10'16 s from the excited M in its ground-state configuration to form M+'. If internal energy remains in M'” after losing e‘, M""* subsequently undergoes a vertical Franck-Condon transition to an excited state, perhaps with electronic internal energy exceeding 10 eV. The electronic transition is vertical because no change in the ground state molecular geometry occurs during the excitation process. In an attempt to restore electronic equilibrium in M‘“, this electronic internal energy in M+'* undergoes fast radiationless transitions between closely situated electronic potential energy surfaces or vibrational levels and eventually becomes randomized intramolecularly into vibrational energy among the active 3N-6 degrees of freedom (N is the number of atoms in M, and the active degrees of freedom are 1/2 to 1/3' of the 3N-6 modes which contain randomized energy). This randomization of internal energy in M“ occurs for a short period of time before unimolecular ion decomposition processes begin because energy redistribution is faster than fragmentation. In the ion 19 source volume, the fastest unimolecular decomposition processes, the simple bond cleavages, begin after 10'13 s and the slowest processes, the cleavages involving a rearrangement process, end around 10'6 s. The low internal energy M+' ions not decomposed in 10‘6 s are extracted and accelerated. After acceleration, some low energy M‘” ions may undergo metastable unimolecular decomposition [42] during their flight to the detector, but the lowest internal energy M+° ions will be detected intact. Unimolecular Ion Decomposition Theory. Fragment ions are formed by unimolecular decomposition kinetics because it is highly improbable that M“ will undergo a bimolecular collision at 1x10'6 torr. The QET and RRKM [43, 44] theories are utilized to describe the rate of unimolecular ion decomposition of the excited M+‘* into fragment ions. In Figure 4, if the internal energy, E, remaining after ionization of M, exceeds the activation energy, E0, for a certain if“ ‘ Kinetic Shift, Bi _ .1211. _ AE E l r 130 F,{*‘ E V 7 -" — " g 150 id 3 Aern ‘5 3 e W“ __y 1 IE (M) * . M —> M+' —— M (Ground state molecule) Reaction Coordinate Figure 4. Thermochemical considerations for unimolecular decomposition of M+'*. 20 reaction pathway, then there is a high probability that M+°* will move through the transition state or critical configuration, Mai, to form the fragment ion, Fn+, and neutral radical, NO, as shown in Equation 4. The critical configuration, Mni, represents the point M‘m‘ —-) Mni —) Fn+ + My (4) in time when enough vibrational energy accumulates in a specific bond or group of bonds to induce either simple bond cleavage(s) or intramolecular secondary interactions and subsequent bond cleavages in M+'. The rate of unimolecular decomposition for M+'* increases as the internal energy, E, exceeds the activation energy, E0, and this difference is defined as the kinetic shift, E1. Thus, the decomposition rate is a function of E and it can be approximated by the expression in Equation 5 k(E) = I) ((E - 150)/E)~°"l (5) where u is the vibrational frequency factor and s is the number of oscillators or active vibrational degrees of freedom. From Equation 5, as E approaches E0, the rate slows to zero while as E approaches infinity, the rate advances to a maximum rate of u. The vibrational frequency factor of the bond is the maximum rate for bond cleavage. Since ‘0 for a simple cleavage, 10'14 s", is greater than t) for a rearrangement, 10-10 5'1, simple cleavages will always dominate over a rearrangement if both pathways have identical activation energies. If the number of oscillators, s, is increased, the internal energy will likely be distributed over more bonds, which will result in slower decomposition rates since the probability of ample vibrational energy localizing at a particular bond(s) is reduced. It is important to realize that the fragmentation pathway with the lowest reaction enthalpy (Aern) does not necessarily produce the most abundant ions. The kinetics often dictate which processes dominate. McLafferty [2] describes the fragment ions that. are formed in El as being either simple cleavage mechanisms as shown in Equation 6 or rearrangement mechanisms in Equation 7. The simple cleavage mechanisms have a "loose" transition state in Equation 6, because the ionized sigma bond undergoes a sigma- 21 47 kcal/mol (26V) CH3CH2‘1'CH2CH20CH3 —> CH3CH2+ + °CH2CH20CH3 (6) 158 kcal/mol 215 kcal/mol -10 kcal/mol H / (Lt/CH3 15 kcal/mol Hcha I)” (0.7 eV) |\,« —> 0CH2CH2CH2CH2+ + CH3OH (7) H2O CH2 C 2 158 kcal/mol 221kcal/mol -48 kcal/mol bond dissociation without regard to critical steric configuration when enough vibrational energy critically displaces the bond length and angle. However, rearrangement reactions in Equation 7 have a "tight" transition state because a critical steric configuration in the transition state is required to bring H and +'OCH3 together intramolecularly to form the H-OCH3 bond in step (a). The second step (b) of the reaction involves a simple inductive cleavage to free CH3OH, and possibly a third step (c) would involve formation of a ionic four-membered ring (not shown). The overall reaction enthalpy is low for a rearrangement reaction because, for example, in Equation 7, the energy required for the cleavage of the C-0 bond is recovered when a new H-O bond(s) is formed. Ironically, in Equation 7, the reaction enthalpy for forming the ionic four-membered fragment is 15 kcal/mol higher than for the linear fragment probably because the cyclic fragment is highly strained and not as stable. For simple cleavage reactions, the reaction enthalpy is usually high because one bond is cleaved and none are formed. The rates of rearrangement reactions are slower than those for simple cleavage reactions because the rate-dependent step is the through space interaction of H and +"OCH3 where simple cleavages just require energy deposition into the bond. Therefore, if a simple cleavage and a rearrangement reaction had identical activation energies, the rate of the simple cleavage would be faster because the vibrational frequency of a single bond is higher than an intramolecular interaction. Thus, simple cleavages are dependent on the average 22 internal energy in M“ while rearrangement reactions are very dependent on the secondary interaction rate and the steric structure of M” in the transition state, provided the thermodynamics are reasonable for both. In Figure 5, an estimated Wahrhaftig diagram for ionized CH3CH2CH2CH20CH3 shows the relationship between the probability distribution function, P(E), in the top plot, and the kinetic rate constants, k(E), in the bottom plot. The P(E) describes the probabilities of M” ions having particular internal energies, B, after ionization and/or activation and the k(E) describes the decomposition rates of M“ as a function of E and structure for a particular fragmentation pathway. Both the P(E) and k(E) are plotted against E, the internal energy in the analyte after ionization with 70 eV electrons. The P(E) was approximated from distribution of abundances for the ionic species M'”, metastable fragment ions, and the fragment ions formed from unimolecular Metastable M4" ions P(E) M-P 0 C2) @ G) @ 13~ I I I I 97 12~ l I | l 11- l I I O €13: : j @l / | M“ = CH3CH2CH2CH26CH3 “ a ll 1 .. + 7" I I I i®®M —> °CH2CH2CH2CH2 + CH3OH + 6""'—'9 I : : @M‘” —> CH2=OCH2 + CH3CH2CH2° :jb- : : l : @M'” —>CH3CH2+ + °CH2CH20CH3 3 I I II I I I I I 7 I o 1 2 3 4 Internal Energy, E, after Ionization (eV) Figure 5. Estimated Wahrhaftig diagram for ionized CH3CH2CH2CH20CH3. 23 decomposition of M“ in the ion source, and it is usually based on the initial abundances of analyte species in case secondary processes would be involved. All ions formed in the ion source are extracted and accelerated to the mass analyzer within 10‘6 s, regardless. Thus, for fragmentation to occur in the ion source, M” must have enough excess internal energy to initiate a fragmentation process having a rate greater than 106 8'1. For this specific Wahrhaftig diagram in Figure 5, the M‘" ions with an internal energy, E, less than 0.7 eV are extracted from the ion source and detected intact without fragmenting because all the fragmentation reactions require more internal energy to occur at an appreciable rate. In region Q), the P(E) indicates that not many M‘” ions acquire an energy less than 0.7 eV because the M” ion abundance is low. In region @, an M+' that acquires between 0.7 eV and 0.9 eV will undergo metastable decomposition after acceleration because the k(E) values for forming metastable fragment ions are less than 106 8’1 but greater than 104-5 8'1, resulting in fragmentation rates too slow to occur in the ion source. The rearrangement fragmentation pathway, M+’—-> ~CH2CH2CH2CH2+, in Equation 7, has an low E0 of 0.7 eV (the E05 are the E values at the start of each decomposition rate curve on the x axis in Figure 5) in the metastable region, but it has an estimated AE of 0.9 eV [45] for ion source fragmentation (the AEs are represented by the dots on each decomposition rate curve in Figure 5). If M” ions have between 0.9 and 1.7 eV of B, then the M“—-) °CH2CH2CH2CH2+ reaction will kinetically dominate in region Q) and undergo ion source decomposition. If M“ has an E between 0.7 and 0.9 eV, M“ will undergo the same fragmentation reaction, but as a metastable ion at a slower rate in region ® outside the ion source after acceleration. The simple cleavage process, M+‘-)CH3CH2+, in Equation 6, has an E0 of 2 eV and an estimated AE of 3 eV [45] for competitive ion source decomposition. If M“ ions have 3 eV of E and up, M” will unimolecularly decompose at a high rate in the ion source to form CH3CH2+ fragment ions which is dominant fragmentation process in region 6). If M“ had an E of 2 eV, it would likely unimolecularly decompose to form CH2=OCH3+ fragment ions as shown in 24 region @ since it is the most kinetically favored pathway in this region. A fragmentation pathway with a low E0 will cause the M” abundance to be low or missing as will a Fn+ ion with secondary fragmentation process having a low E0; both of which are not the cases here in Figure 5. Thus, all fragmentation pathways here originate from an ionized site on CH3CH2CH2CH20CH 3 and are kinetically competitive as shown in the Wahrhaftig diagram depending on E. If multiple sites on M“ are ionized, fragmentation becomes kinetically competitive between ionization sites in addition to the competition between local decomposition reactions at each ionized site. The M“ ion lifetimes are dependent on E and equivalent to the inverse of k(E) values for a particular value of E. For ion source decomposition, the simple cleavage reaction dominates from 10'13 s to 10' 3-5 3 while the rearrangement reaction are slower reactions that dominate from 103-5 s to 10'6 5. Outside the ion source, metastable M“ ions decompose between 10‘6 and 104-5 5 while intact M4” ions are stable for at least 104-5 s or longer. Thus, M+‘ ions that undergo metastable transitions usually have a narrow distribution of low internal energies and follow slow rearrangement reactions having low Eos. However, if both a simple cleavage and rearrangement reaction have identical Eos in the metastable regime, then simple cleavage will dominate over the rearrangement processes in the metastable transition, as explained above. The above explanations give a simple, brief overview of the physical processes occurring during ionization and unimolecular decomposition in E1, but the events that do occur are more complicated than portrayed above and important questions still remain as explained by Lorquet [46]. The unimolecular ion decomposition theory explained above can be utilized also to account for the competitive fragmentation processes following ionization of analytes in other ionization techniques. 25 CHEMICAL IONIZA TION (CI) Ion/molecule reaction products were often observed and considered a nuisance in the early days of mass spectrometry because the vacuum systems were primitive and the pressures were usually above 1x10“4 torr. Chemical ionization (CI) is an outgrowth of fundamental gas-phase ion/molecule studies of protonated methane, CH5+, during the late 19505 and early 19605. In 1966, Munson and Field [47] introduced CI as a "soft" ionization technique for mass spectrometry which means CI provided abundant molecular weight and limited structural information for most analytes in contrast to El; that is, there is little fragmentation in CI as opposed to E1. CI has become popular because it is routinely used to complement the analyses in which EI fails to provide molecular weight information for certain analytes. CI is presented here because the ion formation mechanisms for D/I techniques indicate that chemical ionization processes produce the majority of the observed protonated analytes. CI Instrumentation. CI is performed in a heated, EI ion source identical to Figure 2 except the ion volume is filled with a reagent gas at medium pressure (i.e., between 0.2 and 2 torr) and the large slits for the electron beam and the ion exit slit are greatly reduced in size to prevent the reagent gas from leaking into the surrounding ion source housing chamber held at low pressure. Differential pumping is needed in CI to keep the mass analyzer pressure low at all times and to keep the ion source housing chamber as low as possible to prevent additional ion/molecule chemistry outside of the ion source volume. Also, in CI, the electron beam is accelerated to energies greater than 200 eV and directed into the reagent gas, held at a medium pressure, in the ion volume as in E1. However, even at higher 6 energies, the medium pressure still prevents the electrons from reaching the electron trap so electronic feedback to regulate the filament current is obtained from the 26 electrons that collide with the ion source volume wall upon entrance. Another reason the openings in the ion source volume are reduced in size is to give the newly formed reagent ions longer residence times in the ion source volume to ensure more complete analyte protonation before their extraction. [Mi-HP Formation in CI. If the reagent gas, methane, is present in excess at medium pressures, when a 200- eV electron collides with a methane gas molecule, as shown in Equation 8, there is a high CH4 (3) + e- (200 eV) —) CH4“ + 2 6' (8) CH4” + CH4 (g) —> CH5+ + CH30 (9) CH5+ + M (g) —> [M+H]+ + CH4 (g) (10) probability that the radical cation, CH4“, will be formed by the E1 process. The CI methane reagent ion, CH5+, is formed by an ion/molecule proton transfer reaction from a bimolecular collision between CH4“ and another gas-phase methane molecule, present in excess as shown in Equation 9 above. The analyte, M, either needs to be volatile at medium pressures or thermally desorbed by DCI techniques because CI, like El, requires M to be in the gas-phase for protonation. If possible, the analyte pressure in the ion volume should be at least lxlO‘4 torr. If the proton affinity of M is greater than CH4, then an exothermic proton transfer from CH5+ to M will occur to form the protonated analyte molecule, [M+H]+, when the two species undergo a bimolecular collision as in Equation 10 above. The proton affinity (PA) is defined as the exothermicity (i.e, PA = -A Hun) of the proton attachment to a particular analyte, M, as shown in Equation 11. In M (g) + H+ —> [M+H]+ PA = -A ern (11) Equation 10, for the collision of protonated methane and M, the reaction enthalpy, A Hm], of the proton transfer is governed by the difference in PAs of the two species (i.e., A ern = PA(CH4)-PA(M)). It is important to realize that under the medium pressure conditions described above, the reactions in Equations 8-10 are occurring simultaneously 27 and they represent the primary pathway for the formation of [M+H]+, although secondary pathways exist. Finally, good yields of [M+H]+ ions are dependent on fast reaction rates which are directly proportional to high collision rates between the reagent ions and M brought about by the medium pressure CI environment. A table of PAs and IE5 for small molecules with various functional groups is presented in Table 1. Fragmentation of [M+H 1+ in CI. The extent of fragmentation in CI depends most on the amount of internal energy in [M+H]+, induced by protonation, and also the types of fragmentation pathways available surrounding the protonation site. Protonation of a gas-phase M usually occurs at sites which have the highest PAs, and the highest PAs indicate the most basic sites in M. For biomolecules, a basic site may either be a single functional group containing a heteroatom(s) with non-bonding electrons or a collection of two or more heteroatoms from different functional groups. Protonation may occur at a single heteroatom basic site or the proton may coordinate with two or more heteroatom sites which increases the stability of the [M+H]+. However, a proton that dicoordinates with two heteroatoms sites essentially undergoes cyclization which reduces the basicity of the site (i.e., Aern) because a negative change in entropy occurs. The exothermicity of the proton transfer reaction determines the exact amount of internal energy imparted in the [M+H]+. If the internal energy is localized in M-H+ bond and/or surrounding bonds, the fragmentation may be localized at the protonation site. If the PAs of the two reactants are similar, then the internal energy may be less than 1 eV (23 kcal/mol) which results in abundant [M+H]+ ions for molecular weight information and some low-energy fragmentation such as neutral losses. In CI, most often, stable even-electron (EE) fragment ions are formed from the low-energy elimination of stable EE molecules from the EB protonated molecule although exceptions do exist since odd-electron (OE) species are sometimes formed depending on the analyte. If structural and molecular weight information is desired, then 28 Table 1. PAs and IE5 for small molecules with unique functional groups. All values were obtained from Lias et al. [40] except when noted. Functional Group Compound PA (kcal/mol) IE (eV) fiydrogen H2 151.2 15.4259 alkane CH4 131.6 10.85 alkane C3Hg 150.0 10.95 branched alkane CH3CH(CH3)CH3 163.3 10.57 water H20 166.5 12.612 alkyl chloride CH3CH2C1 169.0 10.97 alkyl bromide CH3CH2Br ~ 171.0 10.28 alkyl iodide CH3CH21 176.0 9.346 alkene CH3CH=CH2 179.5 9.73 benzene C6H6 181.3 9.2459 aldehyde CH3CH2C(O)H 189.6 9.953 toluene C6H5CH3 189.8 8.82 l°-alcohol C3H7OH 190.8 10.22 thiol CH3CH2CHZSH 191.6 9.195 carboxylic acid CH3CH2C(O)OH 191.8 10.525 nitrile CH3CH2CN 192.6 1 1.84 NBA N02C6H4CHZOH 194.0a phenol C5H50H 196.3 8.47 ether CH3CH20CH3 196.4 9.72 ketone CH3C(O)CH3 196.7 9.705 ester CH3CH2C(O)OCH3 200.2 10.15 thioether CH3SCH2CH3 203.5 8.54 ammonia NH3 204.0 10.16 l°—amide CH3C(O)NH2 206.2 9.65 glycerol HOCH2CH(OH)CH20H 209.0a aniline C6H5NH2 209.5 7.720 glycine NH2CH2C(O)OH 21 1.6 8.80 2°-amide CH3C(O)NHCH3 212.7 9.3 alanine NHzCH(CH3)C(O)OH 214.8 8.88 phenylalanine NHzCH(CH2C6H5)C(O)OH 216.5 8.4 l°-amine CH3CH2NH2 217.0 8.86 valine NH2CH(CH(CH3)2)C(O)OH 217.0 8.53 dipeptide CH3C(O)NHCH2C(O)OCH3 217.7 leucine NH2CH(CH2CH(CH3)2)C(O)OH 218. 1 8.5 l proline NHC4H7(C(O)OH) 220.2 8.3 2°-amine CH3NH2CH3 220.6 8.23 pyridine C5H5N 220.8 9.25 methionine NH2CH(CH2CH2SCH3)C(O)OH 22 l .4 8.3 tyrosine NH2CH(CH2C6H40H)C(O)OH 222.3 8.0 dipeptide CH3C(O)NHCH(CH3)C(O)OCH3 224.5 9.2 piperidine C6H10NH 226.4 8.05 lycine NH2CH(CH2CH2CH2CH2NH2)C(O)OH 230.3 8.6 histidine NH2CH(CH2C3H3N2)C(O)OH 231.9 triethanol amine N(CH2CH20H)3 233.0a gimme NHZCH(CH2CH2CH2NHC(=NH)NH2)C(O)OH 245.2b 3‘ Obtained from reference [48]. '3 Obtained from reference [49]. 29 the internal energy imparted into the [M+H]+ needs to be greater than 1 eV. Extensive fragmentation equivalent to or greater than EI can be achieved in CI if the reagent gases such as hydrogen or methane, which have PAs of 101.2 and 131.6 kcal/mol, respectively, are utilized. For example, protonated peptides, having PAs equal to 220 kcal/mol, would have as much as 5.2 and 3.8 eV of internal energy if hydrogen and methane were used, respectively, which is necessary for abundant fragment ion formation. Thus, in C1, the extent of fragmentation from the protonated molecule can be controlled by the PA of the reagent gas. For protonated analytes with high internal energy and two or more basic functional groups having the potential for protonation, the fragmentation processes observed are not well rationalized because the two fragmentation pathways are in direct competition as are the local fragmentation pathways. Thus, the QET theory, developed to rationalize the appearance of El spectra, may be employed in CI to explain the observed abundances of fragment ions assuming protonation activates M and the vibrational internal energy in [M+H]+ ions eventually concentrates in certain bond(s) causing intramolecular rearrangements and/or bond dissociation(s) [50]. In medium pressure C1, the rates of ion source unimolecular decomposition are faster than the rate for collisional de-excitation of the activated [M+H]+ by the excess reagent gas, and the fragment ions observed are formed in 10'7 s or less. 2. Potassium Ion Ionization of Desorbed Species (K+IDS) Potassium ion ionization of desorbed species (K+IDS) [51] is a novel D/I technique developed in the Allison research group 10 years ago that complements the current D/I techniques. K+IDS is essentially a chemical ionization experiment using K+ as the reagent ions because the K+ ions (cat)ionize both intact analyte molecules (M) and neutral thermal degradation (Dn) products that are thermally desorbed into the gas phase, forming [M+K]+ and [Dn-i-K]+ adduct ions, respectively. The [M+K]+ and [Dn+K]+ 30 adduct ions conveniently yield molecular weight and structural information, respectively. K+IDS illustrates the basic, well-understood principles of desorption and ionization. K+IDS is similar to DEI, DCI and FD in that analyte molecules are thermally desorbed and subsequently ionized by K+ in the gas phase. The K+ adduct ion signals are still transient in K+IDS as in other thermal D/I techniques and very thermally labile analytes may still undergo facile degradation preventing intact desorption. In contrast to current D/I techniques, FAB, LSIMS, MALDI, and E81, K+IDS does not require a matrix, so matrix effects and interferences are avoided. K+IDS is a novel mechanistic tool for studying the thermal decomposition of analytes because K‘t simply adds to the desorbed intact neutral M and neutral degradation products, and does not induce any new fragmentation. Thus, K+IDS allows the condensed-phase, thermal decomposition mechanisms of thermally-activated analyte molecules to be elucidated which complements the gas-phase, charge—induced unimolecular decomposition of intact, activated analyte ions in the other thermal and particle beam D/I techniques. K+IDS is introduced here because it was used in the FAB experiment to investigate the ion formation mechanisms and was a starting point for performing M+IDS experiments. K+IDS Instrumentation. The K+IDS experiment is performed on the end of a modified direct insertion probe (DIP) in the ion volume of an El ion source as shown in Figure 6. The K+ glass emitter bead is formed by melting a small amount of a refractory mixture that contains high purity potassium nitrate, alumina and silica in a 2:1:2 mole ratio, onto a rhenium wire loop at very high temperatures. It has a final molar composition of leO: lAle3z28102 after firing since N02 gas is released [52]. In a K+IDS experiment; the K+ glass emitter bead is rapidly heated to 1000° C or more by passing 2 to 3 amperes of current through the underlying Re wire. High voltage wire feed-throughs in the DIP connect the K+ glass emitter to an external power supply outside the instrument. Once 31 Repel] er Ion Source Volume + K+ 16:4- MD2 M M + K+ D D2131 [Mm] —>M Anal ’ ' " + ass yzer K+ Glass / [Dn+K] Emitter (1000° C) Sample holder (0—> 300° C) .... ..I.'..-'.‘ . - . . . . . - . - - ' Ion Optics Figure 6. Schematic diagram of the K+IDS experiment. Desorption Favored l Degradation Favored | 3 I r: a F | i: :1 '° | < .5 O > I '5' .9. 9 | a: l I High T l/T (K) Low T Figure 7. An Arrhenius plot that demonstrates the competition between desorption and degradation processes in thermal heating. 32 the K+ glass emitter is at high temperatures, low-energy K+ ions are thermionically emitted into the gas phase [53]. The heat generated from the production of K+ radiatively heats the analyte "dried" on the adjacent Re wire loop sample holder to approximately 300° C which induces thermal desorption of intact M and/or thermal degradation and desorption of Du neutral products. A rapid heating rate will generate abundant [M+K]+ ions, while a slow heating rate will provide more abundant [Dn+K]+ ions. The two- filament DIP design [54] in Figure 6 is based on a similar DIP developed by Rollgen et al. [55]. This two-filament DIP performs better than the original method of placing the analyte directly on the K+ glass emitter [56] because it delays desorption of the analyte until abundant K+ ions are present in the gas phase. The K+ glass emitter potential can be biased equal to or above the E1 repeller potential so the energy of the K+ ions can be varied. K+IDS is best interfaced with a quadrupole mass analyzer because all the analyte is desorbed and ionized within 10 s after rapid heating is initiated. This means an extremely fast scan rate (i.e., 2 scans per second, full mass range) is required to capture as many mass spectra of the transient ion current as possible for an accurate reconstruction of the analyte desorption profile. Also, K+IDS has the potential to form K+ adduct ions with a range of KEs; fortunately, the quadrupole mass analyzer will mass filter these ions despite the wide KE range. Thermal Desorption and Degradath of M. Desorption of an analyte molecule from a surface occurs when all the non- covalent, intermolecular bonds between the molecule and the surface and/or neighboring molecules are broken. Because a highly polar biomolecule such as a peptide has many functional groups, there is a high probability that more non-covalent, intermolecular bonds will have to be broken for desorption of these intact species as opposed to the case for nonpolar molecules. Thus, for highly polar peptides, substantial thermal energy, equal to the heat of vaporization, needs to be deposited into the 3N-6 degrees of freedom 33 to induce non-covalent, intermolecular bond breaking. In K+IDS, the rapid heating concept, first introduced by Beuhler et al. [57], is utilized to make the rate of thermal desorption for intact neutral analyte molecules competitive with the rate of decomposition so both molecular and structural information can be obtained. In Figure 7, desorption is kinetically favored at higher temperatures while degradation is favored at lower temperatures assuming the relative ion abundances are proportional to the k(E)s. Therefore, if the analyte is heated at a high rate, less time is spent at lower temperatures, which induce degradation, so more analyte molecules will survive and can desorb intact at higher temperatures. If the heating rate is low, then more neutral degradation products will be formed because the temperature will be below the crossover point for a longer period of time. Quite often, thermal decomposition pathways for a biomolecule will lie at a lower activation energy than thermal desorption so neutral decomposition products (Dn) will be formed by a intramolecular bond cleavage(s) instead, owing to the thermal liability of biomolecules. Thermal decomposition of a biomolecule results in the formation of two stable molecular fragments. For peptides, usually one molecular fragment is a six-membered cyclic product that contains the two residues from the N- terrninal side and the other is a linear peptide product that contains the remaining residues from the C-terminal side, assuming a central amide bond was cleaved. Studies of peptide deuterium-labeled peptides indicated that 1, 2-elimination mechanisms do not occur and that cyclization mechanisms dominate. For K+IDS analyses of peptides containing only neutral amino acid residues, the cyclization mechanism with the lowest reaction barrier dominates the thermal decomposition processes [58]. K4“ Ionization of Gas-Phase M and D". The thermal desorption of many neutral molecules in the presence of K+ ions creates a medium pressure region similar to that under CI conditions. In this region, ion/molecule reactions from bimolecular collisions between K+ and M or D“ in the gas 34 phase result in the formation of either [M+K]+ or [Dn+K]+ adduct ions as shown in Equation 12. The higher the medium pressure region and the better the temporal and K+ + M (9 (DH (9) —) [M+K]+ ([Dn+K]+) (12) [M+K]+* ([Dn+K]+*) + M (Dn) -> IM+K1+ ([Dn+K]+) + M* (Dn*) (13) spatial overlap of K+ ions with M and D“, the higher the collision rate between K” and M or D“, which is essential for adduct ion formation. In Equation 13, the medium pressure region also aids in providing de-activating collisions to remove the excess energy from the excited adduct ions so no fragmentation occurs [59]. The order in which alkali metals induce gas-phase fragmentation is Li+ > N a+ > K+ > Rb+ > Cs+ [60]. This order is based on the affinity of M or Dn for the particular alkali metal, in a manner similar to PAs. Of the alkali metals, K+ is a compromise because the electrostatic interaction between M or Dn and K+ is stronger than with C54“ or Rb+, yet it is not as strong as the "near" covalent interaction between M and Li+. For Li+ and Na+, interaction is more "covalent" than with K+ meaning more of the charge is passed over to the neutral molecule which results in more charge-induced fragmentation, similar to the covalent bond between M and H+ [61]. Thus, the process of cationization of neutral molecules with alkali metal ions produces stable adduct ions that do not readily fragment as do certain protonated molecules, so more intact analyte ions are obtained for molecular weight information [62]. 3. F ast-Atom Bombardment Mass Spectrometry and Related D/I Particle Bombardment Techniques In 1981, Barber and co-workers [63, 64] reported the first high quality D/I mass spectrum of a peptide, Met-Lys-Bradykinin (MW 1318), using fast-atom bombardment (FAB) as a novel ionization technique on a double-focusing instrument. However, it was not until later in 1981 when they revealed their unique sample preparation secret, which 35 was the real significance of their work, since the differences between keV-atom and ion guns were well understood in the surface sciences. Instead of "drying" the analyte on the target as in "static" SIMS, Barber and co-workers [65] dissolved the analyte into a polar, viscous matrix, glycerol, which enhanced analyte sensitivity and signal stability, eliminated analyte degradation, and extended the lifetime of the analyte ion currents from seconds to minutes. Today, the D/I techniques FAB and LSIMS (i.e., SIMS incorporating the liquid matrix) either utilize a fast atom or an ion gun to bombard an analyte dissolved in a liquid matrix with keV-Xe or Cs+ particles, respectively, to induce sputtering (i.e., emission) of ions, electrons and neutral molecules from the condensed- phase into the gas-phase. Thus, they are essentially similar techniques that yield identical results. Both FAB and LSIMS are commercially available and have essentially opened the door to routine mass spectrometric analyses of biomolecules and polymers allowing for major advances in the biological, biomedical, and polymer industries. However, the full potential of FAB and LSIMS may be unrealized because the D/I mechanisms are still unclear, which results in the inability to understand the mass spectrum/condensed-phase analyte relationship and resolve matrix effects that still preclude classes of analytes from this analysis. FAB and LSIMS Instrumentation. "Static" SIMS and FAB are essentially an outgrowth of the surface science technique of "dynamic" secondary-ion mass spectrometry (SIMS) that is utilized for depth profiling of metal, inorganic and semiconductor materials [66]. "Dynamic SIMS" involves using a fast ion gun to bombard the sample with keV-particles at high ion beam current densities on the order of 1.0 [IA/cm2 in comparison to 1.0 nA/cm2 from the fast. atom and ion guns used in "static" SIMS and FAB. The fast ion gun for "static" SIMS was introduced by Benninghoven in 1970 to obtain mass spectral information of inorganic surfaces [67] and of organic molecules adsorbed on metal surfaces [68]. The 36 fast atom gun technology is as well known as that of ion guns [69], and during the 19605 and 19705, it was used to analyze inorganic surfaces, and organic and biological molecules adsorbed on metal surfaces [70]. Bombardment of the these surfaces with keV-atoms instead of ions supposedly eliminated sample charging effects [71]. In FAB and LSIMS, the highest secondary ion yields are obtained when Xe and Cs+ are the bombardment species, respectively, because they deliver the greatest momentum transfer to the sample [72]. However, LSIMS is generally more sensitive than FAB, because the Cs+ beam can be more tightly focused on the target at much higher bombardment energies [73]. This deposits more energy into the analyte/matrix which enhances the desorption of analyte species, increases the analyte ion currents, and essentially lowers the detection limits for most analytes. In FAB, the Xe beam defocuses between the gun and target which results in a large area being irradiated on the target and ultimately reduces analyte sensitivity. The angle of incidence (i.e., the angle between the axis of the keV-beam and the normal to the target surface) of the fast atom/ion gun that results in the highest secondary ion yield is approximately 70° [72, 73] and instrument dependent. Stainless steel targets are most common and least expensive, but gold- and silver-plated targets have been found to give greater sensitivity [74]. Both FAB and LSIMS instrumentation can be implemented on sector, quadrupole and TOF mass analyzers. EnsLAtnnLQuns. Barber and co-workers developed the FAB ion source for analyzing biomolecules on a double-focusing instrument, which took advantage of its high mass and resolution capabilities. Barber and co-workers overcame the high accelerating voltages by utilizing a neutral keV-beam of atoms for their experiments because it was not influenced by the high accelerating potentials in the ion source, meaning fewer precautions (or modifications) needed to be taken in contrast to the application of ion beams in LSIMS. The fast atom gun for the JEOL HX-l 10, illustrated in Figure 8 to scale, produces a 6- 37 Ion Source Housing & Filament (3050 V, Xe 10'5 torr) Xe (g) + e' (200eV) A5 Xe+' + 2e' \ Beam Steering (0.50 V & 0 V) Collision Chamber (0 v, Xe 10-4 torr) . CE 6+ (3 keV) + Xe (g, thermal) ' “'5 +0 CE conversion Xe (~3 keV) + X6 (thermal) to Xe atoms Exit Nozzle & Xe Gas Inlet (-3000 V, Xe 10'3 torr) I Fast Xe Atoms (6 keV) Secondary Ions to Mass A FD/F AB Target DIP \_ :l l I FD/F AB Ion Source Figure 8. Schematic diagram 0 the fast atom gun (actually a SIMION model drawn to scale) and the F-AB experimental configuration on the JEOL HX- 110. J 38 keV-beam of Xe atoms at 1.0 nA/cm2 beam densities in a series of discrete steps. In the ion source of this fast atom gun, Xe gas is ionized by El using a highly efficient ion source design identical to that of a residual gas analyzer [75]. The electrons are thermionically emitted from a large Re filament (3050 V, not shown) positioned between the "anode" (3250 V) and the ion source housing (3050 V), and are accelerated to the anode with 200 eV of energy from the 200-V potential difference between the anode and the ion source housing/filament. The electrons circulate in and out of the cylindrical wire mesh anode held at 3250 V because the electrons are repelled by the opposing ion source housing circular walls (3050 V), which enhances the ionization efficiency of this ion source . The Xe“ ions are formed with the highest probability near the top of the anode, and they are subsequently extracted and accelerated to a final KB of approximately 3 keV from the potential difference between the ion source housing (3050 V) and the extraction lens element held at zero V. The ion focus (2600 V) and x- and y-beam steering (~0 V) lens elements comprise an einzel lens that shape Xe“ ions into a well-defined beam and steer it towards the exit nozzle, which is biased at -3000 V in the collision chamber. After the einzel lens, the well-defined 3-keV Xe“ beam undergoes further acceleration to a final KB of 6 keV because the potential difference between the einzel lens (0 V) and the exit nozzle (-3000 V) is 3000 V. In the collision chamber or exit nozzle, at least one-half of the 6—keV Xe+' ions acquire an electron from thermal, gaseous Xe atoms, present at medium pressures, by resonant charge exchange to produce neutral ~6 keV-Xe atoms with minimal reduction in KE. Because the recombination energy of monatomic ions equals the ionization energy of monatomic atoms, charge exchange between Xe't‘ atomic ions and Xe atoms requires zero energy. After charge exchange, the 6-keV Xe neutral beam that emerges from the exit nozzle of the gun is unaffected by electric fields in the ion source, and it impinges on the FAB target, attached to the end of a DIP, with a diverging beam shape. The secondary ions formed as a result of bombardment are extracted and accelerated by the ion source optics system. The Xe” ion (black) and 39 neutral Xe trajectories (gray), shown in Figure 8, were estimated using the electrostatic lens design and analysis computer program, SIMION 4.0, which will be described in more detail in Chapter three and Appendix one. There are other commercially available fast atom guns that are based on the same principles as above, but incorporate slightly different designs [76]. W A fast ion gun for LSIMS is somewhat more difficult to implement on a double- focusing instrument and more precautions need to be taken. The fast atom gun should be mounted as close to the target as possible plus the beam needs to be highly shielded from the accelerating potentials in the ion source because they can change the Cs+ ion beam shape and path. In addition, to generate a 10-keV beam of Cs+ ions, the gun potential needs to be 10 kV higher than the accelerating potential, which may be as high as 10 kV on a sector instrument, so Cs+ reaches the 10-kV target. These two problems are alleviated in FAB because the keV neutral beam is not influenced by electric fields which allows the fast atom gun to operate at much lower potentials. A LSIMS ion source, based on the design by Aberth et al. [77], is available for the JEOL HX-l 10 and a schematic diagram of it is illustrated in Figure 9 to scale. The principles of operation for a fast ion gun in LSIMS are conceptually more simplistic than those for the fast atom guns above because the Cs+ ions generated in the ion source of the gun are extracted, focused and accelerated to bombard the analyte/matrix target directly, attached to the end of a DIP, in the ion source of the mass spectrometer. Specifically, the Cs+ ions are generated by thermionic emission from a heated Cs+ glass emitter (15 kV). They are then are extracted and shaped into a well-defined beam by the ion extractor (14995 V), ion focus 1 (14250 V) and ion focus 2 (12750 V) lens elements which comprise an einzel lens. Finally, they are accelerated to the final, desired KE for bombardment on the sample target (10 kV) between ion focus 2, the C54“ shield lens element (10600 V) and the target itself. I I u ".1:- u .- Cs+ Pellet (Cs‘I' Glass Emitter) & Housing (15 kV) Ion Extractor (14995 V) ‘/ Ion Focus 1 (14250 V) ‘ [on Focus 2 (12750 V) / Cs+ Gun Shield (10600 V) I l/ Cs+ Beam (5 keV) Seconda Ions r to Mass Emlyw< Target (10 kV) mpg __l I LSIMS Ion Source Figure 9. Schematic diagram of the fast ion gun (actually a SIMION model drawn to scale) and the LSIMS experimental configuration on the JEOL HX—110. 41 Sample Preparation and Matrix Selection in FAB and LSIMS. The successful and accurate analysis of an analyte in FAB and LSIMS depends on the correct choice of matrix and additives for that particular analyte so analyte response is maximized and sample effects, matrix effects and interferences are reduced or eliminated [78, 79]. In most cases, the sample preparation for FAB and LSIMS involves initially applying a small, consistent amount of the matrix to the target surface and then mixing a known amount of analyte, usually nanomolar amounts for FAB and picomolar amounts for SIMS, either in solid form or dissolved in a co-solvent such as water or methanol, into the matrix. The sample on the target is then inserted into the mass spectrometer for analysis by keV particle bombardment. The following criteria should be considered when selecting a matrix for a particular analyte [80]. The matrix should dissolve the analyte, possess low volatility under vacuum to preserve the analyte signal, and resist reaction with the analyte. The standard matrix in FAB and LSIMS is glycerol (MW 92) because it readily dissolves most polar analytes, it is viscous and inert, and it provides strong, long-lived secondary ion yields under keV bombardment preventing analyte degradation. However, glycerol is not always the best—suited matrix for each analyte and application. Choosing a matrix for a particular analyte is not an exact science and trial and error is required. The differences in the physical and chemical properties of matrices such as PAs, solubilities, vapor pressures, heat of vaporization, viscosities, reduction potentials, dipole moments, etc. WM. Several reviews and papers have described the advantages, disadvantages, physicochemical properties, and applications of alternative matrices such as thioglycerol, triethanolamine, and 3-nitrobenzyl alcohol [81, 82, 83]. If the analyte does not dissolve in the matrix, no analyte signal will be observed by bombardment. To aid in the dissolution of an analyte, additives, such as acids/bases and/or polar/nonpolar co-solvents, can be dissolved into the analyte/matrix to either preionize and/or solvate the analyte, respectively. 42 Typical FAB and LSIMS Spectra. The typical and reproducible LSIMS mass spectrum for 1 ug/uL of triglycine (M) dissolved in the matrix glycerol (G) is shown in Figure 10. The two sets of peaks observed for M and G are overlaid in one mass spectrum which may potentially complicate the interpretation of analyte ions if any glycerol peaks interfere. The peak at m/z 93 represents protonated glycerol while the peaks at rn/z 29, 45, 57, and 75 represent . fragment ions of glycerol. The peak at m/z 185 represents the proton-bound dimer of glycerol. Peaks representing these glycerol "cluster" ions extend out to approximately m/z 921, which represents 10 glycerol molecules bound to a proton (i.e., m/z=92n+1 where n=10), with a decreasing RI distribution. The peak at m/z 190 represents protonated triglycine and yields the MW information for this analyte. The structurally significant fragment ions for triglycine are labeled and identified on the triglycine structure in which the charged fragment is "identified" to by the arrow. The mass spectrum essentially resembles a CI mass spectrum owing to the protonated species. Most fragmentation involves loss of even-electron (EE) neutral fragments from EE protonated molecules to form EE fragment ions by simple cleavages reactions and rearrangements reactions that involve hydrogen shifts. However, some analytes such as triglycine display a mix of EB and odd-electron (OE) fragment ions. The peaks at every mass or so-called chemical noise is attributed to glycerol radiation-damage products from Xe beam-induced high energy processes [84]. This mass spectrum demonstrates the disadvantage of assessing the RI of a peak representing an analyte fragment ion in the presence of a matrix. The R1 of the peak at m/z 76 represents the ion current from both a 13C-isotope fragment ion of glycerol and the y 1 fragment ion of triglycine. Thus, for most analytes, FAB and LSIMS can provide molecular weight and significant structural information, but for some analytes, these techniques are ineffective owing to the unique chemical properties of the analyte in the presence of a matrix. In the analyses of certain 43 .555: Shook—M 05 E volcano 83 32v 050wa we 43w:— uo 8.58% an: meg—mi— .3 Emmi as es en ...s_.._...em.r.-oe _.__.._.__s.“_________$.1.; __ ____._._______.r.________n___..__________._.33.______.._:=_____._.._._______._.__.__:___._._:j ___.__ _______ :____ i __ _ m_m mom ____ a m: a 2 a . a 0+H=+Ma§~ Nfi VON— m: . . mm~ NQ n» m 9. : . D I ma .ov a I e w a n__u W2, 30/ o \zmm hm rm H mm .8 e :o\~ U ,0 :z N 0 ..~__ . >_ n__u -EU 12.39.20) 2 i __ m. .9: e+HI+m£ Nn Na 3» mm 0 . . o— +E+§ +5.6". +5.é fl um um . m o2 Lem: mm: o2 44 classes of analytes, understanding the current mechanistic details of desorption, ion formation and fragmentation in FAB and LSIMS are important in resolving chemical incompatibilities between an analyte and its associated matrix and additive(s) . Mechanistic Aspects of FAB and LSIMS. In FAB and LSIMS, analyte ions may arise at various points in time and space by different mechanisms during the keV-bombardment process as determined by their chemical nature and/or position relative to the keV-beam impact site [85]. For example, the analyte or matrix molecules may be (I) initially preionized in solution and then desorbed intact, (2) immediately ionized and fragmented by uptake of too much energy in the bombardment process if too close to the impact site, or (3) desorbed intact and ionized by gas-phase ion/molecule chemistry after desorption. Traditionally, in FAB and LSIMS, desorption, ionization, and fragmentation are considered as sequential, separate processes in time, and they will be described that way here [86]. Although this section describes events that are highly controversial, debated, and not well understood, the following discussion is based on the available information in the literature and is a good attempt to provide insight into the mechanisms responsible for formation of the ions typically observed in the mass spectra of biomolecules. Understanding these mechanisms is required for successful sample preparation and accurate spectral interpretation. t' im . Historically, the first theoretical desorption (or sputtering) models were developed to gain a better understanding of the sputtering mechanisms responsible for production of atoms and ions from solid surfaces (i.e., metal, amorphous, or crystal lattice) under SIMS keV-bombardment so that secondary ion yields could be improved [87]. The now widely accepted collision cascade model formulated by Sigmund [88] in 1969 was first to quantitatively describe the sputtering of atoms (not ions) from metal surfaces which is 45 based on elastic (not inelastic) momentum transfer between the keV-projectile particle and each atom in its random path. Qualitatively, in this model, sputtering is initiated when a keV-projectile ion or atom enters the solid and is gradually slowed by binary inelastic collisions with various atoms until all of its translational KE has been transferred to the surrounding atoms in the interfacial region along its track. Then, these initially excited atoms in the track undergo second generation binary inelastic collisions with neighboring atoms and so on, forming a "collision cascade". The initially-deposited energy is essentially dissipated radially away from the projectile track by momenta transfers. Energetic binary collisions near the surface may sputter atoms into the gas phase if the surface binding energies are overcome. The atomic ions, the more important species in a SIMS experiment, are probably formed by electronic excitation and a subsequent electron transfer process near the initial keV-particle impact site, but unfortunately, the sputtering of atomic ions is not as well understood as that for atoms [89, 90]. The collision cascade model is believed to principally energize the solid, and sputter atoms away from the target with up to 20 eV of KE . The thermal spike desorption model [91, 92] evolvesifrom the collisional cascade later in time when the high translational energies therrnalize the lattice structure and sputter atoms with KEs less than 1 eV [93]. The thermal spike model describes the collisionally activated region radially surrounding the track of the keV-projectile as being a high temperature region that behaves like a dense gas because this lattice structure contains excessive vibrational (thermal) and translational energies. In this hot, high pressure gaseous region, the thermal spike model predicts that a large number of atoms/molecules in the upper surface layers will undergo an explosive expansion into the cool vacuum similar to that in a hydrodynamic process [94]. The collision cascade and thermal spike phenomenon for- atoms are widely accepted, although the subject is still debated because much remains to be learned. The two desorption models for atoms described above have been successfully 46 extended to provide insight into the sputtering processes for keV bombardment of organic molecules dissolved in liquid matrix. Fast Xe atoms bombard and penetrate the glycerol matrix, which contains a dissolved analyte, from a large angle and deposit keV energy into the upper layers of the matrix as illustrated in Figure 11. Initially, the impinging keV—Xe atom induces a collisional cascade in glycerol which is shown by the "black" translational energy flow paths. The desorption events are considered isolated since the beam density and flux are such that one keV-Xe atom impinges the same area every 1-5 seconds. Because the ratio of matrix to analyte is greater than 1000: 1, glycerol molecules uptake most of the energy deposited by the keV-Xe projectile, which essentially protects the analyte from damage by collisional excitation. Hoogerbrugge et. al. [95] have observed this phenomenon because sputtered N 3*, polyatomic ions, and molecules from a glycerol matrix were energy-analyzed and had KEs of only 5 eV or less, and no high energy tail. This indicates a temporary, less violent collision cascade occurs in liquids in comparison to that in solids. In Figure 11, the top-layer of glycerol molecules at or near @ Mass Analyzer \\ Figure 11. Energy deposition by a keV-Xe beam induces desorption of atoms, fragment ions, preionized molecules, and intact molecules (not drawn to scale). 