,.2.y.._....$ ‘ . ‘ :"u. . :3 .ihflu... 3 . 1.11, , . aka v r1. . . .v. ..I .4! E . .3... 2.0?53. its $55.»... . E a 5...: 31h.» .u . 1.3%.... , In... to" udhfi. 2:... u...‘ 91') . Ln A x . 4....th L»... 2.: 53.3, 3...: .2... it 89.3%.}? K. . .3206: n “r . c . . :.. fiesnhafi: our... 3. nib gin-3;! :5. . .rwizflt. 5, $5 . in}: “by. . 32?”??? .2? 2.2.3."; .’0\ 5.. i .. .. N? :5?” .3 lamb. Vi}: .. 5.65.)! a . diviinfir a. .u ”1,433.39" Igfls!’ mm , LIBRARY 100 ,1 Michigan State University This is to certify that the thesis entitled STUDIES ON THE FRAGMENTATION REACTIONS OF FIXED CHARGE SULFONIUM ION CONTAINING PEPTIDES BY TANDEM MASS SPECTROMETRY presented by MAHASILU AMUNUGAMA has been accepted towards fulfillment of the requirements for the MS. degree in CHEMISTRY “—QK 27’3/ Major Professor's Signature ”70] 4 I/ 200 5 Date MSU is an Affirmative Action/Equal Opportunity Institution .---:-u— -._.- _ PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/05 p:lClRC/DaleDue.indd-p.1 STUDIES ON THE FRAGMENTATION REACTIONS OF FD(ED CHARGE SULFONIUM ION CONTAINING PEPTIDES BY TANDEM MASS SPECTROMETRY BY MAHASILU AMUNUGAMA A THESIS Submitted to Michigan State University In partial fulfillment of the requirements For the degree of MASTER OF SCIENCE Department of Chemistry 2006 ABSTRACT STUDIES ON THE FRAGMENTATION REACTIONS OF FIXED CHARGE SULFONIUM ION CONTAINING PEPT IDES BY TANDEM MASS SPECTROMETRY BY MAHASILU AMUNUGAMA In order to enable the development of improved tandem mass spectrometry based methods for selective proteome analysis, the mechanisms, product ion structures and other factors influencing the gas-phase fragmentation reactions of methionine side chain derivatized 'fixed-charge' phenacyl sulfonium ion containing peptide ions have been examined. Dissociation of these peptide ions results in exclusive loss of the derivatized side chain, thereby enabling their selective identification. The resultant product ion(s) are then subjected to further dissociation to obtain sequence information for subsequent protein identification. Molecular orbital calculations (at the BBLYP/6-31+G** level of theory) performed on a simple peptide model, together with experimental evidence obtained by multistage dissociation of a regioselectively deuterated methionine derivatized sulfonium ion containing tryptic peptide, indicate that the initial fragmentation occurs via 8N2 reactions resulting in the formation of cyclic five- and six- membered hydrofuran and oxazine product ions. Further, molecular orbital calculations and experimental evidence obtained from various para-substituted phenacyl sulfonium ion peptide ions have determined that the ratio of neutral versus charged losses of the derivatized side chain exhibits a linear dependence on the proton affinity of the side chain fragmentation product, as well as the proton mobility of the peptide product ions. ACKNOWLEDGMENTS No dissertation is ever successful due to one persons efforts, and certainly this one was no different. It would never have become produced without the help and suggestions of many people. Dr. Gavin E. Reid has been a great source of guidance and motivation. He is more than an advisor. His leadership in both personal and scientific endeavors has been an inspiration to me throughout my graduate career, and I am grateful for having had the opportunity to pursue my master’s degree under his supervision. Special thanks are also extended to Dr. Merlin Bruening and Dr. Gary J. Blanchard who are on my thesis committee. I would also like to thank Dr. Kade D. Roberts for all the support, advice and encouragement he provided me throughout my graduate studies. I am grateful to James Sierakowski, who is a wonderful friend, for his help in many ways to make my research successful. I am also deeply indebted to my dear friends Jennifer Marie Froelich and Gwynyth Scherperel for their continuous support and encouragement. They were always with me during the good, difficult and distressing times in my graduate career. Special thank also goes to all the group members of the Reid research group for their support. I especially thank the free education system in Sri Lanka without which I would have never been able to come to graduate school. I am eternally grateful to my parents who, though living oceans apart, have shown me how to live life with grace and compassion. Finally, I am thankful to all who have made this journey both possible and pleasurable. TABLE OF CONTENTS LIST OF FIGURES ................................................................. viii LIST OF TABLES ................................................................. xi LISTOFSCHEMES............................ ................................... xii 1. CHAPTER ONE: Introduction .................................................. 1 1.1 Proteomics .................................................................. 1 1.2 Mass Spectrometry ......................................................... 1 1.2.1 Electrospray Ionization (ESI) ........................................ 2 1.2.2 Matrix-assisted laser desorption ionization (MALDI) ........... 4 1.2.3 Mass Analyzers ........................................................ 5 1.2.3.1 Quadrupole Mass Analyzers .................................... 5 1.2.3.2 The Quadrupole Ion Trap Mass Analyzer .................... 10 1.3 Current Approaches Employed for Proton Identification and Characterization by Mass spectrometry ................................. 13 1.3.1 Database Search Algorithms for Protein Identification ......... 15 1.4 Understanding and Controlling Peptide Fragmentation Reactions... 17 1.4.1 Mechanisms for the Fragmentation of Protonated Peptide Ions... 18 1.4.2 Controlling Peptide Fragmentations ................................... 22 1.4.2.1 N-terminal Fixed Charge Derivatization and Fragmentation 22 1.5. Side Chain Fixed Charge Derivatization and Fragmentation ............... 23 1.6 Aims of this Thesis ......................................................... 27 iv 2. CHAPTER TWO: Experimental ................................................. 2.1 Materials ..................................................................... 2.2 Synthesis of (D, L)—, (D)- and (d4-D, L)-Fmoc-Methionine .......... 2.3 Synthesis of 4-Bromoacetylbenzoic- acid ................................. 2.4 Synthesis of Methyl-4—bromoacetylbenzoate ............................ 2.5 Synthesis of 4-Methylphenacy1bromide .................................. 2.6 Synthesis of 4-Ethylphenacylbromide...................................... 2.7 Side chain Fixed-Charge Derivatization of Methionine-Containing Peptides ...................................................................... 2.8 Mass Spectrometry ......................................................... 2.9 Computational Methods ................................................... 2.9.1 Computational Methods Employed for Mechanistic Studies. 2.9.2 Computational Methods Employed for Proton Affinity Calculations 2.9.3 Proton Affinity Calculations ......................................... 3. CHAPTER THREE: Mechanisms for the Selective Gas-Phase Fragmentation Reactions of Methionine Side Chain Fixed Charge Sulfonium Ion containing Peptides ................................................ 3.1 Introduction .................................................................. 3.2 Molecular Orbital Calculations to Examine Mechanisms for the Loss of CH3SR from Methionine Fixed Charge Sulfonium Ion Containing Peptides .......................................................................... 28 28 29 29 3O 30 3O 31 31 33 33 33 34 37 37 38 3.3 Potential Ring Opening of the Five and Six Membered Cyclic Product Ions via Intramolecular Proton Transfer Reactions .............. 3.4 Obtaining Experimental Evidence for the Mechanisms Responsible for the Loss of CH3SR from Methionine Fixed Charge Sulfonium Ion Containing Peptides. . . . . . . . . . . . .. ..................................... 3.4.1 Regioselective Deuterium Labeling Experiments to Examine Mechanisms for the Side Chain Fragmentation Reactions of Methionine Fixed Charge Containing Peptide Ions ................. 3.4.2 Multistage Tandem Mass Spectrometry Experiments to Examine the Structure(s) of the Product Ions Formed by Side Chain Fragmentation of Methionine Fixed Charge Containing Peptides ............................................................................... . ...... 3.5 Conclusion ................................................................... 4. CHAPTER FOUR: Substituent Effects on the Fragmentation Reactions of Methionine Side Chain Fixed Charge Phenacylsulfonium Ion Containing Peptides ....................................................................................................... 4.1 Introduction .................................................................. 4.2 Proton Affinity Calculations ............................................... 4.3 Determining the Influence of Side Chain Fragment Proton Affinities on the Relative Abundance of Neutral versus Charged Side Chain Losses from Methionine Fixed Charge Containing Peptide Ions ..................................................................................................... vi 47 53 53 57 62 63 63 65 4.4 Conclusions and Future Directions ................................................. . 78 REFERENCES ........................................................................ 8O vii LIST OF FIGURES FIGURE PAGE 1.1 Components of a mass spectrometer .................................................................. 3 1.2 Stability areas as a function of U and V for positive ions with different masses where ml [ Source ]C>[ Analyzer jd LDeteotor jé Vacuum Pump Data Handling [ Data System J 11 Data | i 1 Output l 7 g l Mass spectrum Figure 1.1 Components of a mass spectrometer (adapted from “What is Mass Spectrometry”. www.asms.org). under atmospheric pressure, enabling a wide range of compounds including proteins and peptides that are sufficiently polar to allow attachment of a charge to be analyzed. ESI can be operated in either an infusion mode or in combination with high performance liquid chromatography (HPLC). In the infusion mode, the analyte is dissolved in a suitable solvent (typically 1:1 methanolzwater) and introduced to the ESI capillary via an injection pump at flow rates as low as few hundred nanoliters per minute. When on-line HPLC is employed, the analytes are introduced to the ESI source as they elute from the chromatographic column. Hence, sample clean up, concentration and separation of the analytes can be achieved in a single step. A striking feature of ESI is the ability to form a series of pseudomolecular ion charge states, including both singly and multiply charged ions. The extent of multiple charging is influenced by factors such as the composition and pH of the electrospray solvent, as well as the chemical nature of the analyte. These multiple charge states can be used to provide independent verification of the mass, and allow large molecules such as proteins and peptides to be observed using mass analysers with limited m/z charge ranges. 1.2.2 Matrix-assisted laser desorption ionization (MALDI) Karas and Hillenkarnp developed matrix-assisted laser desorption ionization (MALDI) technique in the late 19805 [7]. MALDI is achieved in two steps. First the analyte is mixed with a solvent containing small organic molecules (matrix) that has a strong absorption at the wavelength of the laser employed for ionization, then allowed to dry on a metal substrate. Second, the resulting solid is placed under vacuum and irradiated by nanosecond laser pulses for a short duration. This generates a rapid heating of the matrix crystals and sublimation of matrix and analyte molecules into the gas phase. Ionization can occur during any time of this process, however, the exact mechanism by which ionization takes place is not yet fully understood [11]. Typically, singly charged pseudomolecular ions are observed by MALDI. 