. u: . 1...?th . .w )4. .0. .5 .1 1...; . $.51“? “I . . do 4.. . 3:51 ,Lwfim . . .knkmwwnun. b! u. 5.. i [5.2. r? , ll 0.521. .033)..- at)... .al: .. 3 Kilns“: .6 L 1.3. \o 51“}..‘1 T Uzi?) .wwfimwfimmmsh I:II..!=..I9 00.. aolo .LIBRARY Michigan State University This is to certify that the dissertation entitled DEVELOPMENT AND APPLICATION OF TANDEM MASS SPECTROMETRY METHODS FOR PHOSPHOPROTEIN ANALYSIS presented by AMANDA M. PALUMBO has been accepted towards fulfillment of the requirements for the Ph.D. degree in Chemistry Major Professor’s Signature 4.4 0C7 ,/¢ ,200‘). Date MSU is an Affirmative Action/Equal Opportunity Emplo yer .-.—u_-_4-.—--o-.—.- -.- _._.- .— 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 5/08 K:IProleoc&Pres/CIRCIDateDue.Indd DEVELOPMENT AND APPLICATION OF TANDEM MASS SPECTROMETRY METHODS FOR PHOSPHOPROTEIN ANALYSIS By Amanda M. Palumbo A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Chemistry 2009 ABSTRACT DEVELOPMENT AND APPLICATION OF TANDEM MASS SPECTROMETRY METHODS FOR PHOSPHOPROTEIN ANALYSIS By Amanda M. Palumbo Protein phosphorylation is involved in nearly all essential biochemical pathways and the deregulation of phosphorylation events has been related to the onset and progression of numerous diseases. Mass spectrometry has been at the forefront of phosphoproteomic characterization due to the plethora of different tandem mass spectrometry (MS/MS) strategies that enable comprehensive phosphoproteome analysis. However, each of these tandem mass spectrometry strategies has a limited set of conditions under which phosphoproteome information may be produced due to the difference in compatible precursor ion charge states and the type (i.e., sequence vs. nonsequence), structure, and abundance of the structurally-informative product ions formed. Accordingly, understanding the underlying chemical principles involved in these techniques that influence the gas-phase fragmentation reactions are critical, as the chemistry directly influences their utility. Therefore, a comprehensive review regarding the gas-phase fragmentation reactions of phosphopeptides and their influence on the various tandem mass spectrometry strategies employed for phosphoproteome analysis is discussed here. Based on this discussion it is deduced that much of the gas-phase ion chemistry involved in the most commonly used tandem mass spectrometry method for phosphoproteome analysis, collision-induced dissociation (CID), has yet to be understood. To address this need, the studies described here demonstrate that the combined role of precursor ion charge state and peptide basic residue content (i.e., proton mobility) heavily influences the gas-phase ion chemistry of phosphorylated peptide ions using CID. The mechanisms and related product ion structures associated with the formation of phosphate-specific reactions that occur under CID conditions are also elucidated and invalidate previously established mechanisms. Additionally, it is shown that intramolecular gas-phase phosphate migration and other competing fragmentation reactions involving the phosphate occurs upon ClD-MS/MS, which may cause ambiguous or erroneous phosphorylation site interpretation. As a result of all of these studies, the applicability of CID for phosphopeptide identification and characterization is discussed. Finally, the development and application of a phosphoproteome analytical strategy aimed to determine the phosphorylated proteins that associate with p65, a protein critically involved in the immune response, from differentially treated monocytic leukemia cells are described. Importantly, the affects of gas-phase ion chemistry of phosphorylated peptides was considered during the downstream data interpretation. In this study, five p65-associated proteins are identified, including two that have not been previously reported to interact with p65. ACKNOWLEDGEMENTS First and foremost, I would like to thank my graduate co-advisors, Professors Gavin Reid and Jetze Tepe for allowing me to complete this collaborative research project which required the expertise from each person's lab. I am so proud of the research that we were able to complete and the results that we achieved together. Gavin and Jetze, I want to thank you both for your guidance and never-ending support, both in and out of the lab. I also must thank all of the current and former members of the Reid and Tepe research