47 the site of Xe impact may be so electronically and vibrationally excited that they are sputtered intact and fragmented completely into atonrs, atomic ions, ion pairs, or small, high energy ions according to the QET theory [96]. At a larger radius from the Xe impact in a less energetically excited region, the desorption of less excited glycerol molecules occurs which would fragment into one of the normally observed glycerol fragment ions (G1:+). In the least excited outer radial region, intact cations, preionized intact analyte ions, and intact glycerol and analyte molecules are desorbed. During the thermal spike regime, the translationally excited molecules in the collision cascade become vibrationally excited, so weak intermolecular bonds (i.e., H bonds and van der Waals forces) between glycerol and analyte molecules are cleaved, ready for low energy desorption. The highly vibrationally excited glycerol molecules (G) in the interfacial region of the surrounding projectile track are vaporized and ionized to form G+‘ in Equation 14. Since a peak representing G+' is not observed in FAB and LSIMS mass XC(5 keV) + G(g) —) G” + e‘ + Xe(..6 keV) (l4) 6*” -> GF+ + N120 (15) Xe(6 key) + 6(2) —9 612+ + A‘ + Xe(..6 key) (16) Gp+ + G(g) —> G-GF+ + H2 (17) spectra, the average excitation energy must be high enough that G+’* fragments completely, according to the QET theory, to form the glycerol-related fragment ions, GF+, by unimolecular decomposition in Equation 15. Gp+ may also be formed by the separation of an ion pair from collisionally activated glycerol as shown in Equation 16. Also, in Equation 17, GF+ may react with gas-phase G to form, G-Gp+, which is essentially a fragment ion of a proton-bound dimer of glycerol [97]. In addition, in this interfacial region, some analyte and matrix combinations undergo undesirable, keV- particle beam induced recombination reactions such as oxidation and reduction [98, 99], and/or additions and substitutions [100, 101]. These reactions may occur in the gas- and/or condensed-phase, may involve ions and/or molecules of glycerol, analyte, matrix 48 additives, etc., and are probably catalyzed by the highly excited matrix molecules in the interfacial region. Undoubtedly, these reactions are most serious when they affect the analyte ions, so the possible particle beam-induced products should be recognized for spectral interpretation [102, 103]. High-energy, particle beam-induced free-radical glycerol chemistry also occurs in the interfacial region surrounding the Xe impact site yielding an abundance of radiation-damage products [84]. At some point in time, the entire hot, activated region surrounding the impact site will "explode" into the vacuum, removing and vibrationally cooling several exterior layers of excited intact analyte/matrix molecules and preionized analyte molecules, and the representative interior ionic species and radiation products in the interfacial region. Sunner et al. [104] calls this a "phase explosion" and discusses at great length how and why the superheated matrix irreversibly "explodes" into the gas-phase. Similarly, in MALDI, a microvolume containing analyte molecules entrained in a principally solid matrix absorbs laser energy and undergoes a "collective" phase change forming 104 more intact gas—phase neutral molecules than ionized intact analyte molecules [105]. Recent results indicate that the top most layers of matrix and analyte are sputtered away completely exposing a new analyte/matrix surface [106, 107] because the rate of analyte diffusion from the matrix bulk to replenish the matrix surface is slow on the experimental time scale [25]. Thus, a matrix such as glycerol will provide interpretable, long-lived analyte ion signals as long as volatile radiation-damage products are produced and sputtered away, and beam-induced reactions between the analyte and matrix are minimized. Otherwise the matrix would likely polymerize and stop functioning. For MALDI, Vertes et al. [108] gives insight into the vibrational energy flow in the crystalline lattice, and also between the analyte and the lattice according to the bottleneck model which is applicable to desorption in other D/I techniques. The desorption of salt cations, preionized analytes and intact neutral analyte molecules from a glycerol matrix dominate in the FAB and LSIMS experiments because 49 glycerol solvates these species and separates the ionic charges which significantly reduces the surface binding energies required for desorption. Benninghoven [109] introduced the precursor model to account for the desorption of preionized analytes and cations directly from the matrix in Equation 18. Here, if a collision from the collision M“ A‘ (aq) -> We) + A'(g) (18) cascade provides sufficient energy, the preionized species on the surface are desorbed into the gas phase usually with little vibrational internal energy content. Here, a large dose of energy that is deposited locally from the collision cascade results in limited vibrational energy gain by the analyte in comparison to global energy deposition in thermal desorption because energy is concentrated in one part of the matrix and its dissipation from this site is very fast (<10'12 5). For alkali and quaternary salts traditionally thought to be desorbed as intact cations and anions, Kidwell et al. [110] suggests that activated ion pairs are desorbed into the gas-phase instead and followed by dissociation in the gas phase. In contrast, Sunner [111] recently suggested that desorbed intact cations and anions pair up during desorption or the disintegration of the matrix. As mentioned above, the molar ratio of glycerol matrix to analyte is 1000:1 so most neutral molecules and ions desorbed by the collisional cascade and thermal phase explosion regimes are related to glycerol. Wong et al. [107] estimated 1000 glycerol molecules are desorbed for each Xe impact. In addition, the glycerol matrix has a high vapor pressure under vacuum. Thus, between desorption and evaporation of glycerol molecules, the liquid-vacuum interface is estimated to be a high pressure region. The observation of protonated analyte molecules and EE fragment ions in typical FAB or LSIMS mass spectra indicates CI processes may be operative in this high pressure area. The coupling of CI with FAB for analysis of postdesorbed neutral species has resulted in [M-I-H]+ sensitivity increases by three orders of magnitude, which indicates a large 50 number of neutral species are present in this region [112]. A similar experiment by Schroder et al. [113] concluded gas-phase glycerol molecules, introduced into the gas- phase from a heated inlet, were initially ionized from collisional activation by the keV—Xe beam to produce the glycerol reagent ions, G‘”, which then protonate other gas-phase glycerol molecules in a process very similar to that for CI. Historically, Rabalais and co— workers [114] noted the formation of cesium iodide clusters by Ar+ bombardment of C51 using SIMS and attributed this phenomenon to a high pressure region above the target surface they termed the "selvedge". Here, the desorbed cations and anions with similar energies moving through this "selvedge" region would aggregate into gas-phase cluster ions. Using SIMS, Cooks and co—workers [115, 116] used the selvedge region concept to explain the cationization of organic molecules with protons and desorbed metal ions from the target in Equation 19. They described the selvedge at the liquid—vacuum interface as a M(s) —) M(g) + H+(g) (Ag+, K+) -) [M+H]+(g) ([M-I-Ag, K]+) (19) high pressure, high temperature, and highly charged region where gas-phase, recombination, CI-like, associativeion/molecule cherrristry occurs [117]. The height of the selvedge region was estimated to be on the micrometer-nanometer level, but its height should be defined as where the gas-phase chemistry ceases. In Figure 12, the neutral and ionic species desorbed into the selvedge region by FAB are shown in addition to the [G+H]+ and [M+H]+ ions formed. Kebarle and co-workers [118, 119] noted if two matrices with different gas-phase basicities (GB) are mixed on the target, with one having the typical concentration of an analyte, and analyzed by FAB, the matrix with the lower GB will be suppressed by an amount relative to its concentration in comparison to the matrix with the higher GB. The GB (-AGm,) is approximately equal to the PA (-AHm.) if the Aern'T term in the Gibbs-. Helmholtz equation Aern=Aern-A8m,°T is negligible, but if a proton dicoordinates or cyclizes, ASrxn cannot be neglected and GB¢PA. Acid-base chemistry in the condensed phase was ruled out because these matrices do not readily ionize or exhibit dependence 51 Mass Analyzer [M +H1+ [G+H]"' Salt '9' [(M+H)-H20]+ Z ’ " ~ ~ ‘ p Z Free Vacuum [(G+H)'H20I+ / ’ ..... 3:- / ®® Selvedge Region + [M+H]+ [Gan] ® ® Salt G // 5 G M + Fast Atoms G M + (j / F" Gas Phase " .’ ~ ’1’ .\“ \“ Figure 12. The neutral and ionic species desorbed into the selvedge region above the FAB target during Xe bombardment. on bombardment time. From the above observation, they proposed the "gas collision model", which is similar to the selvedge region model. The majority of the species desorbed into the gas phase are glycerol (G) molecules. Also, "trickles" of analyte molecules (M) are desorbed as are trickles of the ionic fragment species such as Gp+ that are related to glycerol and formed near keV-Xe impact site as presented in Equations 14- 17 above. An additional, fraction of ionic species may be formed in the gas phase when intact gas-phase glycerol molecules are collisionally activated by a colliding keV-Xe atom and ionized by E1 or charge transfer/CE processes (see below) [120]. Primarily, the G1:+ ionic species, the fragment ions related to glycerol (and also to a lesser degree G- GF+ and 6"“), undergo many collisions with G and initially protonate the intact, gas- phase glycerol molecules in the high pressure selvedge region by a process similar to that 0f CI as shown in Equation 20. A protonated glycerol molecule may undergo a 52 GF+ (G+') + G(g) —9 [G+H]+ + [Gp+-H+] ([G+’-H+]) (20) bimolecular collision with a gas-phase analyte molecule which will result in a proton transfer if the GB of the analyte is higher than that of glycerol as shown in Equation 21. [G+H]+ + M(g) —> [M+H]+ + G(g) (21) Thus, this model portrays FAB essentially as a glycerol chemical ionization experiment using protonated glycerol molecules as the reagent ions for protonating the analyte molecules. In addition, the analyte may be protonated by G“ or the ionic fragments of glycerol if the analyte GB is less than that for glycerol as shown in Equation 22. Once 01:4” (64") + M(g)( —> [M+H]+ + [Gp+-H+] ([G+‘-H+]) (22) protonated analyte molecules are formed in the selvedge region, they may undergo a nonreactive collision to remove excess energy, a reactive collision with another G to produce a "cluster" ion, [M+G+H]+, or an immediate extraction towards the mass analyzer which will probably result in fragmentation if they were not vibrationally "cooled" in the selvedge [121]. In Equation 23, if two analytes M1 and M2 are dissolved M1(g) + [M2+H]+ —> [M1+H]+ + M2(g) (23) in glycerol, and M1 has a higher GB than M2, but M2 has a higher solution pr, [M1+H]+ will have a higher RA than [M2+H]+ supporting CI gas-phase chemistry in the selvedge since proton transfer from [M2+H]+ to Mug) will occur assuming both analytes are initially protonated equally by glycerol ionic species [122]. The formation of "cluster" ions in FAB and LSIMS represents another selvedge region reaction type in which protonated glycerol molecules collide with G to form a proton-bound dimer of glycerol as shown in Equation 24. The decreasing RI distribution [G+H]+ + G(g) -> [2G+H]+ (24) 0f peaks representing the increasing size of glycerol cluster ions according to the formula, m/z=92n+1, in a typical FAB mass spectrum indicates cluster ions are most likely formed from a sequential clustering process in the selvedge region. Because the larger clusters require more collisions in the selvedge, they are not formed as often as smaller cluster lllI f0 53 ions that require fewer collisions as indicated by the peak RIs. This distribution may also indicate that larger clusters are not as thermodynamically stable and are not readily formed. The other school of thought considers the possibility that both small and large neutral and ionic cluster species are desorbed into the gas-phase instead of individual molecules from the thermal expansion (phase explosion) of the matrix. These small and large ionic cluster species then desolvate in the free vacuum to give the distribution of peaks observed in the mass spectrum [94, 123]. The formation of a bare protonated analyte or glycerol molecule from a large cluster ion seems unlikely as the dominant process because experimentally, [G+H]+ and [2G-i-H]+ are the dominant glycerol ions observed and a decreasing RI distribution for peaks representing the larger glycerol cluster ions is observed. If declustering was a major process, a gaussian RI distribution of peaks representing the "surviving" glycerol cluster ions would be expected since complete desolvation seems unlikely under FAB and LSIMS experimental conditions [122]. However, dissociation of cluster ions probably occurs to some extent depending on the analyte and matrix. For preionized analytes or cationic salts, R4N+X', in positive mode, the respective cations, [M+H]+ and R4N+, would be desorbed into the selvedge region. The high RIs of intact cationic substances indicates either they are desorbed in staggering numbers, desorbed with low internal energy that prevents fragmentation and/or are initially excited and undergo nonreactive collisions for de-excitation. If the R4N+X‘ ion pair is desorbed intact or undergoes aggregation in the selvedge, the reaction in Equation 25 will yield R4N+X' + [G+H]+ —> R4N+ + G(g) + HX (25) R4N+ in the selvedge, and R4N+ may subsequently undergo an ion/molecule collision in the selvedge with glycerol to form [R4N+G]+. In addition, preionized [M+H]+ ion may be neutralized to M if its GB is less than that of another desorbed molecule as in Equation 23. Also, if the matrix is d0ped with KCl salts, either K+ or KCl will be desorbed into the selvedge region to form K+ adduct ions with glycerol and analyte molecules as shown in 54 Equations 26 and 27, respectively [97]. The exchange of K+ between G and M will occur K+ + G(g) (M(g)) —> [G+K]+ ([M+K]+) (26) KCl(g) + [G+H]+ ([M+H]+) —> [G+K]+ ([M+K]+) + HCl (27) as in Equation 21. For polyaromatic hydrocarbons (PAH), the M“ ion is observed instead of a protonated molecule which is expected because the PAs of these compounds are very low (~18O kcal/mol) in comparison to that of glycerol (209 kcal/mol). Because protonation is not a readily available option and PAHs are highly volatile, ionization may occur by El, CE or collisional activation in the selvedge region from sputtered secondary electrons during Xe bombardment as in Equation 28 [117], from a charge transfer/charge exchange Mrs) '7 Mrg) + e'(>12 eV) -> M” (28) process where keV-Xe ionizes residual Xe (Xer) in the ion source which then ionizes M Xeawv) + Xer -—> XC(~keV) + Xe” + e' (29) Xe(kev) + M(g) —> XC(~keV) + M” + e' (31) by CE as in Equations 29 and 30 [120], and from collisional activation of M by a keV-Xe particle as in Equation 31, respectively [120]. Thus, the high collision rate between glycerol and analyte molecules, and the dependence of the chemistry on GBs in the high-pressure selvedge region allow CI-type ion formation mechanisms to explain the presence of various protonated molecules, cations, clusters ions, and molecular ions in the FAB and LSIMS mass spectra. Recently, in the work of Rouse and Allison, desorbed analyte molecules were detected in the so- called selvedge region using the K+IDS-by-FAB technique, which supports the mechanisms set forth by the Kebarle and Cooks research groups (see Chapter three and Appendix two for details) [124]. Cooks and co-workers [86] introduced a unified D/I model based on the similarity of mass spectra from FAB, LSIMS, PD and LD analyses for a wide range of analytes. Basically, the energy deposited by particle bombardment is 55 isomerized (i.e., redistributed) in the analyte-matrix system resulting in the desorption of analyte species, subsequent ionization and stabilization in the high pressure "selvedge" region by C1 processes and/or unimolecular dissociation in the free vacuum. Recently, in MALDI, an ion formation model was proposed from fundamental experiments in which the matrix is first photoionized and subsequent ion/molecule reactions between these desorbed, photoionized matrix species and gas-phase analyte molecules occur in the gas- phase [125]. Il'lll! "8' Both the neutral and preionized species contain an undetermined, but sufficient, amount of initial internal energy from the desorption event to induce unimolecular decomposition (after ionization, if required). Species desorbed from the collision cascade and thermal desorption regimes likely have high and low internal and kinetic energies, respectively. In the high pressure selvedge region, the initial internal energies of these species will ultimately be determined by reactive (i.e., protonation) and nonreactive collisions which increase and decrease internal energies, respectively. The longer an ion or molecule spends in the selvedge region, the greater the probability for more stabilizing collisions that ultimately decrease the initial internal energies. The protonated molecules, either preionized or formed in the selvedge region, are gradually extracted into the free vacuum towards the mass analyzer by the ion source ion optics. In FAB and LSIMS, the observation of metastable transitions that yield EE fragment ions following protonation of the analyte indicates that gas-phase unimolecular decomposition occurs in the free vacuum above the selvedge region [86]. The [M+H]+ internal energies are sufficient to induce either fast simple cleavages or slower rearrangement fragmentation since fragment ions are observed on the "normal" and metastable timescales. The [M+H]+ fragmentation appears to be kinetically competitive, and is expected to be in accordance with the QET and RRKM theories discussed above for El (see Chapter 4 for peptide fragmentation 56 mechanisms in LSIMS). Attempts to estimate the internal energy content of desorbed intact neutral molecules and preformed ions emerging from the selvedge region have been made previously as described below. Hoogerbrugge et al. [126] determined the internal temperature of glycerol molecules sputtered from a target of 267 K to be 190 K, which indicates the desorbed glycerol molecules are cooled by expansion into the vacuum and/or by nonreactive collisions in the selvedge. This internal temperature of sputtered glycerol molecules corresponds to an internal energy of 0.044 eV assuming only 15% of the total degrees of freedom are active. Derwa et al. [127] and Williams et al. [128] have estimated, using benzylpyridinium salts in glycerol, that the internal energy content of preionized secondary ions in FAB and LSIMS is between 0 and 4 eV; this energy distribution is most populated between 0 and 2 eV, with a high energy tail that extends out to 4 eV. Derwa et al. [127] used Hoogerbrugge's internal temperature value of 190 K to calculate an internal energy of 1.13 eV for desorbed benzylpyridinium ions assuming all 69 degrees of freedom were active. Kebarle and co—workers [118] and Hamdan [129] found the distribution of peaks representing fragment ions in FAB and FAB—CAD (see next section for CAD discussion) mass spectra for typical matrices (PAz200 kcal/mol) corresponds almost identically with the RI distribution of peaks representing fragment ions in methane CI (PA=132) and CI-CAD spectra. This indicates the protonated molecules formed in FAB and in methane CI have very similar internal energy distributions that extend out to 3 eV owing to the 68 kcal/mol difference in PAs. Sample and Matrix Effects in FAB and LSIMS. The inability to identify important analyte peaks from matrix interferences, and/or the partial or complete analyte signal suppression often observed in FAB and LSIMS mass spectra are referred to as sample and matrix effects [72, 78, 79, 80, 130]. A peak representing a matrix ion that unfortunately overlaps with an important analyte peak in 57 the mass spectrum is a common occurrence. The m/z values of the analyte or matrix peaks may be "shifted" by derivatizing the analyte or employing a different matrix, respectively. An MS/MS product ion scan (discussed below in Sections 4 and 5) will provide fragment/product ions free of matrix interferences. Cationization of the analyte molecules with NaCl or KCl salts, placed in the matrix, results in a new peak 23 or 39 mass units higher than the analyte MW, respectively, that represents the highly stable adduct ions, [M+Cat]+. These stable adduct ions do not fragment and they provide confirming MW information, especially if the exact location or form of the MW peak (i.e., [M+H]+, M+', a fragment ion, or [M+K]+ adduct ion) is unknown [131, 132]. If the concentration of salts is high, as in biological samples, additional adduct ion peaks such as [M-I-2K-H]+ will appear in the mass spectrum which is less (interpretable at higher mass, and quite possibly the analyte peaks may become partially suppressed in the mass spectrum . HPLC is usually used to remove salt impurities from biological samples. Analyte suppression is a serious matrix effect, especially in analyte mixtures, because the R15 of each component in the mass spectrum may be a poor representation of their actual concentrations. Suppression can be caused by poor analyte solubility (see "sample preparation" above), poor surface activity, or an unfavorable difference between the analyte and matrix PAs (see "mechanisms" above). A "surface active" analyte preferentially concentrates in the matrix surface which results in preferential desorption and thus strong ion signals. Preionized analytes and hydrophobic analytes are usually surface active while hydrophilic analytes readily dissolve and disperse in the matrix which results in poor ion yields. The addition of acids/bases and surfactants to the matrix may bring the hydrophilic analytes to the surface as will preionization of the analyte by fixing a charge using known derivatization techniques to the analyte [133, 134]. Some analytes respond better in negative ion detection mode owing to a highly acidic group, while other analytes respond better in positive ion mode owing to a highly basic group. 58 4. Principles and Utility of Collisionally-Activated Dissociation (CAD) For accurate determination of analyte structure, abundant and informative fragment ions in the mass spectrum are required, as are routinely obtained by El. However, the mass spectrometric analyses of thermally labile, high molecular weight biomolecules require a "soft" ionization technique, either FD, FAB, LSIMS, MALDI, PD, or ESI, which on average only impart a small amount of internal energy into the protonated molecules in relation to their large size. In contrast, a large amount of internal energy is imparted into small molecules by El. Thus, structurally informative fragmentation for these high molecular weight biomolecules is limited because the small amount of internal energy imparted into the protonated molecules is distributed over more vibrational degrees of freedom than is the case in E1. As a result, the probability of sufficient internal energy concentrating in a bond or group of bonds to induce fast fragmentation in the ion source is lowered. Collisionally-activated dissociation (CAD) [135] can be employed to impart a range of sufficient internal energies into low energy [M-i-H]+ ions through energetic bimolecular collisions with an inert gaseous atoms to stimulate both high- and low-energy fragmentation processes so that a wide range of structural information can be obtained. The amount of energy imparted depends on the number of collisions, KB of the [M+H]+ projectile ion, and the atomic size of the target gas. In fact, CAD product ions may be obtained from any user-selected precursor ion, which is any intact ionized analyte molecule or fragment ion formed in the ion source before acceleration that appears in the "normal" mass spectrum. Often times, protonated biomolecules with low internal energies also undergo metastable transitions which may produce all or some of the complementary structural information required. However, metastable transitions usually represent the low-energy fragmentation options for a poorly activated protonated molecule meaning that the metastable fragment ions observed underwent primarily rearrangement reactions. If the identical fragment ions are formed 59 in CAD and metastable decomposition, it is important to realize that the metastable mechanisms induced by lower average internal energies may be different from the CAD mechanisms induced by higher average internal energies. Since CAD fragmentation occurs after the acceleration of the user-selected precursor ions, which are formed in the ion source, tandem mass spectrometry is required for detection; this technology is discussed in the next section as will the applications of CAD. High-energy CAD is performed on sector instruments because the ion source accelerating voltage and ion KEs are in the keV range while low-energy CAD is performed on instruments containing quadrupoles that ultimately mass analyze the CAD product ions since the accelerating voltages and ion KEs are restricted to the eV range. Low-energy CAD was not performed in this work and is discussed elsewhere [136]. High-Energy CAD Instrumentation and Parameters. In high-energy CAD [137] on a double-focusing instrument, a lO-keV-ion beam traverses a collision cell, positioned between the ion source and the mass analyzer entrance (see Figure 14), containing an inert gas such as He, Ar, or Xe at thermal energies. The ion beam undergoes one or more energetic collision(s) with the gaseous atoms, depending on the magnitude of the cell pressure, which in turn activates the ions as shown for [M+H]+ in Equation 32. If the [M+H]+ is user-selected, meaning [M+H]+1o keV + He —-) [M+H]+*(..1o keV) + He (32) [M+H]+* -> [M+H]+1' —> Pn+ + Nn (33) fragmentation is desired from this [M+H]+, then the mass analyzer will detect the product ions, Pn+, formed from the unimolecular decomposition of the [M+H]+-’F in the critical configuration following collisional activation of [M+H]+* in Equation 33. In high- energy CAD, each successive collision imparts more energy into the colliding [M+H]+ ion which in effect opens up fragmentation pathways requiring high energy such as simple bond cleavages and the successive decomposition of primary product ions into 60 secondary or higher product ions [138]. The number of collisions is proportional to the target gas pressure in the collision cell. Since the pressure in the collision cell cannot be measured directly, the number of collisions is adjusted by increasing or decreasing the target gas pressure in the cell so the user-selected ion signal is attenuated by a certain percentage . For example, the attenuation of the ion signal by: (1) 10% results in 100% single collisions; (2) 50% results in 70% single, 20% double, and 10% triple collisions; and (3) 70% results in 50% single, 30% double, 16% triple, and 4% quadruple collisions, respectively [138]. In addition, the energy uptake by user-selected precursor ions can be controlled by the size of collision target gas [139]. Helium is best for high-energy collisions because it minimizes scattering and neutralization of [M+H]+ ions while still yielding abundant CAD fragmentation from moderately efficient [M+H]+ activation. However, the increased atomic size of Xe and Ar allows more efficient activation of [M-I-H]+ ions, but the competing scattering and neutralization processes are increased resulting in sensitivity losses. High-Energy CAD Energy Deposition and Fragmentation. The first step in high—energy CAD involves the conversion of a small portion of the [M+H]+ ion's translational energy into additional internal energy as the result of a bimolecular collision with the target gas. The [M+H]+ ion is excited electronically from the electron repulsion generated as the fast moving ion approaches and separates from the short-lived (i.e., approximately 10‘14 s), inelastic ion-target molecule collision complex. In CAD, the maximum KE available for conversion into [M+H]+ ion internal energy, Ecom, is calculated using center-of-mass (COM) considerations as shown in Equation 34. Beam = Elab (mt/(mt+mion)) (34) For example, a user-selected precursor ion, [M+H]+, with a m/z value of 500, set equal to man, has the potential to gain 79 eV of internal energy, Ecom, assuming He is the target gas (m = 4) and Eiab is the 10 keV accelerating voltage. However, in high-energy CAD, 61 assuming single-collision conditions exist in a He-filled collision cell, the average energy uptake by the [M+H]"’ from the bimolecular collision in Equation 32 is only between 1 and 3 eV with a high-energy tail extending out to 15 eV [140]. The average internal energies imparted into [M+H]+ are far less than Beam and have been shown to be proportional to the mass of the analyte ion [14]]. The second step after [M+H]+ activation is its unimolecular decomposition into a product ion(s), Pn+, as shown in Equation 33, which depends on its internal energy content for determining the rate of decomposition, k(E). According to the QET and RRKM theories discussed above, the internal energy is randomly distributed among the active 3N-6 vibrational degrees of freedom in [M+H]+* until it concentrates in certain vibrational mode(s) to cause bond cleavage(s) in [M+H]+*. Since the energy imparted into protonated molecules by CAD is on average high, both simple bond cleavages and rearrangement product ions are observed on fast and slow timescales. However, as the degrees of freedom increase with the number of atoms in [M+H]+, the internal energy imparted into [M+H]+ is increasingly spread over more vibrational modes and eventually it becomes ineffective in inducing competitive CAD fragmentation rates [142]. For analytes with MW 5 greater than 1500 Da, CAD is essentially ineffective for inducing abundant, useful fragmentation. In addition, the CAD ion decomposition mechanisms may be different from ion source mechanisms for the same product/fragment ion because CAD may impart a different distribution of internal energies into the [M+H]+ ions, or low energy [M+H]+ ions for CAD activation may have different protonation sites than higher energy [M+H]+ ions that fragment in the ion source. CAD is only one method of activation, but its simplicity makes it most popular. Surface-induced dissociation (SID) [143, 144] and photo-induced dissociation (PID) [145, 146] techniques have demonstrated that more internal energy than that of CAD can be deposited in large biomolecule [M+H]+ ions, but their set up and operation are more complicated. 8f. 3B 0p§ 11161 and Will] i002. 62 5. The Mass Analyzers Used for D/I Experiments D/I techniques are most often performed on commercially available mass spectrometers that employ either a double-focusing, quadrupole, or time-of-flight (TOF) mass analyzer. As is true for ionization techniques, no one mass analyzer has the ideal combination of mass range, mass accuracy, resolving power, sensitivity, scanning speed, etc. for all mass spectrometric experiments, and compromises are common [147]. The double—focusing sector mass analyzer provides high mass range, high resolution (at the expense of sensitivity), and accurate mass determinations for ions generated by the ionization techniques EI, CI, FI, FD, DCI, FAB, LSIMS, and CF-FAB. Unfortunately, sector instruments are expensive, difficult to maintain, non-portable, possess slow scan speeds (in comparison to quadrupoles and TOF over the identical mass range) and utilize high acceleration and detector voltages. The quadrupole mass analyzer provides unit resolution over the entire mass range, fast scan speeds, strong focusing properties and no slits which enhances sensitivity, usability at higher pressures, easy maintenance, portability and affordability. However, the quadrupole mass analyzer has a low upper mass range limit in comparison to those of sector and TOF instruments, and it discriminates against ions of high mass. Quadrupoles are most often used as mass analyzers for the ionization techniques EI, CI, APCI, DCI, FAB, or ESI. TOF mass analyzers almost represent the ideal mass analyzer because they boast the fastest scan speeds, unlimited mass range, high sensitivity (no slits), easy maintenance and simple operation. Unfortunately, the resolution is often less than 500, which severely limits the identification and mass accuracy of high-mass ions. The addition of a reflectron [148] and utilization of high accelerating voltages greater than 10 kV improves the resolution to an upper limit of 1500 which is still too low for accurate mass assignments of analytes with molecular weights above 10 kDa. TOF is the ideal mass analyzer for the pulsed ionization techniques PD, LD, and MALDI. These ionization techniques are all capable 63 of producing ions with m/z values greater than 10 kDa which is well suited for TOF. In addition, EI and LSIMS can be set to pulsed mode for TOF analyses as well. The fundamentals of operation for these three analyzers will be described below, but first extraction and acceleration of ions will be discussed which is universal to any mass analyzer. Ion Extraction from Ion Source. The ions (e.g., M“) formed in the ion volume of the ion source are extracted and accelerated by the ion optical lens system on the ion source as shown in Figure 13a. The potential energy surface in Figure 13b is generated by placing the specified electrostatic potentials on the lens elements of the ion optics in Figure 13a. The ion of mass, In“, and charge, 2, in Figure 13a will move through a potential difference, V, of 10000 V between the exit slit and L2, and, therefore, gain 10000 eV of kinetic energy (KE) if z is one. Thus, the KB gained by the ion will be equal to the potential energy (PE) lost in Equation 35 as shown in Figure 13b. The repeller in Figure 13a simply "pushes" the newly formed Repeller Exit Slit Re 11 er . . (10910 V) (10008 V) Ion Optics PC EXlt Sllt 5 \l I Q A ‘3 > , Mass :73 I; ) M"""> <( . Analyzer g g) é’ 3’ L2 m (0 V) 0 Ion Source L1 Distance (cm) Volume (10000 V) (8000 V) a) b) Figure 13. The extraction and acceleration of M“ a) from an ion source and b) the representative potential energy surface for this ion source. KE = 1/2mnvn2 = zeV = PE (35) ions towards the exit slit for extraction by the ion optics. All ions are formed in the same general area in the ion source volume, so each ion will have nearly identical final KEs when they emanate from the ion source, but their velocities, V“, will be different and inversely proportional to their m/z ratio. By fine tuning the potentials on the ion optical elements, extraction of the maximum number of ions to enhance sensitivity and the generation of a well-defined ion beam to enhance the peak shape can be achieved. Sector Mass Analyzer. A Nier-Johnson, double-focusing mass spectrometer is shown in Figure 14. The forward-geometry mass analyzer contains an electric (E) and magnetic (B) sector in tandem, often expressed as EB. The ion beam shown in Figure 14 is isobaric meaning all ions in the beam have identical m/z values. In the E sector, a fixed electric field, E), is created by the potential difference between the two electrostatic plates, which is linked to the accelerating voltage. For ions to traverse the E sector of a fixed radius, R, and fixed electric field strength, E0, the ions need to have a KE equal to the accelerating voltage, V, as shown in Equation 36. However, isobaric ions that emanate from the main slit with R = mnan/Eo = 2 KE/Eo = 2 WED (36) angular divergence and/or ions with greater or lower KE in comparison to the accelerating potential are spatially refocused, if possible, which enhances sensitivity. Thus, there is a certain tolerance of ion kinetic energies that are passed through B because the beta slit deterrrrines the finite range of radii to be passed onto B, so R varies with ion KE in practice. The alpha slit width determines the maximum angular divergence of the ion beam that enters E, and the beta slit width determines the velocity spread of the ions; resolved by E, allowed to enter the magnetic sector, B. In B, the magnetic field extends out of the page, denoted by "x"s in Figure 14, and is perpendicular to the velocity vector, V“, of an ion. The force exerted by the magnetic field on an ion changes the direction of 65 Magnetic Sector (B) Collector Slit I-Collision Cell Direct Insertion Probe Volume Ion Source Figure 14. Schematic diagram of a double-focusing mass spectrometer. w-RFV Figure 15. Schematic diagram of a quadrupole mass analyzer. 66 its velocity, but not its magnitude. Thus, the basis of mass separation for the magnetic sector analyzer, as shown in Equation 37, is Newton's second law which states that ions rn =mnvn/zeBo (37) of different momenta, mnvn, but same 2, will have different radii, rn, in a uniform magnetic field, B0. In the early days of mass spectrometry, ions with different radii, having essentially different momenta, were detected simultaneously using a photographic plate before sophisticated electronic control systems and sensitive detection systems were developed [149]. Today, as illustrated in Figure 14, ion flux detectors are positioned at a fixed radius, r, such that ions of different momenta are detected individually by scanning (i.e., increasing or decreasing) the magnetic field. The ion flux detectors are more sensitive and convenient for routine use than photographic plates. Since the ions entering B have the same velocity, V“, as obtained from acceleration in the ion source, the velocity term in Equation 35 can be substituted for the velocity term in Equation 37 and this expression can be solved for any mn/z value as shown in Equation 38, assuming the magnetic field, B“, is scanned, and the radius of ion detection, r, and the accelerating mn/z = Bn2r2/2V (38) voltage, V, are fixed. Thus, by scanning the magnetic field, B", to higher fields at a fixed accelerating potential, V, each group of ions from the total ion beam with unique and increasing mn/z values will be sequentially focused onto the collector slit before the detector in Figure 14 when the radius, rn, of the group of ions equals the fixed radius, r, of the collector slit. If E and B work in tandem, the double-focusing properties of the mass spectrometer are achieved when the isobaric ions with diverging trajectories and/or velocity differences from inhomogeneous energies are corrected spatially in both sectors as shown in Figure 14. As a result, the beam of isobaric ions with various velocities and directions initially are (doubly) focused to a single point on the collector slit. In 1953, Nier and Johnson [150] designed the most sensitive and accurate, high resolution EB double-focusing magnetic sector instrument at that time by refining the 67 various instrumental geometries to achieve true double-focusing of the isobaric ion beams with inhomogeneous energies and/or angular deviations. The double-focusing mass spectrometers commercially available today, such as the JEOL HX-l 10, are based on the Nier-Johnson geometry and they include such refinements such as electrostatic quadrupole lenses to reduce ion beam dispersion in the horizontal and vertical planes (2- direction) before both E and B sectors. The FAB experiments described in this dissertation were conducted on the JEOL HX-110 because of its high mass range, mass accuracy, and sensitivity. Quadrupole Mass Analyzer. The quadrupole mass analyzer [151] consists of a set of 4 parallel, precision- machined rods arranged radially, equally distant from each other and a center point as illustrated in Figure 15. A quadrupole mass analyzer electronically "filters" low KE energy (i.e., 100 eV or less) ions directly by their m/z values through a combination of time-independent (dc) and time-dependent (ac, i.e., radio frequencies (RF)) potentials applied to the four rods. Even if the ions possess a wide range of energies or angular directions upon entering the central space between the four rods, the unit resolution remains unaffected unlike that in sector and TOF instruments because ions are filtered by their m/z value. The stability and paths of ions can be calculated from the Mathieu equations, but here, only a qualitative description will be presented. In Figure 15, the RF potentials applied to the yz plane are 180° out of phase from the RF potentials in the x2 plane while the dc polarities remain fixed. The quadrupole mass analyzer is scanned incrementally by increasing both the RF and dc potentials at a fixed ratio. When the RF is positive in a particular xz or yz plane, a positive ion will be focused to the center axis of the space between the rods, while a negative RF potential will defocus the ion towards the rods. As the RF voltage in the yz plane is increased during scanning, the positive ion will 68 eventually strike the rod when the RF negative potential becomes too large with respect to the m/z value and the position of the ion in the quadrupole assembly such that the positive RF fails to restore the ion's straying trajectories. Light ions are primarily affected by the RF voltage while heavy ions tend to feel only the average RF force. The dc potential in the yz plane is positive which contributes to focusing both small and large ions in the quadrupole assembly. Heavy ions are gradually defocused by the negative dc potential in the xz plane. As the negative dc voltage is increased, heavy ions are eliminated when the respective RF potentials fail to restore or stabilize their trajectories owing to their increased momentum. Thus, the yz rods pass high-mass ions, but filter low-mass ions because the RF has negligible filtering effect on high-mass ions. The rods in the xz plane pass low-mass ions and filter high-mass ions because the negative dc potentials have little filtering effect on low-mass ions. By scanning or increasing the dc and RF potentials at a constant ratio, a group of isobaric ions at a particular m/z value, depending on the potentials, will remain stable in the xz and yz planes while the rest of the ions are lost to the four rods. The quadrupole mass analyzer was developed by Paul et al. [152] and recently reviewed by Dawson [153]. The K+IDS experiments described later in this dissertation required the fast scanning capabilities of a quadrupole mass analyzer and were performed on a Hewlett Packard quadrupole mass spectrometer. Time-o -Flight (TOF) Mass Analyzer. In TOF, ion packets are generated by pulsing the ion source accelerating voltage, V, on and off rapidly. The ions in each ion packet after acceleration gain nearly identical final KEs as explained above. As demonstrated in Equation 39, each group of ions of a KE = l/2mnvn2 (39) unique mass, In“, travel at unique velocities, v... TOF takes advantage of this phenomenon by measuring the time, t, it takes for an ion packet composed of ions of mass, In“, to translate the a hollow flight tube of distance, D, and reach the detector. The 69 mn/z values of ion packets can be determined as shown in Equation 40 by knowing V, tn, mn/z = 2eV(t,,/D)2 (40) and D. The TOF instrument was originally described by Wiley and McLaren [154] in 1955 and Cotter recently reviewed TOF instrumentation for applications in the biological sciences [155]. Tandem Mass Spectrometry (MS/MS). Until now, the mass analyzer discussion above has only considered scanning a single mass analyzer to separate, as a function of mass, the ions formed in the ion source volume before acceleration. Instrumentally, true tandem mass spectrometry (MS/MS) involves interfacing two mass analyzers on either side of a field-free region that contains a collision cell, stainless steel surface, laser excitation cell for the activation techniques CAD, SID and PID, respectively, although MS/MS on sector instruments is more obscure. Commercially available instrumentation capable of MS/MS are the triple quadrupole (QqQ), double-focusing (EB or BE), hybrid instrument (BEqQ) and four- sector (EBEB, two double-focusing mass spectrometers linked to either side of a collision cell) mass spectrometers where B is a magnetic sector, E is an electric sector, q is an RF- only quadrupole, and Q is a RF/dc quadrupole. MS/MS on each instrument will be discussed in further detail below. I! 1' EMS“: Tandem mass spectrometry instrumentation is necessary for observing the fragmentation of intact ionized analyte molecules and fragment ions formed in the ion source after acceleration, specifically, low-energy metastable transitions or product ions' from higher-energy processes formed by CAD, SID, and PID. Both types of ions arise from a single, user—selected precursor ion, which is formed in the ion source before acceleration and would appear in the "normal" mass spectrum. The popular MS/MS scan IN 70 type is the product-ion scan. Here, the tandem mass analyzer detects all metastable fragment or CAD product ions that arise from the user-selected precursor ion. Most often, the product ion scan is used in conjunction with CAD to obtain additional, complementary structural information for a highly stable ionized analyte ion that yields little informative fragmentation in the ion source or for obtaining primary fragmentation from a poorly activated [M+H]+ formed by a "soft" ionization technique such as FD. Additionally, if a mixture of analytes was present in a sample, a product-ion scan of each component coupled with CAD will yield only the structural information for each individual component. Another important MS/MS scan is the precursor ion scan which involves user-selecting the mass of a metastable fragment or CAD product ion and scanning the MS/MS instrument to detect all the metastable or CAD product ions of that mass that arise from all the respective decomposing precursor ions. Thus, the identity of each precursor ion is reconstructed by the respective metastable fragment or CAD product ion transitions because the precursor ions are impossible to detect directly. Product and precursor ion scans can be utilized to obtain structural and mechanistic information for the low-energy metastable and higher-energy CAD fragmentation processes. In fact, a "fragmentation map" can be constructed and possible mechanisms can be identified once all product ions from each precursor ion and all precursor ions generating each product ion are deterrrrined. However, this "map" only suggests the possibilities and should not be interpreted literally. A fragment ion produced by a low- energy metastable transition may not be the identical mechanism by which the same product ion is formed in high—energy CAD or in the "normal" mass spectrum because the precursor/[M+H]+ ions in each case have different average internal energies which ultimately determines which kinetically favorable mechanisms dominate. WWW Tandem mass spectrometry on the triple quadrupole instrument (QqQ) is straight 71 forward [156], so it will be introduced first. The RF-only quadrupole, q2, located between the two mass analyzer quadrupoles, Q1 and Q3, is the collision region for CAD or unimolecular decomposition region for metastable transitions. An RF-only quadrupole simply transmits ions of all masses from end to end with high efficiency owing to its excellent focusing properties. The "norm " mass spectral scan on the triple quadrupole is achieved by setting qz and q3 to RF-only mode and scanning Q]. If collision gas is present in qz, then low-energy CAD can be performed. If no collision gas is present, then metastable transitions will be observed. Only low-energy CAD is possible on this instrument because the accelerating voltage is less than 200 eV. However, it is very efficient, meaning the conversion of precursors to product ions is high, since the scattered ions are transmitted owing to the high focusing properties of the qg. For a product ion scan, the precursor ion is mass-selected by Q1 and the metastable/CAD fragment/product ions formed in q2 are mass-filtered by Q3. For the precursor ion scan, the fragment/product ion of interest is mass-selected by Q3 and Q1 is scanned for each precursor ion that yields that user-selected fragment/product ion in the collision region, q2. One advantage of the triple quadrupole for MS/MS work is that unit mass can be achieved for both the precursor/product ion selection and precursor/product ion scanning. This is important for accurate structural and mechanistic work because Q1 only transmits the analyte 12C isotopes of the analyte ion into q2 so only monoisotopic product ions are detected. The double-focusing and TOF instruments do not have this built in luxury. W MS/MS on an EB double-focusing instrument is not as straightforward as on the triple quadrupole. The plot of E versus B, and the example fragmentation map/mechanism in Figure 16 should aid in understanding the "normal", product and precursor scans on an EB instrument. For each of these scans, the accelerating voltage, V, is held constant. Along the normal scan line, the m0+ ion represents the [M-I-H]+ ion, 72 while my’ and m2+ represent the fragment ions from m0+. A portion of the m2+ ions are formed by a secondary fragmentation process from m1+. Since m2+, m1+, and mo+ ions along the normal scan line in Figure 16 are produced in the ion source and have full accelerating potential, they are transmitted through the electric sector, held constant at a potential of E0. Next, these ions are mass-dispersed in B and individually focused onto the detector at the appropriate magnetic field strength, Bn, during scanning. Thus, the ions along this scan line appear in the normal mass spectrum. However, if during normal operation an ion, mo+ undergoes a metastable transition after acceleration to form m2+ in the first field-free region between the ion source and E, the fragment ion, m2+, will have the same velocity, v0, as the precursor ion, mo+, but m2+ will most likely be filtered out in E because its KE is less than that of moi: because the mass of m2+ is less than mo+. To Metastable _(m2‘*‘)2 m2+ m1+ m0+ Transition " E0 B/E Scan Line Electric Field (V) Magnetic Field (T) Figure 16. A plot of the B, E plane for linked scans at constant B/E and BZIE on double- focusing instruments. 