1.2.3 Mass Analyzers The mass analyzer is one of the most important components of the mass spectrometer. Once the ions have been produced, the mass analyzer separates ions according to their mass to charge (m/z) ratio. Key analytical figures of merit in proteomics include resolution, sensitivity and mass accuracy, which are all dependant on the performance of the mass analyzer. There are four basic types of mass analyzers currently used in proteomics. They are (i) quadrupole (ii) quadrupole ion trap, (iii) time of flight (TOF) and (iv) Fourier transform ion cyclotron resonance (FT- ICR) analyzers. The quadrupole and quadrupole ion trap mass analyzers are described in more detail below. 1.2.3.1 Quadrupole Mass Analyzers Quadrupole mass analyzers are made using four parallel hyperbolic or cylindrical rods, upon which is imposed a variable amplitude oscillating AC (RF) voltage superimposed with a variable DC voltage to separate ions according to their m/z ratios. The potentials applied to opposite sets of rods ((D0) are given by the following equations. (D0 = + (U-Vcos wt) and - (D0 = - (U-Vcos wt) where U is the amplitude of the DC potential, V is the amplitude of the AC potential, (1) = 27w is the angular frequency (v is the frequency of the radio frequency) and t is time. The force on the ions in the x and y directions of the quadrupole rods are given by dzx 3(1) Fx—m?——ze$ .................... (I) 2 3(1) Fy=m dtzy =-ze3y- ..................... (2) Where (I) is a function of (D0 2 2 _ 2 2 2_(x -y )*(U-Vcosax) (DU—(Dow —y )/r0 — 2 ........... (3) ’0 After differentiating and rearranging this leads to the following equations 2 $.55]. 2282 (U ~Vcosax)x = O ................... (4) dt rm 0 d2 y 2ze 2 — 2(U-Vcosa)t)y=0 ..................... (5) dt er In this equation, x and y are the distances of the ion from the center of the quadrupoles in the x and y directions, respectively. U and V are variables. As long as the values of x and y do not reach re, the distance between the center and the quadrupole, the trajectories of the ions are stable. These two differential equations are similar to the Mathieu equation written below dzu (1:2 + (au — 2qu cos 2§)u = 0 ..................... (6) In equation (6), ‘u’ is similar to either x or y in equation (4) and (5), where f = % Therefore, the previously determined solutions for au and qu from the Mathieu equation can be used to obtain solutions for U and V. (I -a — a - 82€U u" x"- y‘— mwzroz q “q _ q _ 4zeV u— x“ y‘— mwzroz After rearranging these equations, we obtain the following equations used to derive the Operating parameters for the quadrupole mass analyzer. 22 ma) r0 = a ..................... (7) 28e ma)2 r02 = q“ ... ................... (8) z4e In these solutions, r0 and a) are constants for a given quadrupole. Therefore, for a given m/z, r0, e and (D are constant. A plot of U and V gives the stability areas for different masses where ions have a stable trajectory through the quadrupole. Figure 1.2 shows the stability areas for positive ions with increasing m/z values from m; to m. To obtain a mass spectrum, we can scan the different ions by changing U linearly as a function of V at a constant UN ratio along a straight line near the apex of the stability diagram. When a given UN ratio is selected near the apex, only one m/z ion corresponding to that UN ratio is stable, whereas other ions are unstable and are lost. Thus, the higher the slope of scanning line, the higher the resolution. If the slope of the scanning line is low, a range of m/z ions will be stable at a given UN value, resulting in lower resolution since the stability arrears overlap each other [12]. Quadrupole mass spectrometers have been constructed to perform tandem mass spectrometry (MS/MS) by coupling three quadrupole mass analyzers together in series. In a triple quadrupole mass spectrometer, the first and the third quadrupoles are used to scan ions while the second is used as a collision cell. A product ion scan mode tandem mass spectrum can be obtained by selecting a precursor ion of interest using the first quadrupole and by fragmenting it in the second quadrupole while scanning the third Scanning at constant UN than the total number of basic residues (i.e. combined number of Arg, Lys and His), or “partially mobile” when the total number of basic residues 2 the total number of protons 2 number of Arg residues. [35, 27]. Under ‘mobile’ or partially- mobile’ proton conditions, migration of an ionizing proton may result in localization at an amide nitrogen along the peptide backbone. 18 a, b1 c1 az b2 c2 a3 b3 c3 —l m ‘- R (i R" =O/ \ \ z: / \ n— / / \ 20/ \ Z: / 0:0 / \ HZN OH o :q :1: o 73 X3 3’3 23 x2 Y2 22 xi 3’1 21 Scheme 1.1 Nomenclature for peptide fragment ions. 19 It is generally considered that fragmentation of the amide bond then occurs by nucleophilic attack from an adjacent amide carbonyl to yield complementary b- and/or y- type product ions, as shown in Scheme 1.2. Oxazolone b-type product ions may then lose CO to form a—type ions [36]. The losses of small neutral molecule such as water or ammonia, via similar mechanisms can also be observed [37]. Note that although Scheme 1.2 shows the amide bond cleavage reaction proceeding via an amide nitrogen protonated precursor, it has been shown using theoretical ab initio and semi-empirical molecular orbital calculations that protonation of the carbonyl oxygen of the amide bond is thermodynamically preferred over protonation of the amide nitrogen in peptide ions [38,39,40,41,42]. Thus, fragmentation via an amide carbonyl oxygen protonation precursor should also be considered as a viable dissociation pathway. The presence of additional basic residues in peptide ions can lead to localization of the ionizing protons at their side chains (i.e., resulting in ‘non-mobile’ proton conditions). Under these conditions, higher energies are required for proton mobilization, and as a result, alternative dissociation reactions driven by “charged remote” fragmentation mechanisms may predominate, such as those involving enhanced ‘sequence’ cleavages at the C-terrninal of aspartic acid residues [35,43] or enhanced ‘non-sequence’ cleavages at the side chains of methionine sulfoxide residues [27], suppressing other ‘sequence’ ion formation. 20 I OH \ l o - o E 2>P_ o=xi \ z "3 n / path b H+ transfer y2 ion -HNCO ‘CO -mcnco R A i . + HzN (If/ +\/ HZNVR O immonium ion an ion Scheme 1.2 Mechanisms for the formation of b-type, y-type, a-type and immonium ions. 21 1.4.2 Controlling Peptide Fragmentation Reactions Understanding the mechanisms for peptide fragmentation has been a key step toward developing methods for controlling or directing peptide fragmentation towards the formation of analytically useful product ions from which peptides or proteins may be more readily identified and/or characterized. A number of different methods for the control of peptide fragmentation, involving the introduction of a fixed charge to the peptide, have been developed. 1.4.2.1 N-terminal Fixed Charge Derivatization and Fragmentation Fixed charge derivatization strategies for peptides have been used previously in conjunction with MS/MS for sequencing applications. However, most prior work has been limited to derivatization of the N- and C-termini, and has largely been focused on directing fragmentation toward formation of a series of amide backbone cleavage ‘sequence’ ions [44,45]. Watson and co-workers have demonstrated that N-terminal peptide derivatization with tris(2,4,6-trymethoxyphenyl) phosphonium (TMPP)+- Ac_SC6F5 bromide results in the formation of a well controlled series of N-terminal a- type product ion series under high energy fragmentation conditions [46]. Adamczyk et. al. have used a similar reagent [tris(2,4,6-trymethoxyphenyl) phosphonium] acetic acid N-hydroxysuccinimide ester (TMPP+-Ac_OSU) to generate a series of N-terminal b-type ions accompanied by a less intense a-type product ion series under low energy fragmentation conditions in an ion trap mass spectrometer [47]. It is believed that the formation of N-terminal b- and a—type ions are due to a ‘charge remote’ mechanism involving a shift of the amide hydrogen to the C-terminal carbonyl carbon of the amide 22 bond [48]. Wysocki and co-workers have demonstrated that the selective gas phase cleavage of the amide bond C-terminal to Aspartic acid residues maybe controlled using a ¢3P+CH2C(=O)- [ —> [M +nn]"‘+”+ 0 CH2 (in Ms 0 CH2 (:iijscnzcocfiii5 ll 1.. s t! in i ,J-r \N/ ‘C/ my JJ-f \N/ ‘C/ ‘11” H II H II 0 0 [M] [M+nH+CH2COC6H5]("+”+ Scheme 1.3 Fixed charge derivatization and selective fragmentation of methionine sulfonium ion derivatized peptide ions. 25 O CH II | + Pathway 1 H I HN H C / “LL. 3 \+/\C 1I I H, C\ q 0 Pathway 2a H‘H Pathway l Pathway 2 '0' .3 Pathway 2 C\ n CH /O° ’\/\I\/ \§/ \C/ CH n H / 2. I CH \Pathway 2a H ILHZ ? f‘ I o. H | ”NH—L, Scheme 1.4 Potential mechanisms for the selective fragmentation of methionine sulfonium ion derivatized peptide ions. 26 Thus, if neutral loss or precursor ion scan mode MS/MS and MS3 methods are to be employed for the comprehensive identification of side chain fixed charge peptide ions, the factors influencing this loss must be more fully understood. 1.6 Aims of this Thesis The aims of this thesis are: l. to identify the mechanisms and product ion structures associated with the neutral loss of CH3SR from the side chains of methionine fixed charge sulfonium ion containing peptides, and; 2. to determine the influence of para substituents on the ratio of charged versus neutral loss product ions formed by dissociation of phenacylbromide alkylated fixed charge sulfonium ion containing methionine peptides. 27 CHAPTER TWO Experimental 2.1 Materials Unless stated otherwise all reagents were analytical reagent (AR) grade and used as supplied without further purification. The synthetic tryptic peptides GAILMGAEK and GAILAGAILK were obtained from Auspep (Melbourne, Vic.). Poly(4- vinylpyridiniam tribromide) and Phenacylbromide (Compound 1 in scheme 2.1) were purchased from Fluka (Switzerland). Fmoc-chloride, methanol (HPLC grade), 4- Acetylbenzoic acid, Ethylacetate, Acetyl Chloride, 4-Acety1benzoic acid, D,L- Methionine and D-Methionine were purchased from Sigma-Aldrich (St. Louis, MO, USA). d4-D,L-Methionine was purchased from CDN Isotopes (Canada). Glacial acetic acid (ACS grade), P-Dioxane, Magnesium Sulfate, Sodium Carbonate and Sodium bicarbanate were purchased from Spectrum Chemicals (Gardena, CA, USA). Acetonitrile (HPLC grade) was purchased from OmniSolv (Gibbstown, NJ, USA). Hydrochloric acid and hexane (ACS grade) were purchased from Columbus Chemical Industries (Columbus, WI, USA). Diethyl Ether and Bromine were purchased from Jade Scientific (Canton, Michigan, USA). Ethyl acetate was purchased from Mallinckrodt Chemicals (Phillipsburg, NJ, USA). 4-Ethy1acetophenone and 4-Methylacetophenone were purchased from Acros Organics (New Jersey, USA). All solutions were prepared using deionized water purified by a Barnstead nanopure diamond purification system (Dubuque, Iowa, USA). 