groups who have supported me, laughed with me, and encouraged me. Many thanks to all of the collaborators who have contributed to this work. Dr. Henry Duewel at Sigma Aldrich and Stacey Hoge at Sigma Genosys for synthesizing many of the phosphorylated peptides for this work. Additionally, Professor Don Hunt and Dr. Erin Jeffery at the University of Virginia and Professor Paul Stemmer at Wayne State University for enabling the analysis of many phosphopeptides by electron transfer dissociation. I would like to also thank Mr. Weihan Wang and Professor Menin Bruening from Michigan State University for developing their phosphopeptide enrichment strategy. In particular, I truly thank Weihan for helping me with data acquisition. The results from this work could not have been achieved had it not been for your dedication. Lastly, the completion of my degree would not have been possible without the hospitality and kindness shown to me by my many interim roommates: Nicole Torres, Troy Knight, Claire Laity, Georgia, Angus, and Sylvester. TABLE OF CONTENTS LIST OF TABLES ................................................................................................ IX LIST OF FIGURES ............................................................................................... X LIST OF SCHEMES .......................................................................................... XVI CHAPTER ONE .................................................................................................... 1 Mass Spectrometry Methods for Phosphoproteome Analysis ............................... 1 1.1 Introduction ................................................................................................. 1 1.2 Mass Spectrometry Methods Employed for Phosphoproteome Analysis ...5 1.2.1 Ionization of Phosphorylated Peptides ............................................... 5 1.2.2 MS Analysis of Phosphorylated Peptide Ions ..................................... 6 1.2.3 Tandem Mass Spectrometry and Gas-phase lon Chemistry of Phosphorylated Peptides .................................................................... 7 1.2.3.1 Collision-Induced Dissociation (CID) ........................................ 13 1.2.3.1.1 Formation of 79 m/z (P03) and [M+nH-79]‘"‘1* Ions ....... 15 1.2.3.1.2 Formation of Phosphotyrosine-Specific Immonium Ions at m/z 216.043 .................................................................... 16 1.2.3.1.3 Product Ions Formed From 80 or 98 Da Neutral Losses. 18 1.2.3.1.4 Chemical Derivatization and ClD-MS/MS ........................ 23 1.2.3.2 MALDI Post-Source Decay (PSD) ............................................ 25 1.2.3.3 Photodissociation ..................................................................... 28 1.2.3.3.1 Infrared Multiple Photon Dissociation (IRMPD) ............... 28 1.2.3.3.2 Ultraviolet Photodissociation (UVPD) .............................. 29 1.2.3.3.3 Femtosecond Laser-Induced Ionization/Dissociation (stlD) ............................................................................. 35 1.2.3.4 Electron-Driven Dissociation .................................................... 36 1.2.3.4.1 Electron Capture Dissociation (ECD) .............................. 37 1.2.3.4.2 Electron Transfer Dissociation (ETD) .............................. 41 1.2.3.4.3 Electron Detachment Dissociation (EDD) ....................... 45 1.2.3.5 Metastable Atom-Activated Dissociation (MAD) ....................... 45 1.2.3.6 Comparison of Tandem Mass Spectrometry Techniques for Phosphoproteome Analysis ...................................................... 47 1.3 Specific Aims ............................................................................................ 50 CHAPTER TWO .................................................................................................. 53 Mechanistic Insights into the Multistage Gas-Phase Fragmentation Behavior of Phosphoserine- and Phosphothreonine-Containing Peptides ............................. 53 2.1 Introduction ............................................................................................... 53 2.2 Results and Discussion ............................................................................ 59 2.2.1 Evaluation of Charge-Remote and Charge-Directed Fragmentation Mechanisms for the Loss of H3PO4 from Phosphoserine- and V Phosphothreonine-Containing Peptides via Multistage Tandem Mass Spectrometry Analysis of Their Regioselectively and Uniformly Deuterated Precursor Ions ............................................................... 