73 sample these metastable transitions in the first field-free region of an EB instrument, a product ion scan can be performed by a linked scan at constant B/E [157, 158] in which both E and B are scanned simultaneously, under computer control, holding the ratio between the 8,, and En fields constant. In this scan type, shown in Figure 16, the precursor ion, mot is user-selected from the "normal" mass spectrum, so the slope of the linked-scan line at constant B/E can be calculated from the Boon ratio for moi”, where E0 is the potential of E at full accelerating voltage at which mo+ is transmitted through E, and B0 is the magnetic field strength required to pass mo" onto the detector. Since m2+ has a lower KE than mo+, the E potential is lowered by scanning from E0 down to E2 so, for example, m2+, with a lower KE than mo+ is allowed to pass through E as illustrated by Equation 41. Likewise, m2+ has a lower momentum than mo+ 50 = movoz/R and 132 = m2V02/R, SO M2 = mo/mz => 132 = szo/mo (41) B0 = movolr and 8;; = mzvolr, so Bo/Bz = mo/m2 =9 B2 = szo/mo (42) B/E ratio = Bo/Eo = constant (43) but the same velocity, v0, so Bo needs to be lowered to B2 so m2+ can be passed onto the detector as shown in Equation 42. Thus, for a linked scan at constant B/E, the slope of the linear scan line in Figure 16 coordinates E, and Bn values for transmitting each metastable ion of mo+ onto the detector. The scan slope is defined by the BolEo ratio in Equation 43 or more accurately, Bo, since the user-selected mass is transmitted by this field strength and E0 is fixed. The magnetic field strength, B2, in Figure 16, required for transmitting m2+ to the detector in the linked scan at constant B/E is the identical field strength in a "normal" scan required to transmit the mo+ —> my“ metastable transition to the detector if it occurs in the second field-free region between E and B. Thus, for metastable transitions after the electric sector, a classic broad peak(s) at an apparent m/z value equal to the (m2+)2/m0+ ratio will appear in the "normal" mass spectrum. Additionally, if the metastable transition mo+ -—) m1+ occurs in the first field-free region, then by reducing E and B to the E1 and B1 fields, respectively, along the B/E scan line, 74 the m1+ fragment ion can be detected. Linked scans at constant B/E can be done for fragment ions as well, such as m1+ in Figure 16, and, in this case, a peak corresponding to m2+ would be detected along the m1+ B/E scan line according to the mechanism. For the precursor ion scan on an EB instrument, a fragment ion, m2+, is user- selected from the "normal" mass spectrum and a linked scan at constant B2/E, which has a parabolic scan line extending from zero to m2+ in Figure 16, is performed to identify the various precursor ions of mzf. The precursor ion scan is confusing because the actual precursor ions of a metastable transition are not directly detected. Instead, the different possible precursor ions of m2+ are represented by the metastable fragment ions from each precursor ion that has the same mass as m2+. In Figure 16, the B2/E scan essentially intersects the respective isobaric metastable ions in the B, E plane precisely at the B/E scan slopes of each precursor ion. For example, in Figure 16, along the B2/E scan line of m2+, the two broad peaks represent two metastable fragment ions equivalent in mass to m2+ formed from the metastable decomposition of the two precursor ions, m1+ and mi: In addition, the fragment ion m1+ has only one precursor ion, mof, in Figure 16, shown by the single broad peak on the m1+ B2/E scan curve. In Equation 44, the Bn2/En ratio is Bn = mnvo/r and E1] = mnv02/R =9 an/En = mnR/r2 (44) proportional to the mass of the user-selected fragment ion mn+ if both Bn and En are solved for v0 and set equal. The BZ/E constant for the parabolic scan line is calculated using the user-selected fragment/product ion mass to determine the appropriate magnetic field strength, B", from a "normal" magnetic field scan, assuming E0 is held constant. Linked scans at constant B/E and B2/E can be also utilized for detecting CAD product ions formed in the first field-free region assuming the collision cell is positioned between the rear of the ion source and in front of the main slit as shown in Figure 14. Poor resolution is obtained with linked scans at constant Bz/E, as indicated by the broad metastable ion peaks on the B2/E scan lines in Figure 16. The wide peak widths are proportional to the translational energy released (i.e., energy equal to the kinetic shift, 75 Eli) and the peak tailing originates from metastable transitions that occur during the late stages of acceleration or in the early stages of the E sector. The resolution for linked scans at constant B/E is at least unit mass because the NE scan line "cuts" these broad peaks on the Bz/E scan line perpendicularly. Unfortunately, the resolution of precursor ion selection for linked scans at constant B/E is low and becomes worse as the m/z value increases since B is nonlinear and the m/z values converge. Depending on the m/z value, a number of the analyte isotopic peaks and possibly a set of peaks from an interfering analyte may unavoidably be user-selected. This will yield a more complex fragment/product ion mass spectrum because the product ions will have isotopic peaks and, if an interfering analyte exists, its mass spectrum will overlay the desired analyte spectrum. A hybrid [159] and a four-sector mass spectrometer [160] are able to select the precursor ion with unit mass so that only 12C isotopic product ion peaks appear in the product ion mass spectrum. Until now, only first-generation CAD product ion spectra have been discussed. With hybrid or four-sector instruments, second-generation product ion mass spectra (i.e., MS/MS/MS) can be obtained [161]. W The detection of metastable transitions and CAD (i.e., assuming the cell is located between the ion source and flight tube entrance) fragment/product ions in linear TOF is not possible because the fragment/product ions have the same velocity as the precursors so both will be detected simultaneously. However, these fragment/product ions can be detected if the TOF instrument is equipped with a reflectron. By scanning the reflectron incrementally over small mass ranges, the fragment/product ions are sequentially focused on the detector since they have different KEs than their precursors. The fragment/product ions detected in each mass range increment can be assembled to form a composite MS/MS mass spectrum. However, for large biomolecules such as proteins, the highly efficient PID and SID techniques have the potential to activate the proteins more than 76 CAD and yield either more or entirely new products ions with better resolution, but unfortunately, these techniques are much more difficult to implement. Recently, two reflectron-TOF mass analyzers were linked to a PID laser excitation cell for high-speed MS/MS applications [162]. If [h. CHAPTER THREE. K+IDS-BY-FAB TARGET DEVELOPMENT AND APPLICATION TO MECHANISTIC STUDIES OF ION FORMATION IN FAST-ATOM BOMBARDMENT (FAB) I . Introduction Fast-atom bombardment (FAB) has served the biological community well by providing routine molecular weight and structural information for high molecular weight, highly polar, thermally labile biomolecules. However, the availability of FAB preceded an understanding of the chemical processes that yield gas-phase ions from condensed phase molecules. Insight into the FAB desorption/ionization (D/I) mechanism would allow for improvements in the technique for the analysis of compounds that are currently difficult to analyze, and the evolution of new applications and new ionization techniques. As described in Chapter two, a typical analysis by FAB involves dissolving analyte molecules (M) in a polar, viscous matrix, usually glycerol (G), with a GM molar ratio of 100021, respectively, and subjecting the G and M mixture on a stainless steel target to 6- keV Xe atoms for sequential desorption and ionization. Briefly, the desorption of intact M and G molecules from either the collision cascade, thermal spike, or final phase explosion events, in addition to the high rate of G molecule evaporation under vacuum, all contribute to a high pressure region that exists above the FAB target called the selvedge region. Here, in this high-pressure selvedge region, extensive gas-phase ion/molecule chemistry similar to that in chemical ionization (CI) is substantiated by a high rate of multiple collisions. Essentially, glycerol-CI conditions are created owing to the large number of G molecules in the experiment. A small fraction of ionic species related to glycerol, GF+, are responsible for initially protonating the glycerol molecules in the selvedge region to form [G+H]+. These GF+ species are likely formed when a keV- 77 78 Xe atom collisionally activates glycerol molecules in the condensed or gas phases (see Chapter two for specific reactions). In the selvedge, the CI reagent ions, [G+H]+, subsequently either (1) protonate the desorbed intact neutral analyte molecules having PAs greater than glycerol to form [M+H]+, (2) undergo subsequent collisions with G to form the [nG-i-H]+ (n22) cluster ions, or (3) undergo extraction towards the mass analyzer. Preionized analytes in the matrix may be desorbed directly into the gas phase without the need for further chemistry in the selvedge region. Last, the protonated molecules formed in the selvedge or the preionized analytes may unimolecularly decompose in the free vacuum above the selvedge to form respective fragment ions. In this chapter, the FAB ion formation mechanism described above will be investigated by injecting low—energy K+ ions into the selvedge region of the FAB experiment to probe the types of neutral species present there following desorption. This work is an outgrowth of potassium ion ionization of desorbed species (K+IDS), explained fully in Chapter two, which takes advantage of the fact that K+ ions only form adducts with gas-phase molecules without inducing fragmentation as in El and CI. Thus, this capability will allow the identification of desorbed neutral molecules by their respective [M+K]+ adduct ions. Since we use K+ here to ionize neutral molecules desorbed by keV Xe in the FAB experiment, we refer to this technique as K+IDS—by-FAB [163]. This chapter was written to supplement the previously published K+IDS-by-FAB paper [124] in Appendix one with additional studies and data not in that publication. In K+IDS-by-FAB, the K+ ions are injected into the selvedge region using a spatially-separated, two-stage FAB divided target as illustrated in Figure 17. One half of the divided target is the K+ alurrrinosilicate glass emitter, identical to the emitter used in the K+IDS experiment, and the other target half is the sample holder for the analyte and matrix as in normal FAB. In K+IDS, when the K4“ alurrrinosilicate glass emitter is heated to high temperatures, a large flux of K+ ions are therruionically emitted into the gas phase, but this same K+ glass emitter will also emit copious amounts of K+ ions when 79 Mass Analyzer / I [G+K]+ “(MD-H 291+ ®/ + Free Vacuum [M+K]+ [M+H]"’ [6:HL_ _ _ 7 ® /, , —-— , ’ [M+H]+ ® ‘:_____/ [G+K]+ I; (36% A [G+H]++G M Fast Atoms Selvedge Region K G K-I- Matrix / Analyte K+Glass Emitter Figure 17. A K+IDS-by-FAB divided target under Xe bombardment demonstrating the injection of desorbed 1K" ions into the selvedge region for the identification of desorbed neutral molecules. bombarded by keV-Xe atoms in FAB W [164]. The FAB beam simultaneously impinges both target halves desorbing K+ ions from the K” glass emitter target half, and desorbing intact analyte and glycerol molecules, preionized analyte molecules, and ionic glycerol species (G1:+) from the analyte/matrix target half as in the normal FAB experiment. The desorbed K+ ions may be manipulated by using both electrostatic extraction fields in the ion source and divided target geometries, swept across the divided target barrier in the gas phase and injected into the selvedge region above the analyte/matrix target half. These injected K+ ions will ionize the various types- of desorbed M and G neutral molecules present in the selvedge to form respective [M+K]+ adduct ions which are then used for the identification of these gas-phase molecules. A divided FAB target is used to separate the analyte/matrix from the K+ 80 emitter to prevent condensed-phase interactions between the two targets and ensure only gas-phase chemistry occurs in these FAB mechanistic studies. Since K+ ions are injected into the selvedge region above the analyte/matrix target, some further comments should be made on the characteristics of this region. To date, no direct experimental evidence has definitely proved the existence, the actual high pressure and the dimensions of the selvedge region above the FAB target. The existence of a selvedge region is validated by: (1) the independence of Xe projectile energy on types of ions formed in contrast to those from El and CI, (2) the observation of protonated intact molecules in the mass spectra, and (3) their correlation with established proton affinities (PA). The high local pressure in the selvedge region is expected based on high vapor pressure of glycerol, 0.13 millitorr, under vacuum, and the high sputtering yield estimated at greater than 1000 glycerol molecules per incident keV Xe atom from the liquid target [107]. A high density of gas-phase neutral molecules exists near the target-vacuum interface; these molecules probably undergo expansion into the free vacuum resulting in an exponential pressure gradient that extends away from the target. Calculations by Pachuta and Cooks [117] have indicated that within a volume of 1010 A3 the pressure exerted by 10 particles with 2 eV of energy is about 1 torr, a pressure able to sustain CI type collisions and chemistry. Thus, if an area of 106 A2 receives one keV particle impact per second and 10 or 1000 particles are desorbed to maintain a pressure of l torr, then the height of this selvedge region is 1 um and 100 um, respectively. The most useful dimensions of the selvedge region extend out beyond the target in a decreasing exponential pressure gradient to the point where ion/molecule chemistry ceases. Therefore, if K+ is injected above the FAB target and an [M-i-K]+ adduct ion is formed, then this can be interpreted as evidence that K+ ions sampled a region of space in which the pressure of the desorbed neutral molecules were sufficiently high that a collision occurred in the time period of a few microseconds, corresponding to the lifetime of K+ ions in the ion source. 81 2. Development and Characterization of the K +IDS-by-FAB Target The function of a divided target in a K+IDS-by-FAB experiment is to inject K+ ions into the selvedge region to sample desorbed neutral molecules. Various divided targets with minor geometry differences were designed and assembled with intuition which resulted in mostly unsuccessful results. In a more rigorous attempt, SIMION, an ion optics design and analysis computational tool, was used to assist in the development of a successful K'I’IDS-by-FAB target. These computational experiments revealed that the success of a K+IDS-by-FAB experiment depends essentially on the divided target design and geometry, and how well the divided target interfaces with the extraction potentials in the ion source of the mass spectrometer. Only after gaining insight to the K+IDS-by-FAB experiment with SIMION computer modeling was a functional divided target able to be designed and assembled for routine and reproducible use. All K+IDS-by-FAB experiments were conducted on the JEOL HX-l 10 mass spectrometer using the field desorption/fast-atom bombardment (FD/FAB) ion source and operating at 10 kV. The K+IDS-by-FAB target was placed on the end of a direct insertion probe and entered into the instrument in an identical manner as a regular JEOL FD/FAB target. The ion source temperature was 25° C and the ion source operating pressure with FAB Xe gas on was 5x10’6 torr. In this section, my experiences and insights for designing a successful K+IDS-by-FAB divided target are discussed. Generation of KI” Ions. The use of a divided target for the K+IDS-by-FAB experiment requires the keV- Xe beam to simultaneously impinge both the analyte/matrix on one target half to create the FAB experiment, and a solid surface on opposite target half that yields gas-phase K+ ions. These K+ ions are then injected in the selvedge region above the analyte/matrix 82 target half. The method of choice for generating K+ ions in a K+IDS-by—FAB experiment is bombarding a K+ alurrrinosilicate glass emitter surface held at room temperature in close vicinity of the analyte/matrix target. The FAB mass spectrum of such a K+ glass emitter surface, comprising one target half on a divided target, is shown in Figure 18a without any analyte/matrix present on the opposite target half. Bombardment of the glass emitter surface desorbs one dominant ion into the gas phase, K+, with an ion current proportional to ~66 counts out of 100 possible. Thus, K“ is the only ion injected into the selvedge region which provides for a straightforward K+ CI experiment identical to K+IDS. The K+ glass emitter is permanently attached to the divided target so it can be used for many experiments. However, it will eventually crack from careless handling before K+ is ever depleted. Depending on the divided target design the K+ aluminosilicate glass emitter is prepared either by melting the tertiary rrrixture of finely ground 2KNO3:1A1203:28102 directly on a divided target surface half with a Bunsen burner, or by dipping and coating the end of 0.007 in. diameter rhenium wire with a "head" of the molten 2KNO3:1A1203:28102 liquid prepared in a platinum dish by heating with an oxygen-acetylene torch [52]. In the latter case, the K+ emitter bead is the target surface half from which K“ is desorbed and the Re wire holding the K+ emitter bead is spot welded to the appropriate site on the divided target. In both cases, the final molar composition of the K+ glass emitter is 1K20:1A1203:2S102 after heating since N02 gas is released. Often times, for direct application to a divided target surface, the spattering of the Kf glass from N02 release during melting can potentially contaminate the analyte/matrix divided target surface. This nuisance is elinrinated by first melting the K+ glass in a platinum dish to release all N02, cooling, re-crushing the solid glass into a fine power with a mortar and pestle, and then melting a portion on the divided target surface. The more common way to generate gas-phase K+ ions is to bombard a KCl film dried on one target half from a saturated solution of KCl in water. The FAB mass spectrum in Figure 18b shows bombardment of a divided target with one target half 83 100‘ 39 If! . a) K + K+ glass emitter (66 counts K“) a 80‘. E 1 IV I e 601 1 I l‘ 40: e g l 201 + i j Na+ (5:; y 0‘ 213; I .__L_ I 0 50 100 150 200 m/z 100‘ 39 If? . b) K+ KC] solid emitter (74 counts 10*) I a 80‘ l i I I {I . r I ‘ ' {I 40; [(KC1)+K]+ Z c ‘ 113 l : 1 l 201 2K, [2(KC1)+K1+C Y ()1 l 713 1?: ' (3 50' T 150' ‘T 150 T r260 m/z 100 R j C) 93 Dual-surface space-divided target : f . [G+H]+ with KCI emitter on one half and - f1 30: glycerol on the other. f i : ; e 60: [(ZG+H)-H20]+ T l i 16712ci+m+ : g 40: 185 - n . S . 10 ’ l , [G+K]+ X . ty 201 31%;; ] 131 [20+K]+ [33;5‘1“; 1.331% ...... we ...23_3.....J... 100 150 200 250 300 m/z Figure 18. a) FAB mass spectra of K+ glass emitter and of b) KCl solid film on a divided target half without matrix or analyte present on the opposite half. c) K+IDS-by-FAB mass Spectrum of a KCl emitter and glycerol on either half of a dual-surface space-divided target. The latter spectrum is presented to show KCl is just as effective as K"' glass emitter in a R‘I'IDS-by-FAB experiment, but the KC] emitter is more difficult to apply cleanly to only one target surface plus it needs to be regenerated for each run. 84 coated with KCl without analyte or matrix on the opposite half. A strong K+ ion current proportional to 74 counts is obtained but the presence of [(KCl)n+K]+ cluster ions may undergo undesirable gas-phase chemistry with analytes because they will also be injected into the selvedge region and/or they may spectrally interfere with analyte peaks. In addition, the KC] solid surface needs to be regenerated for each experiment to insure integrity of the results which is quite tedious considering the small size of the target halves and high probability of contaminating the analyte/matrix target half from spattering. The sensitivity obtained from the reactants mixed on single stage target are much greater than the reactants separated on a divided target. The most popular method used to generate K+ adduct ions of analyte molecules with much higher sensitivity than K+IDS- by-FAB is to place KCl or KI salts directly in the analyte/glycerol matrix and perform FAB. The high sensitivity achieved with this method is great for molecular weight confirmation studies, but not for mechanistic studies because it is not certain whether the [M+K]+ ions are desorbed intact from the matrix or formed by gas-phase ion/molecule chemistry similar to KflDS-by-FAB. However, KflDS-by-FAB has some analytical advantages over FAB when KCl salts are used which will be discussed in Chapter five. Previous Experiments with Divided Targets. The first type of K+IDS-by-FAB experiment involved bombarding a single-stage target shown in Figure 19a in which the K+ glass emitter was melted on a stainless steel surface and the analyte/glycerol was placed on top of the glass [164]. As a result, abundant [M+K]+ adduct ions were detected for analytes, such as fructose, which do not respond well in standard FAB, with detection limits 2-3 orders of magnitude lower than the [M+H]+ ions formed from FAB of fructose in glycerol. Apparently, glycerol leached Ki" ions from the K+ glass emitter, and the K+ adducts of the analyte were formed either by desorption of intact [M+K]+ ions from glycerol, or in the gas phase by Kf CI of 85 3) mm. V/l/l/l/l/l/l/l/A c) Top Top Ke d) ‘\ y ‘\\\\ \ \\ K+ Glass Emitter ..‘\\\\\\\\\\\\\\\\\“\ - Analyte/Matrix Figure 19. a) The first K+IDS-by-FAB type target employing a K+ glass emitter reported by Akerrnann et al. [164], b) the wall and slit targets reported by Munster et al. [166, 168], c) the dual-surface target reported by Miller et al. [169], d) the K+IDS-by-FAB "set screw" target reported by Schultz et al. [172], e) the K+IDS-by-FAB "wall-divided" target reported by Kassel [171], f) the K+IDS-by-FAB "K+ bead pop-up" target, g) the K+IDS-by-FAB "vertical arrangement" FD/FAB type divided target, and h) the vertical, dual-surface space- divided target, all which provided unsatisfactory gas-phase K+ adduct ion formation for mechanistic studies. 86 desorbed intact analyte neutral molecules. Since the K+ and analyte were not separated in the condensed phase, no mechanistic conclusions could be made. Analytically, the analysis of fructose was straightforward and yielded only [M+K]+ adduct ions, but less desirably cholic acid (MW 408) yielded both [M-i-K]+ and [M-H+2K]+, which is the potassium adduct ion of the potassium salt of cholic acid. This study was an outgrowth of an archaeological study involving 19th century ceramic pottery in which elemental compositions and lead isotope ratios were obtained by direct keV-Xe bombardment of the ceramics attached to a standard FAB target. A comparison between the ceramic pottery and the National Bureau of Standards "common" lead glass led to the identification of their geographic origin [165]. The use of wall- and slit (space)-divided FAB targets shown in Figure 19b are utilized by FAB practitioners for exact mass measurements in which the analyte/matrix is placed on one target half and a reference compound such as C51 is placed on the opposite target half to reduce the number of prepared samples and to prevent unwanted sample/matrix effects from mixing both the analyte and the reference compound together in the same matrix [166]. As a result, exact analyte masses of the reference and analyte can be measured with high precision and accuracy because the reference target half can be bombarded individually before and after analyte bombardment, on the opposite target half, by simply rotating the respective divided target half in the Xe beam. Divided targets were also employed previously in conjunction with D/I techniques to explore gas-phase ion formation in the high pressure selvedge region. Chiarelli and Gross [167] studied gas-phase Na+ cationization of sucrose in the selvedge during laser desorption using a divided target with NaCl and sucrose spatially-separated by 30 pm to ensure the laser irradiated both halves. At low laser powers of 106 W cm'1 , they concluded that Na+ ions and neutral species were desorbed from both target surfaces in the thermal spike regime and a long-lived selvedge region existed above both targets since [M+Na]+ was observed, but at higher laser powers, 1010 W cm'z, no [M+Na]+ ions were observed owing to a 87 short-lived selvedge region in which the desorbed Na+ and sucrose molecules had high translational energies and little opportunity to overlap spatially and/or undergo three- body collisions for stabilization. Munster et al. [168] used the wall- and slit- (space—) divided targets in Figure 19b to show that two different matrices A and B placed on opposite target halves undergo ion/molecule reactions in the gas phase during Xe bombardment of both target halves, similar to a CI experiment, to form [A+B+H]+ ions. Miller et al. [169] placed a volatile crown ether in 3-nitrobenzyl alcohol (NBA) on the top half of the divided target, shown in Figure 19c, and placed either KCl salts in NBA or solid KCl deposited from solution on the bottom target half, which was spatially separated by 0.6 mm from the top target half. The bottom potassium salt target was selectively bombarded bottom with FAB to demonstrate that gas-phase [crown+K]+ adduct ions could be formed analogous to [M+H]+ in CI processes such as protonation of gas-phase analyte molecules or proton exchange between a protonated molecule and gas- phase neutral analyte molecule with a higher gas-phase basicity (GB). Michaud et al. [170] placed two different analytes, A and B, in glycerol on opposite divided target halves separated by a 200 um gap and did not observe the expected [A+B+H]+ adduct ions unless A was an alkali metal ion. This experiment indicates only single-site, electrostatic interactions are formed in the gas-phase. The [A+B+H]+ adduct ions were only observed when both A and B analytes were mixed in glycerol on the same target which indicates the multiple ligand-ligand interactions either occur in the condensed- phase or possibly in the selvedge region. Since they claim three-body collisions in the gas-phase are not likely, [A+B+H]+ may be desorbed preionized from the matrix, [A+H]+ may be desorbed preionized and react with desorbed B in the gas-phase selvedge, or the [AB] cluster may be desorbed intact and become protonated in the selvedge. In each mechanistic study above which employs a divided target for the study of gas-phase ion formation in FAB, there were many experimental difficulties and precautions that needed to be taken into account to ensure the observed product/adduct 88 ions were formed from gas-phase ion/molecule reactions and not from those in condensed-phase chemistry. Condensed-phase cross-contamination between the two target halves as a result of condensation, splattering, sputtering and/or creeping of reagents to the opposite target halves under keV-Xe bombardment had to be checked after each run. Usually, each reactant on each divided target half were individually analyzed by FAB for traces of the opposite reactant. If the condensed-phase reactants originated from separate divided target halves, then the ion currents of the product/adduct ions formed in gas-phase are much lower than the ion currents obtained when both condensed-phase reactants are mixed together on the same target stage. Depending on the divided target geometry, ions from one of the target halves may be preferentially formed/detected if keV-Xe bombardment is skewed and/or if the potentials on the ion source ion optics favor extraction from the "right" or "left" target. Moreover, just the right ion optical tuning parameters are needed to draw out the ions formed in the gas- phase above both target halves. Kassel in 1988 [171] and Schultz in 1989 [172] performed preliminary K+IDS-by- FAB experiments on the JEOL HX-l 10 mass spectrometer using the wall-divided target in Figure 19c and "set-screw" target in Figure 19d, respectively, for the initial investigation of the FAB gas-phase ion formation mechanisms. The wall in the wall- divided target prevents cross-contamination of the K+ glass emitter by glycerol during bombardment to ensure only gas-phase [M-I-K]+ adduct ion formation. The "set-screw" target allowed the height of the K+ glass emitter, located on top of the set-screw, to be adjusted in an attempt to probe the height of the selvedge region while keeping the glycerol at the base of the set-screw separate from the emitter. Irreproducible K+ adduct formation and unpredictable condensed-phase glycerol contamination of the K+ emitter were major problems for both of these early K+IDS-by-FAB targets. As discovered later in SIMION modeling, the wall- and set-screw target designs geometries did not interact well with the characteristics of the ion source preventing formation of [M+K]+ ions. 89 KHDS-by-FAB Target Development-The Next Generation. The current K+IDS-by—FAB work began by completely characterizing and understanding the simple wall-divided target beyond what was done previously. The wall of the wall-divided target was constructed using stainless steel or platinum ribbon (~0.007 in. thick) formed as shown in the center - w — and spot welded to the center of a standard JEOL FAB target. With the analyte/glycerol and the K+ glass emitter placed on either half of the wall-divided target, [M+K]+ adduct formation with desorbed neutral molecules was usually irreproducible and unpredictable, and often times, there were spectral and visible signs of analyte/glycerol contamination of the K“ glass emitter. In addition, optimization of the potentials of the ion optics for the extraction of the K+ adduct ions formed in the gas-phase above the wall region was not straightforward, and usually ions were extracted from either target half instead resulting in either an analyte/glycerol or K+ glass emitter FAB mass spectrum. The wall height was incrementally lowered from 1.5 mm to 0.1 mm (the lowest practical limit with simple magnification), similar to the "set-screw" target, without enhanced performance which indicated at the time that perhaps the selvedge region was on a smaller scale, it did not exist, or K+ ions were not successfully being injected into the FAB selvedge region above the analyte/matrix with this target. In Figure 19f, using the same concept as the "set- screw" target, the Re support on a K+ "bead" emitter was spot-welded to the bottom of the target in such a way that it emerged from a hole drilled through a standard JEOL FAB target. The analyte/matrix was placed on the JEOL target surface surrounding the K+ emitter bead. While this K+ emitter was flexible for adjustment purposes, this target also exhibited the same drawbacks as the wall-divided and set-screw targets. llM': "CII+EIE'- Following these initial studies, the first problem area in K+IDS-by—FAB to receive 9O attention was the condensed-phase cross-contamination of the K+ glass emitter by analyte/glycerol. In Figure 20a, the normal FAB mass spectrum of glycerol is shown in which the peak at m/z 93 represents protonated glycerol, the peaks at lower mass represent the fragment ions of glycerol and the peaks at higher mass represent the proton- bound cluster ions of glycerol (i.e., m/z=92n+1). In Figure 20b, a typical K+IDS-by-FAB mass spectrum of glycerol using the wall-divided target is shown in Figure 1%. The appearance of a peak representing K+ suggested that K+ was successfully injected into the selvedge region, and the intense peaks representing the K+ adduct ions of glycerol indicated there was a high conversion of desorbed glycerol molecules into K+ products. However, the peak at m/z 169 represents [G-H+2K]+, the potassium adduct ion of the potassium salt of glycerol, and its appearance is indicative of a condensed-phase reaction between glycerol and the K+ glass emitter surface under bombardment in which glycerol is sputtered or creeps from the opposite target half onto the K+ emitter. If the 10' glass emitter is purposely contaminated with glycerol, an identical mass spectrum is obtained as in Figure 20b. Most likely, the condensed-phase chemistry also contributed to the observed intense [G+K]+ peak since these ions may be desorbed preionized from the contaminated K+ glass emitter. Thus, if [G-H+2K]+ or [M-H+2K]+ peak appears in the KflDS-by-FAB mass spectrum, then K+ adduct ion formation cannot be guaranteed to be solely result of gas-phase ion/molecule reactions. In addition, after each KflDS-by-FAB experiment, the divided target was examined under a microscope for visible signs of K+ emitter contamination by glycerol. This method is less reliable for identifying small amounts of contamination but it does allow some glycerol flow paths to be identified. Therefore, to solve this contamination problem, the K+ glass emitter cannot be physically attached (i.e., spot welded) to the same surface as the liquid analyte/matrix because under bombardment, glycerol "creeps" around, up and/or over the barriers separating physically connected surfaces. In the "set-screw", wall-divided, and "pop-up" K+ emitter bead targets KflDS-by-FAB targets in Figures 19d, e, and f, respectively, the set-screw, the Ill REla‘t.I\ .f\ I‘ll“. Rela\rlv e Intens.11n\vc 91 1 R . 93 f 1 a) [G+H]+ 126+HI+ 9“ i1 801 185 HO-CHz-CH-Cflz-OH i, 60-: Glycerol on FDIFAB target it 4o-j I€oommm minced—b. 2: mean .ccccbm 5 Ag 35 secs .6 538% as: $139.0. ram .9 came mm mm -o . an . u . m . ON m . a o . a . ow : . _ . o .8 > A _ . u . a tow _ n o . ~— noo— 99 definite K+ adduct ion formation with glycerol and analyte molecules, a more reproducible and predictable K+IDS-by-FAB experiment was still desired in addition to stronger K+ adduct ions of desorbed neutrals. W+W A set of experiments was conducted to investigate the possibility of increasing the yield of K+ ions from the K+ glass emitter in the KflDS-by-FAB experiment by heating the K+ emitter similar to K+IDS in addition to Xe bombardment. However, only low heating currents could be used to heat the K+ emitter because if the K+ emitter became too hot too quickly, it would radiatively heat the analyte/matrix target and induce unwanted thermal desorption and/or degradation of the analyte. These experiments were challenging to implement on the JEOL HX-l 10 with kV-accelerating voltages because the K+IDS power supply unit is designed to operate in conjunction with quadrupole instruments at low accelerating voltages [173]. However, the K+IDS power supply is best for these experiments because it allows fine control of the heating current directed through the K+ glass emitter to prevent possible analyte thermal desorption and/or degradation while also providing circuitry for biasing the K” bead relative to the potential on the FAB target. The accelerating voltage on the HX-110 was reduced to 4 kV in order to prevent damage to the K+IDS power supply. In Figure 24, the input for the reference voltage power supply in the K+IDS power supply unit was connected to the 4 kV:t50 V repeller reference site #5 in the surge filter of the HX-110 in order to bias the K+ emitter and the heater circuitry equivalent to the accelerating voltage. In FAB on the HX-l 10, the repeller voltage, i50V of accelerating voltage, is normally applied to the FAB target. This potential is measured at reference site #10 in the surge filter, but the EI/CI repeller voltage at reference site #5 tracks the FAB repeller at #10 exactly and was used instead because it was remote from the FAB target. The heating current outputs from the K+IDS power supply unit were connected to the EI/CI filament heating current reference sites #3 Figure FABC Ion Source Housing of JEOL HX-l 10 #3 EI/CI Filament #4 EI/CI Filament #5 EI/CI Repeller #6 Source Chamber #7 EI/CI Trap #8 Thermocouple #9 Thermocouple #10 FD/FAB Repeller #11 Accelerating V FD/FAB Ion Source (4 kV) : (7) 7 Surge Filter (4 kV) 3 4 5 6 7 8 91011 f‘?OOOOOO Reference I Power Supply 0-20 V Heater Current Power Supply I Y3] 0-4 A 4—-—- A K+IDS Power Supply Unit 120VAC 10 kV Isolation Transformer A wooden crate and plexi-glass shield protect the operator since the K+IDS power supply is floated at 4 kV. Figure 24. Schematic diagram for the implementation of thermally-assisted K+IDS-by- FAB on the JEOL HX-l 10. 101 and #4 in the surge filter following removal of the existing connections to sites #3 and #4, so a current loop was isolated between the K+ emitter and the K+IDS power supply. Thus, when the accelerating voltage of mass spectrometer is turned on, the entire K+IDS power supply will operate at 4 kV. To protect the operator and the mass spectrometer from exposed high voltages, the K+IDS power supply was placed in a wooden crate, fitted with a plexi-glass shield. The K+IDS power supply unit was also plugged into a 10-kV isolation transformer to prevent the 4 kV from leaking to ground or the wall AC voltage. The K+IDS power supply was operated by a wooden rod and rubber hose though port holes in the plexi-glass. The K+IDS-by-FAB space-divided target without the K+ glass emitter attached was placed on the end of the FD/FAB direct insertion probe (DIP) and inserted into the FD/FAB ion source. A K‘t emitter bead, routinely prepared as in K+IDS (see Chapter five), was spot-weld mounted to the E1 filament leads in the E1 filament housing block and adjusted so the K+ emitter bead was centrally positioned in the front, right of the cathode lens element slit in the FD/FAB ion source. The two leads on the El filament block were connected to the #3 and #4 feed-throughs as is routinely done in El with Re filaments to complete the K+ emitter heating current circuit. In Figure 25a, with a heating current of 1.8 A, just above the on-set of K+ thermionic emission, 62 counts of K“ were detected. In Figure 25b, glycerol was applied to the K+IDS-by-FAB space-divided target and inserted into the ion source so the angled glycerol target was left of the K4" emitter 1 m away. At a lower heating current of 1.7 A, stronger K+ signal was detected because the analyte/matrix target surface helped focus K+ towards the mass analyzer. In Figure 25b, 0.1 counts of [G+K]+ were also detected which can arise if K+ adds to glycerol that was thermally desorbed or evaporated from the sample holder. In Figure 25c, the K+ emitter was held at 1.7 A while the FAB beam was activated which resulted in a thermally-assisted KflDS-by-FAB mass spectrum of glycerol. Here, the [G+K]+ was 20 times more abundant than in Figure 25b owing to more desorbed glycerol molecules and K+ ions in the selvedge. Also with this experimental set-up, a normal § 102 I 39 - . FAB Off . 1% 80: a) K+ Heating Current: 1.8 A L a I Bias Voltage: 0 V . t 1 Accelerating V: 4 kV i {, 60‘ K+ Counts: 62 - e ‘ . l. 40‘ : I ' t C : i 3 201 . i . 41 , yc'. ......--.' ...-.' 0 20 40 60 80 i9; 100 R 1 b) 39 FAB Off I I 80: K+ Heating Current: 1.70 A i a . Bias Voltage: 0.5 V , t : Accelerating V: 4 kV : it 60‘ K+ Counts: 98 - c ‘ b I : [G+K]+ I n 40‘ 131 - E. I . - - n. n n ,: g 20.: x 50.0 L { ‘ P y 0‘ ' I 1 1 1 . I i . . V I fl 1 v ' r 0 50 100 150 200 m/z R100: c) 185 FAB On I I + . 'f 80- 93 ”6"“) Same as b - [l 1 [G+H]+ . 1,. 60< ; e l : i401 e e ‘ > n ' [3G+H]+ . is 20'. 55 15 [5131?” 167 [20+K]+ 277 E I I 50 100 150 200 250 300 m/z Figure 25. Thermally-assisted K+IDS-by-FAB mass spectra of a) thermally-heated K+ glass emitter, FAB off, and no analyte on FD/FAB sample target; b) thermally-heated K+ glass emitter, FAB off, and glycerol on FD/FAB target; and c) same as b except FAB is on. 0P 065 Jen: TM: gene “'Orf‘ W1mg into h 103 K+IDS-by-FAB mass spectrum of glycerol could be obtained without heating the K+ emitter. When both FAB was activated and the K+ emitter was heated, the K+ ion current was approximately twice that obtained in keV-Xe bombardment of K+ glass alone. However, biological analytes were very difficult to analyze using this method because the analyte ion signals were transient similar to K+IDS because the heated K+ emitter radiatively heated the sample holder on the K+IDS by-FAB target to some degree and unfortunately induced rapid analyte thermal desorption from glycerol. In conclusion, by thermally heating the K+ emitter, more undesirable variables were introduced into the K+IDS—by-FAB experiment which reduced the effectiveness of KHDS-by-FAB for informative mechanistic studies for FAB. I n ti 1 ter 11' rim nt 'h I N SIMION computer modeling was pursued in an attempt to optimize the injection of all desorbed K+ ions by FAB into the selvedge region above the analyte/matrix target half because the "brute force" methods for increasing K+ adduct formation, such as thermally heating the 1K+ emitter or changing its composition, did not give the desired results. The K+ ion trajectories were apparently affected by divided target design and geometry, and the influence of the extraction fields generated by the electrostatic potentials applied to the ion source ion optics. For years, SIMION [174], a respected ion optics computer modeling program, has provided mass spectrometn'sts with the insight needed to understand how ions traverse specific ion optical transport systems, which have lens elements with applied potentials, for designing novel or modifying existing systems. This software package is well known in the mass spectrometry community and it is generally accepted that the theoretical models generated correspond quite well with real- world systems. Up until now, K+IDS-by-FAB targets were designed using basic intuition without stunning success, but the SIMION 4.0 software package [175] provided insight into how the various K+IDS-by-FAB targets interfaced with the FD/FAB ion source f0 101 p01 refi C3101 each 104 mounted in the JEOL HX-l 10. As a result, this additional insight assisted in the design of a functional K+IDS-by-FAB divided target design and realizing a required minor ion source modification which enabled K+ to be efficiently injected into the selvedge region and FAB mechanistic studies to proceed. The first step in SIMION modeling of KflDS-by-FAB targets was to input the scaled dimensions and electrostatic potentials of a two-dimensional "slice" of the target and ion source optical elements into a small array size using the "editor" of SIMION. In SIMION, the lens elements created in the editor are referred to as "electrodes". After data input, the small array (e.g., 50x75 or 3750 points), containing the target ion source electrodes and potentials, was doubled in size (i.e., 100x150 or 15,000 points), using the "double" command, to provide the model with sufficient resolution to insure "real world" accuracy and integrity. The upper limit for array sizes in SIMION are 16,000 points. The same set of ion source electrode dimensions were used in each SIMION model in Figures 26 and 27 and only the target electrode geometries changed. As a point of reference, the slits between the cathode (L1) and lst focus (L2) electrodes are 1 and 2 mm, respectively. Additionally, the same set of electrode voltages were used for each SIMION model in Figures 26 and 27 and these values were obtained by measuring the actual potentials on the ion source elements on the HX-l 10 mass spectrometer during a typical KflDS-by-FAB experiment. The JEOL FD/FAB ion source elements and measured potentials are: FAB target (repeller), 10050 V; cathode (Ll), 10010 V; lst focus (L2), 9000 V; 2nd focus (L3), 600 V; deflector (L4, not shown), 30 V; TIM (total ion monitor) plate (L5), 0 V. Following data input and array size "doubling", this 15,000 point two-dimensional array of ion source "electrodes" having fixed potential values is refined meaning a potential energy surface is calculated in which all non-fixed potential array points in the open spaces between the electrodes are assigned a respective calculated potential. In Figures 26 and 27, the equipotential lines C1-C6, identical in each model, trace a specific potential value across the potential energy surface revealing 105 dew—28?: =2 +M A0 93 :2 “Egg—5832a 8 9.365 .038 823—060an mica—shaman. was .835 Bv_>_?oommm 88.8-33 3 .695 83:62—55 3 .8qu m 88 > 28— > Sea u 90 S 3 >§an8 8 6 >22: "5 «U > 32: u no > eng— u «U 1199 > 6'2: 11.. #9 Nu > em:— 954 3:32—35 6 .... AQ > 88 > 38— 3 «out; no «D 335.8an 6 méééEE > 3...: / I. > 88 > 22: 11 . «A 1— Sneak. m0 $9635.00an 8 82.535 / / e0 _/ '4' .'I‘ 3135-....3 / >88 m.— m0 >§a > 22: 3 no no 5 my > 82: 3.55-8an 5.23:5. / > :33 .motoauoqmb =2 +v. 8 ES :2 xEmEBEmca Q. 9:305 898 coEz—Vooaam m 83 > 3.5— > 83 > 32: NJ 1— > Sea u 00 NA —‘— > cows .1. m0 > 32: u 9U > 3.x: u no > one: u «U > :33 .1. —D 3:: _a=:82__=_.fl 823:3an 31.—adhere. V///A . A Q Fatwa—53% 107 the contour of this surface. The great utility of SIMION for modeling ion optical systems is ions with initial KEs and directions can be placed on this calculated potential energy surface and their responses to it, based on their initial trajectories, can be observed which reveals the subtleties of the potential energy surface. Thus, ion trajectories can be manipulated by altering the potential energy surface which is derived from divided target and the ion source potentials, geometries and dimensions. For simplicity, all ion trajectories in the SIMION models in Figure 26 and 27 were calculated for ions originating on the target surfaces and having initial kinetic energies of 0 eV. Under real world conditions, all ions have an initial KEs and directions, however ions with zero initial KB are most sensitive to potential energy surface and will yield more information about it. Additionally, ion trajectories in or near the selvedge region are not accurate because SIMION does not take into account ion/molecule reactions and non-reactive collisions. The results extrapolated from these SIMION models were used to assist the KflDS-by-FAB target development process by providing insight into the general ion extraction trends from a variety of divided targets, and they should not be interpreted literally. Last, the fast atom beam is not shown in these SIMION models, but the beam would originate from above the Figures and strike the divided targets with an average angle of incidence of 70°. The first course of action with SIMION was to model each KflDS-by-FAB divided target interfaced to the FD/FAB ion source. For reference, the "normal" FAB experiment on the JEOL HX-l 10 is shown in Figure 26a in which the JEOL FD/FAB target is interfaced with the FD/FAB ion source under normal operating conditions. In Figure 26b, the KHDS-by-FAB wall-divided target is interfaced with the FD/FAB ion source unsuccessfully because ions extracted from either target half crash into L1 with the applied "center-extracting" ion optics potentials. Ion extraction from each target half in the KflDS-by-FAB dual-surface space-divided target was more successful as shown in Figure 26c. The purpose of a divided target is to inject [G into the selvedge region above 108 the analyte/matrix target. According to the SIMION models in Figures 26b and c and assuming the "top" divided target half is the K+ glass emitter surface, K+ ions would be extracted by the ion optics without ever passing though the "selvedge" region above the opposite target half. Experimentally, this observation probably accounts for the lack of K+ adduct ions formed in the gas phase observed with this target. Since the wall-divided target is notorious for condensed-phase contamination of the K+ glass emitter and the wall is the primary location of this contamination, the contamination-indicating ions [G+K]+ and [G-H-1-2K]+ may be successfully extracted from this region if they were formed near or on wall region and desorbed preionized. In Figures 26d and e, the analyte/matrix and 10‘ ions are extracted from their respective KflDS-by-FAB space- divided target halves, respectively, when the target was interfaced to the FD/FAB ion source. The KflDS-by-FAB space-divided target in Figure 26c indicates that 10‘ ions are injected into the presumed selvedge region above the analyte/matrix target which is consistent experimentally, because the definite gas-phase K+ adduct ion formation without condensed-phase contamination is observed with this target in comparison to the other divided targets. However, the weak K+ adduct ion formation typically observed using this divided target may be attributed in part to K+ ions traversing the presumed selvedge region with KEs greater than 2 eV, which are too high for K” adduct ion formation. K+ ions traversing the high pressure selvedge region should ideally have low KEs and long residence times which increases the probability K+ forming adducts with desorbed molecules through ion/molecule collisions. Additional SIMION modeling was performed in an attempt to lower the KB of the K+ ions in the presumed area of the selvedge while maintaining their favorable trajectories. Success was obtained by placing a thin lens element electrode, referred to as L0 between the K+IDS-by-FAB space-divided target and L1 electrodes as shown in Figure 27e. Since L0 has a center slit of less than 1 m, it further limits the extent of penetration by the extraction field, which is represented by the equipotential lines Cl-C4, 109 because these equipotential lines are now positioned between L0 and L1. Thus, the presence of L0 essentially creates a field—free region above the target area. Without the influence of the extraction field in the field-free region above the target area, the potential energy surface in this region has a much shallower depression in comparison to the deep depression formed without LO. Thus, K+ ions do not gain as much kinetic energy as rapidly during extraction with LG present. In Figure 28c, for the KHDS-by-FAB space- divided target with L0 present (corresponding to Figure 27e), the typical K+ ion desorbed from the K+ emitter and injected into the selvedge will gain 2 eV of KB within this field- free region, the upper limit for K+ adduct formation, in 4.5 yrs and will remain in the selvedge region for approximately 1.9 us as estimated by SIMION. Without LO present, for the same target (corresponding to Figure 26e), the typical 10 ion will gain 2 eV in 0.45 us and reside in the selvedge region for only 0.3 us. Thus, with the addition of L0, K+ ions are still injected into the presumed area of the selvedge, gain minimal KE in the field-free region, and essentially have much longer residence times, 6 to 10 times longer, in the selvedge, all which in theory should substantially enhance the efficiency of ion/molecule collisions for K+ adduct formation. In Figure 27a, the electrode L0 was successfully placed in the "normal" FAB experiment. It was important for L0 to be compatible with this configuration since all analytes are analyzed by FAB before K+IDS- by-FAB to determine the natural [M+Na]+ and [M+K]+ levels, to ensure [M+K]+ adduct ions in KHDS-by-FAB are real. In Figures 27b and c, L0 dramatically improved the extraction efficiencies from these targets in comparison to Figures 26b and c in theory, ' but these models still did not indicate that K+ ions would interact favorably with the selvedge region above the analyte/matrix target. Additionally, the wall-divided target in Figure 27b would still exhibit glycerol contamination of the K+ glass emitter regardless of L0 presence. In Figure 28a-c, the K+IDS-by-FAB space-divided target with or without L0, corresponding to Figures 26c and 27e, respectively, allows longer K+ residence times below 2 eV in the presumed area of the selvedge comparison to the FAB and dual-surface 110 E 20 la) . ' III/Z 39 W/O LO 3 15 ' - —-D—m/21000w/oLO h I E 10 / - , . mlzB9w/LO .3 5 _ _" / ——<>—m/leOOw/LO 0 I 0 o E ' 0' ...o° .- O00:00.oocooooooo00909000000009.0000...0000‘.‘°°°'.°--°. I : I I :4 0 1 2 3 Extraction Time (microseconds) E 20 b) u ' m/z 39 w/o LO ‘5 15 .' —D—m/21000w/0L0 h I 5 10 . _ . . 1:1/239me .g 5 l I —o—m/z1000w/Lo E O 0.35:: ...... 0.. ooooooooooooooooooooooooooooooooooooooooooo ‘ vo°‘°.r.. 1 I : O 1 2 3 4 Extraction Time (microseconds) m/z 39 w/o LO N O 6 V + m/z 1000 w/o LO u—a LI! O m/z 39 w/LO —a O ——O—-m/21000w/LO Kinetic Energy (eV) 012 3 45 67 89101112131415 Extraction Time (microseconds) Figure 28. Estimated extraction times for typical ions from the a) FAB target, b) dual- surface space-divided target, and c) KflDS-by-FAB space-divided target SIMION models, with and without L0 present, shown in Figures 26 and 27a, c, d, and e, respectively. The K+IDS-by-FAB space-divided target combined with L0 is best technique for injecting low energy K+ ions, with longer residence times, into the selvedge region. 