28 2.2 Synthesis of (D, L)-, (D) - and (d4-D, L)-Fmoc-Methionine D-Methionine (100 mg), D,L-methionine (100 mg), or d4-D,L-Methionine (100 mg), were dissolved in 5 ml of 10% N32CO3. Then 2.5 mL of P-Dioxane was added and finally, 170 mg of Fmoc-chloride dissolved in 5 mL of P-Dioxane was added dropwise while stirring at 0°C. The reaction was allowed to proceed for one hour at 0°C and then continued for 24 hours at room temperature. The resulting mixture was poured onto 75 ml of ice-water and chloroformate was removed by extracting twice with diethyl-ether (40 mL). The pH of the resulted aqueous layer was adjusted to 2 by adding conc. HCl at 0°C. Fmoc-Methionine was extracted twice into diethyl-ether (40mL) and dried with MgSO4. The solvent was evaporated under reduced pressure and Fmoc-Methionine was recrystallized with EtOAc-hexane. The GAIL-(D)-MGAILK, GAIL-(DL)-MGAILK and ' GAIL-(d4-DL)-MGAILK peptides were then synthesized by Syn Pep Company (Dublin, CA, USA). 2.3 Synthesis of 4-Bromoacetylbenzoic acid (Compound 2 in Scheme 2.1) 4-Acetylbenzoic acid (503 111, 3.37 mmol) was dissolved in acetic acid (20 mL) and the reaction mixture was heated at 70 °C until dissolution was achieved. The reaction solution was then cooled to room temperature and Bromine (190 uL, 3.7 mmol) was added in one portion with stirring which was continued over night at room temperature. The reaction mixture was then filtered, the precipitate washed with water and dissolved in Ethylacetate (20 mL). The insoluble material was filtered and the filtrate washed with water (20 mL x 1). The Ethylacetate was then removed under a gentle stream of N2, then completely dried under high vacuum to yield a white solid. 29 2.4 Synthesis of Methyl-4-bromoacetylbenzoate (Compound 3 in Scheme 2.1) Acetyl chloride (1.6 mL, 2 mol) was added drop wise to Methanol (10 mL) with cooling and stirring in an ice-bath. 4’-Acetylbenzoic acid (328 mg, 2 mmol) was then added in one portion and the reaction mixture was left to stir at room temperature for 3 days. The solvent was allowed to evaporate off overnight at room temperature to yield a white/yellow solid, which was then collected and dried under vacuum. The solid was then dissolved in Acetic acid (10 mL) and Bromine (190 uL, 3.7 mmol) was added in one portion with stirring at room temperature, which was continued over night. The reaction solution was then diluted with H20 (50 mL) and extracted with Diethyl ether (3 x 20 mL). The combined organic extracts were washed with 0.1M NaHCO3 (2 x 20 mL) and H20 (1 x 20 mL), then dried over MgSO4 and concentrated invacuo to give a white solid which was dried under high vacuum. 2.5 Synthesis of 4-Methylphenacylbromide (Compound 4 in Scheme 2.1) 4'-Methylacetophenone (0.5g, 3.7 mmol) was added to poly(4-vinylpyridinium tribromide) (1.86 g , ~3 meq Br3' lg resin) in 20 mL Methanol. The reaction mixture was then stirred at room temperature for 4 hours. The reaction mixture was filtered, and the methanol was evaporated off under a gentle stream of N; to yield a white solid. 2.6 Synthesis of 4-Ethylphenacylbromide (Compound 5 in Scheme 2.1) 4-Ethylacetophenone was dissolved in acetic acid and bromine was added in one portion with stirring at room temperature which was continued over night. The reaction 30 solution went from a dark brown/orange color to a light yellow color over this time. The reaction solution was diluted with H20 (50mL) and extracted with Diethyl ether (3 x 20mL) then H20 (1 x 20mL) then dried over MgSO4 and concentrated under reduced pressure to give a dark yellow liquid. 2.7 Side Chain F ixed-Charge Derivatization of Methionine-Containing Peptides Side chain fixed-charge sulfonium ion derivatives of synthetic methionine- containing model “tryptic” peptides were produced by the addition of 10 [.tL of a 1M solution of alkylating reagent (Compounds 1-5 in Scheme 2.1) to 100 11g of either GAILMGAILK, GAIL-(D)—MGAILK, GAE-(D,L)-MGAILK, GAIL-(d4-D,L)- MGAILK, VTMAHFWNFGK or VTMGHFDNFGR dissolved in 100 uL of aqueous 20% HOAc containing 30% CH3CN. The reactions were allowed to proceed for 24 h at room temperature after which the samples were diluted and introduced to the mass spectrometer without further purification. 2.8 Mass Spectrometry Peptides were introduced to a linear quadrupole ion trap mass spectrometer (Thermo model LTQ, San Jose, CA) by electrospray ionization (ESI). Samples, (0.02 mg/mL) dissolved in 50% MeOH, 1% AcOH were introduced into the mass spectrometer at 0.5 [LL/min. The spray voltage was set at 1.8 kV. The heated capillary temperature was 200°C. CID MS/MS and MS3 experiments were performed on monoisotopically mass selected ions using standard isolation and excitation procedures. 31 O 0 Br OH O 0 Br 0 3 \ O O O Br\/n\©\/ Scheme 2.1 The structures of sulfonium ion derivatization reagents. 32 2.9 Computational Methods 2.9.1 Computational Methods Employed for Mechanistic Studies The model peptide N-acetyl methionine (S,S—dimethy1 sulfonium)-N-methyl amide (CH3CONHCH(CHzCHzS+(CH3)2)CONHCH3) was used for computational calculations. Low energy transition state structures related to the possible mechanisms for dissociation of the methionine sulfonium ion side chain, were initially found at the PM3 semi empirical level of theory, followed by further optimization at the B3LYP level of theory using the 6-31+G(d,p) basis set using the GAUSSIAN 98 (version 5.2) molecular modeling package. Intrinsic reaction coordinate (IRC) searches were then performed, followed by geometry optimizations to locate the appropriate reactant and product ion structures associated with each transition state. All optimized structures were subjected to harmonic vibrational frequency analysis and visualized using the computer package Gauss View 2.1 to determine the nature of the stationary points. Zero point energies were obtained from harmonic frequency calculations without scaling. 2.9.2 Computational Methods Employed for Proton Affinity Calculations Full conformational searches on neutral and protonated para—substituted S-methyl phenacyl derivatives (Compounds 6-15 in Scheme 2.2) were initially performed at the PM3 semi empirical level of theory, followed by further optimization of low energy structures at the B3LYP level of theory using the 6-31 + G (d,p) basis set using the GAUSSIAN 98 (version 5.2) molecular modeling package. All optimized structures were subjected to harmonic vibrational frequency analysis and visualized using the computer 33 package Gauss View 2.1 to determine the nature of the stationary points. Zero-points energies were obtained from harmonic frequency calculations without scaling. 2.9.3 Proton Affinity Calculations Consider the following equation for the molecule M and its protonated structure, MH+. M + H+ —-) MH” (Eq.1) The proton affinity (PA) of a molecule M can be calculated according to the negative of the enthalpy (AH) of Eq. 1 via the equations below, where ARH is the enthalpy of the reaction. PA = -ARH°293 (Eq. 2) The enthalpy of the reaction can be written in terms of the translational (AtE), rotational (AB), and vibrational (AVE) energy of the species, ARH°298 = -AxE + A(AEI.298) + A(AE,,398) + MAE/.298) + APV (Eq. 3) AXE = Eclcc(MH+)- Eclcc(M) + ZPVE(MH+)— ZPVE(M) (Eq. 4) , where APV = -RT, Eclcc is the electronic energy, and ZPVE is the zero point vibrational energy correction (i.e., the difference in translational, rotational and vibrational energy 34 from OK to 298K). The value of the translational energy A(AE(,293) = 3/2RT and the rotational energy A(AE,,293) = 0- Eclec(MH+)r Eelec(M)r AEv.298 (MW), Ev.298(M) and ZPVE values may all be obtained from the output of the structural optimizations described above. Thus, the proton affinity of a molecule can be calculated using the final equation below. ARH 298 = Eelcc(MH+)' Eclec(M) + ZPVE(MH+)‘ ZPVE(M) + AEv,298 (MH+)' Ev,298(M) ' 5/2RT (Eq. 4) 35 o +0H \S \s fir 6 /il\©\'( 11 OH OH 0 o o +OH \S \s 7 12 O\ o\ 0 o o +0H NCO ° EKG ‘3 0 ton \SJKQ 9 \S)K©\ 14 O +on \S/u\©\/ 1o \SJ\©\/ 15 Scheme 2.2 The structures of neutral and protonated para—substituted S-methyl phenacyl derivatives. 36 CHAPTER THREE Mechanisms for the Selective Gas-phase Fragmentation Reactions of Methionine Side Chain Fixed Charge Sulfonium Ion Containing Peptides 3.1 Introduction The gas-phase fragmentation reactions of peptide ions containing a phenacylsulfonium fixed charge on the side chain of methionine residues have recently been examined in order to develop a novel strategy to selectively identify methionine containing proteins by tandem mass spectrometry [50]. It was shown that dissociation of these fixed charge derivatives under low energy CID MS/MS conditions results in exclusive loss of the derivatized side chain, with the formation of a single characteristic side chain cleavage product ion (Scheme 1.3 in Chapter 1). It was further demonstrated that the characteristic loss of the fixed charge sulfonium ion methionine side chain could be used to selectively identify methionine containing peptides from within complex mixtures by selective neutral loss scan mode MS/MS methods. Although several mechanistic possibilities involving intramolecular E2 or 8N2 elimination reactions (Scheme 1.4 in Chapter 1) were proposed for these selective fragmentation reactions in the above mentioned work, extensive studies to determine the operating mechanisms, and the influence of the resultant product ion structures on subsequent fragmentation reactions, were not performed. Determining these mechanisms and product ion structures is of particular importance, as characterization of the amino acid sequence of the methionine containing peptides identified via the diagnostic loss of the side chain may potentially be achieved by further subjecting the initial MS/MS product ions to multistage 37 MS/MS (MS3) in a quadrupole ion trap mass spectrometer, or by energy resolved 'pseudo' MS3 in a triple quadrupole mass spectrometer. Here, a combination of transition state calculations at the B3LYP/6-31+G** level of theory performed on a simple peptide model, as well as experimental studies based on the multistage dissociation of a regioselectively deuterated methionine derivatized sulfonium ion containing tryptic peptide, have been carried out to obtain evidence to support the proposed SN2 reaction mechanism. 3.2 Molecular Orbital Calculations to Examine Mechanisms for the Loss of CH3SR from Methionine Fixed Charge Sulfonium Ion Containing Peptides. As shown in Scheme 3.1, the neutral loss of CH3SR from the side chains of methionine fixed charge containing peptide ions could occur via several different mechanistic pathways. Pathways 1 and 2 of Scheme 3.1 shows neighboring group participation reactions involving either nucleophilic attack from the carbonyl group of the C-terminal amide bond, or the carbonyl group of the N-terminal amide bond, respectively. These reactions pathways would result in the formation of protonated five membered hydrofuran or six membered oxazine product ions, respectively. Alternatively, pathways 3 and 4 of Scheme 3.1 show E2 elimination reactions involving intramolecular proton transfer from the [3 methyl group of the methionine fixed charge containing side chain to the N or C-terrninal amide carbonyl oxygens to yield protonated acyclic 3- amino-l-butenoic acid (vinyl glycine) product ions. Furthermore, the five membered hydrofuran and the six membered oxazine product ions formed from Pathways l and 2 in Scheme 3.1 could potentially undergo intramolecular proton transfer ring opening 38 / 14.3 kcal mol ' 312:0 Pathway 1' cl)" 1 27.0 kcal mol'1 Pathway 1/\C .. C< 45.0 kcal mol" /C\§ C/O Pathway 3 \i‘l C\\ " l HN;\ HN\ \t/R \g/ Pathway 2 S) Pathway l Pathway" 4 ) ii)? ICI,‘ 0)" HPathway 3 /c<‘. 2° / 0 5% H H .1; .13 Pathway 2 $6.31"? Pathway 2' [if I Pathway 4 c " o 16.2 kcal mol ‘ / \ c; 45. 3 kcal mol ' /C \N /<|3l/OH 28. 