59 2.2.2 Neutral Loss of H3PO4 from Both Phosphoserine- and Phosphothreonine-Containing Peptides ls Highly Dependent on Precursor Ion Charge State and Amino Acid Composition (i.e., Proton Mobility) ............................................................................................ 76 2.2.3 Experimental Evidence for the Neutral Loss of H3PO4 from Phosphoserine- and Phosphothreonine-Containing Peptides via an 8N2 Neighboring Group Participation Reaction Mechanism. ............ 90 2.3 Conclusions .............................................................................................. 95 CHAPTER THREE .............................................................................................. 96 Evaluation of Gas-Phase Rearrangement and Competing Fragmentation Reactions on Protein Phosphorylation Site Assignment Using Collision-induced- dissociation-MS/MS and MS3 .............................................................................. 96 3.1 Introduction ............................................................................................... 96 3.2 Results and Discussion .......................................................................... 104 3.2.1 ClD-MS/MS of Protonated Phosphopeptides Results in Gas-Phase Phosphate Group Transfer Reactions. ........................................... 104 3.2.2 CID-MS3 of [M+nH-98]n+ Ions May Not Be Used for Unambiguous Phosphorylation Site Localization. .................................................. 145 3.3 Conclusions ............................................................................................ 156 3.4 Other Recent Studies Demonstrating Intramolecular Gas-phase Phosphate Group Transfer ..................................................................... 157 3.5 Additional Commentary .......................................................................... 159 CHAPTER FOUR .............................................................................................. 161 Experimental Methods for Gas-Phase Fragmentation Studies of Phosphopeptides (Chapters Two and Three) ................................................... 161 4.1 Materials ................................................................................................. 161 4.2 Phosphopeptides .................................................................................... 163 4.2.1 Fmoc Solid—Phase Peptide Synthesis (SPPS) ................................ 164 4.2.1.1 Apparatuses ........................................................................... 164 4.2.1.2 Synthesis ................................................................................ 164 4.2.1.2.1 Fmoc Deprotection ........................................................ 165 4.2.1.2.2 Amino Acid Coupling ..................................................... 165 4.2.1.2.3 Resin Cleavage and Deprotection of Acid-Labile Orthogonal Protecting Groups ....................................... 166 4.2.2 Synthesis of GAlL(d3-D,L)pSGAILK ............................................... 167 4.2.2.1 Synthesis of (d3-D,L)-Fmoc-Serine-OH (4.a) ......................... 168 4.2.2.2 Synthesis of (d3-D,L)-Fmoc-Serine(O-TBDMS)-OH (4.b) ...... 168 4.2.2.3 On-resin SPPS of Boc-GAIL(d3-D,L)pSGAlLK(Boc) (4.c) ...... 169 4.2.3 Peptide Purification ......................................................................... 170 4.3 Multistage Tandem Mass Spectrometry Analysis of Phosphopeptides ..171 4.3.1 Phosphopeptide Sample Preparation ............................................. 171 4.3.2 Mass Spectrometry ......................................................................... 172 vi 4.3 Quantitative Analysis of Percent Product Ion Abundances ..................... 174 CHAPTER FIVE ................................................................................................ 175 Application of Phosphoproteomic Analysis Strategy to Elucidate p65-Associated Phosphoproteins in the NF-KB Signaling Pathway From Treated and Untreated THP-1 Cell Nuclear Extracts ............................................................................. 175 5.1 Introduction ............................................................................................. 