111 space-divided targets with and without L0 present, corresponding to Figures 26a and c and 27a and c, respectively. Thus, the geometry of the KflDS-by-FAB space-divided target without L0 is better optimized for injecting K+ ions into the selvedge. The L0 focusing lens element was constructed from two stainless steel ribbon strips (0.125 x 0.005 in) precisely arranged side to side in the same plane less than 1 mm apart. Tungsten wire (.007 in) held the two lens halves together and was used in the connection between the L0 lens and the FAB repeller voltage source, which is positioned on the FD/FAB ion source near the DIP. The focusing lens was firmly mounted to the FD/FAB ion source using existing screws and ceramics. The installed focusing lens could accept any type of FD/FAB target, both stock and modified. In practice, the experimental results obtained with the K+IDS-by-FAB space- divided target in conjunction with L0, as modeled in Figure 27e, were significantly improved in comparison to without L0 present, which allowed the KHDS-by—FAB experiment to proceed as originally envisioned, that is, by efficient injection of 10‘ ions into the selvedge region. In Figure 20c, a KflDS-by-FAB mass spectrum of glycerol obtained with the K+IDS-by-FAB space-divided target and L0 installed is shown as proof that this new target/ion source modification combination does achieve the desired results. In this spectrum, K4" is injected into the selvedge region as shown by the peak representing K+ at m/z 39, to yield peaks representing the [G+K]+ and [26+K]+ adduct ions that have parallel abundances to their respective protonated species. No peak representing [G-H+2K]+ appears at m/z 169 which indicates the K4” emitter is not contaminated by glycerol and the K+ adduct ions observed are truly gas phase products. The similar peak intensity ratios of K+ to the K+ adduct ions of glycerol indicates there is extensive conversion of K+ into product ions in the high pressure selvedge region. I ' ' 'inrii rt Divided targets are notorious for providing ion source ion optical tuning problems 112 because ions are extracted from both right and left target halves instead of from a single central target surface. In Figure 29, the SIMION model of the dual-surface space-divided target and entire FD/FAB ion source is shown to illustrate this point. The potentials in this particular SIMION model have an ion optical "center tune" which means essentially, extraction of ions from either target half is equivalent and an inverted mirror-like image of the divided target surface is reproduced on L5, although some ions do traverse L5. By adjusting the potentials on L1, L2 and L3, so one lens element half is higher than the other, a "right" or "left" ion optical tune can be achieved and the ions generated from the right or the left target halves of this divided target. can be extracted, accelerated and focused out of the ion source, through L5, preferentially over the opposite target. While it is possible to extract ions from both target halves using a "center" ion optical tune, SIMION and experimental results suggest the ion source geometry and range of applied potentials will not readily allow the "dual ion beam" in Figure 29a to unify and exit L5 together, although sometimes ion source conditions will exist, such as cleanliness, which allow unification. Experimentally, even with a center tune, tuning problems from having a dual beam as in Figure 29a, were experienced for the wall-divided target, dual-surface space-divided target, and to a lesser degree with the KHDS-by-FAB space-divided targets. These tuning problems were predicted to occur by SIMION in Figures 26b-e because the ion beams were poorly defined. Experimental evidence of preferential extraction is shown in Figures 29b-e from a K+IDS-by-FAB study with a wall-divided target. In sequential order, ions were extracted from: (1) the K+ emitter target half in b, (2) the glycerol matrix target half in c, (3) both target halves in (1 without glycerol contamination of the K+ glass emitter which was a rare occurrence, and (4) both target halves in as (1 except the K+ emitter was contaminated which most often was the case' with this target. The addition of L0 to the K+IDS-by-FAB dual-surface divided target SIMION model in Figure 29f, having a "center tune", produces a well-defined ion beam, composed of ions extracted from both target halves, that exits L5. Moreover, the 113 \ \ \ Dual-Surface Space- I DividedTarget .1 - _. - - 10050 V F 10010 v 9000 v 600 v 0 v loo—39 100 93 [ZG+H] E 80‘ b) K+ 801C) [G+H]+ 185 E so . 4o 20 0o 93 [261-HF: 801d) [G+H]+ 185 : so : 4o ; 20 IO“ 57 75 tG+K1+ i 0.9241511 113111, “E OWTWS'O' ' '160"""15'O'm '2 m/ Dual-Surface Space- 1 DividedTarget ,1 1 l \. l . n u- j *— l at: . 10050 V C] 1 I— L3 L1 C5 L2 L5 10010V 9000V 600V 0V Figure 29. SIMION model of dual-surface space-divided target and full FD/FAB ion source a) shows K" and glycerol ions, extracted from each respective target half, leave the ion source in two separate beams, b) mass spectrum of K+ beam from K+ glass emitter target half, c) mass spectrum of the glycerol beam from glycerol target half, (I) "rare" mass spectrum of both K+ and glycerol ion beams, e) typical mass spectrum of both K"' and glycerol ion beams showing glycerol contamination of K+ glass emitter, and 0 same SIMION model as in a, except with LO installed, shows a single well-defined beam leaving the ion source which eliminates ion beam focusing problems. 114 combination of L0 with all K+IDS-by-FAB targets in the SIMION models shown in Figures 27 result in well-defined ion beams being accelerated. Experimentally, for a typical run using K+IDS-by-FAB space-divided target with LO installed, the K+ ions from the right target half, K+ adduct ions from the gas-phase, and regular FAB ions from the left target half all appear in a KHDS-by-FAB mass spectrum with a normal "center tune". In theory, the well-defined ion beams should increase the sensitivity for a typical analyses but unfortunately, the FAB beam is partially blocked by L0 resulting in a narrower portion of the divided target being irradiated. V ' i ' l t. Experiments were conducted to substantiate and optimize the use of L0 in the K+IDS-by-FAB experiment after it proved successful in (1) increasing K+ adduct ion formation of desorbed neutral molecules for a variety of analytes, and (2) correcting tuning problems for divided targets. In Figure 30, sequential KHDS-by-FAB experiments with glycerol show the [G+K]+ ion relative abundance was 30 counts when LO was installed and 11 counts when LO was removed in the next experiment which suggests that L0 actually does increase K+ adduct formation in the K+IDS-by-FAB experiment. In Figure 31, increased [G+K]+ adduct formation using glycerol was demonstrated when the K+IDS-by-FAB space-divided target was close to L0 rather than distant as indicated by the 4 and 0.7 count difference, respectively. Perhaps this experiment indicates a local high pressure region exists in small volume between target and lens which enhances adduct ion-forming collisions and/or termolecular stabilization. However, there were never any signs of condensation of vapor-phase glycerol on the LO lens element. In Figure 32, the K+ conversion efficiency of desorbed neutral molecules and the effect of L0 slit width on K+ adduct formation is demonstrated for sequential experiments with the analyte mixture thevetin A (MW 872) and B (MW 858). Consistent and identical amounts of sample and matrix were mixed on the K+IDS-by—FAB target for 115 100 ‘ ) 93 185 . R : a [G+H]+ I f 30. L01nstalled [26+H]+ _ a ; mlz 131,30Counts ; l ‘ . v 60- . e ‘ . 1 [G+Kl+ : p 40- 131 - e 75 ~ 3 20' K.» [ZG+K]+ Z i 1 39 45 223 [3gzm+ ; y 0* 29 167 _ .l . o 50 100 150 200 250 300 m/z - I 1004 b) 75 93 135 R 1 [G+H]+ [ZG+H]+ _ f 80- LORemoved ~ f 2 W2 131,11 Counts ; {I 604 57 ' e l [3G+H]+ . I . . n 40- 45 277 t . S ‘ . ‘ G+K] ' S 20.. [ . i . 1291;; 131 1.7 .1 [ZG+K]+ ~ y I , ; 223 2 9 I 0- ‘ * - o 50 100 150 200 250 300 m/z Figure 30. Sequential KfIDS-by-FAB mass spectra of glycerol using the K+IDS-by-FAB space-divided target a) with L0 installed initially and b) then with L0 removed which demonstrates the effectiveness of L0 for enhanced [M+K]+ adduct formation. 116 100 g j a) Divided target 185 1 80: away from L0 93 [ZG+H]+ é‘ . m/z 131, 0.7 counts [G+H]+ 1 I X 60: I ‘ [G W n ‘ + t8 40: 13 1 167 g . 75 AW 1 20* X5 g, 3 K+ 57 l O: 39 j ‘ F Lfi * .‘ v ‘v v ‘: :lw 1 v . v f 0 50 100 150 200 250 m/z 100 R 1 b) Divided target 185 ]e 80‘ close to L0 + [2G+H]+ ii : m/z 131, 4 counts [694?] ‘v i e 60‘ [G+K]+ }, 1 131 t 40‘ f1 3 167 'S 20" 75 X5 i 1 Id y ‘ 39 57 J 0+ ' " A: {if-l" ' ‘ ' 1 ' v '1 v ‘1 v "v A-‘ v u v # r r 0 50 100 150 200 250 m/z Figure 31. The KHDS-by-FAB mass spectra of glycerol from sequential experiments in which the KflDS-by-FAB space—divided target was positioned a) distant from, and b) then close to the LO lens element resulting in an increase in the intensity of the peak representing [G+K]+ in the latter experiment owing to the possible creation of a higher pressure region between the target and LO. 117 3. R . ‘f ‘ a) FAB a , 859 t . [Thevetin B+HJ+ 1 24 C 1 I t It! ‘ [Thevetin A+H]+ e 1: 873 n s 1 1 y 850 860 870 880 890 900 910 19112/0 2 1 1; j b) Narrow L0 slit width 897 a]! 1 859 K‘HDS-by-F AB [Thevetin B+K]+ E ' [Thevetin B+H1+ g [Thevetin A+K]"’ 911 [I] ‘ [Thevetin A+H]+ g 873 II S j 1. i 1 1 y 01 A 14‘ ‘LLA -1 A- a A 850fi ' 8'60 H 870' ' ' 880' ' - 8'90 ' ' 900 ' 'fi9‘16 ' ' 9'20 m/z 1 R j ) Wide L0 slit width 6 ‘ c 859 + ,1, 1 [Thevetin B+H]+ K ms'bY'FAB t I 897 . , + 2 1 [Thevetin A+Hl+ [Thevetin B+K] I I 873 [Thevetin A+K]+ n j ‘ 911 t . e d n 1 5 1 1 111 11 l 1 111 y0.1.11l111 1111111111 11....111111111111 11..ii..111 1111 850 ' ' 860' ”T8116 " 880' ' 8'90 ” 900 h 910' ' ' 230 2 Figure 32. The K+ cationization efficiency of desorbed neutral molecules and the effect of L0 slit width on K+ adduct formation is demonstrated in sequential experiments using thevetin A (MW 872) and B (MW 858). A comparison between the a) FAB mass spectrum, and the KflDS-by-FAB mass spectra b) with a narrow LO slit width (out of ion optical tune) and c) with a wide LO slit width reveals 30% of total desorbed neutral molecules are K+ cationized, if [M+H]+ and [M+K]+ are summed from b and c, averaged, and compared to [M+H]+ in 8. Slightly more K+ adducts are obtained from a narrow slit width. 118 these sequential experiments, and the spectra were normalized to the same level. In Figure 32b, at a narrow L0 slit width, the peaks representing [M+K]+ adduct ion at m/z 897 and 911 are more intense in comparison to the peaks representing the protonated species. However, in Figure 32¢, a wide LO slit width resulted in a reversal of the peak intensities from Figure 32b which suggests the K+ gas-phase ions have longer residence times, resulting in enhanced K+ adduct ion formation, when the LO slit width is very narrow. By summing the normalized peaks representing the [M+H]+ and the [M+K]+ ions at m/z 859 and 897, respectively, for thevetin in Figures 32b and c, taking an average, and comparing this average to the normalized peak representing [M+H]+ at m/z 859 in Figure 32a, ~30% of the desorbed neutral molecules or [M+H]+ ions are lost in the KflDS—by-FAB experiment. Of these remaining desorbed neutral molecules, only half undergo ion/molecule chemistry with K+ to form [M+K]+ adduct ions, as indicated in Figures 32b and c, which suggests H” and 10‘ may be in competition for analytes in the selvedge region under certain situations. Thevetin is a cardiac glycoside and the PA of cardiac glycosides were estimated to be 1ess than the PA of glycerol which means glycerol cannot readily protonated thevetin [176]. Instead, the less abundant fragment ions of glycerol, GP“, which are estimated to be 1% of the 1000 desorbed glycerol molecules, are believed to protonate thevetin A or B in this case. Thus, because the abundant [G-i-H]+ present in the selvedge is not an effective protonating agent and the 612+ species are as populated as K+, K+ has a better chance of forming adduct ions with thevetin as shown by roughly equal [M+K]+/[M+H]+ ratios. The K+IDS-by-FAB data presented in this thesis support this phenomenon because [M+K]+/[M+H]+ ratios are larger for steroids and oligosaccharides than for peptides and other nitrogen containing compounds. Next, the well-defined ion beams observed in the KflDS-by-FAB SIMION models with L0 present in Figures 27a-e are believed occur in practice because ion tuning is enhanced as explained above. Additionally, the resolution of the peaks representing [M+H]+ and [M+K]+ for a high molecular weight peptide such as kassinin (MW 1334) is 119 H-Asp-Val-Pro-Lys-Ser-Asp-Gln-Phe-Val-Gly-Leu-Met-NHZ R j ) [M+H]+ FAB f 1 a 1335 FD/FAB target i‘ ‘. 1 . V a e . I n t e n .S 1 t y “A. . . . LLLA - ..... - . 1320 1330 1340 1350 1360 1370 1380 m/z g j b) [M+H]+ K+Il)S-by-FAB 1 1 3 3 5 space-divided target a j with L0 present i i v u e . I - [M+K]+ n . 1373 t 1 1 e i {J 1 l I § I‘llalnllll‘h ‘lnnhhllll‘h‘llh511‘klt111111311“ 1111 1320 1330 1340 1350 1360 1370 113/80 2 R 1 e 1 C KHDS-by-FAB l j ) [NI-1H]+ space-divided target f‘ 3 1335 without L0 present 1. C I n g . [M+K]+ 2 3 1373 i «1 t 4 y W 1320 ' T1330 1 ' 1340 I ' 150 ' ' 1360 FT '13'70 ‘ f ‘1380 m/z Figure 33. a) A FAB mass spectrum, b) K+IDS-by-FAB mass spectrum, using with the KflDS-by-FAB space-divided target with L0 installed, and c) same as b, except L0 was not installed, were obtained for Kassinin (MW 1334) at a resolution of 1000. A comparison between b and c demonstrates that L0 enhances resolution of the peaks. 120 increased. In Figure 33b, for L0 installed in the ion source, baseline resolution is obtained for K+IDS-by-FAB of kassinin. In comparison, in Figure 33c, L0 was removed from the ion source which resulted in poorer resolution. Both of these spectra were obtained with the KflDS-by-FAB spaced divided target at a resolution of 1000 which is too low to adequately resolve the high mass kassinin peaks at m/z 1335 under normal circumstances. In Figure 34, as a final test, the L0 lens element was used in conjunction with a regular FD/FAB target in three successive experiments with the peptide methionine enkephalin (MW 573) in glycerol, as shown, which further suggested that L0 delays the extraction of ions responsible for protonation and a high pressure region exists between the target and L0. In Figure 34a, the presence of L0 allows the intensity of the protonated molecule at m/z 574 to exceed the intensity of the peak representing the F immonium ion at m/z 120, which always has a high intensity independent of [M+H]+. In a successive experiment shown in Figure 34c, the L0 lens element was removed and the peak representing the [M+H]+ ion dropped to "normal" intensity which resulted in a normal FAB mass spectrum of methionine enkephalin with the expected ion distribution. A successful K+IDS-by-FAB experiment with methionine enkephalin was performed midway between these two experiments, with L0 installed, to ensure the L0 lens element was installed and working correctly. In this KfiDS-by-FAB mass spectrum of methionine enkephalin, the intensity of the peak representing [M-i-K]+ ions is one-third of [M+H]+ which further supports the fact analytes with higher PAs than [G+H]+ are protonated more competitively than K+ cationization. The K+IDS-by-FAB experiment works best if the ion source is clean, L0 is clean, and the edges of L0 are finely sanded clean. These prerequisites enhance sensitivity, ion extraction and ion optical tuning. In addition to the simple planer L0 lens element modeled (i.e., — —) in SIMION, other L0 designs and geometries were attempted such as angled lens element with a center slit (i.e., / \), a lens element "wrapped" around the target (i.e., 1— __l), etc. In general, these more complex geometries usually resulted in greater chances of 121 1 IF FAB with L0 installed 574 g I a) 120 [Mi-HP ; 80: [G+H]+ a] t 1 93 136 ii 60- [26+H]+ :3 1 185 1‘ 40: 612 g1 1 75 b2 b3 1’2 1 201 (357 147 221278297 )3 a4 c 05 XS y 01 3O . : 354 397 442 529 0 100 200 300 400 500 600 mlz R1 1b) 120 KflDS-by-FAB with Loinstalled [M1H]+ ‘f 803 F 574 a . 136 1 I 01 v 60- e . I I 1‘ 40. e n : K+ [G+H]+ GF—28 b b3 1 20. 39 3 11471772221278 ’2 a4 1 1 75 297 )3 397 y 0 1.1 1 3§4 1 1 1 1 100 200 300 400 500 """ R1 1C) 93T 11,201? FAB withoutLOimtalled f 801 lG+H1+ [26+H]+ a '. i l 75 g 60: I j 57 136 1‘ 49. 45 “I e . n , all-rib ‘ + i 20'. 14761548152 b3 1’2 [M+H] X5 t . 177 278297 1’3 04 574 Y 0+ 221 354 397 0 100 200 300 400 500 600 m/z Figure 34. The utility of the LO lens element in front of the FAB target for delaying ion extraction and possibly providing a higher pressure region above the FAB target is demonstrated, in sequential scans, for the analysis of methionine enkephalin (MW 573) in glycerol. With L0 present, a) an intense peak representing [M+H]+ was observed in the initial FAB mass spectrum, b) a successful KflDS-by-FAB mass spectrum validated a, and last, c) L0 was removed resulting in a decrease in the intensity of [M+H]+. 122 misalignment. A fine mesh grid was placed between the target and L1 without success and actually resulted in an overall decrease in ion current. In an another experiment, metal CI volume was mounted in place of L0. It encased the KflDS-by-FAB target during FAB, but simultaneous alignment of the CI volume with the FAB beam inlet and with the slit of L1 was difficult to achieve plus an increase in K+ adduct ion formation was not observed even when installed correctly. E . E . . 53! r 1 After discussing and solving the analyte/glycerol contamination of the K+ glass emitter above, potassium contamination of the analyte/matrix liquid through the gas phase may also occur in theory but experiments and observations have indicated it does not occur. First, the K+IDS-by-FAB experiment is very sensitive to divided target geometry and positioning in relation to the center axis of the ion source and LO. Since the FD/FAB DIP, on which the KflDS-by-FAB target is mounted, is entered from the front, there is no view of the KflDS-by-FAB experiment and all positioning and tweaking has to be done before evacuating the ion source chamber as opposed to under vacuum with a see-through flange and a special DIP to adjust and optimize LO, etc. If significant amounts of potassium were sputtered into the liquid analyte/matrix during KflDS-by-FAB, then the target design and geometry, and the presence or absence of L0 would not be important factors in the experiment. Moreover, abundant K+ adduct ion formation would have always been observed! Next, a wire coated with potassium salts would emit more K atoms than ions upon heating even though the ionization energy of K is low. However, the work function of a heated K+ aluminosilicate glass emitter is higher than the heated wire which results in thermionic emission of approximately 90% ions‘ [52]. Atom emission or matrix retention accounts for missing 10%. When a K+ glass emitter is bombarded by keV-Xe particles, the K+ ion current is stable and continuous without evidence of charging. Thus, even without heating, K”r ions migrate through the 123 crystal lattice to the surface for desorption into the gas phase while the respective electrons migrate to the underlying Re wire or metal surface to maintain matrix charge neutrality. Clearly, if vapor-phase K or K+ contamination of the analyte/matrix occurred in each KflDS-by-FAB experiment, K+ adduct ions of the analyte would have been observed consistently from day one eliminating any need for target development. Last, the best method for checking the analyte/matrix for potassium contamination is through experimentation. A normal KflDS-by-FAB experiment was performed by: (l) exposing the space-divided target for 2 minutes to Xe bombardment with either a bare analyte/matrix sample surface or a glycerol matrix, (2) removing the DIP and detaching the K+ glass emitter bead from the target, and (3) reanalyzing the bare surface or liquid matrix for possible K+ contamination. For the K+IDS-by-FAB experiment, with a bare analyte/matrix surface target and with LO not installed, the bare target surface initially emitted 1.5 counts of K+ without a K“ emitter present when exposed to the FAB beam, and 18 counts of K+ with the K4“ emitter installed and bombarded. After 2 minutes of exposure to FAB, the K+ emitter was detached, and 1.7 counts of K+ were detected from the bare surface, a slight increase. With L0 installed, the same experiment resulted in a 0.5 count decrease in K+ detected, since 1 count of K+ was detected initially and only 0.5 counts of K+ were detected after a 2 minute exposure to FAB. Thus, these results suggest this particular experiment is highly dependent on the exact target area from which K+ is extracted by the ion optics and, for the most part, K+ does not readily adsorb to the analyte/matrix surface. With L0 present, the lower K+ KEs are more controlled than at high KEs resulting in more efficient extraction instead of adsorption to the analyte/matrix surface as indicated by the decrease in K“ counts. However, in Figure 35, the same type of quantifying experiment was performed but with the glycerol matrix applied to the KflDS-by-FAB space-divided target and LO installed in the ion source. The normal KHDS-by-FAB mass spectrum of glycerol is shown in Figure 35a and the peak representing [G+K]+ adduct ion at m/z 131 has a relative intensity of 6 counts. The 124 50 t? : a KflDS-by-FAB of glycerol, a I Rlofmlzl3lis6wunts i . [26+H] g I [G+H]+ I 301 93 n 2 a 20: 3 3 + . 1 [G+K] { 10: K-l- 75 131 [ZG+K]+ Y 1 39 57 J l 223 C . 2 319111;. . . .+.7. . -1 .‘24. :L . 0 50 100 150 200 250 ' m/z R 3: b FAB of residual glycerol from above experiment after Z the K+ emitter was removed from KflDS-by-FAB target, 1 1 g 2 ‘ Cr+ 56 [644-1]... [2G+H]+ I 1 Fe“ 93 1‘ I K+ [G+K]+ ,6, 11 39 131 f 1 [20+K]+ t J 147 223 1’ 01 239 1 O 50 100 150 200 250 m/z 20 R 3 C FAB of new glycerol on same target used in above experiment, Li: R] of m/z 131 is 0.09 counts 1 151 93 185 g , [2G+H]+ 1 I 10. n i 1% j 147 is 5, Cr+ 75 lxlsln-fllk—ns. 1 2 E; : 455I '57 I 04 4». , L.—.—-. :--, . .fi . ..... 0 50 100 150 200 250 m/z Figure 35. Sequential experiments to prove K+ forms adduct ions in the gas phase and K is not readily deposited into the upper layers of glycerol in KflDS-by-FAB. First, a) K+IDS- by-FAB mass spectrum of glycerol on KflDS-by-FAB space-divided target which yields 6 counts for [G+K]+, then b) the K+ glass emitter is detached from this target and the residual glycerol is reanalyzed by FAB to probe K+ contamination which yields 0.6 counts, and last c) the residual K+ in new glycerol placed on the same target in b is probed. 125 K+IDS-by-FAB target was removed from the mass spectrometer, the K“ emitter bead was detached from the target, and this same target with residual glycerol was reinserted into the mass spectrometer for analysis by FAB which resulted in the mass spectrum in Figure 35b. Here, the peak representing the [G+K]+ adduct ion has a relative intensity of 0.6 counts, 10 times less than the R1 of the peak representing [G-l-K]+ in Figure 35a. In Figure 35c, a new glycerol sample was analyzed showing glycerol contains no potassium nor does the target. The mass spectra in Figure 38 were normalized to the same level to make these comparisons valid. Thus, this experiment indicates approximately 10% of the potassium emitted reaches the matrix, which is not unexpected since the two target surfaces are close together, but 90% of K+ cationization of desorbed glycerol molecules occurs in the gas phase. Additional quantifying K+IDS-by-FAB experiments were conducted by noting the initial and final ion currents on the JEOL HX-l 10 oscilloscope. Here, the [M+K]+ adduct ions of glycerol and analytes never showed a time dependence by which the abundances would increase during the course of a run. From these experiments, it is concluded that only small amounts of potassium are deposited into the matrix and at least 90% of the observed K+ adduct ion formation in KflDS-by-FAB is a gas phase process. Thus, the major thrust of the K+IDS-by-FAB research was to develop an effective divided target with minimal ion source modification to make the technique simple, predictable, robust, and reproducible for a wide variety compounds. This approach ensured a portable platform to other mass spectrometers and a valid technique, with minimal experimental variables, for the investigation of the ion formation mechanisms in FAB as discussed below. 126 3. Mechanistic Studies of Ion Formation in FAB using KfiDS—by-FAB The real utility of K+IDS-by-FAB is detecting the specific types of neutral molecules (M) desorbed in the FAB experiment. An [M+K]+ adduct ion detected in the K+IDS-by-FAB mass spectrum indicates M was desorbed into the gas phase providing the K+ glass emitter and analyte/matrix target halves are not contaminated with each other. This capability will indicate whether intact neutral analyte molecules or neutral molecular fragments are desorbed in the FAB experiment, a feat not previously attempted by the mass spectrometry community and central to further understanding of the highly debated, not well understood ion formation mechanism(s) in FAB. For a detailed discussion of the types of neutral molecules desorbed in FAB for a variety of analytes, the reader is referred to the published KflDS-by-FAB paper in Appendix one. The KflDS-by-FAB concept for investigating the desorbed neutral molecules in a FAB experiment is briefly discussed for the analyte cholic acid (MW 408), a bile acid, dissolved in glycerol. Because this compound is acidic in nature, the [M+H]+ peak at m/z 409 is not very intense in the FAB spectrum as shown in Figure 36a, and it is better suited for negative-ion mode FAB mass spectrometry. The peaks at m/z 355 and m/z 373 represent the dominant even-electron (EE) three and two water-loss fragment ions, [M+H-nH20]+, of cholic acid, respectively, which is not unexpected for a molecule with four hydroxy sites as shown in the structure in Figure 36a. With K+IDS-by-FAB, the origin of the ions observed for cholic acid can be put in perspective mechanistically. In Figure 37, the possible mechanisms are shown for the formation of the [M+H- 3H20]+ fragment ion appearing as a peak at m/z 355 in the cholic acid FAB mass spectrum. In the first pathway, cholic acid molecules (M) are desorbed intact into the gas phase by FAB from the glycerol matrix. If M traverses the selvedge region, there is a high probability that it will be protonated in the protonating environment of the selvedge 127 a) , [M+H-3H201+ '2' 355 [M+H-2H20]+ a s .i>.’ E :12 350 ' ' ' '460' b) [M+H-3H O]+ g 355 2 § § [M+H-2H20]+ .33 D m 3 - f ' '500 1er C) g [Mm-314201“ FAB of 0.01 M 2 355 [M+H-2H201+ KCl in glycerol :3 373 '3 ..>. a; [M+H-H20]+ [M+H]+ lM+K1+ [M-H+2K]+ °‘ 3.21409 _. it: - 485 360 380‘ ' '400' ' '420' 'F440' ' '4607 'F4éo' ' 151/00 2 d); [M+H-3H20]+ [WK]; FAB of0.1 M g 355 |_”M+H-2H20]+ 447 KCl in glycerol 2.; ,373 .3 [M+H-H201+[M +Hl+ [M-H+2K]+ 0‘ 391 499 411 IL 4815 360 380 '400‘ ' '4‘20' ' ‘440 '460' ' ‘480‘ * '500 m/z e) g} FAB of glycerol/cholic acid [M-H+2K]+ g? contaminated K+ glass emitter 485 3. £1 33 “g: [M+K]+ 3.; 447 “I..3?.1-..§7.7_ 7360' ' '3807 F 400 ‘ '4‘20' ' 440* 460 480 15130 2 Figure 36. Cholic acid (MW 408), in glycerol, a) FAB mass spectrum, b) KflDS-by-FAB mass spectrum, c) FAB mass spectrum with 0.01 M KC1 salts, d) FAB mass spectrum with 0.1 M KCl salts, e) FAB mass spectrum of analyte/matrix contaminated K+ glass emitter. 128 [(M+H) - 3H20]+(m/z 355) V3H20 [(M-3H20)+H]+ [M+H]+ (m/z 355) T— (tn/z 409) " — — _ — “l I Protonating Environment 1 I. _______ <—K+ [M'3H201mw 354) M(MW408) Desorption Cholic acid/Glycerol matrix Figure 37. Two possible mechanisms for the formation of the [M+H-3H20]+ fragment ion in FAB of cholic acid (MW 408). to form the protonated molecule which appears at a peak at rn/z 409 in the mass spectrum. Since the average internal energy content in these particular [M+H]+ ions are high enough to induce the low activation energy water-loss fragmentation pathways, three waters are lost by three successive gas-phase unimolecular decomposition processes to yield the [(M+H)-3H20]+ fragment ion peak appearing at m/z 355 in the mass spectrum. The loss of H20 from hydroxy—containing analytes in CI is a prominent unimolecular decomposition pathway. A second, equally valid mechanism for forming the [M+H- 3H20]+ fragment ion involves M undergoing keV Xe beam-induced dissociation in the glycerol matrix to form the degradation product [M-3H20] (MW 354). A cholic acid molecule is solvated by glycerol molecules in the matrix by hydrogen bonding and the weaker van der Waals forces. For intact desorption to occur, all these non—covalent intermolecular bonds between polar functional groups need to be cleaved which is not an entirely favorable process when the matrix and analyte is extremely polar. Degradation 129 of M would not be unexpected because there is much vibrational and thermal energy in the matrix during keV-Xe bombardment which could induce the facile l, 2-elimination cleavages of the polar hydroxy-groups and neighboring hydrogens to form a less polar analyte molecule that would be easier to desorb. If [M-3H20] is desorbed into the gas phase, it would be likely protonated in the selvedge region and detected as [(M- 3H20)+H]+ fragment ion at m/z 355 also. Thus, if K+ is injected into the selvedge region by K+IDS-by-FAB, the types of analyte molecules desorbed, M or [M-3H20] can be identified from gas-phase K+ adduct formation. However, if the ions appearing in cholic acid mass spectrum are desorbed preionized, injection of K+ into the selvedge region will not yield any adduct ion formation since two ions of same polarity will not react. The K+IDS-by-FAB mass spectrum of cholic acid is shown in Figure 36b. One new peak is observed in the mass spectrum at m/z 447 representing the [M+K]+ adduct ion of cholic acid. There is no evidence of a peak at m/z 393 representing the [(M- 3H20)+K]+ adduct ion that would indicate [M-3H20] beam-induced neutral degradation products are desorbed. Thus, cholic acid molecules are Miami and protonated in the gas phase by glycerol CI since injection of K+ ions into the selvedge region yielded an [M+K]+ adduct ion of cholic acid. The [(M+H)-3H20]+ fragment ion is then formed by unimolecular decomposition of [Mi-H]+ in the gas phase following protonation of M in the selvedge. For all non-preionized analytes, only desorption of intact analyte molecules have been observed with KHDS-by-FAB. As shown in Figure 36a, the cholic acid [M+H]+ ions are quite unstable since the fragment ions at m/z 355 and 373 are very intense. In Figure 36b, the K+IDS-by-FAB mass spectrum of cholic acid demonstrates the great stability of the [M+K]+ adduct ion since there is not an [(M+K)-nH20]+ ion series and the intensity of this peak representing [M+K]+ is more intense than [M+H]+. The K+ ion forms primarily long range, electrostatic bonds with molecules because its large size inhibits strong interactions with analytes. The potassium affinities of most molecules are approximately ten times less 130 than corresponding proton affinities for the same M so the binding energies of molecules to the K+ ion are much less than the covalent-like proton binding energies [61]. Thus, the [M+K]+ ions are stable because the K+-M is usually the weakest bond in the [M+K]+ adduct ion so K+ cannot induce much if any fragmentation of M upon adduct formation because stronger bonds cannot be broken. Additionally, the ionization energy of K+ is half of most analytes so the charge remains on K+ which limits K+ to inducing chemistry only by polarization rather than with a charge. The higher intensity of the [M+K]+~ adduct ion peak in comparison with [M+H]+ is consistent with a molecule with a PA less than glycerol as explained above. Last, while it is expected that the PA and potassium affinity of the [M-3H20] will be less than M, K+ adduct formation with these less polar neutral degradation products is still expected to occur. Alkali metals will cationize less polar, small molecules such as monofunctional alcohols [60], unsaturated hydrocarbons [177], and hydrocarbons [178] owing to the polarizing ability of K+ on the molecule. As another follow up study to ensure potassium contamination of the analyte/matrix was not occurring, the molecular weight region of glycerol and several analytes such as cholic acid, the cardiac glycoside digoxin (MW 780), the oligosaccharide stachyose (MW 666), the peptide kassinin (MW 1334), and the peptide bradykinin (MW 1060) are compared in Figure 38 following the analysis of these substances by FAB, KflDS-by-FAB, and FAB of 0.01 M and 0.1 M KCl salts in glycerol. The appearance of the [M-H+2K]+ peak in a mass spectrum indicates potassium has contaminated the glycerol matrix surface. In addition, analyte/matrix was placed on the K+ glass emitter purposely to demonstrate extreme potassium contamination of the glycerol or glycerol contamination of the K+ glass emitter as occurred routinely when using the wall-divided target. For glycerol in Figure 38a, the intensities of the peaks representing the K+ ion. parallel that of the respective [G+K]+ adduct ion for each K+ cationization method. In the same Figure for K+IDS-by-FAB of glycerol, nearly the same amount of K+ is injected into the selvedge region as is from 0.1 M KCl salts in glycerol and as a result, ‘5'38 Glycerol E 70 = 60 50 40 30 a) It iv fl Rela 088 39 K+ 93 [G+H]+ 1 3 l 1 69 [G+K]+ [G-H+2K]+ Cholic Acid 447 N+Kl+ 409 [M+H]+ 485 [M-H+2K]+ Digoxin 9) 05888883888 Relative Intens ty_ 781 [M+H]+ 819 [M+K]+ 857 [M-H+2K]+ I FAB D K+IDS-by-FAB .0.01 M KCl 00.1 M KC1 K+ Glass Emitter é Relative Intensity 0588888388 667 705 743 [M+I-I]+ [M+K]+ [M-l-l+21(]+ 1m o 0 £38 Kassmln U1 '5 50 0 40 E. 30 .3 20 I a? 1335 1373 141 1 [M+H]+ [M+K]+ [M-H+2K]+ 3388 »_l Bradykinin '5 80 § 70 5 60 50 o 40 E 30 2 20 a 18 1060 1098 1 136 [M+H]+ [M+K]+ [M-H+2K]+ Figure 38. A comparison between the molecular weight region peak intensities for the analytes: a) glycerol, b) cholic acid, c) digoxin, d) stachyose, e) kassinin, and f) bradykinin dissolved in glycerol and analyzed by FAB, by KHDS-by-FAB, by FAB in 0.01 M KC] also, by FAB in 0.1 M KC] also, and on the surface of the K+ glass emitter by FAB. 132 both methods produce more intense peaks representing [G-l-K]+ adduct ions than 0.01 M KC] salts in glycerol. The [G-H-l-ZK]+ ion only appeared in the spectrum obtained from 0.1 M KC] salts in glycerol and not KflDS-by-FAB which indicates the [G+K]+ formed in K+IDS-by-FAB is a gas phase process. The bar graph for cholic acid, a surface-active analyte in glycerol, in Figure 38b, is more sensitive to potassium surface contamination since 0.01 M KC1 salts resulted in the appearance of [M-H+2K]+. In the same Figure, K+IDS-by-FAB of cholic acid resulted in a more intense [M+K]+ peak and no [M- H+2K]+ peak which indicates [M+K]+ is formed in the gas-phase. Oddly, stachyose, in Figure 3.22d, a non-surface-active analyte, meaning it dissolves well in glycerol, also has a similar RI distribution to cholic acid above because the [M—H+2K]+ peak is observed for 0.01 KC] salts is high, too. However, in contrast, the [M+K]"’/[M-l-H]+ adduct ion ratio is higher for stachyose than in the mass spectrum of cholic acid. In Figure 38c, K+ cationization of digoxin, another non-surface-active analyte, by KflDS-by-FAB, KCl salts, and the K+ glass emitter, are very similar at both low (e.g., K'l'IDS-by-FAB and 0.01 M KC1 salts) and high (e.g., 0.1 M KC1 salts and on the K+ glass emitter) potassium concentrations. The peaks representing the [M-H+2K]+ ion are very low in comparison to the [M+K]+ for all the methods which indicates digoxin does not readily coordinate multiple K+ ions. Thus, the peaks representing [M-l-K]+ ions formed in KHDS-by-FAB for these three analytes are more intense or equal in comparison to the respective [M+K]+ formed from 0.01 M KC] in glycerol and no [M-H+2K]+ formation is observed in KflDS-by-FAB, which highly suggests a gas-phase K+ cationization process in K+IDS- by-FAB. Each of these analytes have low PAs in comparison to glycerol and perhaps the [M+K]+ peaks are more intense than the [M+H]+ because glycerol cannot protonate these analytes but gas-phase K+ can. In Figures 38e, the peptide kassinin shows the peak representing the [M+K]+ ion formed in K+IDS-by-FAB is more intense than 0.01 M KC1 salts in glycerol, and in both cases, there was no [M-H+2K]+ formation. However, another peptide, bradykinin, in Figure 38f, peaks representing [M+K]'*' and [M-H+2K]+ 133 are only formed from 0.1 M KC] in glycerol or from contamination of the K+ bead, but not K+IDS-by-FAB or the 0.01M KC] solution. The lack of a [M+K]+ ion from K+IDS- by-FAB for bradykinin may suggest that neutral bradykinin molecules are not desorbed in great quantities. The behavior of these two peptides in the KflDS-by-FAB experiment will be discussed in more detail below. Thus, for all these analytes, [M-l-K]+ ions formed by KflDS-by-FAB undoubtedly appear to be different than the [M+K]+ ions formed by the addition of small amounts of KC] to the glycerol matrix which further suggests in K+IDS-by-FAB, K+ forms adduct ions with desorbed molecules in the gas-phase and potassium does not contaminate the glycerol matrix. Because glycerol CI is responsible for the protonation of desorbed neutral molecules in the selvedge region, limiting the amount of glycerol on the target would reduce the number of glycerol-related species that protonate digoxin which would ultimately reduce the abundance of the digoxin [M+H]+ ion and the fragment ions derived from it. In a KHDS-by-FAB experiment, 5 ug of digoxin (MW 780) was added to and mixed with a very minimal amount glycerol on the target. As a result, a low intensity peak representing [M+H]+ and no related fragment ions were observed as shown in Figure 39. However, the peak representing the [M+K]+ adduct ion was more intense than the [M+H]+ because K+ was the dominant reagent ion. No fragment ions related to the [M+K]+ ion were observed owing to its great stability. In FAB, either intact analyte molecules, preionized intact analyte molecules, or both are desorbed into the gas-phase, and K+IDS-by-FAB can distinguish between these two analyte forms since K+ will not interact with other positively-charged analytes. For example, in Figure 40a, the analysis of the salt, benzyltriethylammonium chloride (MW 227), in glycerol by FAB shows an intense peak at m/z 192 indicative of the benzyltriethylammonium cation. If the benzyltriethylammonium cation was desorbed directly as a neutral molecule [R4N-H] or a R4N+Cl' ion pair from the glycerol, then the K+IDS-by-FAB mass spectrum of this same salt in Figure 40b would display a peak 134 .86on .556 Sec coinage saunas—swat 0: m_ 205 van +213: 9 gangs—8 E 8:25 22: fl :2 8:23 +573: 05 wet—.8058 mama 05 85m 83 32V Exowfi «o 8.58% 39: m> DJ] > [b2, b3, yz] > [02, 04, b4, cz, C3, C4, y3, y4]; b1, and C] ions are absent. In contrast, the sequence ions account for 92% of the TIC while the neutral loss ions account for 8% in the LSIMS spectra of the simpler pentapeptide, pentaglycine, (Figure 47a). The lack of side chains eliminates the possibility of side chain ions and side chain loss ions and reduces the possibility of immonium and internal ions because abundant immonium and internal ion formation usually require a large side chain. Here, 157 W” a a‘ a. a 304 R I: a) [M+H]+ a 80.H-GGGGG-0H ! 6 keV Cs+ LSIMS 'v - s Man-1x Ions/Cs" Adducts e 60* 1 n t 40- e b [03+H1+ [cud-IF [a tart, 2 “I 1125 10 c3 c i 20‘ 30 . a; c 133 b31171“ 147 l 87 172.190 y ' ”5. ‘ i. t 202 2291259 288 . “Ill II J ill-I1d Ill I .1 1 1. L 50 100 150 200 360 1311/0 Z "I R b H-GGGGG-OH b4 N+élofi 1c Metastable Ions (MI) 2 29 f 100%T 1 v 1* ’3 l 286 {1 b2 e 115 n m ”3 i x5 0 yz "2 14 I ‘ c 133 ‘ 47 y K :4 [as-HIP ““3 h 1 . 1 259 1 U 50 160 150 260 250 31012 304 R C H-GGGGG-OH 229 [M+H]+ e Single Collision. High Energy CAD Product ’3 r1, 4 ; Ions and Metastable Ions (Ml/CAD) 1 )0 E 100% I v 1‘ -H20 e 286 1 n l C 2 w i 247 ‘ lu+fll+ a + + -NH_ y 202 1 c4 [2559"] ii 11 1 _J_ .11 J 200 250 :30 Z R 0.4 - e E d H-GGGGG-OH ’3) 1 : "True" CAD Product lom (Ml spectrum 1 a E subtracted from MIICAD spectrum) 1 0.3g 2 D) V _ e : 115 ’2 : 133 D} l 0.25 172 n 2 b4 1 g 229 .1120 c : ’4 286 n 0 IE «2 + MW 247 _S : a) y 87[a2+H]* [rfgm 202 4 1 E 30 1:17; a/ c u [a5+fl]* .an ; = J \1 c2 3 219 0 _ 6 5'0 160 150 260 2E0 360 mlz Figure 47. a) LSIMS b) MI, c) Ml/CAD, and (1) "true" CAD mass spectra for pentaglycine (MW 303). See text for description of spectra and peptide sequence ion nomenclature 158 the sequence ion abundances for pentaglycine are ordered [b2, )7] > [a 1, b3, )3] > [a2, b4, c2, C3, C4, y1, y4]; b 1, and C} ions are absent. In addition, the [an+H]+' ion series appears in the LSIMS spectra of both methionine enkephalin and pentaglycine. mm Metastable protonated molecules have initial internal energies which result in sufficiently low decomposition rates (occurring between 1 and 10 microseconds after ionization) such that fragmentation occurs after acceleration. For the MI mass spectrum of methionine enkephalin (Figure 46b), the sequence ions (only an, bn, C", and yn) comprise 72% of the TIC, and the internal, [b4+OH+H]+, side chain loss, and neutral loss ions comprise the remaining 28% of the TIC. The sequence ion abundances for metastable methionine enkephalin [M+H]+ ions are ordered [a4, b4] > [b 3, y;, y 3, y4] > [a], b2, C2, C3, C4, y1, y4]; the b1, and C] ions are absent. For pentaglycine in Figure 47b, the sequence ions comprise 81% of the TIC while the remaining 19% is from neutral loss ions. The sequence ion abundances for metastable methionine enkephalin [M+H]+ ions are ordered [b4, y3] > [b 3, y2, y4] > [b2, C2, C3, C4]. In both MI spectra, the immonium, side chain, and [an+H]+' ions are absent which indicates these are higher-energy product ions. W In single-collision, high-energy CAD, a single energetic collision between a helium atom and an [M+H]+ ion adds additional internal energy to the ion which results in unimolecular decomposition of the energized ion into CAD product ions. These [M+H]+ ions subjected to CAD are those that have not undergone prompt dissociation in the ion source, and thus, they are somehow different from the [M+H]+ ions that give the fragment ions observed in the LSIMS and MI spectra. The differences could be structural and/or energetic. In the "true" CAD mass spectrum for methionine enkephalin (Figure 159 46d), the distribution of sequence product ions is nearly identical to the MI spectrum in Figure 46b, while for pentaglycine in Figure 47d, the distribution of sequence product ions more resembles the LSIMS spectrum in Figure 47a than the MI spectrum in Figure 47b. For both methionine enkephalin and pentaglycine, the appearance of immonium, internal immonium, side chain, and [an+H]+' ions in the CAD spectra suggests that some [M+H]+ ions are activated to relatively high internal energies. A Model for the LSIMS Experiment. In LSIMS, both neutral peptide molecules (e.g., methionine enkephalin) and preionized peptides (peptides that contain very basic side chains such as R, K, H appear to be protonated in the condense phase), are desorbed intact into the gas phase by keV- Cs+ particle bombardment [124]. The neutral gas-phase peptide molecules are subsequently protonated in the high-pressure selvedge region by glycerol chemical ionization (CI) [85, 122], while the preionized peptides require no further chemistry for mass spectrometric detection. Both the neutral and preionized species contain an undetermined but sufficient amount of internal energy from the desorption event to undergo unimolecular fragmentation to some extent. Attempts to estimate the internal energy content of desorbed intact neutral molecules and preformed ions have been made previously. Hoogerbrugge et al. [126] determined the internal temperature of sputtered glycerol molecules to be 190 K which corresponds to an internal energy of 0.044 eV assuming only 15% of the total degrees of freedom are active. Derwa et al. [127] and Williams et al. [128] have estimated, using benzylpyridinium salts in glycerol, that the internal energy content of preformed secondary ions in FAB and LSIMS is between 0 and 4 eV; this energy distribution is most populated between 0 and 2 eV, with a high energy tail that extends out to 4 eV. Derwa et al. [127] used Hoogerbrugge's internal temperature value of 190 K to calculate an internal energy of 1.13 eV for desorbed benzylpyridinium ions assuming all 69 degrees of freedom are active. Using a_ i 160 Hoogerbrugge's internal temperature for desorbed neutral molecules, methionine enkephalin and pentaglycine will have approximate average internal energies of 1.8 eV and 0.91 eV from desorption, respectively, assuming half of the degrees of freedom are active. If protonated glycerol is the CI reagent ion in the selvedge region that protonates the analyte molecules, gas-phase peptide molecules may gain an additional 23 kcal/mol (1 eV) of internal energy, assuming the proton affinity (PA) of glycerol is 209 kcal/mol [118], and the experimentally determined PA of pentaglycine is 232 kcal/mol [190]. Following ionization of the peptide molecules by protonation or desorption of preionized peptides, the greater amount of internal energy an [M+H]+ ion possesses, the greater the chance it will undergo unimolecular decomposition to form fragment ions on the mass spectrometric time scale [38]. The different types of peptide fragment ions that appear in a mass spectrum at various m/z values are determined by unimolecular fragmentation pathways with favorable reaction enthalpies overall, and by unique structural features of the protonated molecule, which include the type of side chains, secondary interactions with groups that are many skeletal atoms away, and hydrogen bonding in the vicinity of the protonation site. The RIs of these fragment ion peaks are presumably determined by the distribution of initial internal energies in the protonated molecules in addition to the transition state structure. The amount of initial internal energy a protonated molecule possesses in excess of the activation energy for the fragmentation pathway determines the unimolecular decomposition rate constant. The highly excited [M+I-I]+ ions quickly fragment in the ion source while the "cooler" [M+H]+ ions either undergo slow, low-energy unimolecular decomposition to form metastable fragment ions after acceleration or are detected intact. In single-collision, high-energy CAD, the low energy [M+H]+ ions are energized by single collisions with. helium atoms to promote fast unimolecular fragmentation. The maximum internal energy, expressed as center-of—mass collision energy, available to impart into either pentaglycine or methionine enkephalin from a collision with helium is 130 eV and 69 eV, 161 respectively, at lO-keV lab energy. However, in a high-energy collision, the average energy deposited into the [M+H]"' ion is only between 1 and 3 eV [135] and, as a result, both low- and high-energy fragmentation pathways will be accessible similar to LSIMS. However, single collision CAD is inefficient overall because to a first approximation, 95% or more of the [M+H]+ ions remain intact while the remaining 5% or less fragment. Protonated Peptide Model. The first step in correlating spectra with analer structure is defining the initially— formed protonated molecule. An [M+H]+ ion containing the basic residues arginine (R), lysine (K) and histidine (H) are expected to be exclusively protonated on these side chains because they have basic sites with gas-phase PAs greater than 230 kcal/mol [49]. The peptides methionine enkephalin and pentaglycine do not contain these basic residues; they have a variety of likely protonation sites available. Single and multiple proton/heteroatom interaction sites on methionine enkephalin, and their estimated PA values, are illustrated in Figure 48. The PA values were estimated by matching each different heteroatom functional group(s) in methionine enkephalin with a small molecule of known PA that contains the exact arrangement of functional group(s) [40, 176]. From Figure 48, the most probable protonation sites on methionine enkephalin and pentaglycine are the amide functional groups along the peptide backbone represented by proton interaction site 2 since the PAs of these sites are highest, extending from 217 to 224 kcal/mol depending on the model compound and side chain presence [40, 191]. In Figure 48, protonation of the N-terminal amine nitrogen may occur in three ways: (1) the proton may be attached solely to N at site 3 which is estimated to have a PA of 217 kcal/mol from a protonated primary amine [40]; (2) the amine proton may undergo a possible hydrogen bond interaction with the carbonyl oxygen, shown by site 5, which is estimated to have a PA of 212 kcal/mol from the protonated structure of glycine [40, 192, 193]; or (3) the proton may be further stabilized by a secondary hydrogen bond 162 1‘1: ..osobw .