3 kcal mol ' .13 ..i Scheme 3.1. Possible reaction mechanisms to account for the neutral loss of CH3SR from methionine side chain fixed charge containing peptide ions. 39 reactions to yield the vinyl glycine product ions (Pathways 1' and 2' in Scheme 3.], respectively). To determine which of the mechanistic possibilities described in Scheme 3.1 are energetically feasible, a series of molecular orbital calculations were performed to determine the transition state barriers associated with each reaction pathway. Potential transition state structures for each of the possible mechanisms shown in Scheme 3.1 were initially examined at the PM3 semi empirical level of theory, using a simple peptide model, N-acetyl methionine-N-methyl amide, containing a fixed charge on the thioether side chain, CH3CONHCH(CH2CH2S+(CH3)2)CONHCH3. Low energy conformers were then re-optimized at the B3LYP/6-31+G** density functional level of theory. Vibrational frequency analysis was then performed to determine the nature of the optimized stationary point structure. Then, intrinsic reaction coordinate searches were performed, followed by geometry optimization at the same level of theory to locate the appropriate precursor and product ion structures associated with each transition state structure. The total energies, zero point vibrational energies (ZPVE) and relative energies (total energies + ZPVE, relative to structure (A)) obtained for each of the optimized low energy transition state structures, as well as those of the optimized precursor, intermediate and product ion structures are given in Table 3.1. Structure TSl in Figure 3.1(A) shows the predicted low energy transition state structure for the loss of CH3SCH3 to form a protonated five membered hydrofuran product ion via an SN2 neighboring group participation reaction involving the amide bond C-terminal to the methionine residue. After optimizing this structure, IRC calculations followed by geometry optimizations were performed to locate the precursor ion (A) and 40 Table 3.1 Total energies (Emmi), zero point vibrational energies (ZPVE) and relative energies (E...) computed for the precursor ions, transition states, intermediate product ions and product ion structures associated with each reaction pathway at the B3LYP/6- 31+G** level of theory. Structure Emmi (Hartree) ZPVE (Kcal mol'r) Em. (kcal mol'])“‘5 A -1012.103755 178.01581 0.0 TS] -1012.080944 176.84406 +14.3 B -1012.1 12208 178.01406 -5.3 C -534. 1521621 177.46909 +0.9 T52 -534.081918 124.64946 +450 D -534.1330484 128.06278 +12.9 E -1012. 100463 177.68801 +2.1 T33 -1012.07791 176.74105 +16.2 F - 1012.1 14764 178.32422 -6.9 G -534. 150836 130.07242 +1.7 T54 -534.0813041 124.88532 +453 H -534.l325096 127.84273 +13.2 I -1012.096112 [77.93925 +4.8 T55 -1012.0607l3 173.96721 +27.0 J -1012.092994 [76.09255 +6.7 K -534. 1324999 127.8488 +13.2 41 Table 3.1 (continued) L -1012.101371 177.91885 +1.5 T86 ~1012.058067 173.63621 +28.7 M -1012.0951 1 175.66895 +5.4 -534. 1330465 128.06403 +12.9 O -1012. 107568 178.4326 -2.3 T87 -1012.058023 173.90797 +28.7 P -1012.09213 175.65843 +7.3 Q -534. 1306281 128.0657 +14.4 -1012.09593 178.13605 +4.9 T88 4012058689 173.865 +28.3 S -1012.088737 175.95845 +9.4 T -534. 1281226 127.75565 +16.0 U -534. 1478423 129.97671 +3.6 T89 -534.0731 1 13 12495592 +505 V -534. 1404751 127. 10571 +8.2 T810 -534.0641064 124.99521 +56.l W -534. 1249397 127.93632 +18.0 8(CH3)2 -477.9502026 47.51316 42 a13ml: total energy + (ZPVE). bEnergy relative to structure A. the intermediate product ion (B) associated with this reaction coordinate. The relative energies of the transition state and the product ion were then calculated with respect to the energy of the precursor ion. The product ion (C) was then obtained by geometry optimization following the removal of CH38CH3 from structure (B). The relative energy of (C) was obtained by summing the energy of the product ion (C) and the energy of the geometry optimized CH3SCH3. Structure (T83) in Figure 3.1(B) represents the low energy transition state structure for the neutral loss of CH3SCH3 via a neighboring group participation reaction involving the amide bond N-terminal to the methionine residue to form a protonated six membered oxazine product ion. After performing IRC calculation on transition state (T83), the optimized precursor ion, (E) the intermediate (F) and the product ion (G) were obtained as mentioned above. The energies of all these structures are given with respect to the energy of structure (A). According to Figure 3.1(A), the predicted transition state barrier for pathway 1, to yield the cyclic protonated five membered hydrofuran ring, is 14.3 kcal mol'1 with respect to the energy of structure (A), whereas the transition state barrier to form the six membered oxazine ring via pathway 2 is predicted to be 16.2 kcal mol'1 relative to the energy of the structure (A). The predicted transition state barriers for pathways 1 and 2 (structures (T81) and (T83) in Figure 3.1) are therefore relatively comparable. The slightly lower energy of transition state (T81) is likely due to better hydrogen bonding between the carbonyl oxygen and the amide nitrogen on the N- and C-terminal amide bonds (1.93 A in (T8 1) versus 2.08 A in (TS3)). The predicted energy difference between 1 these transition states is not considered significant enough to suggest the formation of one 43 1.85 A \ "’8‘ .4 Pathway 1 2 2 A 2. 49 A 85A Pathway 2 2. 05 A 2.48 A 2. 27 “A 2.08 A 2.02 A 1.93 A (A) (T81) (T83) 0.0 kcal mol“1 14.3 kcal mol'l 7 2.1 kcaEl) mol ' 16. 2 kcal mol l 1.48 A 3 02 A\- 1 47A 1. 75 A 1.99 A 1 74 A (G) 0.9 kcaCl )'mol ’ -.5 3 kcal mol ' 1.7 kcal mol‘l 6. 9 kcal mol ' lPathway l' l Pathway 2' 1.27 A 1.24 A 1-34 A 2.19 A 139 A—. 1.4.49 A 1.87 A 1.70 A (T82) (D) (T84) (H) 45.0 kcal mol" 12.9 kcal mol'l 45.3 kcal mol’I 13. 2 kcal mol’l Figure 3.1 Optimized precursor, transition state and product ion structures (at the B3 LYP/ 6-31+G** + ZPVE level of theory) for the loss of CH3SCH3 from the simple model peptide N-acetyl methionine-N-methyl amide, due to a neighboring group participation reaction involving (A) the C-terminal carbonyl group to form a 5 membered hydrofuran product ion (Pathway 1), or (B) the N-terminal carbonyl group to form a 6 membered oxazine product ion (Pathway 2). The transition state and product ion structures for ring opening of the product ions formed from Pathways 1 and 2, involving intramolecular proton transfer (Br-H) to either the N-terminal carbonyl group (Pathway 1’ in A) or the C- terminal carbonyl group (Pathway 2’ in B) to yield protonated vinyl glycine product ions, are also shown. The energies of all molecules are given with respect to structure A. Images in this thesis/dissertation are presented in color 44 particular product over another for the loss of CH3SCH3 via either reaction pathways 1 or 2. Thus, both the five and six membered rings could potentially be formed. Figure 3.2(A) shows the predicted low energy transition state structure (T85) found for the E2 elimination reaction involving intramolecular proton transfer of one of the two hydrogens from the B methyl grOup of the methionine side chain to the C- terrninal carbonyl group to subsequently yield a vinyl glycine product ion. IRC calculations were again performed to obtain the appropriate precursor and intermediate product ions (D and (J), respectively. The structure of product ion (K) was obtained by optimization of intermediate structure (J) after removal of CH38CH3, as mentioned above. The energies of these structures are given with respect to the energy of structure (A) in Figure 3.1. Figure 3.2(B) shows the predicted low energy transition state structure (T86) associated with the E2 elimination reaction involving intramolecular proton transfer of one of the two hydrogens from the [3 methyl group of the methionine side chain to the N-terminal carbonyl group. The energies of this transition state relative to structure (A) in Figure 3.1 was predicted to be 28.7 kcal mol'l. Furthermore, as there are two hydrogens on the [3 methyl group in the methionine side chain, transition-state calculations predicting the barriers to the E2 elimination reactions involving the alternate hydrogens to those shown in Figure 3.2 were also determined. The structures associated with these transition states (structure (T87) in pathway 3-2 and structure (T88) in pathway 4-2) are shown in Figures 3.3. The relative energies of these transition states were predicted to be 28.7 and 28.3 kcal mol", respectively. Thus, the four E2 elimination reaction pathways all have comparable activation barriers. 45 2.30 A ‘ LSSQ Pathwa 3 1-4' A Pathwa 4 1.22 2.23 A —yo r 1.24 A 3.30 A 2.28 A __Z 2'07 1.93 A 11) (T55) (L) (T56) 4.8 kcal mol'l 27.0 kcal mol " 1.5 kcal mol'| 28.7 kcal mol " .49 A + 5((‘H312 o——— L73 A +S(CH3)2 *— (é 1.7011 “87A LssA (K) U) (N) (M) 13.2 kcal mol " 6.7 kcal mol'l 12.9 kcal mol " 5.4 kcal rnol‘l Figure 3.2 Optimized precursor, transition state and product ion structures (at the B3 LYP/ 6-31+G** + ZPVE level of theory) for the loss of CH3SCH3 from the simple model peptide N-acetyl methionine-N-methyl amide due to intramolecular proton (Br-H) transfer to (A) the C-terminal carbonyl group (Pathway 3) or (B) the N-terminal carbonyl group (Pathway 4) to yield protonated vinyl glycine. The energies of all molecules are given with respect to structure A in Figure 3.1. 46 Pathways 1 and 2 in Figure 3.1 can be compared with pathways 3, 3-2, 4 and 4-2 in Figures 3.2 and 3.3 to predict the favorable mechanistic pathway for the neutral loss of CH38CH3. The transition states associated with the neighboring group participation processes (T81) and (T83) have energies about 10 kcal mol'1 lower than those for the E2 elimination reactions (T85), (T86), (T87) and (T88). Thus, given that the significant difference in transition state barriers between the neighboring group and E2 elimination reaction pathways suggest that pathways 1 and 2 are likely to be more favorable, it is expected that the loss of CH38R from the side chain of the methionine fixed charge sulfonium ion peptides will result in the formation of the five membered hydrofuran and/or six membered oxazine product ions. 3.3 Potential Ring Opening of the Five and Six Membered Cyclic Product Ions via Intramolecular Proton Transfer Reactions. Having established above that the initial loss of CH3SR from the side chain fixed charge sulfonium ion containing peptides following MS/MS is likely to occur via neighboring group participation mechanisms involving pathways 1 or 2 in Scheme 3.1, it was then considered whether the resultant cyclic product ions could undergo ring opening to yield the acyclic vinyl glycine containing product, via intramolecular proton transfer reactions according to pathways 1' and 2' in Scheme 3.1, prior to their further dissociation by M33. Structure (T82) in Figure 3.1(A) shows the predicted low energy transition state for the ring opening reaction of the five membered hydrofuran product ion via pathway 1’ in Scheme 3.1. The activation barrier for pathway 1' was predicted to be 45.0 kcal mol'1 47 1.86A 2.33A 1.86A 241A A Pathway 3 -2 1.361 1.4011 1Q; . 1.29A Pa‘hwaY“ 125A.\ 2.12A 1.87A 82.27A (T57) -.(24kcal mol' 28.7 kcalmol‘l 4.9kc§l)'mol' 28.3(1tcalmolI l l 2.64 A + S(CH3)2 ‘— +S(CH3)2‘_ 9» 1.76 A ‘ fl 2.01 A 2.47 217A 1.73 A "86 A (Q) (P) (T) (S) 14.4 kcal mol‘l 7.3 kcal mol'l 16.0 kcal mol'| 9.4 kcal mol'I Figure 3.3 Optimized precursor, transition state and product ion structures (at the B3 LYP/ 6-31+G** + ZPVE level of theory) for the loss of CH38CH3 fi'om the simple model peptide N-acetyl methionine—N-methyl amide involving the alternate intramolecular proton (Bz-H) transfer to (A) the C-terrninal carbonyl group (Pathway 3-2) or (B) the N- terrninal carbonyl group (Pathway 4-2) to yield protonated vinyl glycine. The energies of all molecules are given with respect to structure A in Figure 3.1. 