175 5.2 Results and Discussion .......................................................................... 185 5.2.1 Development of the Phosphoproteome Analysis Strategy Employed for Identification of p65-Associated Phosphoproteins From Differentially Treated THP-1 Cells .................................................. 185 5.2.1.1 Proteolytic Digestion ............................................................... 185 5.2.1.2 Phosphopeptide Enrichment .................................................. 186 5.2.1.3 Mass Spectrometry Analysis and Database Searching .......... 186 5.2.2 Differentially Treated THP-1 Cells and Resultant Differential Levels in Nuclear p65 .................................................................................... 190 5.2.3 Identification of p65-Associated Phosphoproteins From TNF—o— Treated THP-1 Cells ....................................................................... 192 5.2.4 Identification of p65-Associated Phosphoproteins From Untreated THP-1 Cells .................................................................................... 215 5.2.5 Relevance of Identified Proteins Co-lmmunoprecipitated With p65220 5.2.5.1 Ribosomal Proteins: 6OS Acidic Ribosomal Protein from the P1 or P2 Isoforms and Ribosomal Protein L19 ............................ 220 5.2.5.2 TLS and TLS-Associated Serine-Arginine Protein (TASR) ..... 222 5.2.5.3 Actin ....................................................................................... 223 5.3 Conclusions and Future Directions ......................................................... 224 5.4 Supplemental Figures ............................................................................ 226 CHAPTER SIX .................................................................................................. 234 Experimental Methods for Applications of Mass Spectrometry Methods for Phosphoproteomic Analysis (Chapter Five) ...................................................... 234 6.1 Materials ................................................................................................. 234 6.2 Cells ....................................................................................................... 235 6.3 Nuclear Extraction of THP-1 Cells .......................................................... 235 6.3.1 Bradford Protein Assay: Determination of Protein Concentration in Cellular Extracts ............................................................................. 237 6.3.2 Western Blot Analysis of p65 from Nuclear Extracts ...................... 237 6.4 Sodium Dodecyl Sulfate-Polyacrylamide Gel Electrophoresis (SDS-PAGE) of lmmunoprecipitated Proteins from Nuclear Extracts .......................... 238 6.4.1 Preparation of Immunoprecipitation (IP) Matrix: Covalently Cross- Linked p65 Primary Antibody to Protein A/G-Agarose .................... 238 6.4.2 Immunoprecipitation of Cellular Extracts Using Covalently Cross- Linked p65 Primary Antibody to Protein A/G-Agarose .................... 240 6.4.3 Western Blot Analysis of p65 From p65-lmmunoprecipitated Nuclear Extracts of THP-1 Cells .................................................................. 241 6.5 ln-Solution Tryptic Digestion and Subsequent Phosphopeptide Enrichment of lmmunoprecipitated Proteins .............................................................. 242 vii 6.6 Synthesis of the lntemal Standard "D5” CDgCDzCO-LFTGHPEpSLEK..243 6.7 Matrix Assisted Laser Desorption/Ionization and Multistage Tandem Mass Spectrometry .......................................................................................... 244 6.8 Peptide Identification by Database Searching ........................................ 245 REFERENCES .................................................................................................. 246 viii LIST OF TABLES Table 2.1 Summary of the MS/MS and MS3 neutral loss product ion abundances from the gas-phase fragmentation reactions of protonated phosphgsebrine- and phosphothreonine-containing peptides. a “Partially mobile” proton.9 b“Mobile” proton. 