«o .EV gunned—:0 05:250.: co 8% £93 got? .8 m “5 § 5“ U 1 2 3 4 5 6 7 8 9 10 1 1 l 2 1 3 Skeletal Bond Position Figure 49. Bar graphs showing the percent of skeletal bond cleavage products versus skeletal bond positions from a) LSIMS, b) MI, c) "true", single collision CAD mass spectra of the peptide pentaglycine. Structure and bond positions are shown at top. OH s: CH3 ‘7': l it I a; .. in II: an, CH2 0 0 CH2 38 CH 2% CH2 NH (Ii ('21-! OH 0- + 5?: HZN/ \(f/ \C£2\N§ \(lf \Cfi \Nfi \(E/ H g a O O ('21-!2 O O 2 © 18 '2 g 16 1 . _ g = 14 a) LSIMS ofH-YGGFM-OH . l CTerrmnal _ — 12 _ g3 10 I NTerminal _ a o 3 353;" 6 “5 a 4 s6 2 0 1 2 3 4 5 6 7 8 9 10 11 12 13 Skeletal Bond Position NM CA 1)) MI ofH-YGGFM-OH I C Terminal ... O\ I N Terminal % of Skeletal Bond Cleavage Products ES 8 4 o 12345678910111213 20 Skeletal Bond Position fig 18 g :1 12 C) Single Collision CAD 33 12 ofH-YGGFM-OH acremnai 2 10 i go 8 EDNrerminal w 6 h- > cg 421 9 o P o 1 2 3 4 5 6 7 8 9 10 11 12 13 Skeletal Bond Position Figure 50. Bar graphs showing the percent of skeletal bond cleavage products versus skeletal bond positions from a) LSIMS, b) MI, c) "true", single collision CAD mass spectra of the peptide methionine enkephalin. Structure and bond positions shown at top. 173 For the model peptides, pentaglycine and methionine enkephalin, all the R18 of sequence, irmnonium, internal, side chain ion, side chain loss, neutral loss, [a3+H]+‘, and [b4+0H+H]+ fragment ions were normalized to 100% so each sequence ion skeletal bond cleavage can be evaluated in light of all unimolecular decomposition processes. Sequence ion formation involves the cleavage of one of the three skeletal bonds, Ca- C(O), C(O)-N, or N-Ca. In Figures 49 and 50, the percentages of the N- and C-terminal skeletal bond cleavage products vs. skeletal bond position (see peptide structures for skeletal bond positions) were plotted such that the intensity of a bar represents how often a skeletal bond was cleaved. For example, in Figure 49a, for LSIMS of pentaglycine, the cleavage of skeletal bond 1, a Ca-C(O) bond, occurs 9% of the time out of the total fragmentation processes, and results exclusively in the formation of an N-terrninal a1 fragment ion. In the same Figure, the amide bond at position 5 is cleaved 23% of the time which results in the formation of both an N-terminal b2 ion 14% of the time and a C- terminal y 3 sequence ion 9% of the time. W In Figure 49a, the skeletal bond cleavage products distribution for LSIMS of pentaglycine is shown. The N-protonation of amide bonds 2, 5, 8, and 11 leads to an overall RI fragmentation distribution of 13%:40%:30%:8%, respectively, if the skeletal bond cleavage products, representing the an, b", C", and yn sequence ions, are summed. It is immediately clear that peptide fragmentation is not as simple as charge-initiated chemistry from N-protonation of the amide group since each protonated amide group would be expected to yield sequence ions in the same ratios with similar RIs. The cleavage of skeletal bonds 2, 5, 8, and 11 results in the formation of C- terminal yn and N-terminal bn 31119.11 sequence ions from each bond. The RIs of peaks representing y 3 and y2 sequence ions, formed from cleavage of skeletal bonds 5 and 8, respectively, are much greater than the RIs of peaks representing the y4 and y 1 sequence 174 ions from cleavage of skeletal bonds 2 and 11, respectively. If the 1, 2-elimination mechanism in Scheme VI was operative, all y” peaks would be expected to have similar RIs because the a-carbon source of a hydrogen atom is local to the N -protonation site for each amide. The yn mechanism in Scheme VII better accounts for the R15 of y" peaks. Either the N-terminal amine nitrogen, or an amide nitrogen, in protonated or native form, can interact through favorable secondary interactions with either of the two central amide nitrogens, at skeletal bond positions 5 and 8, and induce yn ion formation. The formation of y4 ions is not as favorable as y 3 and yz ions because a strained three-membered cyclic neutral product is formed in the former process. The formation of an ions and b" ions in Schemes I and II are the result of simple cleavages. The RI distribution for these ions should be similar at each amide bond because the chemistry is local to the protonation site, but this is not the case. Cleavage of amide skeletal bond 5 yields more b2 ions than cleavage of skeletal bonds 8 and 11 (b3 and b4 ions, respectively). The a2 ion, from cleavage of skeletal bond 4, is the only an ion formed as in Scheme I since [a3+H]+' and [a4+H]+' ion formation account solely for cleavage of skeletal bonds 7 and 10. Using Scheme 11, the a1 ion formed by cleavage of skeletal bond 1 is really a secondary product from the absent b1 ion formed by cleavage of skeletal bond 2. Perhaps the combined RIs of peaks representing an and b" ions show the effects of competitive fragmentation from yn ion formation, involving secondary interactions, at a particular protonated amide N site. Also, in some cases, the thermodynamic stabilities of the products may determine which process is dominant. Cleavage of skeletal bonds 6, 9, and 12 results in Cu ion formation, which involves a less favorable C-terminal secondary interaction with a protonated amide nitrogen. The C2 and C3 ions are more favored than C4 because, again, more stable six-membered cyclic neutral products can be formed as opposed to a three-membered ring for C4. The cyclo-bz ion in Scheme VIII may in part account for the large R1 of the b2 peak at skeletal bond position 5 since a stable six-membered cyclic b2 ion is formed. No xn or 2,, ions are 175 observed for pentaglycine. The 1, 2-elimination mechanism for x" ion formation does not occur to any extent which further confirms the use of alternative mechanisms. HIMSELF M M l° . E! H. In the pentapeptide methionine enkephalin, the side chains influence the charge- site induced chemistry and secondary interactions in contrast to pentaglycine, probably through stabilization effects. In Figure 50a, the normalized LSIMS skeletal bond cleavage product distribution for methionine enkephalin is shown. For methionine enkephalin, most fragmentation originates from cleavage of each N-protonated amide bond similar to that for pentaglycine described above. The RI distribution for sequence ion fragmentation, summed at each amide bond 2, 5, 8, and 11, is 18%:3%:3%:4%, respectively. The RI distributions resulting from the cleavage of amide bonds 5 and 8 in methionine enkephalin are nearly identical to the same bond cleavages in pentaglycine. However, the presence of the tyrosine and methionine side chains in methionine enkephalin has dramatically enhanced the R15 of peaks representing 61] and y 1 fragment ions from cleavage amide bonds 2 and 11, respectively. The cleavage of skeletal bond 7 represents [a3-l-H]+‘ solely whereas cleavage of skeletal bond 10 yields the a4 ion. Estimation of Reaction Enthalpies for the Mechanisms. For pentaglycine in its linear form, the amide nitrogens are the most probable protonation sites. Initially, each amide nitrogen may be protonated in an even distribution, but secondary interactions may redistribute the protons (protons may "hop" throughout the molecule). The fragmentation mechanisms presented assume that the amide nitrogen protonation initiates charge-induced chemistry, and hydrogens shift according to previously reported deuterium labeling studies. If simple bond cleavages such as homolytic, l, 2-elimination or inductive cleavages were prominent in pentaglycine without any side chain influences, the RI distribution of skeletal bond 176 cleavage products originating from each protonated site would be expected to be very similar. Such a uniform RI distribution is not observed! Instead, the non-uniform RI distributions suggest, at each protonated amide N, secondary interactions are in competition with simple cleavages. As a next step, the reaction enthalpies for each mechanism in Schemes I-X were estimated to gain more insight into why one particular fragmentation process dominates over the other possible pathways at a protonation site and/or why fragmentation from one, particular protonated site dominates over that of other sites in these two model pentapeptides, with and without side chains to influence fragmentation. Since the heats of formation are not known for protonated methionine enkephalin or pentaglycine and their ionic and neutral fragments, the exact reaction enthalpies cannot be determined. However, for small analogous neutral molecules, radicals, and ions, the heats of formation, PAs, and ionization energies (IE) are available [40, 191, 210]. The reaction enthalpies of peptide fragmentation mechanisms were estimated by matching the chemical structure of smaller species of known thermochemistry with the chemical structure of the protonated site and limited surrounding area on the peptide. The outer regions extending away from the protonation/reaction site were ignored. The same process was done for the bond cleavage and formation sites in the ionic and neutral fragments. See Appendix Two for data and methods used to calculate the reaction enthalpies for sequence ion fragmentation, and Figures 51 and 52 for a graphical representation of these estimated energies. i 1e Homolytic bond cleavage requires the most energy because it results in the. formation of two radicals. The reaction enthalpies for homolytic bond cleavages of the peptide skeletal bonds, Ca-C(O), C(O)-N, and N-Ca were estimated to be 87, 98, and 103 kcal/mol, respectively. Peaks representing ions from homolytic cleavages of C(O)-N and 177 .38on E mowgao—o 252 .8283 ovum—on .8 82.35:“. 5:82 Bsfizmm Am Eswfi :5 .22 .5 = 2.2.... 2% on .22 5-222 m 62.5.. 22... ES .22 .2222 m 62.5.. 2.6 :5 .22 .5 2 62:2 2% ES .22 .2222 a 62:2 .25 ca .22 :2 = 2.22 2.5 on .22 ..2 m 2...... 22m 5 .22 :2 Aomwdmmeu E on .22 5.222 m 23.. 2.5 .22 to . ‘Ill a a a . :ogmmeo 2-2 a .22 :2 onflmzuo e 2.22 :2on £5 6 a. a .22 ..e aged E .22 :2 22.69% E .22 .5 £6 32 ml 6226.2.J coemcoaoi .8582 e a: .22 E soummz bN $ 2 9 .22 .5 9:2qu5 o c> .22 .5 82:2 an e a: .22 :2 unmwaovu 3%..“ 5. ex .22 2+ :2 8V .8 mm 6 N 2 mango N 8. 6&9: finial—mam 178 .22 +E+Sz ...—Egon 3:88 ..8 £2.58.— Ev 3335:... 5:82 335:3 :05 was $9553 52.0.3888 EBB—oEED .Nm v.53”— m 0 =2888£ .965on + |. =_U\ max 5% O .m 530825 gap—ovum m=V 5 ”DID I 35.50.70 .. 2m 2m: ...2 \\o WU\$ + \(z “Mylo WW _ W...“ M / Immly . O ... O OImUdIZIIUKI A: «m \\ _+ :9 __ -o .... o .. IIIIIIIIIIIIIIIIIIIIIIIIIIIIIII .mlilllllul I: ...m.” \\0 ...”— om + m z Emu/m I) S + a: \ w \\ W2 o ..m o ..z OIIII. + :U/ 2:. 4.. ..~__ m \(U U N22 20 U N: \ / + Wu 3:2 :02 /U\ mix Wu». 179 N-Ca bonds are not observed in LSIMS, MI, or CAD mass spectra. However, homolytic cleavage of Ca-C(O) is observed which results in the low abundance [an+H]+’ ion series in LSIMS and single collision CAD spectra, but not in MI spectra—probably because Ca- C(O) is the weakest skeletal bond to cleave and its energy requirement is within the high energy tail of [M+H]+ internal energies. The mechanism for [an+H]+‘ ion formation is shown in Scheme XI. The [an+H]+' ion either exists as a distonic species or more likely it is probably stabilized from a hydrogen atom transfer from the protonated nitrogen to the a-carbon because a positive charge on nitrogen is more stable. We Simple inductive bond cleavages require less energy than homolytic cleavages because more stable products are formed, but they require more energy than 1, 2- eliminations because no new bonds are formed. In Scheme 1, the reaction enthalpy of bn ions was estimated to be 54 kcal/mol when the amide bond is inductively cleaved. The high abundance of b" ions in the LSIMS spectra of peptides is good evidence that this mechanism is operative. If additional energy is available, the Ca-C(O) bond in b" ions can be cleaved to form an an ion and CO as shown in Scheme 1. If the (IL-carbon is part of a glycine residue having no side chain, then the formation of an an ion requires an estimated, exothermic reaction enthalpy of -5 kcal/mol. However, if a stabilizing side chain is present on the residue, an reaction enthalpies fall to -15 and -20 kcal/mol. A small activation barrier must exist in the transition from an to bn ions because the abundances of bn ions are comparable to an ions even though the reaction enthalpies for forming an ions are exothermic. The a 1 ion shown in Scheme H is a special case for an ions. From ionization energies, the a1 ion structure is approximately 10 kcal/mol more stable than the an ion structure because the amide carbonyl group delocalizes the amide nitrogen lone pair in the an ion structure inhibiting full stabilization of the charge whereas, in the a] structure, the lone pair on the amine N can fully stabilize the charge. Scheme X] R’ o I HHxaHn /C'H~__> \(S HzNA flc\NH Linear [M+H]+ peptide Scheme XII Linear [M+H]+ peptide 180 ’I‘ e 'I‘ on /CH~ H21~{ (“2 H2 O=C=NH O R" [an+H]+° i" 0 ll RI" CH NH ——> / \ / \ £9 + I V‘HN q CH NH=CH @ 0 5*" ’1‘. 0 ll NH —> /CI\1 / \ Q + NH3 V‘HN % (IEH 0 RN - R @911 - R' HN V‘CH—C— l \ CH n 5§\>"CHR _,mHN/\ /OH + l/CHR C II 1H)" C/\ EBOH o” Secondary interaction Obn+on+m+) 181 Thus, for an a1 ion, the estimated reaction enthalpy is now -30 kcal/mol exothermic if a stabilizing side chain is present on the a-carbon and -15 kcal/mol exothermic if the residue is glycine. Formation of the 2,, ion in Scheme IV requires inductive cleavage of the N-Co, skeletal bond. For 2,, ion formation, reaction enthalpies were estimated to be 62 kcal/mol if a side chain is present on the a-carbon to provide charge stabilization, and 79 kcal/mol if the residue is glycine with no side chain. The high reaction enthalpies calculated for 2,, ions formation are supported the absence 2,, ions in the LSIMS spectra of these model peptides. The b,, ions are probably more abundant than 2,, ions because the b,, ion structure is more stable than 2,, and the C(O)-N bond is weaker than the N-Ca bond. The l, 2-elimination mechanism for peptides requires less energy than either homolytic or inductive bond cleavages even though two bonds are broken, because energy is released in the formationof a new single bond involving the shifted hydrogen and the upgrade of a skeletal single bond to a double bond. In Scheme III, assuming the amide nitrogen is protonated, the estimated reaction enthalpy required to form an a,, ion by a l, 2-elimination is 42 kcal/mol. However, the reaction enthalpy is reduced to 27 kcal/mol for the formation of an a 1 ion if the terminal amine is protonated, because a 1 ion structures are more energetically stable as explained above. In Scheme V, if the amide nitrogen is protonated, the reaction enthalpy for formation of an x,, ion through a l, 2- elimination mechanism was estimated to be 35 kcal/mol. For the y,, ion formed by a 1, 2- elimination mechanism in Scheme VI, the reaction enthalpy was estimated to be 38 kcal/mol. While these 1, 2-elimination mechanisms are more favorable than homolytic. and inductive cleavages, deuterium labeling studies have indicated y,, ions are not formed by this mechanism. The x,, ions are rarely observed in small peptides without basic residues and this may be additional evidence that 1, 2-e1imination mechanisms are not ‘I “/_U 182 operative. No plausible l, 2-elimination mechanisms exist for b,,, c,, and 2,, ions. MW Skeletal bond cleavages through secondary interactions involve breaking two different bonds and forming two new, very similar, bonds which results in very low reaction enthalpies without large energy deficits. The formation of y,, ions in Scheme VII first requires the formation of an intermediate through a favorable secondary interaction so a hydrogen can shift onto the ionic fragment for the formation of a stable cyclic neutral fragment. Specifically, the y,, ions have very favorable overall reaction enthalpies because two bonds are cleaved—the protonated amide bond on the peptide backbone and the interacting N-H bond. Subsequently, energy is returned when new two bonds are formed-a new N—H bond to the amine on the y,, ion fragment and a new amide bond that forms the cyclic neutral fragment. Depending on the origin of the shifting hydrogen, the overall reaction enthalpy for y,, ions was estimated to be —5 kcal/mol exothermic, or thermoneutral if the hydrogen originates from an amide or amine nitrogen, respectively. The 5 kcal/mol difference indicates the N—H bond is stronger in an amine nitrogen. For c,, ions, in Scheme IX, which are formed through C-terminal secondary interactions, the estimated reaction enthalpies are 5 kcal/mol if a hydrogen shifts from an amide nitrogen or 10 kcal/mol if the carboxylic acid hydrogen transfers. Again, the 5 kcal/mol difference indicates that a hydrogen atom is more difficult to remove from C-terminal carboxylic acid moiety than from the amide nitrogen. Assuming amide hydrogens shift, the overall reaction enthalpies for c,, ions are 10 kcal/mol greater than y,, ions because the N-Ca bond is stronger than the amide bond, less energy is returned upon formation of a new N - CO, bond to form the cyclic neutral product, and the PA of the c,, ion structure, similar to acetamide, is lower than the PA for the terminal amine structure in y,, ions. In addition, the y,, ions should be more abundant than c,, ions since the secondary interaction for c,, ions is less favorable than y,, because the partial positive charge on the (It-carbon is less 183 than that on the carbonyl carbon, and for c,, ions, the presence of side chains and/or on— hydrogens on the a-carbon imposes more steric hindrance for the interacting C-terminal amide nitrogens or carboxylic acid oxygen than encountered between the interacting N- terminal nitrogens and the open carbonyl groups for y,, ions. For cyclo-bn ion formation in Scheme VIII, a favorable N-terminal secondary interaction is required, but instead of a hydrogen transfer, the protonated skeletal amide bond is cleaved and a new amide bond is formed during cyclization to form the ionic cyclic fragment . The overall reaction enthalpy for cyclo-b,, ion formation is -5 kcal/mol if the secondary interaction involves an amide nitrogen or O kcal/mol if an amine nitrogen is involved. The energy difference indicates that the basicity of the amide group is higher. The formation of the cyclo-2,, ion in Scheme X requires a C-terminal secondary interaction involving an amide nitrogen in which a subsequent cleavage of a skeletal N- CO, bond and formation of a new N-Ca bond forms the ionic cyclic fragment ion. This process requires an overall reaction enthalpy of -5 kcal/mol, which is 10 kcal/mol lower than that for a c,, ion involving the same amide N interaction, because the PA of the amide nitrogen in the cyclic fragment ion is greater than that of a terminal amide. However, if the acid oxygen is involved in the secondary interaction for cyclo-2,, ions, the reaction enthalpy increases to 20 kcal/mol because the PA of the cyclic C(O)O-Ca ester site is much less than the PA of the cyclic C(O)N-Ca site formed when an amide nitrogen interacts. In general, cyclo-bn ions should be more favorable than cyclo—2,, ions because the secondary interactions in these processes parallel those for y,, and c,, ions above, respectively. 1 0d n mi 71 i r ti n The estimated thermodynamic data presented above describe the energetic requirements of the fragmentation mechanisms in Schemes I-XI, which allowed these mechanisms to be placed in perspective as illustrated in Figures 51 and 52. All sequence 184 ions have reaction enthalpies below 4 eV (70 kcal/mol), the upper limit of internal energies for protonated molecule in the LSIMS experiment, which indicates fragmentation through these mechanisms is thermodynamically accessible. These reaction enthalpies provide additional insight into the RI information for pentaglycine and methionine enkephalin in the LSIMS experiment in Figures 49a and 50a. The resulting y,, and b,, ion formation from amide skeletal bond cleavages is indeed thermodynamically and kinetically competitive because y,, ions have reaction enthalpies that are slightly exothermic overall and result in the formation of a stable cyclic neutral intermediate, while b,, ions evolve from simple bond cleavages, that require approximately 2 eV without the need for a rate-limiting secondary interaction required for y,, ions. For pentaglycine and methionine enkephalin in Figures 49a and 50a, no peaks representing a b; ion are observed at amide skeletal bond position 2, but intense peaks representing the 01 ions are observed at skeletal bond position 1 which supports the idea that b] loses CO to form a 1. The estimated exothermicity of this b1—>a1 fragmentation process corresponds well to the exclusive appearance of intense (11 ions in both peptides and, in fact, the a 1 peak for methionine enkephalin is much more intense than that in the mass spectrum of pentaglycine, which supports even lower reaction enthalpies through side chain stabilization. However, 0,, ions are generally less intense or missing in comparison to d] ions which coincides well with the less exothermic reaction enthalpies for a,, formation. Further evidence that local side chains lower the reaction enthalpy is observed for the intense peak at skeletal bond position 10, which represents an a4 ion (in comparison to the peak at skeletal bond position 11 representing a b4 ion), in Figure 50a for methionine enkephalin. In LSIMS of pentaglycine, at the same skeletal bond position 10, there is no peak representing the (14 ion, but instead an [a4+H]+'. In Figure 49a for LSIMS of pentaglycine, the secondary interaction mechanisms required for y,, ion formation are more thermodynamically favorable, and they account for the two intense peaks at skeletal bond positions 5 and 8, which represent the )3 and yz ions, and the two 185 less intense y,, peaks at skeletal bond positions 2 and 11 better than those that result from 1, 2-elimination mechanisms. The intense yz peaks in both model peptides at amide skeletal bond position 8 suggest amide hydrogens rather than amine hydrogens are more favored to shift, which corresponds well with the calculated y,, reaction enthalpies. However, the enhanced RI of y2 may also suggest that hydrogens shift from either amide or amine N sites. The presence of the methionine side chain appears to enhance the y 1 peak at amide skeletal bond position 11 in methionine enkephalin in comparison to the small y 1 peak for pentaglycine. For the RI distribution of peaks representing c,, ions, formed from cleavage of N-Co, skeletal bonds 6, 9 and 12, are low for both peptide models, which indicates b,, and y,, ion formation mechanisms involving amide bond cleavage(s) and formation(s) are more kinetically favored over cleavage of the N-Ca bond for c,, ions. The calculated reaction enthalpy for c,, ion formation is slightly endothermic overall in comparison to that for y,, ions, which further indicates that the C- terminal secondary interaction is not as favorable as the N-terminal interaction for y,, ion formation. Since RIs of peaks in the mass spectrum contain both thermodynamic and kinetic information for the formation of fragment ions, the estimated reaction enthalpies for the a,,, b,,, c,,, and y,, ion fragmentation mechanisms allow the purely thermodynamic attributes of a particular mechanism to be identified and more insight into the complex and competitive unimolecular decomposition kinetics. Experimentally, these estimated reaction enthalpies correspond well with the stabilities of the derived fragment ion structures, and quantitatively rank how favorable the simple bond cleavages and the rearrangement reactions are with respect to all the fragmentation options of the protonated peptide as shown in Figures 51 and 52. The trends above for these pentapeptide models give strong evidence for secondary interactions being in competition with simple bond cleavages, and the extent to which these postulated mechanisms produce b,, and y,, ions, respectively, were probed by comparing the pentaglycine RI 186 distributions to those of oligoglycines with three, four and six residues. Comparison of LSIMS Oligoglycine Peptide Relative Intensity Data. The extent of secondary interactions was evaluated in LSIMS spectra of triglycine (MW 189), tetraglycine (MW 246), pentaglycine (MW 303), and hexaglycine (MW 360) by considering the R18 of b,, and y,, ions peaks in Figures 53a-d, respectively. In tetraglycine, pentaglycine and hexaglycine, most b,, and y,, ion fragmentation results from cleavage of the one (i.e., amide skeletal bond 5), two (i.e., amide skeletal bonds 5 and 8 in Figure 49) and three (i.e., amide skeletal bonds 5, 8, and 11) central amide bond(s), respectively. Triglycine is an exception because it does not have a central amide bond surrounded by two outer amide bonds. For each oligoglycine, two intense peaks representing b2 and y2 ions, respectively, are observed independent of the number of glycine residues. Cleavage of the one central amide bond in tetraglycine in Figure 53b results in the competitive formation of b2 ions through a simple amide bond cleavage and yz ions through a favorable N—terminal amine secondary interaction with elimination of a stable six-membered cyclic neutral product, respectively. For all oligoglycines in Figure 53, the R1 of the b2 peak may be enhanced if the N-terminal amine undergoes a secondary interaction to either simply protonate or polarize the amide skeletal bond 5 to form additional b2 ions or cyclo-bz ions, respectively. If only tetraglycine and pentaglycine are considered in Figures 53b and c, yz ion formation appears to be straightforward because y2 ions are most likely formed by secondary interactions involving the N-terminal amine in tetraglycine and the skeletal bond 2 amide nitrogen in pentaglycine, which has a lower reaction enthalpy. Moreover in hexaglycine in Figure 53d, if formation of six-membered cyclic neutral products is prominent, a secondary interaction involving amide nitrogen at skeletal bond 5 would induce cleavage of amide skeletal bond 11 for y; formation and likely result in a decrease of R15 for amide bond 5 cleavage products, ()2 and y4 . However, the RI distribution of the peaks representing the b2 and y4 ions for hexaglycine 187 a) Triglycine l I LSIMS e Ilhdl J j Normalized Relative db III- CI:- 1 I it. uence Ion 7O Seq 0 23': b) Tetraglycine T.» a: '8 .23. E e Z y3 b2 y2 b3 )7 Sequence Ion .L 50 - g -I 3 40110) Pentaglycine: I LSIMS *- G a: g 30 IMI . '8 s - :2- : 20 a I- q 2 I'll , '5 10q ‘ a z .l'I-IIFIHIKI y4 b2 y3 b3 yz b4 y] Sequence Ion 03% VIC 1d) Hexaglycine i w 0 -—-—-N omom Normalized Relative Intensity OUI y5 b2 M '93 ya b4 )2 b5 y} Sequence Ion Figure 53. Bar graphs comparing normalized RIs of b,, and y,, peaks in the LSIMS and MI spectra of a) triglycine, b) tetraglycine, c) pentaglycine, and d) hexaglycine. 188 are still as intense as in pentaglycine, which means the b2 and y2 ion formation is independent of the degree of the secondary interaction that induces their formation. In Figure 53a for triglycine, the peak representing the )7 ion is also intense similar to those for the other oligoglycines, which means an N-terminal amine secondary interaction induces the formation of this yz ion and a three-membered cyclic neutral product. This process in triglycine is in competition with b; formation if the N-terminal amine induces a b2 ion through intramolecular protonation and a simple cleavage, or a cyclo-bz through a polarizing secondary interaction. In Figure 53 for each oligoglycine, y 1 formation should be more than favorable through secondary interactions, but the peaks representing the y1 ions are of low intensity, which suggests the formation of protonated glycine is not a favorable ionic product perhaps owing to its small size and limited stability in the high internal energy environment of the LSIMS experiment. Thus, the oligoglycine series shows diglycines, both H-GG$ and @NH3CH2C(O)-G-0H, representing b2 and y2 ions, respectively, are very stable ionic products because their R15 are intense regardless of number of residues in the oligoglycine and the degree of the secondary interaction(s). Under the energetic conditions of LSIMS, the formation of y,, ions and a stable six- membered cyclic neutral products is not a requirement for all secondary interactions because evidence shows three-membered rings are also formed. Because a y,, ion is essentially a shorter, complete N-terminally protonated peptide, the possibility of a subsequent secondary interactions in y,, ions may induce a secondary fragmentation processes leading to the smaller ionic products such as 172 and y2 if enough internal energy remains. Secondary Fragmentation Processes in LSIMS. Because the internal energies in LSIMS are estimated to extend out to 4 eV, the preferential formation of b2 and y2 ions from oligoglycines and the intense (1] peak from methionine enkephalin shown in Figures 53a-d and 503 above, respectively, may be due 189 in part to secondary fragmentation processes. In El, mechanisms were elucidated for the most important ions resulting from primary and secondary fragmentation processes, but the formation of ions such as 0* (carbon) were not considered. For example, in Scheme XII, a simple inductive cleavage of the ionized amide bond in an a,, ion can lead to a b,, ion [211]. The reaction enthalpy for this process was estimated to be 51 kcal/mol, similar to the energy required for the formation of normal b,, ions. In Figure 54a and b, the fragmentation barrier profile, based on the calculated reaction enthalpies for b,, ions, a,, ions and a,, to b,, transitions, is shown for methionine enkephalin and pentaglycine, respectively, to demonstrate the amount of energy needed for secondary fragmentation processes to occur. In the LSIMS mass spectrum of methionine enkephalin, the R15 of peaks representing a,, and b,, ions follow this potential energy diagram well because the RI b4 is less than that of a4 which corresponds to the a4 exothermic pathway enhanced by the presence of phenylalanine side chain in Figure 54, and the R1 of a1 is very intense which corresponds to the highly exothermic b] to a] transition enhanced by the tyrosine side chain. Assuming internal energies of the [M+H]+ ions have an upper limit of 3 eV (70 kcal/mol) and no reverse activation energies exist for these simple cleavages in LSIMS, the transition from b4 ions to b3 ions will take 71 kcal/mol, so b4 is an unlikely precursor to all other ions. However, if b3 is formed by a simple amide bond cleavage, the transition of b3 ions to b2 ions requires 56 kcal/mol and the transition of b3 to a] requires only 59 kcal/mol, so secondary fragmentation may account in part for the very intense a 1 peak in the LSIMS analysis of methionine enkephalin. From the pentaglycine fragmentation barrier profile in Figure 54b, secondary fragmentation can readily occur because a transition from b4 to a; requires only 61 kcal/mol. However, without a very thermodynamically favorable fragmentation process induced by side chains, the R15 of peaks representing a,, and b,, ions are relatively uniform with the slight exception of b2 and a] which represent the smallest, most stable ionic fragments and the most intense peaks in the LSIMS spectrum of pentaglycine . In Scheme XIII, the loss of ammonia 190 6O -- Ia) Methionine Enkephalin l i ’ . Energy (kcal/mol) O I. Peptide b4 04 b3 a3 b 2 a 2 b 1 a1 Sequence Ion 60 -r I b) Pentaglycine I 50‘ d Energy (kcal/mol) U) o Peptide b4 a4 b3 a3 b 2 a2 b 1 a 1 Sequence Ion Figure 54. Fragmentation barn'er profiles based on estimated reaction enthalpies for a,, and b,, ions from unimolecular decomposition of a) methionine enkephalin and b) pentaglycine protonated molecules. 191 from a c,, ion will produce a b,, ion, which was estimated to require 42 kcal/mol. Perhaps c,, ions are initially formed in high abundance through the favorable secondary interactions and subsequently lose NH3 to form b,, ions, because c,, ions are generally less abundant than b,, ions. Thus, for a LSIMS experiment, in some cases there should be enough internal energy remaining in the primary fragment ions to induce subsequent secondary fragmentation assuming the activation energies are negligible. The high abundance of non-sequence, low-mass ions in the LSIMS methionine enkephalin spectrum such as side chain ions [65, 187], immonium ions [212], and internal ions [213] indicates secondary and higher processes are occurring to some extent. By accounting for the formation of these non-sequence ions, 3 more accurate description of [M-t-H]+ unimolecular decomposition can be obtained for peptides. Assuming the R15 of peaks representing the F, G and M immonium ions, the GF and GGF internal ions, the GF-28 and GGF-28 internal immonium ions, and the side chain ions are normalized, along with the sequence ions, and all ions are the result of secondary fragmentation processes that also originate from a protonated amide nitrogen, the revised RI distribution for methionine enkephalin fragmentation is 30%:13%:43%:10% from amide bonds 2, 5, 8, and 11, respectively. The primarily difference between this LSIMS fragmentation RI distribution and the distribution from Figure 50a, which shows only the RI distribution of sequence fragment ions, is amide bond 8 is cleaved 43% of the time when secondary fragmentation processes are tallied in comparison to only 3% in Figure 50a. The major secondary fragmentation process at amide bond 8 is the phenylalanine side chain stabilizes the formation of the abundant F immonium ion, a nonsequence ion. In methionine enkephalin, the tyrosine and phenylalanine side chains have much influence on local fragmentation because the revised distribution indicates that the fragmentation pathways in the vicinity of the two aromatic side chains are more thermodynamically favorable because these side chains stabilize the intermediates and final products. Thus, if a peptide contains side chains, low-energy pathways will exist and then secondary frag sic' 111'. sin 192 fragmentation processes may occur more frequently. Under the high-energy conditions of LSIMS, the fragmentation mechanisms for side chain ions, immonium ions within the peptide chain and internal ions are not well understood. For internal ions, MS/MS/MS studies on a hybrid mass spectrometer, using high/low energy CAD and leucine enkephalin, have shown GGF can be formed from y4 ions and GF can be formed from )3 ions [214]. Because a y4 ion is essentially a smaller but complete peptide that is protonated on the N-terminal amine, €BNH3CH2C(O)-GFL- 0H, a conceivable mechanism for forming the GGF internal ion from ya, would involve a subsequent secondary interaction in which the protonated amine nitrogen protonates the amide nitrogen between F and L to either induce a simple inductive amide bond cleavage and form a b,, type fragment, H-GGFQ, or form a cyclic internal ion structure identical to a cyclo-bn ion. Because amine and amide PAs are very similar, intramolecular transfer of a proton between the amine and amides would be very favorable and would require approximately zero kcal/mol. In the LSIMS, NH and CAD mass spectra of the peptide H- GGFL-OI—I, the low intensity peak representing the a3 ion and the dominant peak representing the b3 ion, H—GGFQ, suggests the H-GGFe fragment is cyclic, which would have the same structure as the internal ion, GGF. Thus, the estimated reaction enthalpies for internal ions would parallel those for the formation of b,, and cyclo—bn ions above and internal immonium ions would parallel a,, ions assuming the overall reaction enthalpies for forming y,, precursor ions are taken into account. The F immonium ions involving the internal phenylalanine residue can be formed by a similar mechanism as internal ions if the protonated N-terminal amide nitrogen on the yz ion would protonate the nitrogen at amide bond 11 in methionine enkephalin to induce an inductive amide bond cleavage to form a b 1-type ion and subsequent CO loss which would result in an a1-type ion, or the F immonium ion [215]. The overall reaction enthalpy for this immonium ion would be similar to the formation of y2, b1 and a1 ions as discussed above. To investigate the secondary interactions in more detail, derivatized peptides were 193 synthesized using well known procedures contained in reference 216. An acetylated peptide is formed by attaching an acetyl group to the N-terrninal amine and, as a result, the N-terminal amine nitrogen is converted to an amide nitrogen. With an N-acetylated peptide, the b1 ion now appears in the LSIMS spectra, unlike that of underivatized peptides, along with (1], that still contains the acetylated group, because thermodynamically, the b] to a] transition here produces an a1 ion with the less favorable a,, protonated imide structure instead of the more favorable a1 protonated imine structure. An a1 ion with the protonated imine structure still exists at m/z 136 for the enkephalin peptides although it is half its normal size, Which indicates perhaps not all of a1 is from b] in the underivatized peptide and/or a ketene is eliminated from the acetylated N- terminus by a 1, 2—elimination mechanism. In addition, acetylation of a peptide produces a uniform RI distribution of intense peaks representing b,, and y,, ions in LSIMS spectra. For b,, ions, this increase can be rationalized if the thermodynamically less favorable protonated imide a1 ion reduces secondary fragmentation processes from b,, ions. For y,, ions in acetylated methionine enkephalin, the ya; and y 3 ions become as abundant as yz and y 1 because a hydrogen can be shifted through secondary interactions from the more thermodynamically favorable acetyl amide nitrogen. Thus, the estimated reaction enthalpies indicate that secondary fragmentation processes are possible and the RI distribution of both sequence and nonsequence ions indicates that secondary fragmentation occurs to some extent, especially if the thermochemistry is very favorable. Evaluation of the MI Relative Intensity Data. Metastable [M+H]+ ions fragment after acceleration because their initial internal energies are too low to induce the required high unimolecular decomposition rates for fragmentation in the ion source. In contrast to high energy simple cleavages in LSIMS, the prominent metastable fragmentation mechanisms for both pentaglycine and methionine enkephalin should consist essentially of low-energy secondary interactions. 194 r r l ' e In Figure 49b, for metastable decomposition of pentaglycine, the RI distribution of the skeletal bond cleavage products shows nearly all fragmentation results from cleavage of the amide skeletal bonds. No a], a,, or [a3+H]+' ions are observed which verifies that metastable fragment ions only represent primary fragmentation from low energy [M+H]+ ions and not secondary fragmentation processes. Summing the fragmentation from cleavage of protonated amide bonds 2, 5, 8, and 11 in Figure 49b results in the RI distribution of 5%:24%:15%:36%, respectively, which is essentially opposite from the LSIMS distribution in Figure 49a. From Figure 49b, the abundant y 3 ion, formed from cleavage of skeletal bond 5, strongly suggests that an N-terminal amine secondary interaction results in the formation of this ionic fragment and the stable six- membered cyclic neutral product. While formation of any y,, ion is indeed a very thermodynamically favorable fragmentation process for low-energy [M+H]+ ions, the thermodynamics for an amide nitrogen secondary interaction and hydrogen shift are more favorable than the amine nitrogen secondary interaction, and y2 ions should dominate over )3 ions as in LSIMS. However, for metastable decomposition of protonated pentaglycine, the peak representing the y 3 ions is greater than that for yz which suggests the secondary interaction involving the amine requires less energy to fold inward to complete an interaction than does an amide nitrogen. The origin of the dominant b4 ion is unclear because it is unlikely that simple inductive bond cleavages are competitive in very low energy [M+H]+ ions. Instead, the intense metastable peaks representing b4 ions are probably the result of slow unimolecular decomposition. Protonated pentaglycine in the gas-phase may exist in a folded state held together by hydrogen bonding and perhaps, a secondary interaction between the N-terminal amine and amide skeletal bond 11 occurs, since they would be close in space to form a symmetrical, low-energy flexible 12- membered cyclo-b4 ion similar to the formation of a six-membered neutral as in Scheme 195 VIII. The low R1 of peaks representing c,, ions at skeletal bond positions 6, 9, and 12, further suggest that the C-terminal secondary interaction required for c,, ions is not as thermodynamically favorable as an N-terminal secondary interaction. Thus, the abundant )3 and b4/cyclo-b4 fragment ions from the metastable decomposition of protonated pentaglycine indicate low-energy, secondary interactions are the driving force in these fragmentation mechanisms that produce the very stable ionic and neutral products. MIMI] M 2!. l . In Figure 53a—d, the RIs of peaks representing the metastable b,, and y,, ions from protonated triglycine, tetraglycine and hexaglycine coincide with the metastable ions from protonated pentaglycine above. In all four cases, the RI distributions differ sharply from LSIMS. In Figure 53a for triglycine, formation of a b2/cyclo-b2 ion and glycine neutral product is more favorable than formation of a protonated glycine molecule, a y 1 ion, and a cyclic diglycine neutral product from the same N-terminal amine nitrogen secondary interaction. In Figure 53b for tetraglycine, the yz ion becomes dominant perhaps because the larger, more stable protonated diglycine molecule is formed in addition to the cyclic diglycine neutral product. The RI distribution for protonated hexaglycine in Figure 53d shows two intense metastable ion peaks representing y4 and b5 ions that parallel the intensity of y 3 and b4 ions for pentaglycine in Figure 53c. For peaks representing the cyclo-bn ions formed from protonated oligoglycine decomposition, higher values of n have more intense RIs supporting formation of larger cyclic ions and small neutral glycine products. The RI distributions of fragment ions from metastable decomposition of acetylated triglycine, tetraglycine, and pentaglycine [M-I-H]+ ions were compared to the. RI distributions of the respective underivatized oligoglycine [M+H]+ ions in Figure 55a-c to further investigate the importance of N-terminal amine secondary interactions. The "bulky" acetylated N-terminal amide nitrogen may either inhibit or enhance the 196 a Triglycine I Underivatized I Acetylated Normalized Relative Intensrty b1 y2 b2 y1 Sequence Ion 60. - g 50%-lb) Tetfaglycine . I Underivatized _ '5 ‘ ’ .3 >. 40? . IAcetylated _ a: 2:: : W -. 2 30. = g .. 2" 20: h .. ° : Z 10, 0' i : : : b1 y3 b2 y2 b3 y] Sequence Ion 50: | 4510 Penta i : yc ne .2 40: ) g] I 3 352 , “and g g 255 ‘ 1. Underivatized: N “ = :5 5 205 Acetylated e 15; 2 10; 5; 0‘ : b1 y4 b2 y3 b3 y2 b4 y] Sequence Ion Figure 55. Bar graphs comparing normalized RIs of b,, and y,, peaks in MI spectra of the underivatized and acetylated forms of a) triglycine, b) tetraglycine, and c) pentaglycine. 197 fragmentation of certain b,, and y,, ions in metastable decomposition. For acetylated triglycine in Figure 55a, the R1 of the peak representing the b2 ion is reduced by two- thirds, and likewise, for acetylated tetraglycine and pentaglycine, the peaks representing the yz and y3 ions, respectively, are reduced which initially suggests the N-terminal amine secondary interaction in the underivatized oligoglycine induced their formation more than the N-terminal acetylated amide nitrogen. In acetylated triglycine, tetraglycine and pentaglycine, the R13 of peaks representing y2, y3 and y4 ions increased dramatically because in one scenario, conditions for a new secondary interaction are created in which a hydrogen can be shifted to the protonated amide nitrogen at skeletal bond 2 (in the underivatized peptide) from the acetyl group to form the respective y,, ion, a ketene molecule and a three-membered cyclic glycine product. However, in the more thermodynamically favorable scenario, a hydrogen may be shifted from the newly created acetyl amide nitrogen to form a three-membered cyclic neutral product and the respective y,, ion. For tetraglycine and pentaglycine in Figures 55b and c, the RIs of peaks representing the b3 ions also increase because perhaps the secondary interaction involving the N-terminal acetyl group itself provides a larger, more stable ll-membered cyclo-b 3 ion in comparison to the nine-membered cyclo-b 3 ion formed in underivatized oligoglycines. The M1 fragmentation from the oligoglycine series indicates in a broader sense that secondary interactions are the preferred thermodynamically favorable mechanisms for producing stable fragments from [M-t-H]+ ions with low internal energies. For hexaglycine, the neutral products formed in conjunction with the dominant y 3, cyclo-b4 and cyclo—b5 ions are the small, highly stable molecules, cyclic diglycine, diglycine, and glycine, respectively, which is similar to low—energy CI. So far, in the MI discussion above, the peptides were assumed to be protonated on the amide skeletal bonds as in LSIMS, but the N -terminal amine nitrogen in protonated form may also interact with the amide skeletal bonds in the peptide to induce fragmentation in metastable decomposition. 198 Perhaps protonation of amide nitrogens results in immediate amide bond cleavages as in LSIMS and protonation of amine nitrogen results in delayed amide bond fragmentation, on the metastable time scale, after an intramolecular proton transfer through secondary interactions. MIRIL [ErM !' . E! H' The RI distribution of MI skeletal bond cleavage products for methionine enkephalin in Figure 50b is very similar to the RI distribution for pentaglycine in Figure 49b. However, the presence of the phenylalanine side chain in methionine enkephalin introduces an additional low-energy pathway to form the a4 ion. Because inductive cleavages would not be favored in metastable decomposition, perhaps the interacting N- terrninal amine from a secondary interaction to amide bond 11 catalyzes loss of CO for the formation of linear 04 or cyclo—a4 ions following cyclo-b4 ion formation through multiple hydrogen bonding. The unimolecular decomposition of acetylated enkephalin pentapeptides in LSIMS, MI and CAD results in a very low a4 ion abundance in contrast to those from analysis of the underivatized enkephalin peptides. This suggests that either the N-terminal acetyl amide group is too "bulky" to interact effectively and act as a "catalyst" in the formation of a4 ions, or the acetyl group interacts very well with amide skeletal bond 11 forming cyclo-b4 ions exclusively. In Figure 50b, cleavage of amide bonds 2, 5, and 8 in protonated methionine enkephalin metastable decomposition results in dominant y4, y 3, and y2 ions, respectively. These y,, ions have similar abundances in contrast to the singly abundant )3 ion for pentaglycine because the tyrosine side chain likely enhances the stability of the cyclic neutral products formed. When methionine enkephalin is acetylated, the y4 ion becomes more abundant similar to the oligoglycines above probably because now a hydrogen can be shifted from the acetyl amide nitrogen which requires less energy. Internal ions and the phenylalanine immonium ion, typically considered high energy fragments, also appear in the MI spectra for methionine 199 enkephalin which further suggests that they can are formed from low energy pathways involving subsequent secondary interactions from y,, precursor ions. In addition, the a1 and y 1 ions which were most abundant in LSIMS, are not observed in the MI distribution indicating significant energy is required for production of these small ionized fragments. In addition, if the C-terminal acid group is converted to a methyl ester, the low mass y,, ions in metastable decomposition increase in abundance which indicates the methyl ester further stabilizes these smaller fragment ions. Last, if the enkephalin C-terminal acid group is converted to a terminal amide, C3 and 2,, ions become prominent peaks in the MI/CAD spectrum because a secondary interaction involving nitrogen is more favorable than oxygen. A metastable decomposition map was created in Figure 56 by selecting the protonated, underivatized methionine enkephalin molecule, each sequence ion and various non-sequence ions in the LSIMS mass spectrum as both precursor and product ions for product ion and precursor ion scans (using linked scans at constant B/E and B2/E), respectively, under metastable decomposition conditions, to investigate secondary fragmentation processes in more detail. A much simpler but similar CAD map for leucine enkephalin was presented by Alexander and Boyd [217]. In contrast to El, many MI product ions are detected from the LSIMS experiment which suggests multiple protonation sites are slowly decomposing, probably through N —terminal amine secondary interactions, instead of one dominant charged site. Some secondary fragmentation is observed on the metastable time scale from primary fragment ions formed immediately after ionization in the LSIMS experiment. For example, when b4 ions were selected as the precursor ion for a product ion scan, an intense metastable peak representing the a4 fragment ion, a secondary fragmentation process, was detected. When a4 was selected as the product ion for a precursor ion scan, an intense peak representing the b4 precursor ion was detected. If y,, ions are the selected precursor ions, intense peaks representing the internal ions, immonium ions or lower series y,, ions are detected as shown in the 200 .m0t0m .053 E 2.2 00.20.58 :0 n83 5000 2 38805 0.8 $95.58 05 :5 .53 $528 E8050. 2.65 @0320 05. .mtmm 28 m5 E8250 3 280m “002:: Boa @0588 5.230520 05:250.: mo +E+S= .8 39553 5289502. 0382208 05 9:32? 92: 5:35.:me .om 05$"— meamem AI R: N ”3:085me ”V .0 $0M a me 3 Al m AIIIIINA N01 1* N2 \\ +E+Sz +E+§\ / / / m. £26 All. 2 \/ a. \ /. . . \.