48 with respect to the energy of structure (A). The structure of transition state (T82) associated with pathway 1’ is similar to the structure of transition state (T86) associated with pathway 4, and result in the formation of identical product ions (compare structure (D) in Figure 3.1A and structure (N) in Figure 3.28). However, it can be seen from Figures 3.1 and 3.2 that transition state (T82) is predicted to have a significantly higher energy (45.0 kcal mol") compared to transition state (T86) (28.7 kcal mol"). Justification for this difference may be obtained by consideration of the bonds being cleaved in each case i.e., transition state (T82) involves cleavage of a strong carbon oxygen bond, whereas transition state (T86) involves cleavage of a weak carbon sulfur bond. The predicted transition state barrier for pathway 2’ (transition state (184) in Figure 3.1), involving ring opening of the six membered oxazine was predicted to be 45.3 kcal mol'l with respect to the energy of structure (A), which is comparable to that predicted for ring Opening of the five membered ring in Pathway 1’. Similar to that discussed above, the structure of transition state (T84) is similar to the structure of transition state (T85) associated with pathway 3, and again result in the formation of identical product ions (compare structure (H) in Figure 3.1B and structure (K) in Figure 3.2A). However, the relative energy of transition state (T85) was predicted to be significantly lower than that for transition state (T84). The transition state barriers for ring opening of the five membered hydrofuran product ion via intramolecular proton transfer pathways 1’-2 and 2’-2, involving the alternate hydrogens to those shown in Figure 3.1 were also determined (structures (T 89) and (T810) in Figure 3.4). The predicted barriers relative to structure (A) for these 49 A B 1.49A 1.19A 1.524 1.4oA1.21A Pathway I‘-2 1.63 A 1119 A Pathway 2’2 [“156 A 2.59 A —" ——’ 2.22 A 5 1.99 A (U) (T59) (G) (T310) 3.6 kcal mol'l 50.5 kcal mol" 0.0 kcal mol'1 56.] kcal mol'l l 1.33 A 2.25 A 2‘ 2.65 A 1.35 A 3.01 A (V) (W) 8.2 kcal mol‘l 18.0 kcal mol" Figure 3.4. Optimized precursor, transition state and product ion structures (at the B3 LYP/ 6-31+G** + ZPVE level of theory) for ring opening of the product ions formed . from Pathways 1 and 2, involving the alternate intramolecular proton (Bz-H) transfer to either the N-terminal carbonyl group (Pathway 1’ in A) or the C-terminal carbonyl group (Pathway 2’ in B) to yield protonated vinyl glycine product ions. The energies of all molecules are given with respect to structure A. 50 pathways were found to be 50.5 and 56.1 kcal mol", respectively. Thus, each of the transition state barriers predicted for pathways 1’, 1’-2, 2’ and 2’-2 are all significantly higher in energy than those associated with pathways 1, 2, 3, and 4. Recently, Paizs et. al. have predicted that the transition state barriers for amide bond cleavage reactions range from 29 to 33 kcal mol", depending on the proton mobility [52]. Thus, the predicted low energy transition states (T8 1) and (T83) for the neutral loss of CH3SR via the neighboring group participation reactions discussed above are around 15 kcal mol'l lower than these predicted amide bond fragmentation barriers. Thus, it is expected that fragmentation of the methionine fixed charge sulfonium ion containing peptide precursors under the ‘slow heating’ collision induced dissociation methods commonly employed for tandem mass spectrometry [53], would preferentially result in formation of the five membered hydrofuran and srx membered oxazine product ions. Furthermore, given that the barriers to amide bond cleavage are expected to be much lower than the barriers predicted for the ring opening reactions of these initial cyclic product ions, it is expected that these ring opening reactions are unlikely to occur, and that further dissociation of the initial product ions would result in fragmentations along the peptide backbone. A summary of the predicted activation barriers associated with the various mechanistic pathways for dissociation of methionine side chain fixed charge containing sulfonium ion peptides discussed above is given in Figure 3.5. 51 Intramolecular Proton Transfer E (45 kcal mol“) ' E2 elimination (27 kcal mol") Neighboring Group (14.3 kcal mol") Vinyl glycine Vinyl glycine I" ' \ I \ \ ‘ . 9 0 ‘ V 7 I l l I D I Hydrofuran / Oxazine Figure 3.5. Summary of the predicted low energy transition state barriers (at the B3LYP/6-3l+G** +ZPVE level of theory) for the dissociation of methionine side chain fixed charge containing sulfonium ion peptides. 52 3.4 Obtaining Experimental Evidence for the Mechanisms Responsible for the Loss of CH3SR from Methionine Fixed Charge Sulfonium Ion Containing Peptides. In order to obtain experimental evidence to support the results from the theoretical calculations discussed above, the multistage tandem mass spectrometry (MS/MS and M83) fragmentation reactions of a sulfonium ion derivatized regioselectively deuterated methionine containing peptide, as well as those of a simpler alanine containing model peptide, have been examined in a linear quadrupole ion trap. It can be seen from Scheme 3.1 that loss of the methionine fixed charge containing side chain via either an E2 elimination reaction would involve transfer of one of the B-methyl hydrogens in the methionine side chain, resulting in ‘mobilization’ of this hydrogen along the peptide backbone. This ‘mobilization’ would also occur if either of the B-methyl hydrogens were involved in intramolecular proton transfer ring opening reactions of the cyclic five or six membered hydrofuran or oxazine product ions initially formed via the neighboring group participation reactions. In either case, facile migration or ‘scrambling’ of this hydrogen (proton) along the peptide backbone would be expected prior to or during further dissociation of the side chain cleavage product ion by MS3 [54]. 3.4.1 Regioselective Deuterium Labeling Experiments to Examine Mechanisms for the Side Chain Fragmentation Reactions of Methionine Fixed Charge Containing Peptide Ions. To determine whether or not facile scrambling of the methionine side chain [3- methyl hydrogens does occur prior to or during M83 dissociation, the MSIMS and M83 fragmentation reactions of the synthetic peptide GAH.(d4-M)GAILK, containing a 53 methionine side chain fixed charged phenacylsulfonium ion derivative, and where the hydrogens at the B and 7 carbon positions of the methionine side chain were re gioselectively exchanged for deuterium, were examined. Figures 3.6A and B show the CID MSIMS product ion spectra of the doubly [M’i-t-H]2+ and triply [M++2H]3+ charged precursor ions of GAIL(d4-M(R)GAILK, where R = CH2COC6H5. From the doubly charged precursor ion product spectrum in Figure 3.6A, exclusive neutral loss of the fixed charge methionine side chain CH3SR was observed. By examination of the product ion spectra obtained from the triply charged precursor in Figure 3.6B, it can be seen that the neutral loss of CH3SR is also accompanied by ‘charged’ loss of the fixed charge methionine side chain, giving rise to complementary —(CH3SR+H+) and CH38R+H+ product ions. The formation of these protonated ‘charged loss’ product ions indicate that the transfer of a deuterium atom via an intermolecular deuterium transfer from the B methyl group of the methionine side chain similar to pathways 1’ or 2’ in Scheme 3.1 (see Scheme 3.2) does not occur. Therefore, this result provides evidence to support the contention that fragmentation of the fixed charge methionine side chain does not result in the formation of a vinyl glycine containing product ion. Figure 3.6C shows the MS3 product ion spectrum obtained by dissociation of the neutral loss [M++H-CH3SR]2+ product ion of the [M++H]2+ precursor of GAIL(d4- M)GAILK from Figure 3.6A, from which a clear series of b- and y-type ions are observed. As mentioned above, if dissociation of the fixed charge containing precursor ion occurs via an E2 elimination reaction, or if ring opening of the initial cyclic product ion occurs via intramolecular proton transfer, a deuterium atom from the [3 methyl group 54 100 ‘ A _. [M++H]2+ 1 l 2110i 4110’ .600197807011'1000 11200 8 100 _ -CH3SR g _ B “E 3 -(CH,SR+H+) q, ‘ [M++2H]3* .g ~CH3SR+Ht l .‘3 l 1 .. --- q) . 1 . . r 3 . . r 4 ~ . 4 4 w . r . 1% CZ 200 400 600 800 1000 1200 _ Y5 100 2 3 C yr;+ 87Da ‘——> r 87Da Y6 Y7 b9 3 F‘B b 3 y] b3 :94 5 b7 8 Y8 l 1;-121 12JL1111-111 .11- -11-- - T11 I 200 400 600 800 1000 m/z Figure 3.6 Linear ion trap CID MSn analysis of a methionine side chain phenacyl sulfonium ion fixed charge derivative of GAIL(d4-M)GAILK. (A) CID MSIMS product ion spectrum of the [M++H]2+ ion. (B) CID MSIMS product ion spectrum of the [M"+2H]3+ ion. (C) CID M33 product ion spectrum of [M++H-CH3SR]2+ ion from Figure 3.6A. 55 "CH3s D ‘1’ f * + CH SR+D 1 9" $6 /“ 3 c N c / \ c Pathway] E H}? H H111\ \ \+/R 5) Pathway 2 Pathway 1 ‘11) \ /C<1§1 céO H 1.11 r _ D 01,516 D Pathwa 2 P o y 9») D II 1 + CH3SR+D+ /C<;i *0 /C\N c¢O H 1 H I HN\ "N\ Scheme 3.2. Potential intermolecular deuterium transfer reactions for the charged loss of CH38R+D+ from methionine side chain fixed charge containing peptide ions. 56 of the methionine side chain would be mobilized in the resultant product ion to a position along the peptide backbone. As a result, statistical ‘scrambling’ of this deuterium along the peptide backbone would be expected. However, it can be seen from Figure 3.6C that a mass difference of 87 Da is observed between the ys and y6 product ions, as well as between the b4 and b5 product ions. These ions represent product ions formed by fragmentation of the amide bonds on the N and C-terminal sides of the methionine residue. This mass difference, as well as the lack of evidence for deuterium incorporation into any of the other product ions formed, indicates that all four deuterium atoms are located in the side chain of the product ion, and are therefore not involved in the initial fragmentation reaction. This data further indicates that the E2 elimination and ring opening reactions are not involved in fragmentation of the fixed charged methionine side chain sulfonium ion peptides, and is therefore strongly suggestive of the formation of cyclic hydrofuran and oxazine product ions. 3.4.2 Multistage Tandem Mass Spectrometry Experiments to Examine the Structure(s) of the Product Ions formed by Side Chain Fragmentation of Methionine Fixed Charge Containing Peptides. The MSIMS and MS3 data obtained from the deuterium labeling experiments above indicate that fragmentation of methionine side chain fixed charge sulfonium ion containing peptides results in the formation of cyclic five and/or six membered hydrofuran or oxazine containing product ions via neighboring group participation reaction mechanisms, and that these product ions do not undergo ring opening via intramolecular proton transfer reactions prior to further dissociation. 