93 ° “Nonmobile” proton. ‘93 d Results obtained by ClD-MS/MS of uniformly deuterium labeled peptide ions. ° Note that these values differ from those shown in Figure 2.1 for the regioselectively deuterated GAIL(d3- D,L)pSGAILK peptide and are likely due to either their acquisition on a different instrument platform (LCQ-DECA versus LTQ), the effect of the D, L- versus L- phosphoamino acid, or deuterium isotope effects on the charge-remote versus charge.directed fragmentation mechanisms. fResults obtained by CID-MS/MS of protonated peptide ions. 9 Calculated as the ratio of % total product ion , abundances of the [M+nH—H3PO41" ions from isomer phosphopeptide Ions (e. g., GAILpSGAILK versus GAILpTGAILK). hResults obtained by CID-MS3 of [M+nH- H3PO4-CH20]"+ (pSer) or [M+nH- -H3PO4-CH3CHO]"+ (pThr) product ions initially formed by MS/MS. .............................................................................................. 70 Table 3.1 Summary of the gas-phase phosphate group transfer product ions observed following ESI- and/or MALDI-ClD-MS/MS of independently synthesized tryptic phosphopeptides. a Proton mobility classifications defined by Kapp, et Non-mobile: number of ionizing protons 5 number of arginine residues; Partially-mobile: number of arginine residues < number of ionizing protons 5 combined number of arginine, lysine and histidine residues; Mobile: number of ionizing protons > combined number of arginine, lysine and histidine residues. b .......................................................................................................................... 133 Table 4.1 Common amino acid building blocks used for solid-phase peptide synthesis. .......................................................................................................... 163 LIST OF FIGURES Figure 1.1 Structures of phosphorylated serine, threonine and tyrosine residues.3 Figure 1.2 Flow chart of previously demonstrated tandem mass spectrometry strategies for phosphopeptide analysis. The phosphopeptide precursor ion charge state (n) requirement for each technique is specified. Some techniques, namely ClD-, PSD-, and IRMPD-MS/MS, may involve formation of “non- sequence” product ions that diagnostically indicate the presence of phosphorylated peptide ions. The observation of these diagnostic product ions has been used to trigger the acquisition of CID-MS3 spectra of the neutral loss product ion, or to trigger the acquisition of MS/MS spectra of the original precursor ion by other methods (e.g., ECD-MS/MS). *Phosphopeptide analysis by UVPD of deprotonated precursor ions has been performed on phosphate- derivatized peptides. ............................................................................................. 9 Figure 1.3 Nomenclature of peptide backbone fragment ions or sequence ions. ............................................................................................................................ 12 Figure 1.4 General structure of an immonium ion, where R is the amino acid side-chain and the phosphotyrosine-specific immonium ion, m/z 216.043 .......... 17 Figure 2.1 Multistage tandem mass spectrometry of regioselectively deuterated GAIL(d3-D,L)pSGAILK. CID-MS/MS of the (A) [M+H]+ and (B) [M+2H]2+ precursor ions. CID-MS3 of the (C) singly and (D) doubly protonated [M+nH- H3PO4]"” neutral loss product ions from panels A and B, respectively. CID-M83 of the (E) singly and (F) doubly protonated [M+nH-HZDPO41'” neutral loss product ions from panels A and B, respectively. A = -H3PO4, ° = -H20, * = -NH;;. ........... 61 Figure 2.2 Multistage tandem mass spectrometry of GAILpSGAILK. CID-MS/MS product ion spectra of the (A) [M+H]+ and (B) [M+2H]2+ precursor ions. CID-MS3 product ion spectra of the (C) singly and (D) doubly protonated [M+nH-H3PO4]n+ neutral loss product ions from panels A and B, respectively. A =-H3PO4, ° = - H20, * = -NH3. ..................................................................................................... 64 Figure 2.3 ClD-MS/MS product ion spectra of protonated and deuterated LRRApSLG. (A) [M+H]*, (B) [M+2H]2*, (C) [M(D19)+D]", and (D) [M(D19)+ZD]2+ IONS. A = -H3PO4. ............................................................................................... 73 Figure 2.4 Multistage tandem mass spectrometry of LRRApTLG. ClD—MS/MS product ion spectra of the (A) [M+H]+ and (B) [M+2H]2+ precursor ions. CID-MS3 product ion spectra of the (C) singly and (D) doubly protonated [M+nH-H3POII]n+ neutral loss product ions from panels A and B, respectively. A = -H3PO4, * = - NH3. .................................................................................................................... 