//.. \ E\ \\ V0\V wmimumv lllll ”HOD TI Fm 201 metastable map. Evaluation of the CAD Relative Intensity Data. Outside the ion source, single collision CAD excites stable [M+H]+ ions from one high energy interaction with He to induce fast fragmentation. The [M+H]+ internal energies from CAD are high enough on average to induce simple bond cleavages as in LSIMS. MW In Figure 49c, the "true" single collision CAD distribution of skeletal bond cleavage products versus skeletal bond positions are shown for pentaglycine from which the MI products were subtracted out. All sequence ions are identical to those obtained by LSIMS above except a small fraction of the a2 ion abundance is [a2+H]+' and a4 is [a4+H]+'. Assuming fragmentation is induced by protonation of the amide nitrogens, the RI distribution of fragmentation induced by CAD at amide skeletal bond positions 2, 5, 8, and 11 is 12%:40%:27%: 13%, respectively, in Figure 49c which is nearly identical to the LSIMS RI distribution for pentaglycine in Figure 49a. The RI distribution of peaks representing b,, ions in CAD is identical to the distribution of peaks from LSIMS in Figure 49a but not the distribution of peaks from MI in Figure 49b most likely because the [M+H]+ internal energies are high enough to competitively induce simple bond cleavages. In contrast to the distribution of peaks in M1, the RI distribution of peaks representing b,, ions in CAD may also indicate an "unzipping" process is occurring in which one precursor ion such as the b4 ion is sustaining formation of the other b3 and b2 product ions, which is energetically feasible from the fragmentation barrier profile in Figure 54b. Alexander and Boyd [217] showed for leucine enkephalin under extreme multiple collisions that the distribution of CAD product ions can be made to resemble a FAB or LSIMS ion distribution indicating either primary fragmentation processes 202 become competitive only at high internal energies, secondary ionization processes can be induced by high internal energies from subsequent collisions, or a combination. The RI distribution of peaks representing y,, ions in CAD resemble the MI distribution instead of that for LSIMS probably because the secondary interactions or rearrangement processes are strictly rate dependent and the average internal energies in LSIMS may be slightly higher, which may account for the y,, RI differences in the distributions. r r i ' In. In Figure 50c, the "true" single collision CAD product distribution for methionine enkephalin, with the MI products subtracted out, is shown and it appears to be quite similar to the MI distribution in Figure 50b above. The addition of three large side chains to the pentapeptide backbone has doubled the number of degrees of freedom in methionine enkephalin which means the internal energy imparted into the [M+H]+ becomes more dispersed and the rates of fragmentation decrease to resemble metastable decomposition. Thus, single collision CAD simply enhances most metastable fragmentation processes for the larger, neutral protonated peptides. However, the reappearance of the a,, ions and [a3+H]+’ ions, at skeletal bond position 7, and the increase in b2, b3, yz, and y 1 ion peak intensities in the CAD distribution indicate a small population of [M+H]+ ions have high internal energies for fast, higher energy decomposition. There is no evidence for conversion of primary products to secondary products in the single collision CAD of protonated methionine enkephalin as in LSIMS. The protonation sites probably differ initially before fragmentation in LSIMS and CAD as indicated in the RI distributions in Figures 50a and c, respectively, because in the LSIMS distribution, the most intense peaks represent the (1] and y 1 fragment ions, but these peaks are not as intense as those in the CAD distribution. Perhaps, in comparison to the LSIMS amide nitrogen protonation, the stable [M+H]+ ions that are excited in CAD are probably protonated on N-terminal amine nitrogen, amide carbonyl oxygens or 203 the side chains inhibiting fast fragmentation processes. CAD simply initiates fragmentation processes from these existing protonation sites or redistributes the proton to the amide nitrogens by secondary interactions which subsequently initiates fragmentation as in LSIMS. Formation of [b4+0H+H 1+ Ions. Recently, for peptides, much published work has focused on exact fragmentation mechanism of [b4+OH+H]+ ions from metastable and CAD [M+H]+ unimolecular decomposition [218, 219]. In Scheme XIV, assuming the peptide is protonated on the basic amide carbonyl oxygen rather than on the less basic amide nitrogen, as suggested by Teesch and Adams [203] for alkali metal cationization, a nucleophilic attack from the C-terminal acid group on the first inside carbonyl group can occur favorably though a five-membered ring. However, some current discrepancies in the [b4+OH-t-H]+ mechanism include: where does the proton reside during the reaction?; how is the -OH group transferred mechanistically?; and what types of neutral products are eliminated? In all the work to date, no one has tested whether the energetics of the various mechanisms are reasonable because the elimination of CO and NH=CHR require that three bonds are broken, and one and half bonds are formed which leaves a very large energy deficit well above the energy available in MI decomposition. However, as shown in Scheme XIV, the formation of a three-membered cyclic neutral reduces this large energy deficit since now two bonds are broken and two similar bonds are formed which results in an estimated overall reaction enthalpy of only 23 kcal/mol. 204 4. Conclusions The sequence and non-sequence peptide fragmentation mechanisms were evaluated using both RI information in LSIMS, MI and single collision CAD mass spectra and estimated reaction enthalpies, similar to practices in 131. In E1, the first step in elucidating fragmentation mechanisms was to describe the structures of the ionized molecules. For E1 of dipeptides, the amine and amide nitrogens are the likely sites of ionization owing to their low IE [220]. From the R13 of fragment ion peaks in the mass spectrum, the distribution of fragment ions originating from these two ionization sites were determined after elucidating the fragmentation mechanisms. E1 was more simplistic for peptide analysis than LSIMS primarily because only simple peptides could be analyzed. In LSIMS, peptides with higher molecular weights and a greater number of functional groups may be protonated, which makes the protonated analyte structure(s) more difficult to describe. For example, the peptides in LSIMS may be protonated exclusively at the most basic site, from which all fragmentation originates, or at a basic site(s) remote from where skeletal bond cleavages occur. However, evidence for these types of fragmentation mechanisms are not observed in the mass spectra neutral pentapeptides here, which contain no highly basic residues for preferential protonation. Instead, the neutral pentapeptides appear to be protonated, with a certain distribution, on the amide nitrogens, amine nitrogens, amide carbonyls, and to a lesser degree on the side chains. If all amide nitrogens were initially protonated equally on a pentapeptide with no side chains to influence fragmentation, such as pentaglycine, then it would be expected that the distribution of fragment ions from each amide nitrogen would be equivalent. However, the fragmentation observed from these protonated amide N sites in pentaglycine is not equal and it appears that secondary interactions are responsible, in 205 part, for either inducing chemistry at each amide nitrogen site and/or redistributing the initial proton distributions to effect fragmentation. The LSIMS RI fragmentation distribution for pentaglycine and the estimated reaction enthalpies suggest that both simple cleavages and secondary interactions are competitive at each protonated amide nitrogen. Although the most favorable fragmentation pathways are not always observed, the most stable ionic and neutral products are formed in all cases, whether a simple cleavage or secondary interaction is required. The M1 RI fragmentation distribution for pentaglycine in contrast to LSIMS indicates that N-terminal amine secondary interaction mechanisms are responsible for each of the fragment ions observed in the mass spectra. The RI distributions observed for methionine enkephalin reveals that each side chain significantly enhances local fragmentation which indicates side chains stabilize the fragments formed. Much work still remains in the elucidation of unimolecular decomposition mechanisms for peptides in LSIMS and FAB. This work represents one of the first real attempts to quantify the energetics of the peptide fragmentation mechanisms. However, the results reported here are preliminary because our knowledge of conformations of peptides in the gas phase, internal energies from the LSIMS experiment, etc., are limited at this time preventing us from exploring other mechanistic possibilities. Neutralization- reionization mass spectrometry (NRMS) is promising for aiding in the identification of the types of neutral products following unimolecular decomposition of metastable [M+H]+ ions and energized [M+H]+ ions in CAD, providing soft ionization techniques such as CI can be implemented for reionization [221]. Gas-phase confirmations of protonated peptides should be investigated initially using computer molecular modeling techniques to learn more about the [M-I-H]+ structural relationship to the R1 of peaks in the mass spectrum and eventually to the fragmentation mechanisms from each protonated site that involve specific secondary interactions. Experimentally, electrospray ionization (E81) and [222] laser desorption (LD) are the two softest ionization methods for 206 producing gas-phase molecules and perhaps gas-phase conformations of peptides could be accurately investigated at first, without fragmentation processes occurring, using spectrochemical methods. Methods should be developed to accurately estimate the appearance energies of the peptide fragment ions using hybrid, ion beam mass spectrometers coupled with desorption/ionization techniques to experimentally confirm the estimated reaction enthalpies presented here. Laser desorption/chemical ionization (LD/CI) coupled with FT -MS detection seems promising for creating breakdown curves to probe the energetics and rates, k(E) functions, of peptide fragmentation mechanisms [223]. CHAPTER FIVE. FUNDAMENTAL AND ANALYTICAL UTILITY OF GAS-PHASE METAL ION/BIOMOLECULE REACTIONS I. Introduction Product ion structures, organometallic bond strengths, equilibrium constants, and rate constants of gas-phase ion/molecule reactions between bare metal reagent ions and small organic molecules with less than two functional groups are typically determined at or near thermal energies on the millisecond time scale using an ion cyclotron resonance (ICR) mass spectrometer [224, 225]. Dedicated ion-beam mass spectrometers [226] are used to obtain appearance energies of organometallic product ions for calculating respective heats of formation (AHf) values. These fundamental studies with ICR and ion- beam mass spectrometers allow both the elucidation of reaction mechanisms for gas- phase organometallic chemistry, and direct correlations to be made between the gas-phase organometallic chemistry, in the absence of a solvent, and the respective condensed- phase chemistry. Gas-phase ion/molecule reactions may also be studied in regions of high pressure in a conventional mass spectrometer such as the enclosed ion source volume utilized in chemical ionization (CI), the volume (or the selvedge region) immediately enclosing the area from which molecules are desorbed into the gas-phase in potassium ion ionization of desorbed species (K+IDS) and fast-atom bombardment (FAB) experiments, or in a collision cell filled with inert noble gases or reagent gases for collisional activation dissociation (CAD) or endothermic reactions, respectively. In comparison to an ICR, the typical mass spectrometer is more restricting because only reactions occurring faster than a microsecond are observed, the ions generally have energies much greater than thermal energies, and lower sensitivity is usually observed 207 208 owing to ion transmission losses. However, mass spectrometers are readily available and provide faster scanning speeds, higher mass ranges and higher resolution than the ICR mass spectrometer. In the last 10 years, the fourier-transform mass spectrometer (FT- MS) [227], has evolved into the state-of-the-art instrument because it incorporates all the inherent advantages of the ICR plus it also incorporates the high mass and resolution advantages of the sector mass spectrometers and the fast scan speeds of quadrupoles. The fundamentals of ion/molecule chemistry, advanced instrumentation and desorption/ionization (D/I) techniques are now available for study of ion/biomolecule gas-phase chemistry, but this area is still virtually unexplored because the observed chemistry is still too complex for accurate mechanistic interpretation just yet owing to the complex interactions of bare metal ions with multiple functional groups of highly structured and functionalized biomolecules. In this chapter, some fundamental and analytical advantages of ion/biomolecule reactions are explored and demonstrated using various metal reagent ions (M+) incorporated into the K+IDS and K+IDS-by-FAB experiments, more correctly referred to as M+IDS and MfiDS-by-FAB, respectively The ionization technique, K+IDS, utilizes gas-phase K+ ion/molecule chemistry in the ion source volume of a mass spectrometer to obtain the molecular weight and some structural information for analytes of biological origin as explained in Chapter two. K+IDS is essentially a K” CI experiment in which low-energy, gas-phase K+ ions (cat)ionize desorbed analyte molecules (M) by simply adding to M without inducing any fragmentation. Thus, K+ is an excellent probe for determining the types of neutral molecules present in the gas phase since the respective [M+K]+ adduct ions are detected for each M without fragmentation. In K+IDS-by-FAB, low-energy K+ ions are injected into the selvedge region, a glycerol-CI environment, to probe the types of neutral molecules that are desorbed in the fast-atom bombardment (FAB) experiment as explained in Chapter three. Metal-containing compounds may be substituted into the aluminosilicate glass matrix instead of potassium, and M+IDS and M+IDS-by-FAB 209 experiments may be performed analogous to those based on 161138 and K+IDS-by-FAB to access whether other metal ions, M+, are as analytically useful as K+. In K+IDS and K+IDS-by-FAB, the K+ reagent ions just "stick" to gas-phase molecules, but by replacing K“ with other metal reagent ions, the outcome of the ion/molecule chemistry will likely be complementary to K+. First, the M+ adduct ion may yield unique fragment/product ions either naturally (exothermic) or when activated (endothermic) which would provide new, informative structural details for M. Second, if the metal ion has an unique isotopic distribution, then peaks in the mass spectrum that represent either an M+ adduct ion or a respective fragment ion will be better identified and confirmed. As an example application, M+ in a CI environment may test the functionality of unknown molecules by reacting or adding to those molecules containing only specific functional groups [228]. Recently, chemical ionization of peptides with Fe+ was initially demonstrated which resulted in the losses of H20, CO and C02 from the [M+Fe]+ ion, and in the cleavage of amide bonds for the formation of meaningful peptide b,, and y,, fragment ions containing Fe+ [229]. This fragmentation is contrary to the well known insertion of Fe+ into Co,- C(O) bonds. Probing gas-phase, highly functionalized molecules such as peptides with various metal ions should produce fragment/product ions that yield complementary structural information to those from other mass spectrometric D/I methods which may help elucidate fragmentation mechanisms and lead to unambiguously identifying the gas- phase binding sites [230]. These gas-phase binding sites, free of solvent effects, may lead to identification of the active metal binding sites on biomolecules in condensed-phase systems aiding in the elucidation of biological mechanisms. 210 2. Reactivities between Metal Ions and Analyte Molecules in M +IDS The gas-phase ion/molecule chemistry utilizing alkali metal reagent ions is well characterized because copious amounts of alkali ions can be thermionically emitted into the gas-phase from aluminosilicate emitters doped with alkali metal substances owing to the low ionization energies of the alkali metals and the high work function of this emitter composition [60]. Cs+ and Rb+ only form adduct ions similar to K4" in K+IDS because the large size of these metal ions prevents sufficient overlap of electron density between M and the metal ion resulting in a very weak bond. In [M+K]+, the bond is electrostatic which means the charge resides external to M because the ionization energy (IE) of K is lower than M. Because the charge is remote from M, charged-induced chemistry cannot occur but charge-enhanced chemistry can if M is sufficiently activated and polarized by the metal ion. The metal ions, Na+ and Li+ in Na+IDS [54] and Li+IDS [173], respectively, form adduct ions with molecules, too, and they also can induce simple neutral losses such as water loss to form fragment ions containing a metal ion. A stronger, more covalent-like bond with better overlap of electron density is formed between these smaller metal ions and M and the IE is greater, which means the charge is nearer M to possibly initiate charge-induced chemistry. In M+IDS, less common metal reagent ions, M+, can be thermionically emitted into the gas phase from a heated aluminosilicate emitter to either form adduct ions, form products from reactive (exothermic) collisions, or undergo charge exchange with gas-phase biomolecules depending on the physical properties of the metal ion. In K+IDS and M+IDS, biomolecules are either thermally desorbed intact and/or thermally decompose through only the lowest energy pathways depending on the heating rate and biomolecule structure. In addition, there is no matrix in M+IDS which means there are no matrix 211 effects as in FAB. Thus, for biomolecules, there are only several analyte-related peaks to interpret in K+IDS and M+IDS mass spectra which makes the extraction of analyte molecular weight and structural information simple. Therefore, M+IDS is ideal to probe and characterize the diverse metal ion/biomolecule chemistry under the normal operating conditions in a typical mass spectrometer, which would be complementary to K+IDS, K+IDS(M+IDS)-by-FAB, and studies using ICR or FT-MS techniques. Below, the preparation and characterization of Al+, Ga+, and In+ emitters is described first. Next, the preliminary results from Al+IDS, Ga+IDS, and In+IDS experiments and their analytical advantages are presented. M+IDS Emitter Preparation. Blewett and Jones [52] initially optimized the molar ratios of an aluminosilicate matrix doped with Li2C03 to obtain the largest amount of low-energy, thermionically emitted Li+ ions when the glass emitter was heated to greater than 1000° C. For preparation of an optimized Li+ emitter, they melted, in powdered form, either Li2CO3 or LiN03, A1203 and Si02 together in a 1 or 2:1:2 molar ratio to form Li20:A1203:2Si02 glass following either C02 or N02 release, respectively. This glass composition is the same formulation as the well known ternary system B-eucryptite. They obtained Li+ ion currents as great at 1 mA, and also discovered that other alkali ion emitters such as 10‘ may be prepared with the same molar ratios as the B-eucryptite Li+ emitter, which was recently confirmed to give the strongest K+ ion currents [231]. Blewett and Jones [52] also described the production of Mg“, Ca+, Sr+, Ba+, Al+, Cat“, and In+ ions from oxide emitters, which provided ion currents less than 1.0 microampere and involved simply melting the oxides MgO, CaO, SrO, BaO, A1203, Ga203, In203 on a filament wire, respectively. Successful Cu“ and Ag+ aluminosilicate emitters were prepared using Cu(NO3)2 or AgNO3 mixed with A1203 and silicic acid in a l or 2:1:2 molar ratio, respectively [232]. Ni+ and Cr+ borosilicate emitters were successfully prepared in a 212 manner similar to that used for the B-eucryptite aluminosilicate emitters described above except that boric acid and silicagel were substituted for A1203 and SiOz [233]. Commercially available alkali and alkaline earth ion emitters are based on the B- eucryptite formula, and are designed to emit copious amounts of ions [234]. They are used in applications such as the Cs+ emitter pellet in secondary-ion mass spectrometry (SIMS) guns. The Cs+ emitters for the SIMS gun cost in excess of 300 dollars, but they can be simply regenerated successfully by incrementally melting small amounts of CsN03 or CszCO3 into the top of the porous tungsten plug. The Ga+ and ln+ emitters were prepared in a manner similar to that used for the B- eucryptite K+ aluminosilicate emitters above instead of melting the respective oxides. Thus, Ga203 or In203 was mixed with A1203 and SiOz using the molar ratios 1 or 1:1:2, respectively. All refractory reagents were high purity and crushed with a mortar and pestle to a powder. An attempted formulation of a C0"' glass emitter with the initial formula, Co(NO3)2:A1203:2Si02, apparently vaporized upon heating in the Co+IDS experiment. Al+ emitters were prepared exclusively with the pure, powdered reagent grade Al(NO3)3-9H20. Acetone was added to the mixed, powdered solids to produce a "slurry" for controlled formation of the emitter. A Pasteur pipette was used to "drop" or "layer" the emitter solid from the "slurry" onto the rhenium heater wire loop which was pre-attached to the direct insertion probe (DIP). After the acetone evaporated, the "dried" emitter solid was heated briefly in a Bunsen burner flame to firmly attach the applied solid mass to the Re wire loop, through crystallization or melting, for successful insertion into the mass spectrometer. If the emitter is heated for extended periods under atmospheric conditions, the Re wire will begin to oxidize which will drastically shorten the lifetime of the emitter. Quite often, N02 or C02 gas was released from the "crystallized" emitter solid upon heating, and this applied solid was splattered off the Re wire. If this occurred, the emitter powder was first melted to a glass to evolve these gases, then the glass was recrushed to powder, and reapplied to the Re wire from the 213 acetone slurry. An alternate, more difficult method for emitter preparation was to dip the Re wire loop into a platinum dish of molten metal-containing aluminosilicate emitter material, heated with the most inner flame (i.e., the hottest part) of the oxygen-acetylene torch, and capture a "bead" of molten emitter at the same time being careful not to oxidize the Re wire or vaporize the Pt dish. The K+IDS, Al+IDS, Ga+IDS, and In+IDS experiments were conducted on a Hewlett Packard (HP) 5985 GC/MS/DS single quadrupole mass spectrometer equipped with a direct insertion probe inlet. Thermionic emission was achieved by passing current through the metal-containing emitter. The K“, Al+, Ga+, and In+ emitters were part of a current loop in which the emitter rhenium wire ends were spot-welded to the nickel heater leads on the end of a standard two-filament K+IDS DIP which were connected to the heating current loop of K+IDS power supply unit described in Chapters two and three. Before the K+, Al+, Ga+, and In+ emitters could be used for M+IDS, they had to be "conditioned" first which involved melting the emitter onto the Re wire. This step was done in the mass spectrometer under vacuum to prevent oxidation of the Re wire at elevated temperatures and the procedure involved raising the emitter heating current 0.2 A every 20 sec until 3 A was reached. Heating the 10, Al+, Ga+ and In+ emitters resulted in emission of only the respective ions without significant co.emission of "impurity" ions with lower ionization energies. Thus, there was no need to "bake out" any "impurity" ions from these emitters for extended periods. In the K+IDS, Al+IDS, Ga‘l'IDS, and In‘l'IDS experiments, gaseous analytes were leaked into the ion source CI volume, into which the emitter was inserted, until the ion source housing pressure was approximately 1x10‘5 torr, and solid analytes, dissolved in MeOH, were "dried" onto the opposite Re wire sample holder after the MeOH evaporated . Liquid analytes were applied directly Re sample holder. 214 M+IDS Emitter Characterization. The A1+, Cal and In+ ion currents were measured from experiments conducted in both a vacuum test chamber attached to a vacuum rack as shown in Figure 57a, and in the ion source of the mass spectrometer. In the vacuum test chamber, an M+ glass emitter, attached to the current loop of the K+IDS power supply unit via the feed-throughs, was heated to induce thermionic emission. A realistic portion of the ions were detected by a faraday metal plate and measured by a picoammeter despite the fact that ions are emitted in a 360° radius of the bead. In Figure 57b, the In+ ion current and emitter bead temperature as a function of heating current is plotted for the In203:A1203:28i02 glass emitter. Microamperes of In+ current are readily obtained between 2.1 and 2.3 A and above 2.9 A of heating current. In comparison, the K20:A1203:2Si02 and Ga203zA1203:2Si02 glass emitters yield 10 11A of K+ and 1 ILA of Ga+ above heating currents of 2.9 A under the same experimental set-up and conditions, respectively. In the HP mass spectrometer, Int, 63*, Al+ and K+ emitters typically yield 9100, 2200, 8000, and 3500 absolute A/D signal counts at heating currents of 2, 2.7, 3, and 2.6 A and at electron multipliers voltages of 1700, 1700, 2000, and 1200 V, respectively. K“ ion currents are strongest because the ionization energy (IE) of potassium is lowest (4.3 eV) in comparison to Al+ (6 eV), Ga+(6 eV), and In+ (5.8 eV) which have [55 slightly greater than lithium (5.4 eV) assuming the work function of the emitter matrices remain constant. Comparable ion currents of transition metal ions such as Co+ (7.9 eV) are more difficult to produce thermionically because the IEs are 2 eV higher than those for the other metals described above. M+IDS Experimental Results and Discussion. The gas-phase ion/molecule chemistry between Al+, Ga+, and In+ and simple, gas-phase analytes was investigated and compared to that of K+. In Figure 58a, the 215 3) Reference 7 To Vacuum :7 I, , _ Rack ' ‘ ‘ :sg'i: Power Supply . ' 7 0-20V' Heater Current Power Supply 0-4 A “ \f\‘ (Re. K IDS Power Supply Umt. Ground Picoammeter E:J(Keithley #417) Ground M+ or K+ Ions Vacuum Chamber 1.6E-06 - 1600 g 14E-06 b) . l_.,-/"'—.’-—. 14‘” 2 1.2E-06 .4" "' . 1200 g: g 1.0E-06 ”An/"M . 1000 j g ’./' I- U 8.0E-07 u" . - ' ' ' ' 800 g E scram . ' - 600 g. 4’: 4.0E-07 ' —D— In+ Ion Current 400 E mam ' . 200 0.0E+00 - 0 Q——NQVW\O§OQO§N-1NMVW\O_OOGM o u— — v—n — —n v— v— v— — N N N N N NN NN Heating Current (A) Figure 57. a) Schematic diagram of the vacuum chamber, K+IDS power supply unit, and picoammeter used for testing M+ thermionic emission ion currents from alurrrinosilicate emitters doped with various metal substances, and b) a plot of In+ emission ion current and emitter temperature versus heating current for the In203zA1203228i02 emitter. 216 [M+K]+ peak in the mass spectrum represents K+ attachment to gas-phase acetone (MW 58). In this experiment, acetone was leaked into the ion source closed CI volume and [M+K]+ adduct ions were formed as K+ CI reagent ions were thermionically emitted from the K+ emitter bead on the DIP inserted into this CI volume. This same experiment was performed with In+ emitted in the open EI ion volume instead which resulted in a significant drop in the total ion current (TIC) as shown in the mass spectrum in Figure 58b in comparison to that in Figure 58a. The peak at m/z 173 represents the [M+In]+ adduct ion of acetone. In fact, In+ bound preferentially to the more volatile acetone component in many solvents such as alcohols than the less volatile solvent itself since acetone is the commonly used denaturing substance in solvents. Contrary to adduct formation with K+ and In+, the mass spectrum in Figure 58c indicates Al+ extracts Cl' from CCl4 (MW 152), which was leaked into the El ion source volume, since CCl3+ is the dominant peak at m/z 117. The reactive Al+ also provided additional peaks in the mass spectrum that are only 5% of the base peak; these represent meaningful fragment/product ions that aid in characterizing this analyte. Under thermal conditions, the reaction between A1+ and CC14 in Equation 45 occurs because the reaction enthalpy Al+ + CC14 (g) -) CCI3+ + AlCl (g) (45) (Aern) is -9 kcal/mol exothermic [235]. If the same reaction is performed with K+ or In+, no reaction is observed because the estimated AHnms are 49.5 and 12.5 kcal/mol endothermic, respectively; these reactions are not possible under thermal conditions [40]. Analyses using Ga+IDS and Al+IDS were also performed with the solid fatty acids, palmitic acid (MW 256) and tridecanoic acid (MW 214), dried on the Re wire sample holder, and compared to those obtained by K+IDS. In Figure 59b, in the Ga+IDS mass spectrum of palmitic acid, the peak at m/z 325 represents the [M+Ga]+ adduct ion, which indicates Ga+ simply forms adduct ions with fatty acids in a manner similar to 217 ‘ a) 39 Metal Ion: K+ K-I- Analyte: Acetone " Heating Current: 2.6A [M+K]+ Bias Voltage: 0 V Electron Multiplier: 1200 V 97 - TIC (Counts): 15849 I l ....I... I....I....I....I....I....I....I....I....I....I....I....I. "I 30 35 40 45 50 55 60 65 70 75 80 85 90 95 100 - b) Metal Ion: In+ 173 Analyte: Acetone ‘ Heating Current: 2 A [M+In]+ I Bias Voltage: 2 V Electron Multiplier: 2700 V . TIC (Counts): 100 llllllllllll lllTTlT 160 162 164 166 168 170 172 174 176 178 180 -— Metal Ion: Al+ l C) Analyte: CC14 CCI3+ Cs+ Heating Current: 3 A 117 133 Bias Voltage: 4 V AIC12+ Electron Multiplier: 2400 V 97 1 TIC (Counts): 13475 . 1 75 x1 x AlCl+ AICHCI+ I. ‘ CC1+ 62 47 AIC+ CC12+ ll 1 J II J I.fi.I....I....I....I....I....Ifi..I.. .....I I.-'... I. .. I 30 40 50 60 70 80 90 100 110 120 130 140 Figure 58. a) K+IDS and b) ln+IDS mass spectra of acetone (MW 58), and c) A1+IDS mass spectrum of CC14 (MW 152) leaked into HP ion source to 1x10'5 torr base pressure. 218 . a) I Metal Ion: K+ ' Analyte: Palmitic Acid - Heating Current: 2.6 A . Bias Voltage: 0 V Electron Multiplier: 2500 V 7 TIC (Counts): 851 295 [M+K]+ ll VVVVVVVVVVVV ‘1'. "" ""l 711‘] 1 #1 1 l l b) [M+Ga]+ 325 Metal Ion: Ga+ Analyte: Palmitic Acid H. Current: 2.9 A [M+Na]+ B. Voltage: 8 V 279 E. Multiplier: 3000 v [M+K]+ TIC (Counts): 1207 295 ‘ H... ....I ..5. .Iu....l ‘. "I. .. .r... 240245 250255260265 270 275 280285 290295 300305 3I0 315 320 325 3'30 c) 295 [M+K]+ ... Metal Ion: Al+ [M+A"C02] Analyte: Palmitic Acid 239 H. Current: 3 A [M+Na]+ B. Voltage: 4 v E. Multiplier: 3000 V 279 TIC (Counts): 794 339 .1 XX) 220 240 260 $0 300 320 340 360 380 d) 237 Metal Ion: Al+ l + Analyte: Tridecanoic [NH-Na] H. Current: 3 A Aci . + B. Voltage: 4 V g [M+Al-C021 E. Multiplier: 3ooov I 197 TIC (Counts): 467 I I 251 279 150160170180190 200 210 220 230 240 250 260 270 280 290 Figure 59. a) K+IDS, b) Ga+IDS, and c) Al+IDS mass spectra of palmitic acid (MW 256), and d) Al+IDS mass spectrum of tridecanoic acid (MW 214) on the HP mass spectrometer. 219 those formed with K+ as shown in the K+IDS palmitic acid mass spectrum in Figure 59a. The intense 716a isotopic peak conveniently indicates which peaks in the mass spectrum represent ions containing gallium. The Al+IDS mass spectra of palmitic acid and tridecanoic acid are shown in Figures 59c and d, respectively. No [M+Al]+ adduct ion is observed for either analyte so if this analyte was an unknown, 3 101138 analysis is still required for molecular weight information. However, a peak representing the [M+Al- C02]+ ion is observed for each analyte indicating Al+ reacts specifically with the carboxylic acid functional group. There is no mention of this reaction in the Al+ ICR literature, probably because carboxylic acid compounds are not very volatile and could not be studied. Al+ may prove to be a good reagent ion to test analytes for the presence of carboxylic acid functional groups. The gas-phase chemistry of Al+, Ga"; and In+ has been characterized previously with compounds having various functional groups [236]. In+ and Ga+ are similar to K+ because they only form adduct ions with M, but Al+ reacts with the fatty acid analytes tested here and previously with small molecules in the literature. The observed chemistry is attributed to decreasing bond energies Al+>Ga+>In+ since the electrostatic interactions and radii increase [237]. Last, in further support of the FAB ion formation mechanism, K+IDS mass spectra of common FAB liquid matrices glycerol (MW 92), triethanolamine (MW 149), thioglycerol (MW 108), and nitrobenzyl alcohol (MW 153), shown in Figure 60a-d, respectively, were obtained from successive experiments to demonstrate that K+ binds more favorably to the highly basic triethanolamine rather than to the other acidic matrices as indicated by the TIC values. Thus, K+ affinities likely parallel proton affinities as observed previously [61]. Additionally, in the 101138 triethanolamine spectrum shown in Figure 60b, the peaks at m/z 114 and 132 represent either gas-phase unimolecular decomposition of protonated triethanolamine molecules formed in a "local" selvedge region, or desorption of protonated degradation molecules of triethanolamine. Further support for a "local" selvedge region in the K+IDS experiment is the K+-bound cluster 220 >\ E ' “31m: Glycero 131 8 ': eating Current: 2.9 A [M+K]+ a) 5 .. ias Voltage: 3 V “:3 4 lectron Multiplier: 1600 V :33 j C (Counts): 4554 £1""I"" "' " "I 'I""1fi"1'l' 70 75 80 85 9095 10010511011512012511)135140 >‘ , g : 132 188 AnaTyte: Triethanolamine g . [M-H20+H]+ [M+K]+ HC: 2.9A b) 3 - 114 BV: 3v ,5 ; [M-2H20+H]+ EM: 1600V g . 11c: 52228 D m I"' '1""I""I " 1"" 1'1 100110 120130140 15016017018019021D210 220230 240250260270 280 2:. Z” l 147 finflyte: Thioglycero 2 I + C: 2.9A 0) E. - [M+K] BV: 3 V g . EM: 1600V [2M-H20+K]+ g : TIC: 1742 237 a: .............................................................. 100110m130140 m1m1'701801'9102mz'10m230240250260 b ‘53 r 192 Analyte: Nitrobenzylalcoho g: j [M+K]+ HC: 2.9A d) .s .. BV: 3v 0 ‘ EM.16OOV + g - TIC: 3829 [212?] 7: ‘ . 1.1 . a: .................................................. 100121)]:10160180200220240260280300320340360 >5 g - 85 [G+K]"’ Analyte: Glycerol/Cholic Aci 447 2 j Rb+ 131 BC: 2.9 A [M+K]+ e) .5 .. BV: 3 V 2 . EM: 2600V ‘5 j TIC: 18613 T) _. 1.141 _ "tut....rul.1 .1. 1.1.; until-1;... ...‘lnkl Lu LLIIJLALIAAJHIILkaj of. 1 l 1 1 1 1 70 110 150 190 230 270 310 350 390 430 470 510 550 590 630 Figure 60. K'l'IDS mass spectra of a) glycerol (MW 92), b) triethanolamine (MW 149), c) thioglycerol (MW 108), d) nitrobenzyl alcohol (MW 153), and e) cholic acid (MW 408) mixed in glycerol. 221 ions represented by peaks at m/z 237 and 345 observed in the K+IDS thioglycerol and nitrobenzyl alcohol spectra in Figure 60c and (1, respectively, although these peaks may also represent intact desorption of this dimer species. In Figure 60c, the K+IDS mass spectrum of cholic acid (MW 408) dissolved in glycerol on the Re sample holder is shown. The appearance of the [M-l-K]+ peaks at m/z 131 and m/z 447 indicate that glycerol and cholic acid molecules, respectively, are desorbed intact from the glycerol matrix into the gas phase identical to the FAB experiment as determined by KflDS-by- FAB. The glycerol matrix essentially protected cholic acid from thermal decomposition in this experiment because cholic acid readily lost water molecules in a K+IDS experiment when cholic acid was dried on the Re wire sample holder. 3. Gas-Phase Metal Cationization in FAB Using MflDS-by-FAB FAB mass spectra of biomolecules tend to be more complex to interpret than those from K+IDS or M+IDS. Spectral interpretation in FAB may be hindered by sample and matrix effects because the analyte or mixture of analytes is dissolved in a matrix for analysis. Also, there are a larger number of fragment ions to interpret in FAB because more energy is imparted into the biomolecule analytes. However, interpretation of all the fragment ions is not always possible because most fragmentation mechanisms are not well known. Often, metal ion cationization of the analytes in FAB can aid spectral interpretation by overcoming matrix effects and producing more informative, complementary fragmentation. M t I a 'x ts. In the FAB mass spectrum, any analyte peak can be potentially blocked by a matrix-related peak. Also, the peak at high m/z, that reveals the analyte molecular 222 weight, may represent a protonated molecule, [M+H]+; a radical cation, M'”; a salt cation, M+; a dehydrated fragment ion, [M+H-H20]+; proton-bound dimer of the analyte, [2M+H]+; or a proton-bound dimer of the analyte and glycerol, [M+G+H]+ depending on the chemical nature of the analyte and its intrinsic chemical relationship with the matrix. For samples of biological origin, large amounts Na+ salts (and/or potassium) are typically present in the sample, which results in an additional peak representing the sodium adduct ion of the analyte, [M+Na]+, appearing 22 Da higher than the [M+H]+ peak. The presence of salts in the matrix sometimes suppresses the other analyte peaks, but this intense [M+Na]+ adduct ion peak actually can aid in identifying the peak that reveals the molecular weight of the analyte amidst all other peaks. The analytes in salt-free samples may be cationized with alkali metal ions with high sensitivity by simply dissolving alkali salts such as NaI, NaCl, KI or KCl in the matrix at concentrations around 0.1 M. If the sample is a mixture of analytes, then the alkali metal will cationize each component, confirming which peaks represent intact analytes. If the sodium contamination or concentration is very high, then additional, intense peaks appear in the mass spectrum representing the sodium adduct ion of the sodium salt of glycerol, [M-H+2Na]+; [M+(NaCl)+Na]+; [M-H+(NaCl)+2Na]+; etc., which may confuse or hinder the identification of the analyte(s) peak(s), especially if there is a mixture of analytes in the sample. Assuming the analyte dissolves in the matrix and no condensed-phase reactions between the analyte and matrix occur, the FAB matrix may still suppress analyte ion formation. For example, if the analyte molecules are not surface active, but strongly "matrix-philic", they will not be readily desorbed from the matrix into the gas phase. If the proton affinity (PA) of matrix is greater than the analyte, then protonated matrix molecules will not protonate the analyte molecules to a great extent, if at all. Presently, it is unclear whether analytes are cationized by metal ions in the condensed phase and subsequently desorbed preionized, or whether gas-phase cationization is the dominant mechanism. 223 In K+IDS-by-FAB, K+ ions are injected into the selvedge region containing desorbed neutral molecules (M) and [M+K]+ adduct ions are formed indicating the types of neutral species present. The KflDS-by-FAB method is not only useful for probing the neutral molecule species in the selvedge region of the FAB experiment, but it can also be used for molecular weight confirmation because only one new peak representing [M+K]+ adduct ion appears in the mass spectrum 38 Da higher than the protonated molecule, even if the [M+H]+ is unstable, eliminating interpretation problems. The [M+K]+ ions formed in K+IDS-by—FAB are generally less abundant than [M-l-K]+ ions formed by adding 0.1 M KI salts to the matrix (see Chapter three) because K+ cationization is achieved solely in the gas-phase. However, gas-phase K+ cationization in KfiDS-by-FAB eliminates any matrix effects and the higher mass K+ adduct ions of KCl salt cluster ions caused by the addition of KCl salts to the matrix which yields much simpler spectra to interpret. As observed in Chapter three, gas-phase K+ is very effective in ionizing analytes with lower PAs than the matrix which results in [M+K]+ ions that have greater abundance than the [M+H]+. For a full discussion of the analytical utilities of K+IDS-by-FAB, see the published paper [124] in Appendix one. The use of metal ions besides K+ in the FAB experiment either using M+IDS-by-FAB or dissolving metal-containing salts in glycerol may complement K+ by providing unique isotopic distributions for easy identification of metal-containing ions as was demonstrated for Ag+ cationization of cholic acid (MW 408) [132]. Ni, Cu, and Ga have only one striking isotopic peak which would provide easy identification of the cationized ions without loss of sensitivity. a - 1 nt ment ti Metastable decomposition and CAD of protonated and cationized analytes formed in FAB require tandem mass spectrometric techniques for the detection of the fragment and product ions formed after acceleration outside the ion source, respectively. Tandem mass spectrometric techniques in conjuction with metastable decomposition and CAD is 224 a method to: isolate the fragmentation from each analyte component in a mixture, separate matrix peaks that interfere with fragment ion peaks in the FAB mass spectrum, enhance existing fragment ion abundances (in CAD), and supply unique and/or complementary, structurally informative fragmentation for an analyte, all of which can increase the chances of accurately defining the structure of the analyte molecule [238, 239, 240, 241]. The fragmentation differences induced by protons and metal ions are attributed to a combination of the metal ion binding site location on the analyte, M+- analyte bond strength, and the changes in conformation the analyte undergoes to host the interacting metal ion. The metal ion binding sites for peptides probably differ from proton binding sites if intact ions are formed in either the condensed [188, 242] or gas phase [203, 206] which would result in different sets of bonds being preferentially cleaved and account for much of the unique, complementary fragmentation and different ion abundance distributions in comparison to those observed following protonation. Assuming the average internal energies of cationized analytes and protonated analytes are similar and the metal ion and proton binding sites are identical, the intrinsic interaction between the metal ion and the analyte molecule may initiate different fragmentation mechanisms than proton-induced chemistry also resulting in complementary fragment/product ions. The analyte-metal ion bond strength represents the energetic upper limit for all complementary metal ion-induced bond cleavages and fragmentation processes observed because if enough internal energy is present to cleave the bare metal ion from the neutral analyte molecule, then the chemistry stops and complementary fragmentation requiring more energy will never be observed. If the metal ion-analyte bond is strong, more structurally informative fragmentation can be potentially obtained since there will be probably other weaker bonds to cleave. The ionic radius of an alkali metal ion, in part, dictates how close it can approach the exposed electron density at the heteroatom site on the analyte molecule before nuclear repulsions occur in order to achieve good overlap for sufficient bond strengths. Larger metal ions such as Cs+ are 225 unable to achieve as good an overlap as Li+, a smaller ion, and the bond energies between Cs+ ions and analytes are much weaker and more electrostatic than those for Li+. The ionization energy dictates, in part, on which species the analyte or metal ion the charge lies. For K+ interactions with analyte molecules, the ionization energy (IE) of K is 4.4 eV, which is half the value than those for most analytes. Thus, the electron density remains shifted towards the analyte and the charge remains on potassium. Thus, K+ can only trigger fragmentation processes by polarizing the local bonds through charge enhancement which almost borders on remote charge-site chemistry [243]. In contrast, Li+ has a smaller radius and a higher ionization energy which results in it forming stronger, more covalent-like bonds with analytes, and the charge, which induces bond cleavages, is likely closer to the analyte. Experimentally, more fragmentation is observed with Li+ than 10. In contrast to cationization-induced fragmentation, a proton has an IE of 13.6 eV, which is higher than that for most analytes, so the charge is transferred onto the analyte in exchange for electron density which essentially makes the analyte-H bond covalent. For a protonated heteroatom site with intermediate basicity, the fragmentation chemistry is believed to be solely charged-site induced, but if the heteroatom site is highly basic and all protonation occurs here, then remote charge-site chemistry may take place. The interactions of transition metals, such as Co+ (3d74sl, not 3d3), with analytes is different than with the alkali metal ions, which have inert, noble gas electron configurations, because the extra remaining valence shell electrons allow Co+ to behave as an ionic radical and insert into C-C bonds [224]. This process is thermodynamically favorable because one C-C bond is cleaved and two Co+-C bonds are formed. Additionally, the interaction and formation of electrostatic bond(s) between analytes and metal ions, that have large ionic radii and diffuse electron clouds, most likely results in. more analyte conformation changes to accommodate the larger metal ion in comparison to minimal perceivable changes in conformation for proton attachment which is small point charge. These new conformations will probably have much influence on the new 226 and/or complementary fragmentation observed with metal ions [244]. For any biomolecule analyte, the possible types of complementary fragmentation can be probed by dissolving different metal substances into the glycerol matrix and selecting the cationized analyte produced by FAB as the precursor for metastable decomposition and CAD. With metal ions in the FAB matrix, perhaps the intrinsic interactions between analyte biomolecules and the metal ions in the condensed-phase can be sampled with FAB which can possibly lead to the determination of the aqueous metal binding site(s). In order to determine the metal binding sites, the cationized analyte biomolecules must yield informative fragmentation by metastable decomposition and CAD. Once a metal is found that yields suitable complementary fragmentation, the metal substance can be inserted into the MflDS-by—FAB experiment to ensure the analyte is only cationized in the gas phase without any matrix effects from having metal ions dissolved in the matrix. With M+IDS-by-FAB, the intrinsic interactions between gas- phase biomolecule and metal ions can be sampled and compared to the condensed-phase M+-analyte interactions in the FAB matrix. Finally, the fragmentation of cationized analyte molecules in M+IDS-by-FAB will not consistently parallel M+IDS gas-phase fragmentation chemistry since the cationized molecules in the FAB experiment have higher average internal energies. All mass spectra in this section were obtained on a JEOL HX-110 double focusing forward geometry mass spectrometer. The M+IDS-by-FAB target was placed on the end of a direct insertion probe. The MflDS-by-FAB experiment was performed identically to that of K+IDS-by-FAB as described in Chapter three and in Appendix one using the same divided targets. The preparation of the metal-containing emitter substances and analytical advantages are described below. 227 MflDS-by-FAB Cationization Results and Discussion. Since our laboratory has considerable knowledge and continuing interest in probing gas-phase transition metal chemistry, transition metals were first inserted into the M+IDS-by-FAB experiment to study ion/biomolecule chemistry. The aluminosilicate matrix doped with Co(NO3)2 from M+IDS was tried initially by melting it onto one half of the wall-divided target in Figure l9e, but no Co+ ions were observed. In Figure 61a, b and c, Co(NO3)2, Cu(NO3)2, and Ni(NO3)2 were melted on the wall-divided target and Cot“, Cu+, and Ni+ ions represent the dominant peaks, respectively, but the ion currents in absolute AID counts are very low in comparison to the 11 counts of K+ which was typically obtained by FAB of the K+ aluminosilicate emitter melted on one half of the wall-divided target shown in Figure 61d. However, no Zn+ ions were obtained from FAB of Zn(NO3)2. The number of counts appears to decrease in the order, Cu+ (7.7 eV) >Ni+ (7.6 eV) >Co+ (7.9 eV)>Zn+ (9.4 eV), as the metal ionization energies (IE) increase. Ti (6.8 eV) and Cr (6.8 eV) have lower IEs than the metals above so the production of Ti+ and Cr"’ ions may be more successful in MflDS-by-FAB, respectively. The compounds Ti02 or Ti2(SO4)3 and Cr(NO3)3 could be used as sources for these ions, respectively. Despite the low Co+ ion currents, CoflDS-by-FAB was still demonstrated to work with glycerol (MW 92) and the peptide H—LGG-OH (MW 245) using the wall- divided target in Figure 1% without any evidence of condensed phase contamination as shown in Figures 62a and b, respectively. However, it was later discovered that Co+ does not ionize these analytes to a great extent owing to a mysterious ion series in the glycerol Co+IDS-by-FAB mass spectrum at m/z 213, 305, etc. separated by 92 Da. In one scenario, protonated glycerol reacts with gas-phase intact Co(NO3)2 to produce the product ion, [G+CoN03]+, appearing at m/z 213 in the mass spectrum as shown in Equation 46 . Co(NO3)2 can either be desorbed intact or formed by Co“ and NO3’ 228 81°C 59 5 8013) (30+ E 60‘ [30+ Emitter: 3 Co(NO3)2 CoOH+ '3 40‘ [30+ Counts: 76 CONO"? '3 20‘ 06 ° 89 . M G - - - ' v - v i 20 4O 6O 80 100 m/z >5 "5 1:310) 58 i 9- Mi+ Emitter: Ni+ . E {3' 60‘ Ni(NO3)2 (171%0)Ni+ 5 '5 40‘ Ni+ Counts: ‘60 NiNO‘fll ’3 20: 0.7 88 . a: l r 62 , o _ 20 4O 60 80 113/0 2 £9100 23 + . ; E 80: e) Na+ a Glass Errutter E SE) 60' Na+ Counts: 3 : '3 A1+K+ : .2 401 27 39 + E '3 20‘ l [45 58 E M O'F—-—-" -'A!—V:L?f 2'77 : 0 20 40 60 80 100 m/z $1003) 23 ‘3 . + Li+ Glass Emitter o 80 L.+ Na E 60: ; Al+ Li+ Counts: 14 ..g. 40' 27 K4" 5 20: 39 Fe+ § 0 ##5‘6 0 20' ' '4'0' ' '6'0' ' '8'0' ”100 m/z 2:100 . '53 ‘ 91+ i 5*: 23b) Cu+ Emitter: 3 T, CU(N03)2 I .> 40‘ 65 . g 20‘ 01+ Counts: CuNO‘l"; 2 0.8 I 913 , G *2? '4‘01 60' '76 100 m/z .2100 2 80: d) 3%. +Glass Emitter ’; 3 K + i 2 4o: ' 'fi 20. 3 ‘ 2.3 u = 00 5'6 100"15o' 200 m/z >~100 .3 + 185 1 § 30' f) [633?” [2G+H]+ t: . E 32 O0 50 100 150 200 250 m/z g: 100'Fh) [69'3“] 1851 5 80‘ + E so 3 40' '3; 20: 00 50 100 150 200 m/z Figure 61. FAB mass spectra of a) Co(NO3)3, b) Cu(NO3)3, c) Ni(NO3)3, d) K+ glass emitter bead on wall-divided target in Figure 1%, e) Na+ glass emitter bead on "pop-up" divided target in Figure 19f, f) Na+IDS-by-FAB mass spectrum of glycerol, g) FAB mass spectrum of Li+ glass emitter bead on KflDS-by-FAB space-divided target, h) LifiDS-by- FAB mass spectrum of glycerol. 229 [G+H]+ + Co(NO3)2 (g) -> [G+CoNO3]+ + HNO3 (g) (46) recombination in the selvedge. Thus, the ion appearing at m/z 150 is [G-H+Co]+ which is formed by subsequent loss of HNO3. Sunner et al. [97] reported a very similar mechanism for the reaction of protonated glycerol and KC] in the gas phase. Oddly enough, if Co(NO3)2 is dissolved in the glycerol matrix, only Co+ ions are observed indicating that Co(NO3)2 is desorbed primarily as a neutral species because it is surface active and essentially blocks desorption of glycerol molecules. For intact analyte species containing Co+, perhaps the Co+ is exchanged from [G-H+Co]+ to the analyte by a subsequent gas-phase collision to account for the [M-H+Co]+ observed in the H-LGG- 0H mass spectrum in Figure 62b. Although more experimentation would need to be done to ensure that this is indeed a gas-phase process, it implies that salts such as KNO3 dissolved in glycerol are desorbed intact into the gas phase and react with protonated glycerol or the protonated analyte to form [G+K]+ and [M-i-K]+ ions, respectively. This K+ adduct ion formation mechanism may account for the high sensitivity observed in the practice of adding KI salts to glycerol since more neutral species are desorbed in the FAB experiment than ions. In Figure 62c, the FAB mass spectrum of Co(NO3)2 and the analyte digoxin (MW 780) mixed in glycerol is shown and the peak at m/z 838 representing the [M-H-l-Co]+ ion is observed. The low-mass peaks normally in the protonated digoxin FAB spectra at m/z 97 and 131 are present without Co+ attachment. The [M-H+Co]+ ions were selected as the precursor ion for a linked scan at constant B/E under metastable conditions as shown in Figure 62d. The distribution of metastable [M- H-l-Co]+ fragment ions is similar to those of the metastable [M+H]+ decomposition of digoxin because the intense peak at m/z 708 and the smaller peak at m/z 579 represent loss of one and two sugar moieties identified in Figure 62d [176], respectively. However, the peaks at m/z 602, 664, and 734 are not present in [M+H]"' metastable mass spectra 230 é 1 93 [2G+H]+ 242 CoflDS-by-FAB £9 804 3) [G+H]+ 185 [ZG-H+C0]+ of glycerol ’. g 1 on wall target I a 601 E T, : [G-H+Co]+ 305 3"” : 2 . + [3 -H+Co] : 33 20-1 294517 J ”(3351] 334 397; G 0' "50' ' ' 166 150 200 250 300 350 4‘30 2 100, 86 r >‘ 1 b) CoflDS-by-FAB of H-LGG-OH : .53 80; a] in glycerol on wall target ; § 6,: a 5 : [G+H]+ [ZG-H+Co]+ [M-H+Co]+ : .L‘a’ 40? 44 :93 [G-H+Co]+ 242 303 .- 35. : G W : 8 2°: : 0* L 0 100 200 300 400 m/z 100 >‘ : c) 131 Co(NO3)2 & digoxin mixed E 553' 80', in glycerol on same target C l 95’ 50‘: 97 [M-H+Co]+ .- 3 : [ZG-H+Co]+ 338 : 7°? 20: 43 186 337 x5.0 i a: 1 I 408 i O: “12* Jul - r“ t .