57 Scheme 3.3 shows the proposed structures of the five and six membered product ion formed via the loss of CH3SR from GAILM(R)GAILK (Structures A and B, respectively) and the structure of a ‘control’ peptide GAILAGAILK (structure C). The fragmentation of product ions containing these cyclic structures (e.g., the y6 and y7 ions from structure A, and the y7 ion from structure B) should differ from the fragmentation of the same product ions formed from a ‘control’ peptide GAILAGAILK lacking these cyclic structures (structure C in Scheme 3.3). Therefore, in order to examine the structure(s) of the initial product ions formed by side chain fragmentation of methionine fixed charge containing peptides, the multistage (MS3 and M84) product ion spectra of the doubly charged neutral loss product ion from the GAILM(R)GAILK peptide, as well as the MSIMS and M83 product ion spectra of the doubly charged ‘control’ peptide GAILAGAILK peptide have been obtained. Figures 3.7A shows the mass spectra obtained by MS3 of the neutral loss [M++H-CH3SR]2+ product ion from the doubly charged precursor of GAILM(R)GAILK, while Figures 3.7B and 3.7C show the mass spectra obtained by MS4 of the y6 and y7 product ions from Figure 3.7A, respectively. Figures 3.7D-F show the MS/MS spectrum obtained from the [M+2H]2+ precursor ion, as well as the MS3 spectra obtained from the resultant yo and y7 ions. Clearly, the spectra obtained by dissociation of the [W+H- CH3SR12+ ion from GAILM(R)GAILK and the [M+2H]2+ ion of GAILAGAILK are quite similar, with a complete series of b- and y-type ions observed in each case. Although y5 and y6 product ions corresponding to fragmentation on either side of the methionine side chain in the GAH.M(R)GAILK peptide are both observed, fragmentation to yield the y5 ion is unlikely for the cyclic five membered hydrofuran containing product ion, 58 o o o H II GAl—C N—CH-ICI C/—r-N—CH2-C—A1LK CH2 :y ”3 o o o 0 || || || 11 ll GAi—C N—CH-C CH— C N—CHz-C—AILK y7CH-CH3 y6 ”3 Scheme 3.3 Structures of (A) the five membered hydrofuran product ion formed via the loss of CH3SR from GAILM(R)GAILK, (B) the six membered oxazine product ion formed via the loss of CH3SR from GAILM(R)GAILK, and (C) GAILAGAILK. The amide bond cleavage sites for formation of the y5, yr, and y7 product ions in each case are indicated. 59 100~ ys A ‘ yfi’ 83Da 1H y‘ Y1 b9 b5 b3 b4 b7 b8 Y8 yl . -l .. - l. 200 400 600 800 1000 100- NH3 6 B E s: :1 .o < 0 -1 .2 E b,+HZO ” b+HO ' a: b,+H,o b 4 2 b3! 11 1412 1 1 .1 I L 250 350 450 550 650 i ‘ -H O 4 \NH, 1‘ 1‘ L J_. . 1 Lry?- I 200 300 400 500 600 700 800 m/z 1’.-l. _‘ .-lyl 600'800‘10'00 y. r b5+HZO b, or y: Y3 . 1 ¥6 —4 AAA- 1 a Y b 4 3.111. 21.1 . 250 350 ‘ r 450 550 650 b6 b or ° 1364-1120 71 Y7 1’1 11 [611,6 b 200 300 400 500 600 700 800 m/z Figure 3.7 Multistage (MS/MS, MS3 and M84) tandem mass spectrometry analysis of GAILM(R)GAILK (where R=CH2COC6H5) and GAILAGAILK. (A) CID Ms3 product ion spectrum of the [M++H-CH3SR]2+ ion from GAILM(R)GAILK. (B) CID Ms4 product ion spectrum of the y5 product ion from Figure 3.7A. (C) CID MS4 product ion spectrum of the y7 product ion fi'om Figure 3.7A. (D) CID MS/MS product ion spectrum of the [M+2H]2+ ion of GAILAGAILK (E) CID MS3 product ion spectrum of the y6 product ion from Figure 3.7D. (F) CID MS3 product ion spectrum of the y7 product ion from Figure 3.7D. 60 due to the double bond character of the amide bond, whereas fragmentation to yield the y6 ion is not likely to occur in the cyclic six membered oxazine containing product ion due to the presence of the oxazine ring. Thus, the fact that both ions are observed in Figure 3.7A suggests that both cyclic products may be formed, consistent with the results from the theoretical transition state calculations described earlier. It can be seen by comparison of Figure 3.7B with 3.7E and Figure 3.7C with 3.7F that the product ion spectra obtained by dissociation of the y6 and y7 ions from the two peptides are very different. It is expected that these differences are due to the presence of the cyclic structures in the GAILM(R)GAILK peptide upon loss of the side chain. The series of product ions observed in Figure 3.7B and F by dissociation of the y6 and y7 ions from the [M+2H]2+ ion of the GAILAGAILK peptide correspond to a relatively complete series of a-, b-, and y-type ions. In contrast, the yr, ion from the [M"+H-CH3SR]2+ ion of the GAILM(R)GAILK peptide may contain only the five membered hydrofuran ring, whereas the y7 product ion may contain either the five membered hydrofuran or six membered oxazine rings. The dominant product ions observed by dissociation of these yrs and y7 ions (Figure 3.7B and 3.7C, respectively) correspond to the loss of NH3 and/or the formation of a yr, ion. The loss of N H3 in Figures 3.7B is consistent with fragmentation of the five membered hydrofuran ring containing y6 precursor ion, while the formation of the y5 ion in Figure 3.7C is consistent with fragmentation of the six membered oxazine ring containing y7 product ion. 61 3.5 Conclusion Taken together, the results from the theoretical calculations and experimental data presented here indicate that the loss of CH3SR from methionine side chain fixed charge sulfonium ion containing peptide ions occurs via neighboring group participation reactions, resulting in the formation of a mixture of both five membered hydrofuran and six membered oxazine product ions. The data also suggests that these cyclic product ions do not rearrange by intermolecular proton transfer reactions prior to or during further dissociation. However, the resultant product ion spectra obtained by MS3 are not significantly different from the spectra obtained from ‘control’ acyclic peptides, particularly with respect to the formation of product ions resulting from cleavage of amide bonds either side of the site of the loss of the fixed charge, and is likely due to the presence of a mixture of both cyclic structures formed from the initial neutral loss. 62 CHAPTER FOUR Substituent Effects on the Fragmentation Reactions of Methionine Side Chain Fixed Charge Phenacylsulfonium Ion Containing Peptides. 4.1 Introduction It was observed in Chapter 3 that fragmentation of the triply charged phenacylsulfonium ion derivative of GAILM(R)GAILK resulted in the formation of protonated ‘charged loss’ product ions in addition to the expected neutral loss product (Figure 3.6 B). These charged loss products are formed via intermolecular proton transfer from the triply protonated peptide product ion to the methylphenacylsulfide side chain fragment following the initial fragmentation reaction. Control over the formation of the characteristic low m/z charged loss product ion could potentially be used to selectively identify the presence of fixed charge methionine containing peptides from within complex mixtures by using a precursor ion scan mode MSIMS experiment. Alternatively, the complementary high m/z charged loss peptide product ion could be used for selective identification of the fixed charge methionine containing peptides by using a ‘neutral gain’ MSIMS scan [55]. It is expected that the ability for proton transfer from the peptide product ion to the side chain fragment will vary according to differences in their respective proton affinities. The proton affinity of the neutral side chain fragment may potentially be modulated as a function of the identity of the para-substituent of the alkylating reagent employed for the initial fixed charge derivatization. Substituents that result in the formation of a more acidic neutral molecule should result in more abundant neutral loss, while more basic substituents should result in more abundant charged losses. 63 However, charged losses are only expected when the proton affinity of the resultant peptide product ion is relatively low i.e., when the number of total charges on the resultant peptide product ion exceeds that of the number of basic sites within the ion. Here, the theoretical proton affinities of a series of para-substituted (-COOH (structure 6 in Figure 4.1), -COOCH3 (structure 7 in Figure 4.1), -H (structure 8 in Figure 4.1), -CH3 (structure 9 in Figure 4.1) and -CH2CH3 (structure 10 in Figure 4.1)) side chain fragments have been calculated by using high level density functional theory calculations of the neutral species and protonated ions at the B3LYP/6-3l+G** +ZPVE level of theory. Then, the influence of these side chain fragment proton affinities on the relative abundance of neutral versus charged side chain losses resulting from the dissociation of methionine fixed charge containing peptide ions have been determined by examination of the fragmentation reactions of a series of para-substituted phenacylsulfonium ion containing peptides. 4.2 Proton Affinity Calculations The structures of the various para-substituted neutral side chain fragments formed via cleavage of the side chain methionine fixed charge sulfonium ion peptides, as well as their protonated ions, were initially subjected to full conformational searches at the PM3 level of theory. Then, low energy conformers were re-optimized at the B3LYP/6-31+G** density functional level. Vibrational frequency analysis was then performed to determine the nature of the optimized stationary point structure, and to obtain values for the zero point vibrational energies (ZPVE) for subsequent proton affinity calculations. Figure 4.1 shows the optimized low energy structures for each of the neutral species and protonated ions determined from these calculationsnThe proton affinities of the neutral structures were then calculated according to the method described in section 2.9.3 of Chapter 2. The proton affinities derived from these calculations are given in Table 4.1. According to these results, the proton affinity of the neutral fragment increases from 210.6 — 219.0 kcal mol'1 in the order of structures 6, 7, 8, 9 and 10. These results are consistent as being due to the expected changes in the electron density of the fragments, as determined from the Hammett equation o-values for the substituents (Table 4.1) [56,57,58]. 4.3 Determining the Influence of Side Chain Fragment Proton Affinities on the Relative Abundance of Neutral versus Charged Side Chain Losses from Methionine Fixed Charge Containing Peptide Ions. Figure 4.2A-C shows the CID MSIMS product ion spectra obtained by dissociation of the singly [M]", doubly [M++H]2+ and triply [M++2H]3+ charged methionine side chain fixed charge phenacylsulfonium ion derivative of the synthetic peptide GAILMGAILK, respectively. It can be seen from these figures that dissociation of the singly and doubly charged precursor ions results in exclusive neutral loss of the side chain (labeled as —CH3SR). In contrast, dissociation of the triply charged precursor ion results in the formation of both neutral loss (—CH3SR) and charged loss (labeled as - (CH3SR+H+) and CH38R+H+) product ions. This result is consistent with the expected proton affinities of the resultant peptide product ions. The proton affinity of the peptide product is expected to decrease with increasing charge state. The ionizing proton on the singly charged peptide product ion is likely to reside at either the 8-amino group of the lysine side chain or along the peptide backbone, while the two ionizing protons on the 65 11 . A 209 I 7 12 2.70:5v 2.10.11, 8 2.72‘A/ 2.72 A; 2.11 A Figure 4.1 Optimized low energy structures at the B3LYP/ 6-31+G** + ZPVE level of theory for the neutral and protonated products of the para-substituted methylphenacylsulfide derivatives employed for proton affinity calculations. 66 Table 4.1 Total energies (Emmi), zero point vibrational energies (ZPVE), vibrational energies (Evrbmgmr) and proton affinities computed at the B3LYP/6—31+G** level of theory for the para-substituted methylphenacylsulfide derivatives. P 31’ 3' S 1111C ture EToml ZPVE EVibrational PI 0101] Affinity substituent (Hartree) (kcal mol") (kcal mol") (kcal mol'l) _ C OOH Neutral (6) -101 1.009419 1 14.375630 121 .249000 210.6 ”08‘3“" -101 1.354606 121.810810 128.692000 _ C o 0013 Neutral (7) -1050.3l664 131.888080 139.810000 212.8 ”0:83?“ -1050.662343 139.323310 147.298000 _H Neutral (8) -822.424349 105.043650 1 10.1 19000 215.0 “0(1):?th -822.776703 112.701460 1 17.755000 CH Neutral (9) -861 .746192 122.103150 128.38900 218.5 " 3 Prom?“ -862. 104304 129.788560 136.04300 -CH2CH3 Neutral (10) -901 .063164 140.288790 147.256000 219.0 ”08;?