74 Figure 2.5 Multistage tandem mass spectrometry of GAILpSGAILR. CID- MS/MS product ion spectra of the (A) [M+H]+ and (B) the [M+2H]2+ precursor ions. CID-M83 product ion spectra of the (C) singly and (D) doubly protonated [M+nH-H3PO4]"" neutral loss product ions from panels A and B, respectively. A =-I"I3PO4, * = -NI"I3. .............................................................................................. 78 Figure 2.6 Multistage tandem mass spectrometry of GAILpTGAILR. CID-MS/MS product ion spectra of the (A) [M+H]+ and (B) [M+2H]2+ precursor ions. CID-M83 product ion spectra of the (C) singly and (D) doubly protonated [M+nH-H3P04]n+ neutral loss product ions from panels A and B, respectively. A =-H3PO4, * = -NH3. ............................................................................................................................ 80 Figure 2.7 Multistage tandem mass spectrometry of GAILpTGAILK. ClD-MS/MS product ion spectra of the (A) [M+H]+ and (B) [M+2H]2+ precursor ions. CID-MS3 product ion spectra of the (C) singly and (D) doubly protonated [M+nH-H3P04]n+ neutral loss product ions from panels A and B, respectively. A =-H3PO4, ° = - H20, * = -NH3. ..................................................................................................... 82 Figure 2.8 Multistage tandem mass spectrometry of LFTGHPEpSLEK. CID- MS/MS 3product ion spectra of the (A) [M+H]+ and (B) [M+2H]2+ precursor ions. CID-MS product ion spectra of the (C) singly and (D) doubly protonated [M+nH- H3PO4]n+ neutral loss product ions from panels A and B, respectively. A =-H3PO4, o = -H20. .............................................................................................................. 84 Figure 2.9 Multistage tandem mass spectrometry of LFTGHPEpTLEK. CID- MS/MS :product ion spectra of the (A) [M-I-H]+ and (B) [M+2H]2+ precursor ions. CID-MS product ion spectra of the (C) singly and (D) doubly protonated [M+nH- H3PO4]"“ neutral loss product ions from panels A and B, respectively. A =- H3PO4, * = -NH3. ................................................................................................. 86 Figure 2.10 CID-MS3 product ion spectra of the (A) singly and (B) doubly protonated [M+nH-H3,PO.I]n+ neutral loss product ions of LRRApSLG from Figure 2.3A and B, respectively. * = -NH3. .................................................................... 93 Figure 3.1 Ion trap CID—MS/MS product ion spectra of the (A) singly, (B) doubly, and (C) triply protonated precursor ions of the model synthetic phosphopeptide LFTGHPEpSLER (TpSR). The spectra are labeled according to the known or expected Ser phosphorylation site. Ions labeled with bold text are unambiguously indicative of the expected site of phosphorylation at the Ser residue. A = —98 Da (-H3PO4 or —(H20+HPO3)); CI = —80 Da (—HP03); ° = -18 Da (-HZO); I = +80 Da (+HP03). ...................................................................... 106 Figure 3.2 lon trap ClD—MS/MS product ion spectra of the (A) singly, (B) doubly and (C) triply protonated precursor ions of the model synthetic phosphopeptide LFpTGHPESLER (pTSR). The spectra are labeled according to the known or expected Thr phosphorylation site. Ions labeled with bold text are unambiguously indicative of either the expected or correct site of phosphorylation xi at the Thr residue, or the unexpected or incorrect site of phosphorylation at the Ser residue. A = -98 Da (—H3PO4 or —(H20+HP03)); CI = —80 Da (—HP03); ° = — 18 Da (—H20); I = +80 Da (+HP03). ................................................................. 109 Figure 3.3 Ion trap CID—MS/MS product ion spectra of the (A) singly, (B) doubly and (C) triply protonated precursor ions of the model synthetic phosphopeptide LFpTGHPESLER (pTSR). The spectra are labeled according to the unexpected (i.e., incorrect) site of phosphorylation at the serine residue. A = —98 Da (—H3PO4 or -(H20+HPO3)); c1 = —80 Da (—HP03); ° = —18 Da (—HzO); I = +80 Da (+HP03). .......................................................................................................................... 115 Figure 3.4 Ion trap CID—MS3 product ion spectra of (A) the y6+80 Da (y6+ HPO3) product ion formed from ClD-MS/MS of the singly protonated pTSR precursor ion in Figure 3.2A and (B) the ye product ion formed from ClD—MS/MS of the singly protonated TpSR precursor ion in Figure 3.1A .................................................. 