‘ c - - c ‘1 ‘ - fi fi. 0 200 400 600 800 1000 m/z 3 Metastable decomposition 708 338' €103 d of [Digoxin-H+Co]+ from c) l [M-H+C0]+ g i [(AOS}OSZOH)-H+Co]+ C: 1 7, : [(AOS}OH)-H+Co]+ .2 5: ‘5 1 734 800 3 j 602 664 a: l 404 541579 00 100 200 300 400 500 600 700 800 m/z Figure 62. a) Co+IDS-by-FAB mass spectrum of glycerol and Co(NO3)2 solid separated on a wall-divided target in Figure 1%, b) CoflDS-by-FAB mass spectrum of H-LGG-OH (MW 245) in glycerol and Co(NO3)2 separated on a wall-divided target, c) FAB mass spectrum of digoxin (MW 780) and Co(NO3)2 mixed in glycerol, and d) [digoxin-H+Co]+ metastable decomposition mass spectrum obtained by a linked scan at constant B/E. 231 and they probably represent ring cleavage products either on the aglycone or sugar moieties . Na+IDS-by-FAB and LiflDS-by-FAB experiments were also conducted using Na+ and Li+ aluminosilicate emitters that were prepared identical to the K+ emitters in K+IDS-by-FAB. In Figure 61c, FAB of the Na+ glass emitter on the "pop-up" target in Figure 19f yielded only three counts of Na+. The NaflDS-by—FAB mass spectrum of glycerol in Figure 61f shows Na+ was injected into the selvedge region to produce the peak at m/z 115 which represents the [G-l-Na]+ without any evidence of contamination. FAB of the Li+ glass emitter on the space-divided target in Figure 21a yielded 14 counts of Li+ in Figure 61g owing to the superior geometry of the divided target. The Li+IDS- by-FAB mass spectrum of glycerol again shows more abundant Li+ adduct ion formation with glycerol than Na+ adduct formation even though more Na+, from impurities in the Li+ emitter, were injected into the selvedge region than Li+ owing to the stronger binding energies of Li+ to alcohols [60]. Metastable and CAD Studies of Cationized Peptides. There has been much attention in the mass spectrometry community concerning cationized peptides in recent years because they provide complementary fragmentation in comparison to that of protonated species. The peptides are usually mixed with alkali metal salts in an appropriate matrix and subjected to analysis by FAB . Metastable and CAD analyses of the cationized peptide are the methods of choice for obtaining the complementary fragmentation because cationized peptides are very stable and they do not usually undergo unimolecular decomposition on the time scale of the FAB mass spectrum. For metastable decomposition and CAD experiments, linked scans at constant B/E were employed. Helium was the collision gas used in CAD to attenuate the ion beam. The conversion dynode was adjusted to maximize abundance of C5312+ (m/z 652.5) at an 232 accelerating potential of 1 keV [245] before mass-axis calibration. Ultramark 1621 (PCR Inc., Gainsville, FL) was used as the mass-axis calibration compound for linked scans at constant B/E compounds because according to the formula for calculating apparent metastable fragment ion masses, m*=m22/m1, the lowest linked scan fragment/product ion mass, m2, obtainable with an [M-l-H]+ precursor ion representing an analyte of molecular weight 800, m1 is m/z 28 if the lowest Ultramark calibration peak, m*, is at m/z 1. For comparison, with CsI calibration, the lowest calibration peak is Cs+ at m/z 133 and the lowest fragment/product ion mass obtainable with a linked scan at constant B/E is m/z 258. The operating pressure of the instrument with FAB Xe gas on was 5x10‘ 6 torr. K+IDS-by-FAB provides the capability to form [M+K]+ ions solely in the gas phase with the bare K+ metal ion. Using the peptide H-VGVAPG-OH (MW 498) as a model, metastable and CAD studies were conducted to probe whether the fragmentation for these gas-phase [M+K]+ ions formed by K+IDS-by-FAB differed from the [M+K]+ ions formed by FAB of both KI salts and analyte dissolved in the glycerol matrix. Currently, it is unclear if these latter [M+K]+ ions are formed in the condensed or gas phase. Any differences in fragmentation in the two methods may indicate a different distribution of K+ binding sites in [M+K]+ ions because [M+K]+ unimolecular decomposition for both methods is observed under the same metastable or CAD conditions. In a FAB mass spectrum, a cationized analyte molecule appears to be quite stable because little or no fragmentation from this species is normally observed. However, unique and complementary [M+K]+ metastable unimolecular decomposition is observed outside the ion source. Thus, the observed metastable fragment ions originate from [M-l-K]+ ions only having low internal energies because, if an [M+K]+ ion had too much internal energy, the electrostatic bond between M and K+ would be cleaved and all fragmentation chemistry would stop. If the internal energies of an [M-l-K]+ ion extended 233 above the range of energies for metastable decomposition into the range for ion source decomposition, but still below the M-K+ bond energy, the fragmentation rates of these "energized" [M+K]+ ions may still be low and extend into the metastable time scale because K4”, which is electrostatically bonded to an analyte, may not be able to promote fragmentation as promptly as a covalently bound proton. The metastable mass spectra of the [M-l-K]+ precursor ions of the cationized H- VGVAPG-OH peptide formed by KflDS-by-FAB and K1 salts in glycerol are shown in Figures 63a and c, respectively. A comparison of these two metastable spectra shows peaks representing the same types of ions with very similar intensities indicating that both [M+K]+ species undergo the same fragmentation mechanisms since both types of [M+K]+ precursor ions have very similar internal energies and represent a similar distribution of isomers with K+ bound to different sites. However, there are some minor exceptions. The K+IDS-by-FAB metastable mass spectrum in Figure 63a shows one dominant peak at m/z 365 representing the [b4-H+K]+ fragment ion whereas in the metastable mass spectrum using KI salts in Figure 63c, this same peak is much less intense. Also, the peak representing the loss of C02 in the K+IDS-by-FAB metastable mass spectrum is also much more intense than that obtained when using KI salts. When KI salts are added to glycerol, K1 is probably desorbed intact into the selvedge region and through multiple, therrnalizing collisions, K+ is attached to glycerol and exchanged multiple times, losing its kinetic energy, and ultimately attached to a variety of basic sites in equal numbers to form the isomeric and very stable, low-energy [M+K]+ adduct ions which may differ slightly from those involved in K+IDS-by-FAB [97]. Here, the bare K+ metal ion may gain up to 2 eV when it is injected in or near the selvedge region from the gas phase which will limit its collision cross section with the peptide; these high-energy K+ ions, on average, will then preferentially only stick to the most basic regions with exposed, high electron densities such as the C-terminus carboxyl group oxygens, or the proline amide nitrogen and the two neighboring amide carbonyl oxygens in the rigid 234 R 521] 317 ‘ [b -H+K]+ i: ‘ a) 4365 [M-l-K]+ a l KHDS-by-FAB of H-VGVAPG-OH l j . . [b5+OH+K]+ 3 E , [M+K]+ metastable decomposrtron -C0 ' v ‘ [y5+H+K]+ 495 j e 438 e 494 l I +. I ‘ yz [04+K] [05-H+K]+ 445 430 -V l {‘ j 173 H [2(163gH+K]+ 338 43: l e , 244 H K + b4 U4+H+K1 n . [8+ Eli 327 381 40322 s . , i 1 t . Y 200 300 400 500 m [a3-H+K]+ gym“ [b H+K1+ Lufmxr, b4 H3C\ {CH3 H3C\ CH3 [05-H+K]+ c 0 [c3+H+K]+ ”2 b) I I 0 cg (I! C H C\H2 / V \ {C HzN (If \Céz Nfi II 1 N 2 OH O 0 A113 0 [y5+H+K]"’ [y4+H+K]+ )3 )7; R 1 521 5397 f j c) H-VGVAPG-OH & K1 in glycerol '1} .t a J [M+K]+ metastable decomposition [NH 1 E . [b5+on+K]+ r v + -V e Us+H+Kl 494 _ + 1 ”’4 Hm] [a5-H+K]+ rt :1 ‘ + 445 , yz [a -H+K] +' e ‘ 173 3266 104;? 365 434 480 n 3 [C3+H+K]+ ; | 408 § , )‘3 311b4 [y4+H+K]+ 438 1 244 32 381 422 I ‘ b y . 3 100 200 300 400 500 m/z Figure 63. The [M+K]+ adduct ions of H-VGVAPG—OH (MW 498) in glycerol formed by a) KflDS-by-FAB and c) FAB of 0.1 M KI salts in glycerol were selected as the precursor ion to undergo metastable decomposition. In b), the structure of H-VGVAPG-OH is shown with the prominent fragmentation sites. The * represents a glycerol contamination peak. 235 molecular region around the proline residue of this peptide. The valine side chains probably interfere with K+ binding at the N-terminal half of the peptide. These preferred sites are supported by loss of C02 and the intense [b4—H+K]+, respectively, assuming K+ charged-enhanced (or -polarized) or -induced chemistry. In K‘TDS-by-FAB, perhaps the attachment of higher kinetic energy K+ ions induces the formation of the abundant [b4- H+K]+ ions. Charged-enhanced or -induced mechanisms have been proposed for amide skeletal bond cleavage and [b4-H+K]+ formation [246]. K+ can polarize the amide bond in such a way to form the intermediate, [b,,-H]---K+---[y,,-H], with the loss of H2 from hydrogens on the two neighboring a-carbons through a six-membered ring, and because the K+ affinity of [b,,-H] molecule is greater than the [y,,-H], [b4-H+K]+ ions are formed. The appearance of the [b5+OH+K]+ rearrangement ion further supports K+ being bonded to the carbonyl oxygens at the C-terminal half of the peptide because K+ will polarize the carbonyl group making the carbon more positive enhancing this site for C terminal -OH nucleophilic attack [188, 203]. In general, the [M+K]+ precursor ions selected for CAD have low initial internal energies and would not normally fragment before the electric sector of the JEOL HX-l 10 without collisional activation which means these ions were probably thermalized after their formation in the selvedge region by subsequent third-body collisions. CAD of the respective precursor [M-rK]+ ions, attenuated by 50%, formed in KHDS-by-FAB and from KI salts for the same peptide H—VGVAPG-OH are shown in Figure 64a and 0. CAD essentially increases the internal energy content of the peptide which initiates the fragmentation that requires higher energy as shown by the intensity increase of the peak representing the [a 3-H-t-K]+ ion in both CAD spectra in comparison to the metastable spectra. In the KflDS-by-FAB CAD spectrum in Figure 64a, the peaks representing [b4—o H+K]+ and loss of C02 are still intense in comparison to those obtained when using KI salts providing more evidence that 10‘ in KHDS-by-FAB may attach preferentially to the amide bond between alanine and proline and to the C—terminus, respectively. For low- w~wwao~=H G Na+ > K+, to sites involving several carbonyl oxygens which may alter the native peptide conformation and induce multiple types of fragmentation mechanisms through polarization. For example, the [b5+OH-t-Cat(ion)]+ is more intense in the metastable spectra of [M+Li]+ and [M+Na]+ than in the metastable spectrum of [M+K]+ because these metal ions can polarize a carbonyl group more effectively than K+ [203]. The peak representing the [b5+OI-I+Cat]+ is not in the metastable spectrum of protonated H- 238 R a H-VGVAPG-OH in glycerol 327 499 f [M+H]+ metastable decomposition b4 [M+H1+ a . t i 1 v 1 e i I {1 y2 e 173 n )3 is . 244 b5 -CH4 1 1 b3 0 424 483 4 Y4 YS -v y 229 e 4101 ...1 '100' 200' ' 300' " 400' 'rrn5/(Z)0 R b H-VGVAPG-OH & Lil in glycerol :4 8 489l_505 fr‘ [M+Li]+ metastable decomposition [b5+OH+L‘] [M+Li]+ a . + E . [C3+H+Li]+ 3135mm] V . [b3—H+Li]+ [a4+Li]+' Ly5+H+L11+ e yz 406 I 1 73 [b4-H+Li]"' n v +H+L 01+56 463 ii . L 3 I LV4+H+L11+ * I n 13 l 7 [a -H+L31]+ b4 i’ j a] 3 J25279 327.33 72 1 1 , “In-1L1. I ..l . 100 W200 300 R C) H-VGVAPG-OH & NaI in glycerol 21 + it [M+Na]+ metastable decomposition [b 5+OH+Na1+ [M+Na] 505 ii 1 [a4-H+Na]+ [b4-H+Na]+ i; . [63+H+Na]+ U4+H+Nal+ ‘ C . I _ Lb3+Na1+ [“5'H+Na]+ I n 418 Lv5+H+Na1+ t Lv3+H+Nal+ ; -V [e] [X4-H+Na]+ l 478 S - -H N + ' i )1.) [a3 +221032149:}5\)11*[:45_8_H+Na]+ ‘ y 173 3 26'6277295 M] M4]: l ‘ 1 . '. 1 '24.? L ILL .jlili." H11. '41 iihl. 100 200 300 400 500 m/z Figure 65 . Metastable decomposition mass spectra of the a) [M+H]+, b) [M+Li]+, and c) [M+Na]+ precursor ions of the peptide H-VGVAPG-OH (MW 498) in glycerol, 0.1 M LiI in glycerol, and 0.1 M NaI in glycerol, respectively. A * is 3 glycerol contamination peak. 239 VGVAPG-OH which suggests that protons are either not readily bonded to this carbonyl oxygen or do not induce this type of chemistry. In Figures 66a-c and Figure 64b, the CAD mass spectra of the protonated and Li+, Na+, and K+ cationized H—VGVAPG-OH precursor ions, attenuated by 50%, are shown, respectively. The [b5+OH+Cat]+ peak becomes less competitive in comparison to the more intense peaks representing the [a 3- H+Cat]+ ions and the product ions not containing a metal cation. This latter point is especially true for CAD of [M-l-Li]+ and it suggests that [M+Li]+ is energized in CAD, the energy gain cleaves the M-Li+ electrostatic bond, a free, "high kinetic energy" proton attaches to M to form the [M+H]+ ion as indicated by the peak representing [M+H]+ at m/z 499 in the CAD spectrum of [M+Li]+ in Figure 66b, and this [M+H]+, if it has ample internal energy, may undergo unimolecular decomposition to form the regular sequence ions. In comparison to Li+ and Na+ cationization, the amount of fragmentation falls off with K+ peptide cationization because the binding energy of K" is much weaker limiting useful fragmentation. In conjunction with the mechanistic and thermochemical studies of fragmentation for protonated peptides presented in Chapter four, fragmentation from metastable decomposition and CAD of K+ cationized peptides were also analyzed for comparison to charged-site initiated [M-l-H]+ fragmentation because the unimolecular decomposition of the [M+K]+ would represent low-energy, remote charged-site fragmentation. The [M+K]+ of methionine enkephalin, H-YGGFM-OH (MW 573), formed using KI salts in glycerol, was the selected precursor ion for metastable and single collision CAD mass spectra shown in Figure 67a and c, respectively. These [M+K]+ spectra represent sequential scans for the same experimental run; the metastable scan was obtained first and then He collision gas was administered for the second CAD scan. The [M+K]+ peaks are on the same scale so the R15 of the peaks could be compared and subtracted. In the single-collision CAD mass spectrum, no new significant peaks are observed in comparison to the metastable mass spectrum, and a subtraction of the metastable peak 240 R a) H-VGVAPG-OH in glycerol 327 499 f CAD of [M+H]+ b4 [M+H]+ E’ ' 3'2 1 173 V C I ."3 n 244 E n b3 5 a, 256 _H o l 72 b2 “3 4b5c V4§1 t p a; 157 223 C3 “4 Y4 -v Y j ”I 129 1 I 273' 499 343 4004 51456 I 0 "100" '200""300"" 400' 500 m/z y2 H-VGVAPG-OH & Lil in glycerol 5 5 7 . § 1b) 3 CAD of [M+Li]+ IM+LII+ L J [C3+l~l+l.i]+ t 4 [V +H+Li]+ _ + i; j a, 3 250 [b3-H+L1]+ _ + [a5-H+Li]+ ”13;” 1 72 -H L. + [04-H+Ll] . + t; [a3 +2 311] [b4-H+Li]+ Ly5+H+L11 . n L“4+H+Li]+ [b5+OH+Ll]+ t e - n S i . t Y 1 - 100 200 300 400 500 “12 C H-VGVAPG-OH & NaI 250 + 5123 g I in glycerol [a3‘H+Na] -V [M+Nal+ 1 CAD of [M+Na]+ [b4-H+Na]+ [i'4-H+Na]+ 47 a -CH4 E I b4+H+Na1 [ H N 1+ 5(‘5 V _‘ H N + 5 05‘ + a e [c3+ + a] _ + [64+H+Na] I418 [b5+OH+Na]+ I [X3-H+Na]+ [“4 H+Nal n 292 32‘ [x4- -H+Na]+ “W t ILi5+H+Na1+ C . N + [s] 1 L H+Nbli+H+ a]\ I722”? H+Na]+ \' ' a i 3 236 \ 307 391 II 448 490 3’ i “I 136152 )2 193 l l 266 I [55913]“ i 4 72 295 78 406 .1, 1 . I ”I 173 I208 I LYII II JLIIL I I'It II 11‘: I111I I!l111.' [J I. III 100 200 300 400 SOOmIz Figure 66. CAD mass spectra of the a) [M+H]+, b) [M+Li]+, and c) [M+Na]+ precursor ions of the peptide H—VGVAPG-OH (MW 498) in glycerol, 0.1 M Lil in glycerol, and 0.] M NaI in glycerol, respectively, at 50% beam attenuation. 241 R 51 a) Methionine enkephalin 612] f I [M+K]+ metastable decomposition £35 [M+l(]+ f 41 [b4+on+x]+ CH3SCH3 1 1 431 550 v 1 -NH3 e 3' 595 I 1 1‘ 2; [a4-H+K]+ c ‘ [C3+H+K]+ 435 n I -5 “WNW [05-H+K]+ { 1 x+ y 1 319 [bz-H-rK]; 2Y2? 0 1 . , r 1 , 39 .1 L 0 100 200 300 400 500 600 on m/z H3 [b4+Ol-I+K]+ [b4-l-1+K]+ [b2-n+x]+ [b3-l-l-l-K]+ [a4-H+K]+ H2 1’) (1:2 ‘1’ 11,1133?“ ... 1121/ 8 wise; [y4+l-I+K]+ [y3+H+K]+ 5 R j C) Methionine enkephalin V -Y 612l+ I: 1 Single collision CAD of [M+K]+ 505 [M+1(] a 4‘ [b4""0H+K]+ 'CH3SCH3 t ‘ 481 550 l I -NH3 z 3: 595 I : [04-H+K]+ " I 435 t 2‘ g 1 [b3-H+K]+ is 1 11:; [a3+K]+' 44 [05-H+K]"" t y [b -H+K]+ 0 1 - - h. c .2 33.9.12 0 160 Y ' 260 I v ' T 300 400 500 600 m/z Figure 67. a) Metastable decomposition mass spectrum and c) single-collision, high-energy CAD mass spectrum of the [M+K]+ of methionine enkephalin (MW 573) formed by FAB of 0.1 M KI salts in glycerol. The methionine enkephalin structure and prominent fragmentation sites are shown in b). Both [M+K]+ precursor peaks are on scale. 242 R15 from the CAD peak RIs results in the same peaks with much lower RIs representing the "true" CAD processes (not shown). The lack of new, structurally informative fragment/product ions, the presence of peaks representing only small neutral losses, and minimal increases in peak relative intensities in the CAD mass spectrum suggest that K+ is only weakly bound to methionine enkephalin and unable to induce structurally significant fragmentation. In contrast to K+ cationization of H—VGVAPG-OH which showed diverse fragmentation, perhaps the large, bulky aromatic side chains prevent strong K” bonding to M. The types of fragment/product ions represented by the spectra indicate K+ is attached to the basic sites in the area of the phenylalanine and methionine residues. The loss of tyrosine (-Y) can be realized by viewing the biologically active form of the peptide made from semiempirical calculations [247]. It was estimated that methionine enkephalin folds inward so the charged N- and C- termini interact and the two aromatic side chain rings and the methionine side chain extend outward in opposite directions from the backbone. Essentially, if K+ was bound to multiple sites on the backbone, it would be in the vicinity of all the side chains. Unfortunately, no exact gas- phase structures or stereochemical information for methionine enkephalin or similar peptides are known. By knowing the secondary and tertiary structures of peptides in the gas phase, the types and relative abundance distributions of fragment/product ions, observed in the mass spectrum, that undergo low-energy fragmentation processes involving secondary interactions may be better explained. Thus, the cationization of biomolecules with M+IDS-by-FAB or metal ions in the FAB matrix have many analytical advantages in FAB such as molecular weight confirmation in one and multiple component samples, biomolecule functional group analysis, and complementary fragmentation to that obtained with protonation. The latter advantage can aid in the structural elucidation of biomolecules, determination of metal ion binding sites on the analyte, and determination of the analyte gas-phase conformation. M+IDS is a good testing ground for metals used in M+IDS-by-FAB and it provides 243 information complementary to gas-phase ion/molecule studies using ICR techniques. M+IDS, M+IDS-by-FAB and metal-ion containing salts may provide additional fundamental information to help solve the mechanistic questions for ion formation and dissociation in FAB. In the quest for obtaining binding site and conformational information for biomolecules, perhaps a combination of multinuclear NMR studies in 90% water and computer molecular modeling studies would provide starting points for extracting such information from the observed fragmentation patterns in the mass spectra. APPENDICES APPENDIX ONE 244 Injection of Reagent Ions into the Selvedge Region in Fast-Atom Bombardment Mass Spectrometry Jason C. Rouse and John Allison Department of Chemistry, Michigan State University, East Lansing, Michigan, USA A divided probe that incorporates a potassium aluminosilicate glass target and an analyte/glycerol matrix target, spatially separated, was used to inject potassium ions (K‘) into the high-pressure ”selvedge“ region formed above the analyte/glycerol matrix target during fast-atom bombardment (FAB); [M + K]* adduct ions that represent the types of gas-phase neutral molecules present in the selvedge region are observed. Computer model- ing assisted in designing the divided target and an additional ion optical element for the FAB ion source to optimize interactions between K ’ ions and the desorbed neutral molecules. The capability of injecting K" ions into the FAB experiment has utility in both mechanistic studies and analyses. Experimental results here are consistent with a model for the desorp- tion / ionization processes in FAB in which some types of neutral analyte molecules are desorbed intact and are subsequently protonated by glycerol chemical ionization. Unstable protonated molecules undergo unimolecular decomposition to yield observed fragment ions. The use of K ‘ cationization of analytes for molecular weight confirmation is demonstrated, as well as its utility in FAB experiments in which mixtures are encountered. (I Am Soc Mass Spectrom 1993, 4, 259—269) ast-atom bombardment (FAB) [1] is clearly a use- ful process for generating gas-phase ions from analytes that were not previously amenable to mass spectrometric analysis. The processes leading to the conversion of condensed-phase analyte to gas-phase ions are generally believed to depend on whether the analyteexistsinneutralorionicformintl'iematrile]. If, for example, a polar, nonionic analyte (M) is placed in the matrix, glycerol (G), with a molar G/M ratio of 1000:1 or more, desorption and ionization are proposed to occur sequentially [3]. The fast-atom beam, imping- ing the sample on the surface of the target, generates gas-phase G and M molecules above the sample sur- face, in the so-called selvedge region. A small fraction of the molecules may be ionized by subsequent fast- atom /desorbed molecule collisions or desorbed di- rectly in ionic form. Because most of the liquid target is composed of G molecules, most of these ions are related to G. Thus, essentially glycerol chemical ioniza- tion (Cl) conditions are created. Initially formed ions related to glycerol undergo many collisions with G molecules to generate the [736 + H]’ ion series, as well as leading to protonation of the analyte, forming [M + H]‘. Thus, a gas-phase selvedge model explains Address reprint requests to John Allison, Department of Chemistry. Miclugan State University, East Laming, M] 68824. C 1993 American Society for Mass Spectrometry 1044—0305/93/5600 the formation of protonated analyte molecules, and it is consistent with the body of literature on proton affinities [4]. If the analyte is a salt, such as a tetra- alkyl ammonium halide, R4N‘X’, which ionizes in glycerol, then fast atoms induce direct desorption of an ion such as [R‘Nlfl No further chemistry is needed to generate an ion representative of the analyte, although adduct formation may occur as this ion passes through the selvedge region, to form ions such as [R,N + 61‘. The analysis of a FAB spectrum is similar to the analysis of any Cl spectrum, and similar questions must be addressed. Does the peak at highest mass-to- charge ratio value (that is not related to the matrix) represent the protonated analyte [M + H]‘ or the rad- ical cation M", or are such forms of the analyte unsta- ble? Does the peak at highest mass-to-charge ratio value represent a fragment of the protonated molecule such as the [(M + H) — H201‘ species commonly seen for alcohols? Such questions become even more exten. sive because biological analytes may contain salts, and one may encounter adduct ions such as [M + Na]‘ as well. Also, when salts are present in high concentra- tions, ions such as [M - H + 2Na]‘ [5] may be ob- served. The choice of matrix in using FAB is a second complication that is obvious when one realizes that the differences in proton affinities of analyte molecules and matrix molecules may determine whether any ions Received October 11, 1991 Revised October 12, 1992 Accepted October 13, 1992 245 260 ROUSE AND ALLlSON representing the analyte will be formed. Third, there are many complications when mixtures are (intention- ally or unintentionally) analyzed by FAB because it is a ”batch" method. The ratio of ion currents represent- ing the two components may be a poor representation of their relative concentrations in the liquid target. In some cases, this may be due to the preferential proto- nation of one component in the gas phase, owing to its relatively high proton affinity. In other cases, the pref- erential response for a compound may reflect the fact that it is surface active and thus is desorbed to a greater extent than other analytes in the matrix. We demonstrate here that K’ ions can be injected into the selvedge region of the FAB experiment and that this capability will allow the neutral molecules formed by particle bombardment to be identified. Such a capability not only provides useful information on the mechanism of FAB, but it also provides useful analytical information. This work is an outgrowth of the technique of K‘ ionization of desorbed species (K ’ IDS) [6] that was developed in this laboratory. The K‘IDS technique uses K‘ as the CI reagent ion for identifying molecules desorbed after the rapid heating of thermally labile analyte molecules. Potassium ions are very useful for such work because they do not induce fragmentation on interaction with organic molecules [7], cannot react by charge transfer owing to the very low ionization energy of potassium atoms. and only form adduct ions. Thus, they provide a straightforward probe of neutral molecules generated in an ion source. Because we use K‘ here to investi- gate neutral molecules desorbed by particle bombard- ment in the FAB experiment, we refer to the technique as K‘lDS-by-FAB [8]. This work is not designed to prove that the selvedge region exists, nor are we attempting to define its di- mensions. Desorbed neutral molecules and desorbed ions are present in the ion source during a FAB experi- ment, and we can use gas-phase chemistry to sample the desorbed neutral molecules. Whether this corre sponds to sampling the selvedge region or not de- pends on individual views and definitions of the term [9-12]. To inject ions into the selvedge region of a normal FAB experiment, we investigated the use of a divided FAB target. One-half of the target would contain the typical glycerol/analyte mixture, and the other half would contain a material that, when subjected to FAB, would yield copious amounts of K‘ ions. Although Munster et al. [13], Miller et al. [14], and Michaud et al. [15] successfully demonstrated the use of divided targets to explore gas-phase ion formation mechanisms in FAB, the interactions of ions derived from one side of their targets with neutral moleoiles derived from the other side usually resulted in ion signals of much lower intensity than those that could be obtained when both analytes were mixed on a single target. We will provide some insight into this observation in the fol- lowing sections. JAmSoc MassSpectmm 1991‘. 259-269 Experimental Instrumental Parameters All mass spectra were obtained on a JEOL FIX-110 double—focusing, forward geometry mass spectrometer equipped with the JEOL FAB gun, the JEOL JMA- DASOOO data system, and the combination JEOL field desorption /FAB (FD/FAB) ion source. For all experi- ments, the accelerating voltage was 10 keV; the resolu- tion was 1000 (10% valley definition); and the scan rate was from m/z 1-1500 per minute. The fast atoms used were 6—keV xenon atoms. A divided target was con- structed by highly modifying a JEOL FAB target, as is described later. It was placed on the end of the JEOL direct-insertion probe. Computer Modeling of Ion Optics The dimensions of the target and ion source optical elements were modeled using the computer program SIMION PC/PSZ Version 4.0 [16]. "Electrodes" were scaled to fit the maximum potential array size of 16,000 points to ensure accuracy. The symmetry used was planar, nonsymmetrical. The set of electrode voltages used for each SIMION model was obtained by measur- ing the actual potentials of the ion optical elements in the JEOL FD/FAB ion source during a typical K ‘IDS- by-FAB experiment. The models were refined until the maximum voltage deviation in the array was 1 X 10“ V. Identical results were obtained with MacSIMION 2.0 [17], in which a more accurate, 330,000—point poten- tial array was defined. Sample Preparation Stachyose, kassinin, cholic acid, digoxin, and bradykinin (Sigma Chemical, St. Louis, MO); polyethy- lene glycol 600, benzyltriethylammonium chloride, and glycerol (Aldrich Chemical Co., Mikwaukee, WI); thevetin (K & K Laboratories, Cleveland, OH); and KNO3, A120,, and SiOz (Johnson Matthey Co., Ward Hill, MA) were used without further purification. Ana- lytes were dissolved either in methanol (J. T. Baker, Phillipsburg, NJ) or a 1:1 mixture of methanol and MilliQ water (Millipore, Bedford, MA). The sample concentrations ranged from 2 to 10 ug/uL. Spectro- scopically pure glycerol was coated on either the K ’ IDS-by-FAB analyte / glycerol target half or the JEOL FAB target, and 1 uL of sample solution was trans- ferred to the respective target and mixed with the glycerol. The FAB spectrum of each compound was obtained with an alkali metalofree JEOL FAB target. Results and Discussion K ‘IDS-by-FAB Target Development The experimental design for injecting K‘ ions into the FAB experiment is conceptually very simple, but the 246 J Am Soc Mass Spectrom 1993, 4, 259—260 most straightforward divided target designs that were investigated gave only limited success. Successful K’ ion injection in the FAB experiment for sampling de- sorbed neutral molecules is highly dependent on both the target geometry and the arrangement of, and po- tentials applied to, the ion source optical elements, because the region of highest pressure presumably occupies a small region of space above the sample target. With the aid of the ion optics computer model- ing program, SIMION PC /PSZ, a functional K’IDS- by-FAB target was developed. Our experiences and insights are summarized here. The design considerations of a divided target for K’ lDS-by-FAB required that the FAB beam strike both a solid surface that would yield K‘ iom and the matrix/analyte target simultaneously, resulting in the injection of desorbed K ’ ions into the gaseous selvedge region above the matrix / analyte target. The [M + K]’ adduct ions formed by this process would then be extracted and mass analyzed. A common way to gen- erate gas-phase K * ions with FAB is to bombard a KCl film. Unfortunately, FAB of KCl yields KnCl; cluster cations in addition to K’, and the KCl film on the target needs to be regenerated after each experiment. One key component of the K’IDS-by-FAB experiment is the generation of K’ ions. In K’lDS, a K’ alumi- nosilicate glass emitter with the molar composition KzozAl203:25i02, generates copious amounts of gas- phase K‘ ions; surprisingly few neutral molecules are emitted from such materials (5 10%) when heated [18]. FAB of the same K’ glass on a target surface at room temperature also yields continuous, copious amounts of gas-phase K’ ions, with no evidence of charging [19]. The FAB mass spectrum of the K' glass (Figure 1) was obtained by exposing the K’ IDS-by-FAB target, without a sample or glycerol present, to the fast-atom beam. The dominant peak represents the reagent ion K‘. The small Na‘ ion signal at m/z 23 in Figure 1 arises from impurities in or on (or both) the K‘ glass emitter and target metal surfaces. The peak at m/z 52 represents Cr’ sputtered from bare stainless steel surfaces. We believe that K’ ion generation via the K‘ glass emitter is the optimal approach because copious amounts of K’ ions are generated exclusively, (“’030‘3— .("'.’.' l l 1 l l .‘a. .i. .. on Figure 1. FAB mass spectrum of the K‘lDS—by-FAB target without matrix or analyte present. REAGENT lONS [N THE SELVEDGE IN FAB MS 261 and a single K‘ glass target can be used for many experiments. The first K’lDS-by-FAB work used a simple wall- divided target [20] that is shown in Figure 2a. We concluded that this wall-divided target was unreliable, regardless of wall height (0.1-1 mm). Most often, this target resulted in little [M + K]‘ adduct ion forma- tion, and either the K ’ IDSby-FAB spectra contained a peak representing the K’ ion only or pealG represent- ing the normal FAB matrix and analyte ions, depend- ing on tuning parameters. In Figure 3, a representative K‘ IDS-by-FAB mass spectrum for glycerol is shown in which adduct ion formation occurred. Such a spectrum is obtained when the K’ glass emitter has been con- taminated by condensed-phase material. An identical mass spectrum is obtained if glycerol is intentionally placed on the K‘ glass emitter and subjected to FAB. Apparently, the analyte/ matrix can be sputtered onto the K' emitter by the fast-atom beam if the wall separating the two targets is too low. Also, the ana- lyte/matrix can "creep" around and over the wall because the two target surfaces are physically con- nected. The abundant [G + KJ’ adduct ion may be due in part to the desorption of the [G + K]’ ion directly from the surface of the K‘ glass emitter. The formation of the [G - H + ZKl’ adduct ion is the primary indicator that a condensed-phase interaction between K’ glass and analyte/ matrix has occurred. If the [G — H + ZKl‘ ion appears in the mass spectrum, we cannot guarantee that the [G + K]’ adduct ion, as well as any K’ adducts of analyte molecules observed, are formed in the gas phase. To decrease the possibility of cross-contamination, as demonstrated by the wall-divided target, we de- signed the divided target shown in Figure 2b. With this “space-divided" target, the K’ emitter and sam- ple support are separated in space, but are close enough so that both targets can be bombarded simultaneously by fast atoms, and desorbed K’ ions can interact with the desorbed neutral molecules above the sample tar- get. The sample stage on a JEOL FAB target was bent to a 30° angle, and rhenium wire (0.007 in.) connects the K’ glass emitter to the target post. Unfortunately, this space-divided target has also performed inconsis- tently in converting a substantial fraction of K’ ions into [M + K]* and [G + K]+ adducts; however, it did ‘ .o e“. . ' . a c ., A A .\ \Ih“ .$ “N3 Figure 2. Schematics of (a) the K‘IDS-by-FAB wall-divrded target and (b) the K’ lDS-by-FAB space-divided target. 247 262 ROL’SE .wo ALLISON . 1 ilooxi' { a I 1 1 It", '. j . t 1 [ I": a 3 ' l n ' . 3' ' ..‘. IMO 1 mm. : I , _ l m-II' :3 t J l .' 204 1 IWJ’ j t ' r I I i l .‘ 1 I ' j _ l ' “ . a; I ‘ ' m ll L A ll. L A a L L 1 DA A l O 1. - ‘ an M Figure 3. K'IDS-by-FAB mass spectrum of glycerol when the K’ glass emrtter is contaminated wrth matrix/analyte. solve the cross-contamination problem that plagued the wall-divided target because [G + K]‘ adduct ions could be formed without [C - H + 2KJ’ ion forma- tion. The ion optics computer modeling program SIMION was used to gain insight into the ion trajectories from K’lDS-by-FAB targets to improve performance. Rep- resentative computational results are presented below. The utility and limitations of the SIMION investiga- tions should be clearly stated. "SIMION is a personal computer program for designing and analyzing charged particle lenses, ion transport systems, various types of mass spectrometers and surface probes that utilize charged particles” [21]. We are using it to gain insights into the influence of ion source modifications on the potential surface on which ions move within the ion source. The results presented here should not be interpreted literally, but are meant to assist in the design and analysis of the experiment. Obviously, the experimental goal is to generate ions from near a solid surface, and have them"'sweep” through a region of space close to the matrix surface. The ionic trajectories are determined by the fields within the source. The best way to appreciate the topology of the electric fields is to follow some ion trajectories. This we do in the figures presented. In these trajectory studies, ions originate near a surface with an initial kinetic energy of zero. Clearly this is not an accurate description of the ions; however, ions with no initial kinetic energy are most sensitive to the potential surface on which they move. The other option would be to compute thousands of trajectories for a variety of possible initial kinetic energy vectors. Although such a rigorous ap- proach may be informative, it must be realized that any trajectories computed cannot be accurate because SIMION will not allow ion-neutral molecule collisions to be considered. The point of the experiment is to inject ions into a high-pressure region of the ion source where collisions occur. Thus, we use SIMION to sug- gest physical possibilities. The SIMION results selected to be shown here are relevant in that they correlate well with what is observed. They correctly predict which variations of the K‘IDS-by-FAB experiment should work; they predict qualitative tuning variations and sensitivity changes that are observed; and the J Am Soc Mass Spectrom 1993. 4. 259-269 combination of SIMION and experiments led to an experimental. design that was effective. Thus, it should be realized that the SIMION results are shown to present insights into what could happen, the shape of the potential energy surface in the ion source, and how it can be distorted to achieve the desired result—the sampling of desorbed molecules by K’ attachment. SIMION was first used to consider the optimization of the spatial overlap between the K’ ions from the glass emitter and the neutral molecules desorbed from the analyte/matrix target. In Figure 4a, a SIMION model of the normal FAB experiment that incorporates an ordinary JEOL FAB target placed in the JEOL FD / FAB ion source is shown. The FD / FAB ion source electrodes and measured potentials are as follows: FAB target (repeller), 10,050 V; cathode (L1), 10,010 V; first focus (L2), 9000 V; second focus (L3), 600 V; deflector (L4), 30 V; and total ion monitor plate (L5), 0 V. The L3, L4, and L5 electrodes were included in the SIMION models, but are not shown in the figures. The FAB beam is also not shown, but it would originate from above the figure and strike the target with an average angle of incidence of 70°. The slits in L1 and L2 are 1 and 2 mm wide, respectively. Here, trajectories for eleven 100-Da cations, with initial energies of 0 eV, are shown starting at various locations near the FAB tar- get. Of the 11 trajectories, only nine ions from the target area are successfully extracted by the penetrat- ing electric fields (shown by the contour intervals labeled Cl-C4). These ions pass through both L1 and L2, but the beam sharply diverges after 1.2. In Figure 4b, a SIMION model of the K’IDS-by-FAB wall- divided target is shown. Twenty-three total trajectories are shown; 11 ions originate on each half of the target. Figure 4. SIMION models of (a) the JEOL FD/FAB target and ion source; (b) the K’ lDS-by-FAB wall-divided target; and the K’ lDSbyoFAB space-divided target showing the fate of (c) ana- lyte/matrix ions and (d) of K’ IOX'LS. Contour lines representing the extraction field for the potentials 10,040 V (C 1 ); 10,030 V (C2); 10,020 V (C3); 10,010 V (C4); 9500 V (C5); and 9000 V (C6) are shown. Source elements and potentials are described in the text for this lO—keV ion source. 248 J Am Soc Mass Spectrom 1993.4, 259-269 and one originates on the wall. Unfortunately, only the ion that originates from the center of the target was successfully extracted. Our experimental results with the K‘IDS-by-FAB wall-divided target coincide with the SIMION results shown here. The unreliable nature of the wall-divided target is illustrated by the SIMION model because the K‘ ions from the K' glass emitter surface diverge away from the selvedge region above the opposite target surface, preventing [M + KJ’ adduct formation. Because the contamination area of this target, discussed previously, is surely centered around the wall, the SIMION model suggests that ions from this central region would be easily exnacted. It is also clear from the SIMION model in Figure 4b why different tuning parameters were required to optimize extraction of analyte / glycerol ions, K‘ adduct ions from the wall area, and K‘ ions in the FD/FAB ion source. Three different sets of potentials were required for the optimum sampling of ions that originate from the three different regions on the target (left, center, and right). Thus, the SIMION models agree well with experimental observations. Figure 4c and d show the SIMION model of the space-divided K‘IDS-by-FAB target. Figure 4c shows the trajectories of 19 cations of mass 1000 Da and with 0 eV of initial kinetic energy originating from various locations near the stainless steel sample target. Twelve of the 19 ions are successfully extracted from the target area. These ions traverse both L1 and L2, but the beam is poorly defined, which is consistent with observed tuning problems. In Figure 4d, the trajectories are shown for K’ ions originating from the K’ glass. All but two trajectories terminate on the source element L1; however, at least these K’ ions do follow favorable trajectories, passing close to the presumed selvedge region. Experimental results with this spacedivided K‘IDS-by-FAB target are consistent with the K’ ions trajectories, as shown in Figure 4d. The [M + K]’ adduct ion formation occurred more often with this design than with the wall-divided target, but the [M + K]' adduct ion signal was still usually weak in com- parison to any FAB matrix/analyte ions. A potential problem with this K’IDS-by-FAB target design is that a typical K’ ion is promptly accelerated to kinetic energies of 2 eV and greater in the presumed vicinity of the selvedge region, resulting in ( 1) short residence times and (2) kinetic energies too high for adduct formation [22]. As a result, many of the desorbed neutral molecules are not sampled by low-kinetic en- ergy K ‘ ions for the set of ion optical element voltages used in this SIMION calculation. In an attempt to correct for the short residence time and high kinetic energy of the K' ions, an electrode, L0, with a center slit less than 1 mm wide and a potential equal to the target potential, was placed equidistant between the target electrode and the cath- ode electrode, L1, in the SIMION model to limit the extent of the extraction field penetration into the target area. The combination of the space—divided target and REAGENT IONS IN me SELVEDGE IN FAB MS 263 the electrode L0 resulted in the ability to perform the K * IDS-bycFAB experiment as originally envisaged, and the reasons are shown in Figure 5a—d. The SIMION model in Figure 5a shows the normal FAB target, as shown in Figure 4a, in the FD/FAB ion source that incorporates the additional electrode, L0, placed be- tween the target and L1. Again, 11 ion trajectories from the FAB target surface are shown. The position and dimensions of the LO electrode decrease the effects of the extraction voltages, as shown by the positions of the contour lines that correspond to the same poten- tials as those shown in Figure 4a. The presence of L0 in the SIMION model still results in the extraction of 9 of the 11 ions, and it produces a well-defined ion beam compared with the ion beam shown in Figure 4a. Experimentally, the FAB spectrum is identical whether [30 is present or not. Figure 5b shows the SIMION model of the wall-divided K’IDS-by-FAB target cou- pled with the additional electrode, L0. The presence of DO in front of the wall-divided target dramatically increases the number of successfully extracted ions from 1 to 15 (7 ions from each target half) by decreas- ing the penetration of the extraction field. This design was never experimentally tested because SIMION re- sults show that the K’ ion overlap with the desorbed neutral molecules above the sample surface may still be insufficient, and previous experimental results showed that K+ glass contamination readily occurs. In Figure 5c and d, the SIMION model of the space-divided K‘IDS—by-FAB target coupled with the additional electrode, L0, is shown. In Figure 5c, 18 of 19 matrix/analyte FAB ions from the target area are accelerated toward the analyzer in a well-defined beam, in contrast to the SIMION model shown in Figure 4c. In Figure 5d, all K’ ions are injected into the pre- sumed selvedge region above the sample target. Figure 5. SIMION models with the additional ion optical ele- ment L0 present (a) the JEOL FD/FAB target; (b) the K' IDS-by- FAB wall-divided target; and the K’lDS-by-FAB space-divided target, with trajectories shown for (c) analyte/ matrix ions and (d) K' ions. Contour lines representing the potentials 10,040 V (C 1); 10.030 V (C2); 10,020 V (C3); 10,010 V (C4); 9500 V (C5); and 9000 V (C6) are also shown. 249 264 ROUSE AND ALLISON SIMION trajectory calculations show that the L0 elec- trode increases the residence time and lowers the ki- netic energy of the K’ ions because the penetrating extraction fields stop sharply at L0, and essentially a field-free region is created between L0 and the target. With the L0 ion optial element present, the typical K ’ ions gain 2 eV of kinetic energy, a realistic upper limit for K’ adduct formation, in 4.5 as and are present for approximately 1.9 us in the presumed area of the "selvedge region." Without L0 installed (Figure 4c and d), typical K’ trajectories achieve 2 eV in 0.45 [1.5 and are present for approximately 0.3 as in the “selvedge region." Thus, there is substantially more opportunity for each K’ ion to react With a desorbed neutral molecule with L0 installed. In addition, a high-pres- sure region would probably be created in the small volume between the target and L0 electrode that would provide stabilization for termolecular ion-molecule adduct formation, if required. Experimentally, there were no visible traces of sample condensation on L0. SIMION results clearly indicate that no preferential tuning problems should occur with the additional L0 ion optical element installed because all K' and ma- trix/analyte-related ions are focused into a well- defined beam. Experimental results below confirm this observation because [M + KJ’ adducts and analyte / matrix ions appear in the same spectrum. We incorporated the design shown in Figure 5c and d into the K‘lDS—by-FAB experiment with successful results. Figure 6b shows the K’IDS-by-FAB spectrum of 3 glycerol sample obtained by using the space-di- ii 1 a M. ‘ O , I l u" M‘ l . V“. l . l I l 0:.- . 3 2.- W I , : < ‘ ”’3 y = Ark A .L 0 I in i. - - ’ .I I “ T r ' M' C I . IOi I I '004 O , . M. n 00-3 W ; wr : r l O o , , 1.1 - - - - - we“! - ~20 . " u m m sea as m b '8 Figure 6. (a) FAB mass spectrum and (b) K‘IDSby-FAB mass spectrum of glycerol (MW 92). J Am Soc Mass Spectrom 1993. 4. 259-269 vided target with the ion optical element L0 mounted in the FD/FAB source, showing that [G + K]' adduct ions can be formed without [G - H 4» 2K]‘ ions being formed. Thus, the analyte/liquid matrix is not being sputtered onto the K’ glass. To confirm that these [G + Kl’ ions are formed in the gas phase, one addi~ tional process needs to be considered. Although glyc- erol does not contaminate the K’ glass, potassium could be sputtered into the matrix, yielding the ob- served adducts from condensed-phase, not gas-phase, processes. The following observations and experiments sug- gest that potassium is not being sputtered from the thermionic emitter onto the liquid target in this experi- ment. First, we note that the K’IDS-by-FAB experi- ment is difficult to perform because it is very sensitive to the target design and positioning in relation to L0 and the ion source. If significant amounts of potassium were sputtered from the glass to the liquid matrix and this led to K’ adducts, the results should be much less sensitive to experimental design. For example, the de- signs shown in Figure 4b-d should yield [G + K]‘ and [M + K]‘ ions as well as those shown in Figure Sb-d; this was not the case. Second, the emission properties of alkali aluminosilicate glasses suggest that the particle-induced desorption of K(g) need not occur to a greater extent than for K'(g) owing to the nature of these materials and their performance as thermionic emitters. Third, if potassium is sputtered into the liq- uid matrix in this experiment, and this is the process by which K‘ adducts are formed, then a time depen- dence should be observed. There should be no K’ adducts formed when the FAB experiment begins, and the adduct ion signals should increase With time as sputtered potassium accumulates in the matrix. This is not observed. The K‘ adduct signals are present when FAB is first initiated, and the adduct ion signals are constant for a period of several minutes. Fourth, exper- iments have been performed to test transfer of potas- sium from the emitter to the sample target. For exam- ple, if no glycerol is added, and the space-divided target is bombarded for several minutes, potassium may be sputtered onto the bare sample surface on which glycerol is usually deposited. If the K' emitter is then removed and the bare surface subjected to FAB analysis, there is no evidence for the accumulation of potassium on this surface. This is also true when glycerol is placed on the second target. If the space-di- vided target is bombarded for a period of time, the K’ emitter removed, and the remaining glycerol analyzed by FAB, there is no evidence for detectable concentra- tions of potassium in the glycerol. In some cases, a very small peak that could represent [G + K]’ could be observed, but it was barely detectable and very much smaller than those observed in the K‘lDS-by- FAB experiment. Also, experiments that involved the addition of various concentrations of KCl to glycerol demonstrated that if potassium was sputtered into the glycerol and this process leads to K‘ adducts, then 250 J Am Soc MassSpectrom 1993. 4. 