“ -901422005 147.935160 154.93900 67 -CH3SR 100' A 938.3 fl M+ ‘ 1T5 400 600 800 1000 1200 1400 o 100 -CH3SR 8 l B 469.6 CG a “U G 3 _ < [M++H]2+ a) 7 522.5 .2 . ~ 1 2 go 200 400 600 800 1000 1200 1400 -CH3SR 100 ‘ 313.4 ‘ [M++2H]3+ ‘ 368‘s -(CH3SR+H+) _ 469.7 CH3SR+H+ ‘167.0 1 1 1 - 200 400 600 800 1000 1200 1400 m/z Figure 4.2 Linear ion trap CID MS/MS analysis of methionine side chain fixed charge phenacylsulfonium ions of GAILMGAILK. (A) CID MSIMS product ion spectrum of the [M+] ion. (B) CID MSIMS product ion spectrum of the [M“+H]+2 ion. (C) CID MS/MS product ion spectrum of the [M++2H]+3 ion. 68 doubly charged peptide product ion are likely to reside at both side chain and backbone positions. In the case of the triply charged peptide product ion however, increased columbic repulsion, as well as a decrease in the local proton affinities of the remaining sites potentially available for protonation, are likely to result in intermolecular proton transfer from the peptide fragment to the side chain fragment following bond cleavage, but prior to separation of the initial ion-neutral complex. The fragmentation reactions of the multiply charged precursor ions of the methionine side chain fixed charge phenacylsulfonium ion containing synthetic peptides VTMGHFDNFGR and VTMAHFWNFGK were also examined (Figure 4.3 and Figure 4.4, respectively). Similar to that discussed above for dissociation of the GAILMGAILK peptide ions, dissociation of VTMGHFDNFGR and VTMAHFWNFGK resulted in the exclusive neutral loss of the side chain from the low chare state precursor ions ([M"+H]2+ and [M++3H]3+), where the ionizing protons in the resultant peptide product ions could be accommodated at the side chains of the lysine, arginine and histidine amino acids within the sequence or along the peptide backbone. For the quadruply charged ([M++4H]4+) precursor ions however, extensive charged losses are observed. Interestingly, when the CID MS/MS product ion spectra of the triply charged GAILMGAILK, and the quadruply charged VTMGHFDNFGR and VTMAHFWNFGK peptide ions are examined, it can be seen that the abundance of the neutral versus charged loss product ions formed are highly dependant on the charge state and amino acid sequence of the peptide precursor. The relative abundance of neutral versus charged loss product ions observed from the fixed charge containing peptides is expected to vary according to the differences in proton 69 -CH3SR 100- 616.9 A _ [M"+H]2+ 6919.8 200 400 600 800 1000 1200 1400 v 100 a CH SR é B 4113.5 1, _ C.‘ = . ..D < - [M++2H]3+ é" 416.8 ‘5 _ '75 ..- l a .- m AI I I T I I I § 200 400 600 800 1000 1200 1400 -(CH3SR+H+) 1001 C 411.5 . -CH3SR 308.9 _ [M++3H]4+ - CH3SR+H+ / 350-6 167.0 ( l l Jm 1. l -a . m 200 400 600 800 1000 1200 1400 m/z Figure 4.3 Linear ion trap CID MSIMS analysis of methionine side chain fixed charge phenacylsulfonium ions of VTMGHFDNFGR. (A) CID MSIMS product ion spectrum of the [M‘“+H]2+ ion. (B) CID MSIMS product ion spectrum of the [M++2H]3+ ion. (C) CID MSIMS product ion spectrum of the [M++3H]4+ ion. 70 1 -CH3SR 100 A 645.4 4 - [M++H]2+ 728.1 200 400 600 800 1000 1200 1400 a) _ -CH SR 8100 B 430.35 5; . c: 3 - < [M++2H]3+ E J 485.7 ‘5 * 1 be 200 400 600 800 1000 1200 1400 -(CH3SR+H+) I -CH3SR - 323.2 [M++3H]4+ 364.6 . CH3SR+H+ / 167.0 f 200 400 600 800 1000 1200 1400 m/z Figure 4.4 Linear ion trap CID MSIMS analysis of methionine side chain fixed charge phenacylsulfonium ions of VTMAHFWNFGK. (A) CID MSIMS product ion spectrum of the [M“+H]2+ ion. (B) CID MSIMS product ion spectrum of the [MW-2H]3+ ion. (C) CID MSIMS product ion spectrum of the [MW-3H]4+ ion. 71 affinity between the peptide product ions and the neutral side chain fragment. Thus, from the results observed above, it can be determined that the proton affinities of the triply charged VTMGHFDNFGR and VTMAHFWNFGK product ions are both lower than that of the doubly charged GAILMGAILK product ion, and that the proton affinity of the triply charged VTMAHFWNFGK product ion is lower than that of the triply charged VTMGHFDNFGR product ion. Previous studies carried out to examine the effect of proton affinity on the ratios of product ion abundances formed by CID have observed a linear relationship between the log of the ratio of the b2 and yr type product ion abundances and the proton affinities of the C-terminal amino acid for a series of protonated tripeptides Gly-Gly-Xxx (where Xxx represents various amino acid residues) [59]. Here, to determine the effect of the leaving group proton affinities on the ratio of neutral versus charged loss from the methionine side chain fixed charge sulfonium ion containing peptides, the fragmentation of a series of para-substituted phenacylsulfonium ion containing peptides have been examined. Figures 4.5A-D show the CID MSIMS product ion spectra obtained from the triply charged [M+2H“]3+ precursor ions of GAILMGAILK derivatized at the methionine side chains using the para-benzoic acid (Figure 4.5A), para-methylbenzoate (Figure 4.5A), para-methyl (Figure 4.5A) and para-ethyl (Figure 4.5A) substituted phenacylbromide reagents (Compounds 2-5 in Chapter 2, respectively). Figures 4.6 and 4.7 show the CID MSIMS spectra obtained by dissociation of the quadruply charged [M+3H+]4+ precursor ions of VTMGHFDNFGR and VTMAHFWNFGK, respectively, each formed by derivatization at their methionine side chains using the same reagents as above. 72 -CH3SR 100 -CH3SR 00 8 313.4 A 313.5 B E ‘2 CH3SR+H* 5 3 225.1 M+ 211* 9 583.1 ' IM*+2HI3* E .CH35R+H* 38’“ § 469.6 .CH35R+H* Q / 469.8 a l I ‘ v V I L L1 200 400 600 800 l000 1200 I400 200 400 600 800 1000 1200 1400 -CH,SR+H' -(CH,SR+H‘) U ‘00 469.7 C 100 D U G CB '0 S + g 3133“" CH3SR+Ht u ‘ [M"+2Hl3* 195.0 .3 .cthsg 373.5 [M++2H]3* T“: 313.5 £11,511 3789 a: 3135 e I | l 200 400 600 800 1000 l200 I400 200 400 600 800 1000 1200 1400 m/z m/z Figure 4.5 Linear ion trap CID MS/MS analysis of methionine side chain fixed charge para-substituted phenacylsulfonium ions of GAILMGAILK. (A) CID MSIMS product ion spectrum of the para-benzoic acid phenacylsulfonium [M++2H]3+ ion. (B) CID MSIMS product ion spectrum of the para-methylbenzoate phenacylsulfonium [M++2H]3" ion. (C) CID MSIMS product ion spectrum of the para-methyl phenacylsulfonium [M++2H]3+ ion, and (D) CID MSIMS product ion spectrum of the para-ethyl phenacylsulfonium [M++2H]3+ ion. 73 -CH3SR 100- 308.8 g, A 5.; -CH3SR+H+ a 411.5 g 5333“” .3 // I 3 ( [M*+3Hl4+ 0 a: . / 361.6 f M 200 400 600 800 1000 1200 1400 ‘ ‘CH38R+H’ @100 411.5 C U 5 : c113sa+ll+ .‘3 .o - 181.0 < 23 303-8 [M*+3H]“* 2'5 353.8 a: 4 / e 1 (1! 200 400 600 800 10001200 1400 m/z 1003 100‘ CH 3SR+H* 1 225\.0 / -CH3SR+H* 41 1.5 -CH3SR / 308.9 [M*+3H]“ .10: 200400600800 CH3SR+H+ ‘ 195.0 ~CH3SR ‘ 301.8 *——r_ ‘CH3SR+H+ 41 1.5 [M’+3H]‘“ / 357.8 200400600800 mlz J 1000 1200 1400 D 1000 1200 1400 Figure 4.6 Linear ion trap CID MSIMS analysis of methionine side chain fixed charge paraesubstituted phenacylsulfonium ions of VTMGHFDNFGR. (A) CID MSIMS product ion spectrum of the para-benzoic acid phenacylsulfonium [M++3H]4+ ion. (B) CID MSIMS product ion spectrum of the para-methylbenzoate phenacylsulfonium [W+3H] 4+ ion. (C) CID MSIMS product ion spectrum of the para-methyl phenacylsulfonium [M“+3H]4+ ion, and (D) CID MSIMS product ion spectrum of the para-ethyl phenacylsulfonium [M++3H]4+ ion. 74 -CH3SR+H* 100- 8 430.6 A C 63 4 '0 g c113512+11+ D 211.0 ‘5 {113512 ‘>_’ \ 323.1 g x [M++3H14+ a: / 375.6 e K 200 400 600 800100012001400 -CH3SR+H* 100‘ 430.5 C 8 G g g c113sa+w .o .1810 <2 9; £11351: 33 323.2 [M*+3H]‘“ - 368.4 é ~ ’ a l 200 400 600 800 1000 1200 I400 m/z 100‘ ‘ 225.0 CH3SR+H‘ -CH3SR 323.1 11/ -CH3SR+H* 430.5 [M‘+3H]‘* / 379.10 ‘ CH3SR+H‘ ‘ -c113sn 323.2 11hr 200400600 195.0 800 -CH,SR+H’ 430.6 [M’+3H]4+ / 37l.8 1200 1400 D 2004006008001000 m/z [200 1400 Figure 4.7 Linear ion trap CID MSIMS analysis of methionine side chain fixed charge para-substituted phenacylsulfonium ions of VTMAHFWNFGK. (A) CID MSIMS product ion spectrum of the para-benzoic acid phenacylsulfonium [M"+3H]4+ ion. (B) CID MSIMS product ion spectrum of the para-methylbenzoate phenacylsulfonium [M++3H]4+ ion. (C) CID MSIMS product ion spectrum of the para-methyl phenacylsulfonium [MW-3H]4+ ion, and (D) CID MSIMS product ion spectrum of the para-ethyl phenacylsulfonium [M++3H]4+ ion. 75 It can be seen from Figures 4.5, 4.6 and 4.7 that the ratio of neutral versus charged losses from the methionine side chain of these precursor ions varies as a function of the identity of the neutral side chain fragment, and that the extent of charged loss product ions increases with increasing proton affinity of the neutral side chain fragment. For example, when the triply charged para-benzoic acid substituted phenacylsulfonium ion derivative of GAILMGAIK was subjected to CID-MSIMS (Figure 4.5A), the neutral loss of the side chain was observed as the dominant fragmentation pathway. This result indicates that the proton affinity of the neutral side chain fragment (calculated as 210.6 kcal mol", see structure 6 in Table 4.1) is lower than that of the doubly protonated peptide product ion. In contrast, CID-MSIMS of the triply charged para-ethyl phenacylsulfonium ion derivative of the same peptide gave rise to almost exclusive charged loss of the side chain. Thus, in this case, the proton affinity of the neutral side chain fragment (calculated as 219.0 kcal mol", see structure 10in Table 4.1) must have a higher proton affinity compared to the doubly protonated peptide product ion. These data therefore provide a means to ‘bracket’ the proton affinity of the doubly charged peptide product ion to between these two values. Similar proton affinity bracketing of the triply charged product ions of VTMGHFDNFGR and VTMAHFWNFGK was obtained by dissociation of the quadruply charged substituted phenacylsulfonium ion derivatives. The plots shown in Figure 4.8 show a linear correlation between the log of the experimentally observed ratios of the charged versus neutral loss product ion abundances and the theoretically calculated proton affinities for the para-subtituted methylphenacylsulfide side chain fragments of GAILMGAILK (Figure 4.8A), VTMGHFDNFGR (Figure 4.83) and VTMAHFWNFGK (Figure 4.8C). 76 41 A A ‘\ + \ =5 x a C "\ \‘c + 0 \208 2r04._ 212 214 \246 218 220 i“ °\., 7“ Proton Affinity (kcal mol") 5 \., ~\ 5) 2 . \..\ I ‘.‘o f ' "\.. R2: 0.9988 8 \ '12., 'R2=0.9851 :1: E.) -6‘ 5 ‘R2=0.9605 -8‘ Figure 4.8. Plot of ln[-CH3SR / (-(CH3SR+H+) + CH3SR+H+)] versus calculated proton affinity for the phenacylsulfide derivatives of (A) the [M+2H]2+ side chain cleavage product ion of GAILMGAILK (B) the [M+3H]3+ side chain cleavage product ion of VTMGHFDNFGR and (C) the [M+3H]3+ side chain cleavage product ion of VTMAHFWNFGK. 77 In order to account for any charge state dependant detector response, and therefore to obtain a closer approximation of the actual abundances of the neutral and charged loss side chain cleavage product ions formed, the product ion abundances were first normalized with respect to charge state (i.e., intensity divided by charge), assuming mass detection response to be linear with respect to the charge state of the ion. The x-axis intercept of these plots may be used to estimate the proton affinities of the doubly charged product ion from GAILMGAILK (216 kcal mol"), and the triply charged product ions of VTMGHFDNFGR (211 kcal mol'l) and VTMAHFWNFGK (207.5 kcal mol'l). The proton affinity of the amide bonds along the peptide backbone may be estimated from a simple model of the peptide amide bond, N-methyl acetarnide, which has a proton affinity of 212.4 kcal mol'1 [60]. Thus, each of the values estimated from the data above are consistent with the ionizing proton being located at a position along the peptide backbone. 