259-269 other ions should be present in these spectra, such as [G — H + ZKI‘; theseare notobserved. Probing the Selvedge Region of the FAB Experiment with K ‘ Ions Injection of K' ions into the selvedge region of the FAB experiment will allow the desorbed neutral molecules to be identified. That is, if K' ions are injected into the FAB experiment and some [M + K]' ion is formed, we will interpret this as evidence for the presence of desorbed M molecules in the selvedge region. We use this capability here to investigate the fate of analyte molecules for a variety of analyte com— pounds. Do they desorb as intact neutral molecules, or does FAB lead to desorbed molecular fragments as well? Glycerol. The ions formed by FAB of glycerol are shown in Figure 6a. It is well known that a sequence of pale representing protonated glycerol clusters, [nC + H]’,dominatesthespectrumandbeginswith n -1 at m/z 93. In addition, two prominent ”fragment iom" are represented by peaks at m/z 75 and 57, which are 18 and 36 Da below m /z 93, respectively. Consider the ion at m/z 75. There are a variety of mechanisms by which this ion could be formed. The ion may be formed as a fragment ion of G", or it may be a fragment formed from [G + H]’. Because it is an even-electron ion, it may well be a fragment formed following protonation and should be written as [(C + H) - H2015 Another possibility is that FAB imparts sufficient energy to the sample to induce dehydration. Thus, FAB could lead to the desorption of G, [G - H20], and [G - 2(H,O)] into a gaseous, protonating environment, leading to the ionic species [C + H]‘, [(G - H20) + H]', and [(G - 2(H20)) + H]’ at m/z 93, 75, and 57, respectively. When K’ ionsare injected into theselvedge region above the glycerol target during FAB, an ion series corresponding to [nG + K]’ is observed that parallels the [nG + H]’ series (Figure 6b). This indicates that K’ ions are successfully being injected into a high- pressure region and are undergoing multiple colli- sions. This process is also indicated by the ratios of the peaks representing K’ and its glycerol adducts; there is fairly extensive conversion of this ”CI reagent ion" into product ions, which can occur in a high-pressure environment The relative intensities of both the [nG + K]’ series and the [nG + H]‘ series correlate with each other and are a function of the amount of glycerol present. The members of both series with small values of n always dominate, and the low relative intensities ofthehighermassionsineachsenessuggestthatthey are formed by sequential clustering processes in a high-pressure environment [10, 23]. That is, it appears that FAB induces the desorption of single glycerol REAGENT IONS [N THE SELVEDGE IN FAB MS 265 molecules and not glycerol clusters. More important, the [nG+K]‘ ions are the only new ions formed whenK’ ionsareinjected.Therearenopeaksob— served corresponding to [(G - nH20) + K]+ ions, suggesting that the fast-atom beam desorbs glycerol molecules, predominantly intact, and does not lead to desorbed degradation products. This is somewhat un- expected because elimination of water from alcohols is a very low-energy process, compared with the other processes that occur, such as desorption (vaporization). If [G - (H,O)l molecules were desorbed, we would expect K" adducts to be formed. Although the K’ affinities for [G‘ - (H20)], a diol, and [C - 2(H20)], an alcohol, would be lower than that for glycerol, a triol. K+ has been observed to readily form adducts with small molecules such as alcohols. Thus, we suggest that the ions observed at m/z 75 and 57 should be considered fragment ions of [G + H]‘; the elimination of water occurs following protonation. Cholic aa'd. Cholic acid [molecular weight (MW) 408], a bile acid, can be considered an example in which positive-ion FAB mass spectrometry ”fails." The FAB mass spectrum of cholic acid in glycerol is shown in Figure 7a. There is a very small [M + H]’ signal at m/z 409. The dominant ions are at m/z 355 and 373. These fragment ions are even-electron ions and corre- spond to a protonated molecule in which multiple water molecules have been lost. Again, there are two extreme situations that could explain the observed C”..."I- .C-‘.'.. : “3' 1 ' as l I t j W... I . | ' l 7. M’ . a “7 1 : mm ...... l I 4 .anL" n f m- are ads as. see b WI Figure 7. (a) FAB mass spectrum and (b) K’IDSby-FAB mass spectrum of cholic acid (MW 4%) in glycerol. 251 266 ROUSE AND ALLISON spectrum. First, cholic acid molecules could be de- sorbed intact by FAB and subsequently protonated in the gas phase. Because the protonated molecule is unstable, the [M + H]‘ ions essentially completely dissociate by losing one, two, and three water molecules. This would not be unexpected for a molecule containing multiple — OH groups under Cl conditions. Also, not unexpected, would be a situation in which the polar molecule simply is not desorbed intact on FAB, but instead the dominant analyte-re- lated molecules that desorb are the less polar, dehy- drated forms [M - 3H20] (MW 354) and [M — ZHZO] (MW 372), which then become protonated to yield the ion currents at m/z 355 and 373. Figure 7b shows the corresponding K ’ IDS-by-FAB mass spectrum for cholic acid. When K’ ions are injected into the selvedge region, only one additional peak (not including those related to glycerol) appears, at m/z 447, which is the K' adduct of the intact cholic acid molecule. There is no peak observed at m /z 393, which would represent [(M - 3(H20) + Kl’. Thus, the K‘lDS-by-FAB results suggest that FAB desorbs intact cholic acid molecules that dissociate completely on protonation. Although the protonated molecule is unstable, the neutral ana- lyte molecules remain intact throughout the desorption process. The results shown in Figure 7b also demon- strate that K’ adducts of molecules are stable [24]; K‘ forms adducts without inducing fragmentation and is thus ideal for such experiments. When a cholic acid /KC1 (low concentration)/ glycerol mixture is sub- jected toFAB,an[M - H + ZKI‘ ionisformed,but in K‘IDS—by-FAB of cholic acid, this ion is not formed (see Figure 7b). Digoxin. As the size of the analyte molecules in- creases. the possibility for fragmentation accompany- ing neutral desorption increases as well. The FAB spectrum of the cardiac glycoside, digoxin (MW 780), is shown in Figure 8a. An intense peak at m/z 781 representing the protonated molecule is observed, as is a rich array of fragment ions. Do all fragment ions result from the unimolecular fragmentation of the pro- tonated molecule, or does the molecule decompose on fast-atom impact to yield an array of smaller, even- mass neutral molecules, each of which become proto- nated following dissociation? For example, the peak at m/z 651 represents the protonated form of digoxin ([(AO$,OS,OS,OH) + H]‘) minus the terminal sugar, 53, due to cleavage of a glycosidic bond, written as [(AOSjOSZOH) + H]+ using the Light-Kassel-Allison notation [25]. The aglycone is represented by A, and the sugars are designated as 5,. Does FAB of the digoxin / glycerol target lead to desorption of [AOS,OS,OH] molecules (MW 650) mat become proto- nated following desorption? Figure 8b shows the K’lDS-by-FAB mass spectrum for this analyte. Again, on injection of K’ ions into the FAB experiment, only one new analyte-related peak is formed, seen at rn/z 819, representing the K’ adduct of the intact digoxin J Am Soc Mass Spectrom 1993. 4. 259-269 1‘-.,.‘3- .C-.-.' U 8 V 8 O O 0 ‘--...‘8— .Q—-.-.. Figure 8. (a) FAB mas spectrum and (b) K’lDS-by-FAB mass spectrum of digoxin (MW 780) in glycerol. molecule. Thus, we conclude that FAB results in the intact desorption of polar digoxin molecules from the glycerol matrix; the analyte is protonated following desorption. The lower mass ions can be interpreted as true fragment ions of the protonated molecule. Benzyltriethylammoniurn chloride. When this salt (MW 227) is analyzed by FAB, using glycerol as a matrix, the resulting spectrum is dominated by the benzyltri- ethylammonium cation [CanNl’ at m/z 192. Frag- ments of this cation are also observed. We assume that this "preformed" cation desorbs directly as an ion, although one could certainly foresee, through a rear- rangement reaction, the formation and subsequent de- sorption of a neutral species, [CanN], which would then be protonated to yield a peak at m/z 192. An- other possibility is the intact salt desorbs and becomes protonated, as suggested by eq 1: [c,,HnNrcng) + [c + H]' i [acufian’CI—i + H] —~ [CanNl’ + H0 (1) When K’IDS—by-FAB is used to probe the selvedge region in this experiment, no new ions representing the analyte are formed. Thus, direct desorption of preformed ions does appear to be the correct mecha- nism for such molecules. 252 I Am Soc Mass Spectrom 1993. l. 259- 209 Peptides. The behavior of peptides in the K’lDS—by- FAB experiment varies, which suggests that the mech- anisms for ion formation should depend on the specific peptide under study. When the peptide kassinin (MW 1334) is analyzed by FAB using glycerol as the matrix, an abundant protonated molecule is formed. A portion of the K ’ lDS-by-FAB spectrum of kassinin is shown in Figure 9. The appearance of the additional peak at m/: 1373, 38 Da above the protonated molecule. repre- sents the potassium adduct of the intact, desorbed peptide. Even for an analyte of this size, FAB appears to induce the desorption of the analyte intact. Bradykinin (MW 1059) also forms an abundant pro- tonated molecule at m/: 1060; however, when the K‘lDS-by-FAB experiment is performed, and K’ ions are injected into the selvedge region, no new ions are formed representing the analyte. Apparently, bradykinin exists in glycerol in protonated form and desorbs directly as an ion. This may be due to the very basic arginine residues in the peptide. Thus, even in a simple glycerol matrix, without addition of acid or base, peptides may behave in a variety of ways, de- pending on the pK,s of the various basic sites in the molecule. Analytical Applications of K +lDS-lry—l-‘AB The obvious analytical utility of K‘IDS-by—FAB, in the context of FAB analyses, is the molecular weight con- firmation for analytes that are not preionized in the matrix. Such capabilities are demonstrated in the cholic acid example in Figure 7a and b. If this were an unknown, the 18-Da difference between m/z 355 and 373 would certainly be recognized as a water 1055: however, one would probably not be able to distin- guish the peals at m/z 391 and 409 above the back- ground. Thus, one may assume that the molecular weight of the analyte was 372. In addition to those cases in which analytes yield very small peaks in the molecular weight region of the spectrum, K‘IDS—by- FAB is also useful in this regard when the mass-to- charge ratio of the protonated molecule coincides with that of a matrix-related ion. Similar cationization stud- ies can be done by adding KCl or KI to the matrix, a marl : egg. 2: “WWW . . I r I l‘ ' s g} | 3: fl“? 0 1373 I I , s H s l i ' Q 1120 1’” 1m 1.“ "C. 1". ‘3“ '8 Figure 9. K‘ lDS-by-FAB mass spectrum of kassinin (MW 1334) in glycerol. REAGENT IONS [N THE SELVEDGE [N FAB MS 267 although this can lead to other ions in addition to an [M + K]‘ ion. A common occurrence that can ”mask" molecular weight information in FAB is the formation of ana- lyte/matrix cluster ions. For example, Figure 8a and b shows a peak at m/z 873, which is the proton-bound adduct ion of digoxin and glycerol. Usually, the 92-Da difference between the peaks at m/z 781 and 873 would be recognized as an adduct ion containing a single-matrix molecule, although the peak at m/z 873 could represent either the protonated analyte or a second component with a molecular weight 92 greater than that of digoxin. Again, the K’ lDS-by-FAB results clearly indicate the molecular weight of the compound and assist in identifying the high-mass ions that are candidates for pseudomolecular ions. Another typical situation arose in obtaining the FAB mass spectrum of stachyose (MW 666), an oligosaccha- ride. The peak representing the protonated molecule is discernible but of low intensity, comparable to the intensities of nearby matrix-related ions in the mass spectrum. Actually, in some experiments with this analyte, no protonated parent was observed at all. Its fonnation was dependent on the glycerol/stachyose ratio used. Also, sodium salt impurities in the sample led to an [M + Na]‘ peak 22 Da above the protonated molecule. Figure 10 shows the K’IDS-by-FAB mass spectrum of stachyose, which clearly allows the molec- ular weight to be verified and the assignments for the peaks at m/z 667 and 689 to be confirmed. K’IDS-by-FAB also provides useful information when mixtures are the subject of a FAB analysis. The use of FAB mass spectrometry for the analysis of mixtures always presents particular challenges, espe- cially if no chromatographic information is available for the sample. A mixture that has received much attention in the desorption/ionization literature is that encountered in polyethylene and polyproplyene gly- cols. For example, polyethylene glycol (PEG) 600 is a mixture of oligomers with the formula [H(OCH2CH2),,OH]. The distribution of oligomers is such that the average molecular weight is approxi- mately 600. One can obtain information on the oligomer distribution by FAB. If a small amount of PEG is added to a glycerol matrix, a series of protonated ‘- ... ' ”Qfifl W2 "f" 90...! '1' M'cs”rs°sl° ‘ ,. ... .... I M. W . as: a Instr M M m ass m’ or W 7:? Jill ...L 4.... ALL * <-I:O-a— o¢—-I-OI s '0' us an us rss rss Figure 10. K’ley-FAB mass spectrum of stachyose (MW 666) in glycerol. 253 268 ROUSEAN'DALLISON oligomers appears in the resulting FAB mass spec- trum. Figure 11 shows the K’lDS—by-FAB spectrum of PEG 600. The K’ adducts of each oligomer clearly stand out and dominate over all other ions formed. Again, as in the single-component analyses, FAB ap- pears to induce the desorption of all of the oligomers present in the mixture intact. In our work to understand the fragmentation of cardiac glycosides [26], we obtained the FAB mass spectrum of thevetin (MW 858) that is shown in Figure 12a. As expected, a peak representing the protonated molecule was present at rn/z 859; unexpected was a second peak 14 Dahigherat m/z 873. It israretofind high-mass ions separated by 14 Da because the loss of CH, is rarely observed in unimolecular dissociatiorts. A more reasonable explanation was that a mixture of two compounds was present with MWs 858 and 872 in approximately a 3:1 ratio, respectively. An investiga- tion revealed that there are two forms of thevetin, thevetin A and B, whose structures, as indicated on Figure 12a, differ by 14 Da. Apparently these two compounds are difficult to separate. The K’lDS-by- FAB mass spectrum of thevetin (Figure 12b) confirms thepresenceoftwocompoundsthatdifferinmassby 14 Da because two additional ions are formed on injection of K’ ions. If one looks more closely at the spectrum in Figure 12a, the relative abundances of the isotopic clusters for protonated thevetin A and B are very different, particularly at m/z 875. In fact, this same isotopic distribution also appears in the K‘IDS- by-FAB spectrum in Figure 12b, suggesting that the sample is at least a three-component mixture. The third compound may be a reduced variant of thevetin A Last, it should be noted that the K’ adduct of the intact analyte can be selected for collision-activated dissociation (CAD) or metastable analysis in the K‘IDS—by-FAB experiment. The product ions of the K‘lDS—by-FAB precursor [M + KY adduct ions, forrnedinthegasphase,arethesameasthoseob- served for the [M + K]+ ions formed when K1 is added to the peptide/ glycerol matrix. The utility of CAD for alkali cationized molecules in providing structural in- formation has been demonstrated [27] and appears to be a method that can provide complementary informa- "I ~tl eat. .73 '1‘-.30‘3- a¢--o. '0' ass sss us no ass - Figure 11. K’lDS—by-FAB mass spectrum of polyethylene gly- col 600 in glycerol. lArnSoc MassSpectrom I993. $259469 I M. Yin-sass heron 7 ~ ran-ass than, t m o—or. i 3 qumQ ' O. on 0 I 4 MM. W ' "I e '0‘ J l ” ms ‘ - l H "‘ V .1!” 1. ll: .11 . . l.. . . i. ”+1. I I“ “0 .0. II. no ‘ m : i‘MMI’ l ”7 s I I 0 maoxr a l l l s 2 {l l t . Y I ii. ““111“ l “I .2. us Figuelz. (alFABmassspectr-um and (b) K‘lDS—by-FAB mass spectrumofthevetinA(MW872)andthevetinB(MW858)m glycerol. tion to that available from CAD studies of protonated molecules. Conclusions We have demonstrated that K’ ions can be success- fully injected into the selvedge region of the FAB experiment to probe the desorbed neutral species pres- ent. The K’IDS-by-FAB technique is both a mechanis- tic and an analytical tool. Mechanistically, the desorp- tion aspects of FAB, for both analyte and matrix molecules, can be probed. Analytically, K’ cationiza- tion of analytes is useful for molecular weight confir- mation. One assumption central to the K’IDS-by-FAB data discussed here is that K’ will form adduct ions with a variety of neutral molecules in the FAB ion source and that there are no dramatic differences in adduct forma- tion rates for the molecules under consideration Cer- tainly, the details of this assumption remain to be explored. For example, in the experiments with glyc- erol, K‘ forms a complex with glycerol but not with dehydrated glycerol, and this is interpreted as suggest- ing that desorbed, dehydrated glycerol is not a major product of particle bombardment. This assumes that K’ will form an adduct with [G — H20]. At this point, we make two observations. First, K’ and other alkali ions form adducts with monofunctional molecules such as alcohols and small molecules such as propene, so that multiple functional groups are not required [7, 254 I Am Soc Mass Spectrom 1993.4. 259-269 28]. Second, polar groups need not be present in a molecule for a substantial binding energy to an alkali ion. Because the interaction is electrostatic, both the dipole moment and the polarizability must be consid- ered. ln field desorption /surface ionization experi- ments, alkali ion adducts with saturated alkanes have been observed [29], showing that such species can be stable, even when there is no dipole moment, if the size and structure of the molecule result in a substan- tial overall polarizability. Obviously, much work re- mains to quantitate this aspect of the experiment. Finally, we note that for the compounds studied to date, it is intriguing that FAB provides a source of gas-phase molecules with molecular weights above 1000 Da. This capability may eventually find applica- tions in fields other than mass spectrometry. For exam- ple FAB may prove to be a useful tool for the spectro- scopic characterization of large molecules in the gas phase in the absence of solvent. Acknowledgments This work was supported by the Biotechnology Research Pro- gram of the Divrsron of Research Resources of the NIH (RR- 00480-20), the Michigan State University Research Excellence Fund. and the Dow Chemical Company Analytical Sciences Group. We are grateful to Drs. Karen Light-Wahl, Dan Kassel. and Cary Schultz for their contributions to this project. References 1. Barber. M.; Bordoli, R 5.; Sedgwrck, D.; Tyler, A. N. I. Chem. Soc. Chem. Commun. 1981, 325. 2. FerLselau, C.; Cotter, R I. Chem. Rev. 1”7, 87, 501. 3. Cooks, R. (3.; Busch. K. L. lnt. I. Mass Spectrom. lon Phys. 1983. S3, 111. 4. Sunner, I. A.; Kulatunga, R; Kebarle. P. Anal. Chem. was, 58, 1312. 5. Mallis. L. M.; Russell, D. H. hit. I. Mass Spectrom. lon Processes 1987. 78, 147. 6. Bombick, D. D.; Allison, I. Anal. Chem. 1987, 59, 458. 18. 19. 21. REAGENT IONS N THE SELVEDGE iN FAB MS 269 . Allison, I.; Ridge, D. P. I. Am. Chem. Soc. 1979, 101, 4998. . Rouse, I. C.; Light, K. 1.; Allison, 1. Proceedings of the 38th ASMS Conference on Mass Spectrometry and Allied Topics; Tuc- son, AZ, lune 1990; p. 455. Pachuta, S. I.; Cools, R. G. Chem. Rev. 1987, 87, 647. . Sunner, I.; Morales, A.; Kebarle, P. Anal. Chem. 1987, 59. 1378. . Schroder, E.; Munster, 11.; Budzikiewrcz, H. Org. Mass Spec- trom. 1986, 21, 707. . Freas, R. 8.; Ross, M. M.; Campana. I. E. I. Am. Chem. Soc. IQS, 107, 6195. . Milnster, H; Theobald. F.; Budzikiewrcz, H.; Schrt'rder, E. Int I. Mass Spectrom. Ion Processes 1987, 79, 73. . Miller, I. M.; Balasanmugam, K.; Fulcher. A. Org. Mass Spectrom. mo, 24, 497. . Michaud, D. P.; Kyranos, I. N.; Brennan. T. F.; Vouros. P. Anal. Chem. 1990, 62, 1069. . Dahl. D. A.; Delmore, I. E. SIMION PC/PSZ Version 4.0. EGG-CS-7233 Revisron 2, April 1988. . McGilvery, D.; Morrison, R. MaélMlON Version 2.0, Octo- ber 1991. Blewett, I. P.; Iones, E. I. Phys. Rev. 1936. 50. 464. Ackermann, 8.; Allison, I.; Musselman, 8.; Pinkston. D; Bombick, D.; Dolnikowski. 6.; Tsarbopoulos, A.; Watson, I. T. Proceedings of the 315i ASMS Conference on Mass Spectrome- try and Allied Topics; Boston, MA, May 1983; p. 600. . Kassel, D. B. Ph.D. dissertation, Michigan State University. 1988. Dahl, D. A.; Delmore, I. E.; Appelhans, A. D. Rev. Scr. lnstnrm. 1990, 61, 607. . Bombick, D. D. Ph.D. dissertation, Michigan State Universrty, 1986. . Honda. F.; Lancaster, C. M.; Fukuda, Y.; Rabalais, I. W. I. Chem. Phys. 1978, 69. 4931. . Davidson, W. R.; Kebarle. P. I. Am. Chem. Soc. 1976, 98. 6133. . Light, K. I.; Kassel, D. 8.; Allison, I. Biomed. Environ. Mass Spectrom. 1909, 18, 177. . Light, K I., Allison, I. I. Am. Soc. Mass Spectrom. 1990, 1, 455. . Teesch, L. M.; Orlando, R. C.; Adams, I. I. Am. Chem. Soc. 1991, 113, 3668. . Weddle, G. H.; Allison, 1.; Ridge, D. P. I. Am. Chem. Soc. 1977, 99, 105. . Borchers, F.; Giessmann, U; Rollgen, F. W. Org. Mass Spec- trom. 1977, 12, 539. APPENDIX TWO APPENDIX TWO. THERMOCHEMICAL DATA FOR ESTIMATION OF REACTION ENTHALPIES FOR PEPTIDE SKELETAL BOND CLEAVAGES In Chapter four, the reaction enthalpies (Aern) were estimated for peptide fragmentation mechanisms by matching the peptide protonation site, the bond cleavage site(s) and the bond formation site(s) on the [M+H]+ ion, fragment ions and fragment neutrals with small neutral molecules, radicals and ions with nearly identical chemical structures and available heats of formation (AHf), proton affinities (PA), and ionization energies (IE) [40, 191, 210]. The AH,“ and AHf values are listed for each species below in kcal/mol. In Equation A1, a AHm, of 87 kcal/mol was estimated for the homolytic C(1- C(0) bond cleavage in Scheme XI using protonated glycine as the model. The AH,“ of y,, ions formed by the 1, 2-climination mechanism in Scheme VI was estimated to be 38 kcal/mol using N-methyl-acetamidc in Equation A2. The formation of b,, ions in Scheme I by an inductive cleavage of the amide bond in N-methyl-acetamidc requires a AHm, of 53.5 kcal/mol as in Equation A3. The formation of (11 ions from b; ions in Scheme 11 was estimated to be -15 kcal/mol exothermic in Equation A6. However, the AHf of NH2CH2C(O)® in Equation A6 is not available and it was estimated to be 166.3 kcal/mol from Equations A4 and A5 assuming the AH,“ for homolytic cleavage of the C(O)-O bond in acetic acid and the ionization of CH3C(O)° in Equation A4 is equal to the identical processes using glycine in Equation A5. The difference in reaction enthalpies for a l and a,, ions was estimated in Equations A7 and A8 using 1133 of amines and amides which are 204 and 214 kcal/mol, respectively. In Equation A8, the amide bond carbonyl in a,, ions (see Scheme 1) delocalizes the electron density of the neighboring amide nitrogen which results 255 256 in the amide N lone pair providing less stabilization for the ionized an structure. In contrast, the amine N in a1 ions as shown in Scheme II and Equation A7 can provide full charge stabilization because the lone pair is localized. As a result, the IE for an ions is 10 kcanol higher, owing to electron delocalization, in comparison to that for a1 ions, so the AH,“ for forming a,, ions with no side chain from b,, ions is -5 kcal/mol, 10 kcal/mol more than a1 ions. Assuming a hydrogen is shifted from a distant amide nitrogen through a secondary interaction, the estimation of the overall AH,“ for y,, ions formed is a four-step process and modeling requires four equations: the protonated amide bond is homolytically cleaved in Equation A9, a distant hydrogen is cleaved from an amide represented by a secondary nitrogen in Equation A10, the cyclic neutral product is formed in Equation A1 1, and the shifted hydrogen is attached to the ionic homolytic product from amide bond cleavage to form a y,, ion in Equation A12. The overall AH,“ is -5 kcal/mol which was obtained by summing the individual AHm in Equations A9-A12. O O NH3CH2C(O)OH -—) NH3CH20 + 0C(O)OI-I Aern = 86.7 (A1) 61 201 -53.3 O O CH3C(O)NH2CH3 -—) CH2 = C = O + NH3CH3 Aern = 37.6 (A2) 97 -11.4 146 O CH3C(O)NH2CH3 —> CH3C(O)O + NHzCH3 AHrmz 53.5 (A3) 97 156 -5.5 CH3C(O)OH —) CH3C(O)O + OOH + 6' AH“: 268.6 (A4) -103.3 156 9.3 NHzCH2C(O)OH —> NH2CH2C(O)O + 00H + e' AHnm = 268.6 (A5) -93 X=166.3 9.3 O NHzCH2C(O)O —> NH2=CH2 + C E O AHnm = - 15 (A6) 166.3 (ESL) 178 ~26.4 $0 CH3CH2NH2 -—> CH3CH2NH2 + 6‘ AH,“ = 204.3 (A7) -1 1.3 193 $0 CH3C(O)NHCH3 —> CH3C(O)NHCH3 + 6' Aern = 214 (A8) -56 158 257 O O CH3C(O)NH2CH3 —> CH3C(O)0 + oNHzCH3 AHm = 98 (A9) 97 -6 201 NH(CH3)2 —) H- + oN(CH3)2 Aern = 91.5 (A10) -4.4 52.1 35 CH3C(O)o + oN(CH3)2 -) CH3C(O)N(CH3)2 AHm=-85 (A11) -6 35 -56 O O H0 + 0NH2CH3 —-> NH3CH3 AB,“ = -107 (A12) 52.1 201 146 In Equations Al-A12, examples of each estimation method for calculating the AHms are demonstrated. Below, the estimated AHms below utilize the above methods to estimate the remaining ions formed from homolytic cleavages, 1, 2-elimination cleavages, inductive cleavages, and secondary interactions in Schemes I-XIV. In Equations A9 and A13, the AHms of the homolytic amide and N-Ca bond cleavages are estimated to be 98 and 103 kcal/mol, respectively. In Equations A14 and A15, the estimated AHms for the formation of Jr", and 01 ions by 1, 2—e1imination cleavages in Schemes V and III are 35 and 26.5 kcal/mol, respectively. The AH,“ of a,, ions formed by the 1, 2-elimination mechanism in Scheme [H was estimated to be 42 kcal/mol, but the AHf of ionized (CH3C(O)NH=CH2)O in Equation A18 is not known. It was estimated to be 137 kcal/mol from Equations A16 and A17 assuming the AHm. for removal of hydrogen in methyl ethyl amine in Equation A16 is equal to the identical processes using ionized N-methyl acetamide in Equation A17. In Equations A19 and A20, the AH,“ of an a1 ion with a side chain from a b1 ion was estimated to be -30 kcal/mol using the identical estimation method as for a1 ions without a side chain shown in Equations A4-A6. In Equations A21 and A22, the AH,“ of an an ion without a side chain from a b,, ion was estimated to be -5 kcal/mol using the same method for (11 ions above and Equations A4, A17 and A18. In Equation A23, the AHm, for conversion of 0,, ions to 1),, ions by amide bond inductive cleavage in Scheme XII is estimated to be 51 kcal/mol. In Equations A24-A27, the overall AH”m for the 258 [174+OH+H]+ rearrangement ion shown in Scheme XIV was estimated to be 23 kcal/mol by employing a similar estimation method as used for y,, ions in Equations A9—Al2, except an 00H is shifted instead of H0. In Equations A13, A10, A28, and A29, the overall AHm, of c,, was estimated to be 5 kcal/mol assuming hydrogen was shifted from an amide nitrogen as shown in Scheme IX using the same method as for y,, ions in Equations A9-A12. If a hydrogen is shifted from the carboxylic acid for c,, ion formation, Equations A10 and A28 were substituted with Equations A30 and A31 to give an overall AHm of 10 kcal/mol. In Equation A32, the AH,“ for conversion of, c,, ions to b,, ions by an inductive cleavage in Scheme XIII is estimated to be 42 kcal/mol. For y,, ions, if a hydrogen shifts from the amine nitrogen instead of the amide nitrogen, Equations A10 and All were substituted with A33 and A26, respectively, to estimate the AH,“ which is 0 chmol. For cyclo-bn ion formation in Scheme VIII, if a hydrogen was shifted from an interacting protonated amine nitrogen (not shown), then the AH,“ was estimated to be 0 kcal/mol simply by negating the AH,“ values in Equations A9, A33, A26, and A12, or if a hydrogen was shifted from an interacting protonated amide nitrogen, then Equations A9-A12 were negated to calculate the AH,“ which is -5 kcal/mol. If the skeletal amide nitrogen is protonated as shown in Scheme VIII for ion cyclo-bn formation, a hydrogen shift is not required since the amide skeletal bond is cleaved and a new, similar amide bond is formed with the interacting amine or amide nitrogen. The AHm, for the latter cyclo-bn mechanism is identical to the former regardless if a hydrogen shifts or not and it can be estimated as follows. If the protonated amide skeletal bond is cleaved as in Equation A9, the positive charge can be moved from the protonated amide nitrogen fragment to the cyclic neutral product in All or A26, depending if an amide or amine N interacts, respectively, providing (1) the amide nitrogen fragment neutralized by retuming its ionization energy, (2) the amide or amine hydrogen is removed from the interacting nitrogen as in Equations A10 or A33. respectively, and (3) this removed hydrogen is ionized and used to protonate the cyclic neutral product. For 2,; ion formation by N-Ca inductive bond cleavage in Scheme IV, the 259 AHm, was estimated to be 79 kcal/mol in Equation A34 if a side chain was not present on the a-carbon and 62 kcal/mol in Equation A35 if a side chain was present The overall AH,“ of cyclo-zn ions was estimated using the same methodology as for cyclo-bn ions using Equation A13 and the Aerns calculated for c,, ions in Equations A28 or A31, and A10 or A30 depending if an amide nitrogen or the carboxylic acid oxygen interacted. The overall AHnns of internal ions and immonium ions were estimated to be 0 kcal/mol identical to cyclo-bn ions assuming a protonated amine nitrogen interacted and protonated an amide skeletal bond. For internal immonium ion formation, the loss of CO from internal ions can be estimated using the AHrms of 0,, ions above. O O CH3C(O)NH2CH3 —-) CH2C(O)NH20 + 0CH3 AHm = 102.8 (A13) 97 165 34.8 O O CH3C(O)NH2CH3 —) CH4 + O=C=NHCH3 AHm, = 35.2 (A14) 97 -17.8 150 O O NH3CH2C(O)OH -—> NH2=CH2 + HC(O)OH AHrm = 26.5 (A15) 61 178 -90.5 O0 - O CH3CH2NHCH3 —> H- + CH3CH2NH=CH2 AHm=3l.1 (A16) 177 52.1 156 O0 O CH3C(O)NHCH3 -9 H- + CH3C(O)NH£H2 AHm, = 31.1 (A17) 158 52.1 X=137 CH3C(O)NH2CH2C(O)OH —) CH3C(O)NH=CH2 + HC(O)OH Aern- — 42 (A18) 4. 5 137 (Est.) -90. 5 NH2CH(CH3)C(O)OH -) NHzCH(CH3)C(O)O + 00H + e' AHrmz 268.6 (A19) -99 X=160.3 9.3 O NHzCH(CH3)C(O)O —-) NH2=CHCH3 + C E O AHrm = -29.7 (A20) 160. 3 (Est. ) 157 -26.4 CH3C(O)NHCH2C(O)OH —) CH3C(O)NHCH2C(O)O + 00H + e' Aern = 268.6 (A21) -l44 X=1 15.3 9.3 O CH3C(O)NHCH2C(O)O —> CH3C(O)NH=CH2 + C E O AHm, = -4.7 (A22) 115.3 (Est) 137 (Est) -26.4 O CH3C(O)NH=CH2 —9 260 CH3C(O)O + NH=CH2 137 (ESL) 156 32 O O CH3C(OH)NHCH3 -) CH3C(OH)- + ONHCH3 97 207 43.6 CH3C(O)OH —) CH3C(O)0 + OOH - 103.3 -6 9.3 CH3C(O)0 + oNHCH3 —) CH3C(O)NHCH3 -6 43.6 —56 O O CH3C(OH)- + 00H -) CH3C(OH)OH 207 9. 3 72 CH3- + -N(CH3)2 —) CH3N(CH3)2 34. 8 35 -5.7 O0 O H- + CH2C(O)NH2 —) CH3C(O)NH3 52. 1 165 103 CH3C(O)OH —) H- + CH3C(O)O¢ -103.3 52.1 -49.6 CH 3- + CH3C(O)OO —) CH3C(O)OCH3 34. 8 -49.6 -98 O CH3C(O)NH3 —) CH3C(O)O + NH3 103 156 -1 1 NH2CH3 —) H- + ONHCH3 -5 .5 52. 1 43.6 O CH3CH2NH2CH2CH3 -> CH3CH2NH2 + OCHzCH3 125 - 1 1.3 215.6 O CH3C(O)NH2CH(CH3)C(O)OCH3 —) CH3C(O)NH2 + CH3CHC(O)OCH3 57 1 15 AHm, = 51 (A23) AHm= 153.6 (A24) AHl-m = 106.6 (A25) AHm = -93.6 (A26) AH,“ = -144. (A27) AHm = -75.5 (A28) AHm, = -114 (A29) AHm, = 105.8 (A30) AH,“ = -83.2 (A31) AHnm = 42 (M2) AHm, = 101.2 (A33) Aern = 79.3 (A34) (A35) AHnm- - 62 LIST OF REFERENCES LIST OF REFERENCES 1. Watson, J. T. Introduction to Mass Spectrometry, 2nd ed.; Raven: New York, 1985; Chapters 1-4. 2. McLafferty, F. W.; Turecek, F. Interpretation of Mass Spectra, 4th ed.; University Science Books: Mill Valley, CA, 1993. 3. Miiller, E. W.; Bahadur, K. Phys. Rev. 1956, 102, 624. 4. Chait, E. M. Anal. Chem. 1972, 44, 77A. 5. Harrison, A. G. Chemical Ionization Mass Spectrometry, 2nd ed.; CRC Press: Boca Raton, FL, 1992. 6. Einolf, N.; Munson, B. Int. J. Mass Spectrom. Ion Phys. 1972, 9, 141. 7. Huang, E. C.; Wachs, T.; Conboy, J. J .; Henion, J. D. Anal. Chem. 1990, 62, 713A. 8. Busch, K. L.; Cooks, R. G. Science 1982, 218, 247. 9. Beckey, H. D. Int. J. Mass Spectrom. Ion Phys. 1969, 2, 500. 10. Beckey, H. D. Principles of Field Ionization and Field Desorption Mass Spectrometry; Pergamon: Oxford, 1977. 11. Prokai, L. Field Desorption Mass Spectrometry; Marcel Dekker: New York, 1990; Vol. 2. 12. Gomer, R. J. Chem. Phys. 1959, 31 , 341. 13. Giessmann, U.; Rollgen, F. W. Org. Mass Spectrom. 1976, II, 1094. 14. Sweeley, C. C.; Soltmann, B.; Holland, J. F. In High Performance Mass Spectrometry - Chemical Applications; Gross, M. L., Ed.; ACS Symposium Series 70; American Chemical Society: Washington, DC, 1978; Chapter 11. 15. Wong, S. S; Giessmann, U.; Karas, M.; Rollgen, F. W. Int. J. Mass Spectrom. Ion Processes 1984, 56, 139. 16. Lattimer, R. P.; Schulten, H. R. Anal. Chem. 1989, 6], 1201A. l7. Baldwin, M. A.; McLafferty, F. W. Org. Mass Spectrom. 1973, 7, 1353. 261 262 18. Cotter, R. J. Anal. Chem. 1980, 52, 1589A. 19. Hunt, D. F.; Shabanowitz, J .; Botz, F. K.; Brent, D. A. Anal. Chem. 1977,49, 1160. 20. Soltmann, B.; Sweeley, C. C.; Holland, J. F. Anal. Chem. 1977, 49, 1164. 21. Anderson, W. R.; Frick, W.; Daves, G. D. J. Am Chem. Soc. 1978, 100, 1974. 22. Torgerson, D. F .; Showronski, R. P.; Macfarlane, R. D. Biochem. Biophys. Res. Commun. 1974, 60, 616. 23. Posthumus, M. A.; Kistemaker, P. G.; Meuzelaar, H. L. C.; Ten Noever de Brauw, M. C. Anal. Chem. 1978, 50, 985. 24. Benninghoven, A.; Sichtermann, W. K. Anal. Chem. 1978, 50, 1180. 25. Barber, M.; Bordoli, R. 8.; Elliott, G. J.; Sedgwick, R. D.; Tyler, A. N. Anal. Chem. 1982, 54, 645A. 26. Rinehart, K. L. Science 1982, 218, 254. 27. Aberth, W. H.; Straub, K. M.; Burlingame, A. L. Anal. Chem. 1982, 54, 2029. 28. Jardine, 1.; Scanlan, G. F.; Tsarbopoulos, A.; Liberato, D. J. Anal. Chem. 1988, 60, 1086. 29. Hillenkamp, F.; Karas, M.; Beavis, R. C.; Chait, B. T. Anal. Chem. 1991, 63, 1193A. 30. Caprioli, R. M. Anal. Chem. 1990, 62, 477A 31. Fenn, J. B.; Mann, M.; Meng, C. K.; Wong, S. F.; Whitehouse, C. M. Mass Spectrom. Rev. 1990, 9, 37. 32. Dempster, A. J. Phys. Rev. 1921, 18, 415. 33. Bleakney, W. Phys. Rev. 1929, 34, 157. 34. Nier, A. O. Rev. Sci. Instrum. 1940, 11, 212. 35. N ier, A. O. Int. J. Mass Spectrom. Ion Processes 1990, 100, 1. 36. Nier, A. 0. Science 1955, 121, 737. 37. McLafferty, F. W. Anal. Chem. 1956, 28, 306. 38. Brenton, A. 6.; Morgan, R. P.; Beynon, J. H. Ann. Rev. Phys. Chem. 1979, 30, 51. 39. Harrison, A. G.; Tsang, C. W. In Biochemical Applications of Mass Spectrometry; Waller, G. R., Ed.; Wiley: New York, 1972; pp 135-156. 40. Lias, S. G.; Bartmess, J. B.; Liebman, J. F.; Holmes, J. L.; Levin, R. D.; Mallard, W. G. J. Phys. Chem. Ref. Data 1988, 17, Suppl. 1. 263 41. 42. 43. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. Rosenstock, H. M.; Wallenstein, M. B.; Wahrhaftig, A. L.; Eyring, H. Proc. Nat. Acad. Sci. 1952, 38, 667. Cooks, R. G.; Beynon, J. H.; Caprioli, R. M.; Lester, G. R. Metastable Ions; Elsevier: New York, 1973. Marcus, R. A. J. Chem. Phys. 1952, 20, 359. . Robinson, P. J .; Holbrook, K. A. Unimolecular Reactions; Wiley: New York, 1972. Franklin, J. L.; Dillard, J. G.; Rosenstock, H. M.; Herron, J. T.; Draxl, K; Field, F. H. Ionization Potentials, Appearance Energies, and Heats of Formation of Gaseous Positive Ions; National Bureau of Standards: Washington DC, 1969. Lorquet, J. C. Org. Mass Spectrom. 1981, 16, 478. Munson, M. S. B.; Field, F. H. J. Am. Chem. Soc. 1966, 88, 2621. Sunner, J. A.; Kulatunga, R.; Kebarle, P. Anal. Chem. 1986, 58, 1312. Wu, Z.; Fenselau, C. Rapid Comm. Mass Spectrom. 1992, 6, 403. Bowen, R. D.; Harrison, A. G. Org. Mass Spectrom. 1981, I6, 159. Bombick, D. D.;Pinkston, J. D.; Allison, J. Anal. Chem. 1984, 56, 396. Blewett, J. P.; Jones, E. J. Phys. Rev. 1936, 50, 464. Reimann, A. L. Thermionic Emission; Wiley: New York, 1934. Kassel, D. B.; Allison, J. Biomed. Environ. Mass Spectrom. 1988, 17, 221. Schmelzeisen-Redeker, G.; Giessmann, U.; Rollgen, F. W. Org. Mass Spectrom. 1985, 20, 305. Bombick, D. D.; Allison, J. Anal. Chem. 1987, 59, 458. Beuhler, R. J .; Flanigan, B.; Greene, L. J .; Friedman, L. J. Am. Chem. Soc. 1974, 96, 3990. Wu, H.; Rouse, J. C.; Leali, K. A.; Allison, J. Proceedings of the 40th ASMS Conference on Mass Spectrometry and Allied Topics, Washington, DC, May 31- June 5, 1992; p 1795. Spears, K. G.; Ferguson, E. E. J. Chem. Phys. 1973, 59, 4174. Allison, J.; Ridge, D. P. J. Am. Chem. Soc. 1979, 101, 4998. Davidson, W. R.; Kebarle, P. J. Am. Chem. Soc. 1976, 98, 6133. Rollgen, F. W.; Schulten, H. R. Org. Mass Spectrom. 1975, 10, 660. 264 63. Barber, M.; Bordoli, R. S.; Sedgwick, R. D.; Tyler, A. N. J. Chem. Soc., Chem Commun. 1981, 325. 64. Barber, M.; Bordoli, R. S.; Sedgwick, R. D.; Tyler, A. N. Nature 1981, 293, 270. 65. Barber, M.; Bordoli, R. S.; Garner, G. V.; Gordon, D. B.; Sedgwick, R. D.; Tetler, L. W.; Tyler, A. N. Biochem. J. 1981, 197, 401. 66. Honig, R. E. Int. J. Mass Spectrom. Ion Processes 1985, 66, 31. 67. Benninghoven, A. Z Phys. 1970, 230, 403. 68. Benninghoven, A.; Jaspers, D.; Sichtermann, W. Appl. Phys. 1976, II, 35. 69. Lew, H. In Methods of Experimental Physics 4A; Hughes, V. W.; Schultz, H. L., Eds.; Academic: New York, 1967; p 155. 70. Devienne, F. M.; Roustan, J. C. Org. Mass Spectrom. 1982, I7, 173. 71. Surman, D. J.; Vickerman, J. C. Appl. Surf. Sci. 1981, 9, 108. 72. Martin, S. A.; Costello, C. E.; Biemann, K. Anal. Chem 1982, 54, 2362. 73. Aberth, W. H.; Burlingame, A. L. Anal. Chem. 1988, 60, 1426. 74. Kloppel, K. D.; Weyer, K.; Von Bfinau, G. Int. J. Mass. Spectrom. Ion Phys. 1983, 51, 47. 75. Gilbert, J. R.; Leroi, G. E.; Allison, J. Int. J. Mass Spectrom. Ion Processes 1991, 107, 247. 76. Perel, J. In Desorption Mass Spectrometry: Are SIMS and FAB the Same ?; Lyons, P. A., Ed.; ACS Symposium Series 291; American Chemical Society: Washington, DC, 1985; Chapter 8. 77. Aberth, W.; Burlingame, A. L. In Ion Formation from Organic Solids; Benninghoven, A., Ed.; Springer-Verlag: Berlin, 1983; Vol. 25, p 167. 78. Fenselau, C.; Cotter, R. J. Chem. Rev. 1987, 87, 501. 79. De Pauw, E.; Agnello, A.; Derwa, F. Mass Spectrom. Rev. 1991, 10, 283. 80. De Pauw, E. Mass Spectrom. Rev. 1986, 5, 191. 81. Gower, J. L. Biomed. Mass Spectrom. 1985, 12, 191. 82. Cook, K. D.; Todd, P. J .; Friar, D. H. Biomed. Environ. Mass Spectrom. 1989, 18, 492. 83. Sweetrnan, B. J .; Blair, 1. A. Biomed. Environ. Mass Spectrom. 1988, I 7, 337. 84. Field, F. H. J. Phys. Chem. 1982, 86, 5115. 265 85. Pachuta, S. J.; Cooks, R. G. Chem. Rev. 1987, 87, 647. 86. Cooks, R. G.; Busch, K. L. Int. J. Mass Spectrom. Ion Physics 1983, 53, 111. 87. Couchouron, M.; Hennequin, J. F. Int. J. Mass Spectrom. Ion Processes 1986, 69, 1. 88. Sigmund, P. Phys. Rev. 1969, 184, 383. 89. Benninghoven, A. Surf. Sci. 1973, 35, 427. 90. Andersen, C. A.; Hinthome, J. R. Anal. Chem. 1973, 45, 1421. 91. Kelly, R. Radiat. Eff. 1977, 32, 91. 92. Lucchese, R. R. J. Chem. Phys. 1987, 86, 443. 93. Kelly, R. Surf Sci. 1979, 90, 280. 94. Mich], J. Int. J. Mass Spectrom. Ion Phys. 1983, 53, 255. 95. Hoogerbrugge, R.; Van der Zande, W. J .; Kistemaker, P. G. Int. J. Mass Spectrom. Ion Processes 1987, 76, 239. 96. Falcone, G.; Sroubek, Z.; Sindona, G.; Uccella, N. Int. J. Mass Spectrom. Ion Processes 1988, 83, 223. 97. Sunner, J. A.; Kulatunga, R.; Kebarle, P. Anal. Chem. 1986, 58, 2009. 98. Visentini, J .; Zidarov, D.; Allard, M.; Bertrand, M. J. J. Am. Soc. Mass Spectrom. 1993, 4, 482. 99. Reynolds, J. D.; Cook, K. D.; Burm, J. L. E.; Woods, C. J. Am. Soc. Mass Spectrom. 1992, 3, 113. 100. Dass, C.; Desidcrio, D. M. Anal. Chem. 1988, 60, 2723. 101. Aubagnac, J. L.; Doulut, S.; Rodriguez, M.; Martinez, J. Org. Mass Spectrom. 1992, 27, 645. 102. Detter, L. D.; Hand, 0. W.; Cooks, R. G.; Walton, R. A. Mass Spectrom Rev. 1988, 7, 465. 103. Vékey, K.; Zerilli, L. F. Org. Mass Spectrom. 1991, 26, 939. 104. Sunner, J .; Ikonomou, M. G.; Kebarle, P. Int. J. Mass Spectrom. Ion Processes 1988, 82, 221. 105. Ens, W.; Mao, Y.; Mayer, F.; Standing, K. G. Rapid Comm. Mass Spectrom. 1991, 5, 117. 106. Ligon, W. V. Int. J. Mass Spectrom. Ion Phys. 1983, 52, 183. 266 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125 126. Wong, S. S.; Rollgen, F. W.; Manz, I; Przybyski, M. Biomed. Mass Spectrom. 1985, 12, 43. Vertes, A.; Gijbels, R.; Levine, R. D. Rapid Comm. Mass Spectrom. 1990, 4, 228. Benninghoven, A. Int. J. Mass Spectrom. Ion Phys. 1983, 53, 85. Kidwell, D. A.; Ross, M. M.; Colton, R. J. Int. J. Mass Spectrom. Ion Processes 1987, 78, 315. Sunner, J. J. Am. Soc. Mass Spectrom. 1993, 4, 410. Freas, R. B.; Ross, M. M.; Campana, J. E. J. Am. Chem. Soc. 1985, 107, 6195. Schroder, B.; Miinster, H.; Budzikiewicz, H. Org. Mass Spectrom. 1986, 21, 707. Honda, E; Lancaster, G. M.; Fukuda, Y.; Rabalais, J. W. J. Chem. Phys. 1978, 69, 4931. Grade, H.; Cooks, R. G. J. Am. Chem. Soc. 1978, 100, 5615. Day, R. J .; Unger, S. B.; Cooks, R. G. Anal. Chem. 1980, 52, 557A. Pachuta, S. J .; Cooks, R. G. In Desorption Mass Spectrometry: Are SIMS and FAB the Same ?; Lyons, P. A., Ed.; ACS Symposium Series 291; American Chemical Society: Washington, DC, 1985; Chapter 1. Sunner, J. A.; Kutlatunga, R.; Kebarle, P. Anal. Chem. 1986, 58, 1312. Sunner, J .; Morales, A.; Kebarle, P. Int. J. Mass Spectrom. Ion Processes 1988, 86, 169 Bojesen, G; Moller, J. Int. J. Mass Spectrom. Ion Processes 1986, 68, 239. Sunner, J .; Morales, A.; Kebarle, P. Anal. Chem. 1988, 60, 98. Sunner, J.; Morales, A.; Kebarle, P. Anal. Chem. 1987, 59, 1378. Vestal, M. L. Mass Spectrom. Rev. 1983, 2, 447. Rouse, J. C.; Allison, J. J. Am. Soc. Mass Spectrom. 1993, 4, 259. Ehring, H.; Karas, M.; Hillenkamp, F. Org. Mass Spectrom. 1992, 27, 472. Hoogerbrugge, R.; Bobeldijk, M.; Kistemaker, P. G.; Los, J. J. Chem. Phys. 1988, 88, 5314. 127. Derwa, F.; DePauw, E.; Natalis, P. Org. Mass Spectrom. 1991, 26, 117. 128. Williams, D. H.; Naylor, S. J. Chem. Soc., Chem. Commun. 1987, 1408. 129. Hamdan, M. Rapid Comm. Mass Spectrom. 1992, 6, 224. 267 130. Lehmann, W. D.; Kessler, M.; Koning, W. A. Biomed. Mass Spectrom. 1984, I I, 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144. 145. 146. 147. 148. 149. 150. 151. 217. Keough, T. Anal. Chem. 1985, 5 7, 2027. Musselman, B. D.; Allison, J .; Watson, J. T. Anal. Chem. 1985, 5 7, 2425. Shiea, J .; Sunner, J. Org. Mass Spectrom. 1991, 26, 38. Ligon, W. V.; Dorn, S. B. Int. J. Mass Spectrom. Ion Processes 1986, 78, 99. Hayes, R. N.; Gross, M. L. In Mass Spectrometry; McCloskey, J. A., Ed.; Methods in Enzymology; Academic: San Diego, CA, 1990; Vol. 193, Chapter 10. Herman, Z.; Futrcll, J. H.; Friedrich, B. Int. J. Mass Spectrom. Ion Processes 1984, 58, 181. Levsen, K.; Schwarz, H. Mass Spectrom. Rev. 1983, 2, 77. Kim, M. 8.; Int. J. Mass Spectrom. Ion Phys. 1983, 50, 189. Laramée, J. A.; Cameron, D.; Cooks, R. G. J. Am. Chem. Soc. 1981, 103, 12. Kim, M. S.; McLafferty, F. W. J. Am. Chem. Soc. 1978, 100, 3279. Neumann, G. M.; Derrick, P. J. Org. Mass Spectrom. 1984, 19, 165. Griffin, L. L.; McAdoo, D. J. Proceedings of the 40th ASMS Conference on Mass Spectrometry and Allied Topics, Washington, DC, May 31-June 5, 1992; p 1831. Dekrey, M. J .; Mabud, M. D. A.; Cooks, R. G.; Syka, J. E. P. Int. J. Mass Spectrom. Ion Processes 1985, 67, 295. Bier, M. E.; Amy, J. W.; Cooks, R. G. Int. J. Mass Spectrom. Ion Processes 1987, 77, 31. Bomse, D. S.; Woodin, R. L.; Beaucharnp, J. L. J. Am. Chem. Soc. 1979, 101, 5503. Gorman, G. S.; Amster, I. J. Org. Mass Spectrom. 1993, 28, 437. Jennings, K. R.; Dolnikowski, G. G. In Mass Spectrometry; McCloskey, J. A., Ed.; Methods in Enzymology; Academic: San Diego, CA, 1990; Vol. 193, Chapter 2. Mamyrin, B. A.; Karatajev, V. J .; Shmikk, D. V.; Zagulin, V. A. Soviet Phys. JETP 1973, 3 7, 45. Mattauch, J. Phys. Rev. 1936, 50, 617. Nier, A. O. J. Am. Soc. Mass Spectrom. 1991, 2, 447. Miller, P. E.; Denton, M. B. J. Chem. Ed. 1986, 63, 617. 268 152. Paul, W.; Reinhard, H. P.; von Zahn, U. Z Phys. 1958, 52, 143. 153. Dawson, P. H. Mass Spectrom. Rev. 1986, 5, l. 154. Wiley, W. C.; McLaren, I. H. Rev. Sci. Instrum. 1955, 26, 1150. 155. Cotter, R. J. Anal. Chem. 1992, 64, 1027A. 156. Yost, R. A.; Enke, C. G. Anal. Chem. 1979, 5], 1251A. 157. Bruins, A. P.; Jennings, K. R.; Evans, S. Int. J. Mass Spectrom. Ion Phys. 1978, 26, 395. 158. Jennings, K. R. In High Pen‘ormance Mass Spectrometry - Chemical Applications; Gross, M. L., Ed.; ACS Symposium Series 70; American Chemical Society: Washington, DC, 1978; Chapter 1. 159. Yost, R. A.; Boyd, R. K. In Mass Spectrometry; McCloskey, J. A., Ed.; Methods in Enzymology; Academic: San Diego, CA, 1990; Vol. 193, Chapter 7. 160. Gross, M. L. In Mass Spectrometry; McCloskey, J. A., Ed.; Methods in Enzymology; Academic: San Diego, CA, 1990; Vol. 193, Chapter 6. 161. Schey, K. L.; Schwartz, J. C; Cooks, R. G. Rapid Commun, Mass Spectrom. 1989, 3, 305. 162. Puzycki, M. A.; McLane, R. D_.; Vlasak, P. R.; Enke, C. G. Proceedings of the 40th ASMS Conference on Mass Spectrometry and Allied Topics, Washington, DC, May 31-June 5, 1992; p 1338. 163. Rouse, J. C.; Light, K. J.; Allison, J. Proceedings of the 38th ASMS Conference on Mass Spectrometry and Allied Topics, Tucson, Arizona, June 3-8, 1990; p 455. 164. Ackermann, B.; Allison, J .; Musselman, B.; Pinkston, D.; Bombick, D.; Dolnikowski, G.; Tsarbopoulos, A.; Watson, J. T. Proceedings of the 3lst ASMS Conference on Mass Spectrometry and Allied Topics, Boston, MA, May 1983; p 600. 165. Dolnikowski, G. G.; Watson, J. T.; Allison, J. Anal. Chem. 1984, 56, 197. 166. Munster, H.; Budzikiewicz, H.; Schr'dder, E. Org. Mass Spectrom. 1987, 22, 384. 167. Chiarelli, M. P.; Gross, M. L. Int. J. Mass Spectrom. Ion Processes 1987, 78, 37. 168. Miinster, H.; Theobald, F.; Budzikiewicz, H.; Schroder, E. Int. J. Mass Spectrom. Ion Processes 1987, 79, 73. 169. Miller, J. M.; Balasanmugam, K.; Fulcher, A. Org. Mass Spectrom. 1989, 24, 497. 170. Michaud, D. P.; Kyranos, J. N.; Brennan, T. F.; Vouros, P. Anal. Chem. 1990, 62, 1069. 269 171. Kassel, D. B. Ph.D. Thesis, Michigan State University, 1988. 172. Schultz, G. A.; Allison, J. Proceedings of the 37th ASMS Conference on Mass Spectrometry and Allied Topics, Miami Beach, Florida, May 21-26, 1989; p 527. 173. Light, K. J. Ph.D. Thesis, Michigan State University, 1990. 174. Dahl, D. A.; Delmore, J. E.; Appelhans, A. D. Rev. Sci. Instrum. 1990, 61, 607. 175. Dahl, D. A.; Delmore, J. E. SIMION PC/PSZ Version 4.0, EGG-CS-7233 Revision 2, April 1988. 176. Light, K. J .; Allsion, J. J. Am. Soc. Mass Spectrom. 1990, I, 455. 177. Weddle, G. H.; Allison, J.; Ridge, D. P. J. Am. Chem. Soc. 1977, 99, 105. 178. Borchers, F.; Giessmann, U.; Rollgen, F. W. Org. Mass Spectrom. 1977 , 12, 539. 179. Harris, D. C. Quantitative Chemical Analysis, 2nd Ed.; Freeman: New York, 1987; p 202. 180. Musselman, B. D.; Watson, J. T.; Chang, C. K. Org. Mass Spectrom. 1986, 21, 215. 181. Denys, L.; Bothner-By, A. A.; Fisher, G. H.; Ryan, J. W. Biochemistry 1982, 21 , 6531. 182. London, R. B.; Stewart, J. M.; Cann, J. R.; Matwiyoff, N. A. Biochemistry 1978, I 7, 2270. 183. Lintner, K.; Fermandjian, S.; St. Pierre, 8.; Regoli, D. Biochem. Biophys. Res. Commun. 1979, 91, 803. 184. Naumenko, V. N.; Koren, A. 0.; Gaponik, P. N. Magn. Reson. Chem. 1992, 30, 558. 185. McLafferty, F. W. Org. Mass Spectrom. 1980, 15, 114. 186. Biemann, K. In Mass Spectrometry; McCloskey, J. A., Ed.; Methods in Enzymology; Academic: San Diego, CA, 1990; Vol. 193, Chapter 25, Appendix 5. 187. Dass, C.; Desidcrio, D. M. Anal. Biochem. 1987, 163, 52. 188. Grese, R. P.; Cerny, R. L.; Gross, M. L. J. Am. Chem. Soc. 1989, 111, 2835. 189. Bean, M. F.; Carr, 8. A.; Thorne, G. C.; Reilly, M. H.; Gaskell, S. J. Anal. Chem. 1991, 63, 1473. 190. Wu, Z.; Fenselau, C. J. Am. Soc. Mass Spectrom. 1992, 3, 863. 191. Meot-Ner (Mautner), M. J. Am. Chem. Soc. 1984, 106, 278. 270 192. 193. 194. 195. 196. 197. 198. 199. 200. 201. 202. 203. 204. 205. 206. 207. 208. 209. 210. 211. 212. Bouchonnet, S.; Hoppilliard, Y. Org. Mass Spectrom. 1992, 27, 71. Jensen, F. J. Am. Chem. Soc. 1992, 114, 9533. Naylor, S.; Moneti, G. Biomed. Environ. Mass Spectrom. 1989, I8, 405. Johnson, R. S.; Martin, S. A.; Biemann, K. Int. J. Mass Spectrom. Ion Processes 1988, 86, 137. Bunk, D. M.; Macfarlane, R. D. Int. J. Mass Spectrom. Ion Processes 1991, 111, 55. Fraenkel, G.; Franconi, C. J. Am. Chem. Soc. 1960, 82, 4478. Cavell, R. G.; Allison, D. A. J. Am. Chem. Soc. 1977, 99, 4203. Hinton, J. F.; Beeler, A.; Harpool, D.; Briggs, R. W.; Pullman, A. Chem. Phys. Lett. 1977, 47, 411. Kollman, P.; Rothenburg, S. J. Am. Chem. Soc. 1977, 99, 1333. Addoud, J .-L. M.; Cafiada, T.; Homan, H.; Notario, R.; Cativiela, C.; Diaz de Villegas, M. D.; Bordejé, M. G; M6, 0.; Yafiez, M. J. Am. Chem. Soc. 1992, 114, 4728. Wagner, D. S.; Gage, D. A.; Watson, J. T.; Allison, J. J. Am. Soc. Mass Spectrom., submitted for publication. Teesch, L. M.; Orlando, R. C.; Adams, J. J. Am. Chem. Soc. 1991, 113, 3668. Tsang, C. W.; Harrison, A. G. J. Am. Chem. Soc. 1976, 98, 1301. Hunt, D. R; Yates 111, J. R.; Shabanowitz, J .; Winston, S.; Hauer, C. R. Proc. Natl. Acad. Sci. 1986, 83, 6233. Teesch, L. M.; Adams, J. J. Am. Chem. Soc. 1991, 113, 812. Kiryushkin, A. A.; Fales, H. M.; Axenrod, T.; Gilbert, E. J .; Milne, G. W. A. Org. Mass Spectrom. 1971, 5, 19. Kenny, P. T. M.; Nomoto, K.; Orlando, R. Rapid Commun. Mass Spectrom. 1992, 6, 95. Yeh, R. W.; Grimley, J. M.; Bursey, M. M. Biological Mass Spectrom. 1991, 20, 443. CRC Handbook of Chemistry and Physics, 65th ed.; Weast, R. C., Ed.; CRC Press: Boca Raton, FL, 1985; p F-188. Fenselau, G; Murphy, C. M.; Cotter, R. J. J. Am. Soc. Mass Spectrom. 1990, I, 336. Madden, T.; Welham, K. J .; Baldwin, M. A. Org. Mass Spectrom. 1991, 26, 443. 271 213 214 215 216 217 218 219 220 221 222 223. 224. 225. 226. 227. 228. 229. 230. 231. 232. 233. . Thorne, G. C.; Gaskell, S. J. Rapid Comm. Mass Spectrom. 1989, 3, 217. . Ballard, K. D.; Gaskell, S. J. Int. J. Mass Spectrom. and Ian Processes 1991, 111, 173. . Renner, D.; Spiteller, G. Biomed. Mass Spectrom. 1986, 13, 405. . Knapp, D. R. In Mass Spectrometry; McCloskey, J. A., Ed.; Methods in Enzymology; Academic: San Diego, CA, 1990; Vol. 193, Chapter 15. . Alexander, A. J .; Boyd, R. K. Int. J. Mass Spectrom. Ion Processes 1989, 90, 211. . Tang, X.; Ens, W.; Standing, K. G.; Westmore, J. B. Anal. Chem. 1988, 60, 1791. . Thorne, G. C.; Ballard, K. D.; Gaskell, S. J. J. Am. Soc. Mass Spectrom. 1990, I, 249. . Svec, H. J.; Junk, G. A. J. Am. Chem. Soc. 1964, 86, 2278. . Cordero, M. M.; Houser, J. J.; Wesdemiotis, C. Anal. Chem. 1993, 65, 1594. . Smith, R. D.; Light-Wahl, K. J .; Ogorzalek Loo, R. R.; Winger, B. E.; Loo, J. A.; Rockwood, A. L. Proceedings of the 40th ASMS Conference on Mass Spectrometry and Allied Topics, Washington, DC, May 31-June 5, 1992; p 637. Speir, J. P.; Gorman, G. S.; Comett, D. S.; Amster, I. J. Anal. Chem. 1991, 63, 65. Allison, J. In Progress in Inorganic Chemistry; Lippard, S. J ., Ed.; Wiley: New York, 1986; Vol. 34, p 627. Lehman, T. A.; Bursey, M. M. Ion Cyclotron Resonance Spectrometry; Wiley: New York, 1976. O'Connor, P. J. Ph.D. Thesis, Michigan State University, 1991. Wilkins, C. L.; Gross, M. L. Anal. Chem. 1981, 53, 1661A. Lombarski, M.; Allison, J. Int. J. Mass Spectrom. Ion Processes 1983, 49, 281.; Lombarski, M.; Allison, J. Int. J. Mass Spectrom. Ion Processes 1985, 65, 31. Speir, J. P.; Gorman, G. S.; Amster, I. J. J. Am. Soc. Mass Spectrom. 1993, 4, 106. Teesch, L. M.; Adams, J. Org. Mass Spectrom. 1992, 2 7, 931. Hogan, M. J .; Ong, P. P.; Ang, J. L.; Cheang, K. K. Int. J. Mass Spectrom. Ion Processes 1992, 116, 249. Holland, P. M.; Castleman, A. W. J. Am. Chem. Soc. 1980, 102, 6175. Volkening, J .; Heumann, K. G. In Advances in Mass Spectrometry; Todd, J. F. J ., Ed.; Wiley: New York, 1985; Part B, p 1059. 272 234. Heinz, 0.; Reaves, R. T. Rev. Sci. Instr. 1968, 39, 1229. 235. Hodges, R. V.; Armentrout, P. B.; Beauchamp, J. L. Int. J. Mass Spectrom. Ion Phys. 1979, 29, 375. 236. Chowdhury, A. K.; Wilkins, C. L. Int. J. Mass Spectrom. Ion Processes 1988, 82, 163. 237. Leuchtner, R. B.; Farley, R. W.; Harms, A. C.; Funasaka, H.; Castleman, A. W. Int. J. Mass Spectrom. Ion Processes 1990, 102, 199. 238. Tomer, K. B. Mass Spectrom. Rev. 1989, 8, 445. 239. Mallis, L. M.; Russell, D. H. Int. J. Mass Spectrom. Ion Processes 1987, 78 147. 240. Leary, J. A.; Williams, T. D.; Bott, G. Rapid Comm. Mass Spectrom. 1989, 3, 192. 241. Tomer, K. B.; Deterding, L. J .; Guenat, C. Biomed. Mass Spectrom. 1991, 20, 121. 242. Grese, R. P.; Gross, M. L. J. Am. Chem. Soc. 1990, 112, 5098. 243. Gross, M. L. Int. J. Mass Spectrom. Ion Processes 1992, 118, 137. 244. Leary, J. A.; Zhou, Z.; Ogden, S.; Williams,T. D. J. Am. Soc. Mass Spectrom. 1990, I. 473. 245. Nakamora, T., JEOL USA, Ltd., Boston, MA, personal communication, 1990. 246. Wagner, D. S.; Allison, J. Proceedings of the 40th ASMS Conference on Mass Spectrometry and Allied Topics, Washington, DC, May 31-June 5, 1992; p 1399. 247. Balodis, Y. Y.; Nikiforovich, G. V.; Grinsteine, I. V.; Vegner, R. E.; Chipens, G. F. FEBS Letters, 1978, 86, 239. "JJJJJJJJJJJJJJ