4.4 Conclusions and Future Directions These data described above indicates that the ratio of neutral loss to charged loss product ion abundances from fixed charge containing peptide ions varies according to the differences between the proton affinities of the neutral side chain fragment and the peptide product ion. Given that the ratio of charged versus neutral loss observed for a given peptide is observed to change as a function of peptide ion charge state, specific information regarding the number of basic sites contained within the peptide may potentially be obtained by having control over this reaction. In the future, the potential for utilizing this information, along with knowledge regarding the presence of a specific 78 amino acid residue (e.g., methionine), proteolytic enzyme cleavage specificity, chromatographic retention time, or accurate peptide mass, will be examined for directly identifying peptides from the sequence databases, without need for further structural information. Several precedents supporting the concept of improving the specificity of database searches for protein identification, by promoting the formation of specific product ions through chemical derivatization e.g., enhanced C-terminal aspartic acid cleavage from singly-charged peptides by conversion of lysine to homoarginine, or enhanced N-terminal amino acid cleavage by phenylthiocarbamyl modification have recently appeared in the literature [61]. 79 REFERENCES 1. Stein, L.D. End of the beginning. Nature, 2004, 431, 915-916. 2. Hanash, S. Disease proteomics. Nature, 2003, 422, 226-232. 3. Barber, M.; Bordoli, R.S.; Sedgwick, D.; Tyler, A.N. Fast Atom Bombardment of Solids (F.A.B.): A New Ion Source for Mass Spectrometry. J. Chem. Soc., Chem. Commun. 1981, 7, 325-327. 4. Rinehart, K.L. Fast Atom Bombardment Mass Spectrometry. Science, 1982, 218, 254- 260. 5. Fisher, D.L.L.; Gross, M.L. Fast Atom Bombardment and Tandem Mass Spectrometry for Sequencing Peptides and Polyamino Alcohols. Anal. Chem. 1985, 57, 1174-1180. 6. Fenn, J.B.; Mann, M.; Meng, C.K.; Wong, S.F.; Whitehouse, C.M.; Electrospray Ionization for Mass Spectrometer of Large Biomolecules. Science, 1989, 246, 64-71. 7. Karas, M.; Hillenkamp, F. Laser Desorption Ionization of Proteins with Molecular Masses Exceeding 10 000 Daltons. Anal. Chem. 1988, 60, 2299-2301. 8. Gomez, A.; Tang, K Charge and fission of droplets in electro static sprays. Phys. Fluids 1994, 6, 404-414. 9. Thomson, B.T.; Iribarne, J .V. Field induced ion evaporation from liquid surfaces at atmospheric pressure. J. Chem. Phys. 1979, 71, 4451-4463. ‘ 10. Iribarne, J.V.; Thomson, B.T. On the evaporation of small ions from charged droplets. J. Chem. Phys. 1976, 64, 2287-2294. 11. Zenobi, R.; Knochenmuss, R. Ion Formation in MAIDI Mass Spectrometry. Mass Spectrom. Rev. 1998, 17, 337-366. 12. Hoffmann, E.D.; Stroobant, V. Mass Spectrometry Principles and Applications, 2nd edition, John Wiley and Sons, New York, Page 69. 80 13. March, R.E. Quadrupole Ion Trap Mass Spectrometry: Theory, Simulation, Recent Developments and Applications. Rapid Commun. Mass Spectrom. 1998, 12, 1543-1554. 14. Hager, J.W. A new linear ion trap mass spectrometer. Rapid Commun. Mass. Spectrom. 2002, 16, 512-526. 15. Schwartz, J.C.; Senko, M.W.; Syka,.J.E. A two-Dimensional Quadrupole Ion Trap Mass Spectrometer. J. Am. Soc. Mass Spectrom. 2002, 13, 659-669. 16. Makarov, A. Electrostatic axially harmonic orbital trapping: a high-perforrnance technique of mass analysis. Anal. Chem. 2000, 72, 1156-1162. 17. Hardman, M.; Markarov, A. Interfacing the orbitrap mass analyzer to an electrospray ion source. Anal. Chem. 2003, 75, 1699-1705. 18. Hu, Q.; Noll, R.J.; Li, H.; Makarov, A.; Hardman, M.; Cooks, R.G. The Orbitrap: a new mass spectrometer. J. Mass Spectrom. 2005, 40, 430-443. 19. Olsen, J.V.; de Godoy, L.M.F.; Li, G.; Macek, B.; Mortensen, P.; Pesch, R.; Makarov, A.; Lange, 0.; Homing, S.; Mann, M. Parts per Million Mass Accuracy on an Orbitrap Mass Spectrometer via Lock Mass Injection into a c-trap. Mol. Cell. Proteomics. 2005, 4, 2010-2021. 20. Eng, J.K.; McCormack, A.L.; Yates, J.R An Approach to Correlate Tandem Mass Spectral Data of Peptides with Amino acid Sequences in a Protein Database. J. Am. Soc. Mass Spectrom. 1994, 5, 976-989. 21. Perkins, D.N.; Pappin, D.J.C.; Creasy, D.M.; Cottrell, J.S Probability-base protein identification by searching sequence databases using mass spectrometry data. Electrophoresis 1999, 20, 3551-3567. 22. Yates, J .R.; Eng, J.K.; McCormack, A.L.; Schieltz, D. Method to Correlate Tandem Mass Spectra of Modified Peptides to Amino Acid Sequences in the Protein Database. Anal. Chem. 1995, 67, 1426-1436. 23. MacCoss, M.J.; Wu, C.C.; Yates, J.R. Probability-Based Validation of Protein Identifications Using a Modified SEQUEST Algorithm. Anal. Chem. 2002, 74, 5593- 5599. 81 24. Hunt, D.F.; Yates, J .R.; Shabanowitz, J .; Winston, S.; Hauer, C.R. Protein sequencing by tandem mass spectrometry. Proc. Natl. Acad. Sci.. 1986, 83, 6233-6237. 25. de Hoog, C.L.; Mann, M. Proteomics. Genomics Hum. Genet. 2004, 5, 267-293. 26. Creasy, D. M.; Cottrell, J.S. Error Tolerant searching of uninterpreted tandem mass spectrometry data. Anal. Chem. Proteomics 2002, 2, 1426-1434. 27. Reid, G. E.; Roberts, K. D.; Kapp, E. A.; Simpson, R. J Statistical and Mechanistic Approaches to Understanding the Gas-Phase Fragmentation Behavior of Methionine Sulfoxide Containing Peptides. J. Proteome Res. 2004, 3, 751-759. 28. Annan, R.; Carr, S. A. Phosphopeptide Analysis by Matrix-Assisted Laser Desorption Time-of-Flight Mass Spectrometry. Anal. Chem. 1996, 68, 3413-3421. 29. Mann, M.; Jensen, O.N. Proteomics analysis of post-translational modifications. Nature Biotechnology, 2003, 21, 255-261. 30. Dongre, R.A.; Jones, J.L.; Somogyi, A.; Wysocki, V.H. Influence of Peptide Composition, Gas-Phase Basicity, and Chemical Modification on Fragmentation Efficiency: Evidence for the Mobile Proton Model. J. Am. Chem. Soc. 1996, 118, 8365- 8374. 31. Cooks, R.G.; Kruger, T.L Intrinsic basicity Determination Using Metastable Ions. J. Am. Chem. Soc. 1977, 99 1279-1281. 32. Armentrout, P.B. Entropy Measurements and the Kinetic Method: A Statistically Meaningful Approach. J. Am. Soc. Mass Spectrom. 2000, 11, 371-379. 33. Kuntz, A.E.; Boynton, A.W.; David, G.A.; Colyer, K.E.; Poutsma, J. C. The Proton Affinity of Proline Analogs Using the Kinetic Method with Full entropy Analysis. J. Am. Soc. Mass Spectrom. 2002, 13, 72-81. 34. Schroeder, O.E.; Andriole, E.J.; Carver, K.L.; Colyer, K.E.; Poutsma, J.C. Proton Affinity of Lysine Homologues from the Extended Kinetic Method. J. Phys. Chem. A, 2004, 108, 326-332. 82 35. Kapp, E.A.; Schultz, F.; Reid, G.E.; Eddes, J .S.; Moritz, R.L.; O’Hair, R.A.J.; Speed, T.P.; Simpson, R.J. Mining a Tandem Mass Spectrometry Database To Determine the Trends and Global Factors Influencing Peptide Fragmentation. J Anal. Chem. 2003, 75, 6251-6264. 36. Schlosser, A.; Lehmann, W.D. Five-membered ring formation in unimolecular reactions of peptides; a key structural element controlling low-energy collision-induced dissociation of peptides. J. Mass Spectrom. 2000, 35, 1382-1390. 37. Paizs, B.; Suhai, S. Fragmentation Pathways of Protonated peptides. Mass Spectrom. Rev. 2005, 24, 508-548. 38. McCormack, A.L.; Somogyi, A. Dongre, A.R.; Wysocki, V.H. Fragmentation of Protonated Peptides: Surface-Induced Dissociation in Conjunction with a Quantum Mechanical Approach. Anal. Chem. 1993, 65, 2859-2872. 39. Vekey, V.; Gomory, A. Theoretical Modeling of Mass Spectrometric Behavior of Peptides: Singly and Doubly Protonated Tetraglycine. Rapid Commun. Mass Spectrom. 1996, 10, 1485-1496. 40. Somogyi, A.; Wysocki, V.H.; Mayer, I. The Effect of Protonation Site on Bond Strengths in Simple Peptides: Application of Ab Initio and Modified Neglect of Differential Overlap Bond Order and Modified Neglect of Differential Energy Partitioning J. Am. Soc. Mass Spectrom. 1994, 5, 704-717. 41. Wu, J.; Lebrilla, C.B. Gas-Phase Basicities and Sites of Protonation of Glycine Oligomers. J. Am. Chem. Soc. 1993, 115, 3270-3275. 42. Zhang, K.; Cassady, C.J.; Phillips, C.A. Ab Initio Studies of Neutral and Protonated Triglycines: Comparison of Calculated and Experimental Gas-Phase Basicity. J. Am. Chem. Soc. 1994, 116, 11512-11521. 43. Lee, S.W.; Kim, H.S.; Beauchamp, J .L. Salt Bridge Chemistry Applied to Gas-Phase Peptide Ions Adjacent to Aspartic Acid Residues. J. Am. Chem. Soc. 1998, 120, 3188- 3195. 44. Roth, K.D.W., Huang, Z, H., Sadagopan N.; Watson J .T. Charge derivatization of peptides for analysis by mass spectrometry. Mass Spectrom. Rev. 1998, 17, 255-274. 83 45. Keough, T.; Youngquist, R.S.; Lacey. M.P. Sulfonic acid derivatives for peptide sequencing by MALDI MS Anal. Chem. 2003, 75, 156A-l65A. 46. Huang, Z.; Wu, J.; Roth, K.D.W.; Gage, D.A.; Watson, J.T. A Picomole-Scale Method for Charge Derivatization of Peptides for Sequence Analysis by Mass Spectrometry. Anal. Chem. 1997, 69, 137-144. 47. Adamczyk, M.; Gebler, J.C.; Wu, J. Charge Derivatization of Peptides to Simplify their Sequencing with an Ion Trap Mass Spectrometer. Rapid Commun. Mass spectrum. 1999, 13, 1413-1422. 48. Sadagopan, N.; Watson, J.T. Mass spectrometric evidence for mechanisms of fragmentation of charge-derivatized peptides. J. Am. Soc. Mass Spectrom. 2001, 12, 399- 409. 49. Gu, C.; Tsaprailis, G.; Breci, L.; Wysocki, V.H. Selective Gas-Phase Cleavage at the Peptide Bond C-Terrninal to Aspartic Acid in Fixed-Charge Derivatives of Asp- Containing Peptides. Anal. Chem. 2000, 72, 5804-5813. 50. Reid, G.E.; Roberts, K.D.; Simpson, R.J.; O’Hair, R.A.J. Selective Identification and Quantitative Analysis of Methionine Containing Peptides by Charge Derivatization and Tandem Mass Spectrometry. J. Am. Soc. Mass Spectrom. 2005, 16, 1131-1150. 51. Reid, G.E.; Simpson, R.J; O’Hair, R.A.J. Method for Analyzing Amino Acids, Peptides and Proteins. PCT Int. Appl. 2004, WO 2004046731. 52. Paizs, B.; Suhai, S. Towards Understanding the Tandem Mass Spectra of Protonated Oligopeptides. l: Mechanism of Amide Bond Cleavage. J. Am. Soc. Mass Spectrom. 2004, 15, 103-113. 53. Mcluckey, S.A.; Goeringer, D.E. Slow Heating Methods in Tandem Mass Spectrometry. J. Mass Spectrom. 1997, 32, 461-474. 54. Johnson, R.S.; Krylov, D.; Walsh, K.A. Proton mobility within Electrosprayed Peptide ions. J. Mass Spectrom. 1995, 30, 386-387. 55. Averbukh, M.E.; Pipkom, R.; Lehmann, W.D. Phosphate Group-Driven Fragmentation of Multiply Charged Phosphopeptide Anions. Improved Recognition of 84 Peptides Phosphorylated at Serine, Threonine, or Tryosine by Negative Ion Electrospray Tandem Mass Spectrometry. Anal. Chem. 2006, 78, 1249-1256. 56. Hammett, LP. The Effect of Structure upon the Reactions of Organic Compounds, benzene Derivatives. J. Am. Chem. Soc. 1937, 59, 96-103. 57. Jaffe, H. A re-examination of the Hammett Equation. Chem. Rev. 1953, 53, 191-261. 58. Hansch, C., Leo, A.; Taft, R.W. A Survey of Hammett Substituent Constants and Resonance and Field Parameters. Chem. Rev. 1991, 97, 165-195. 59. Morgan, D.G.; Bursey, M.M. A Linear Free-energy Correlation in the Low-energy Tandem Mass Spectra of Protonated Tripeptides Gly-Gly-Xxx. Org. Mass Spectrom. 1994, 29, 354-359. 60. Hunter, E.P.L.; Lias, S.G. Evaluated Gas Phase Basicities and Proton Affinities of Molecules: An Update, J. Phys. Chem. Ref. Data, 1998, 27, 413-656. 61. Sidhu, K.S.; Sangvanich, P.; Brancia, F.L.; Sullivan, A.G.; Gaskell, S.J.; Wolkenhauer, 0.; Oliver, S.G.; Hubbard, S.J. Bioinforrnatic assessment of mass spectrometric chemical derivatisation techniques for proteome database searching. Proteomics, 2001, 1, 1368-1377. 85 11111111111111111111111111111