120 Figure 3.5 lon trap ClD—MS/MS product ion spectra of the singly protonated precursor ion of the model synthetic phosphopeptide LFpTGHPESLER (pTSR) using (A) 10 msec and (B) 2000 msec ion activation times at an activation q value of 0.25. The spectra are labeled according to the known (i.e., expected) site of phosphorylation (i.e, Thr). Ions labeled with bold text are unambiguously indicative of the unexpected site of phosphorylation (i.e, Ser), resulting from the gas-phase rearrangement reaction. A = -98 Da (—H3PO4 or —(HZO+HP03)); CI = —80 Da (-HP03); ° = —18 Da (—H20); I = +80 Da (+HP03); RA = relative abundance. ....................................................................................................... 124 Figure 3.6 Triple quadrupole CID-MS/MS product ion spectra of the (A) singly, (B) doubly, and (C) triply protonated precursor ions of the model synthetic phosphopeptide LFpTGHPESLER (pTSR). The spectra are labeled according to the known (i.e., expected) site of phosphorylation (i.e, Thr). Ions labeled with bold text are unambiguously indicative of the expected site of phosphorylation at the threonine residue. A = —98 Da (—H3PO4 or -(HZO+HP03)); CI = —80 Da (- HPOa); ° = —18 Da (—H20). ................................................................................ 126 Figure 3.7 Expanded regions of the product ion spectra obtained by (A) ion trap (Figure 3.2A) and (B) triple quadrupole (Figure 3.6A) ClD-MS/MS of the singly protonated precursor ion of the model synthetic phosphopeptide LFpTGHPESLER (pTSR). The spectra are labeled according to the known (i.e., expected) site of phosphorylation (i.e, Thr). Ions labeled with bold text are unambiguously indicative of the unexpected site of phosphorylation (i.e, Ser), resulting from the gas-phase rearrangement reaction. a = —80 Da (—HP03); ° = — 18 Da (-HzO); I = +80 Da (+HPO3). ................................................................. 129 Figure 3.8 lon trap CID—M83 product ion spectra of the (A) singly, (B) doubly, and (C) triply protonated ([M+nH—98]"*) neutral loss product ions from the TpSR peptide in Figure 3.1A—C, respectively. Ions labeled with bold text correspond to the presence of an unmodified serine residue rather than the expected xii dehydrated residue. The inset in panel B shows the expanded m/z region from 600-630. ° = —18 Da (—H20); . = +18 Da (+H20). ............................................. 149 Figure 3.9 lon trap CID—M83 product ion spectra of the (A) singly, (B) doubly, and (C) triply protonated ([M+nH—98]"+) neutral loss product ions from the pTSR peptide in Figure 3.2A—C, respectively. Ions labeled with bold text correspond to the presence of an unmodified threonine residue rather than the expected dehydrated residue. The inset in panel B shows the expanded m/z region from 600-630. ° = -18 Da (—HZO); - = +18 Da (+H20). ............................................. 153 2[figure 4.1 Flowchart to determine cleavage solution composition. Adapted from . ..................................................................................................................... 167 Figure 5.1 The NF-kB pathway. Adapted from 254. .......................................... 178 Figure 5.2 Summary of the known phosphorylation sites of p65 and the corresponding [kinases] and [phosphatases*]. The kinases responsible for Ser205, T254, and Thr435 phosphorylation are at present unknown, indicated with [?]. This information was compiled from 256' 257; and adapted from 256. RHD is the Rel homology domain and TAD is the transactivation domain ................. 181 Figure 5.3 Schematic of the phosphoproteome analytical strategy employed for identification and phosphorylation site determination of p65-interacting proteins from the nuclear fractions of untreated and TNF-o-treated THP-1 cells. ........... 184 Figure 5.4 Western blot detection of p65 (total, C-terminal epitope) and actin from replicate untreated and TNF-a-treated THP-1 nuclear extracts. Western blot of the actin loading control indicates that the untreated and TNF-cr-treated THP-1 nuclear extracts contain comparable amounts of protein; and therefore, that the amount of nuclear p65 is greater for the TNF-a-treated nuclear extract than the untreated nuclear extract. This gel was edited to enable facile comparison between the untreated and treated samples. The unaltered image is shown in Figure 5.15. ........................................................................................ 191 Figure 5.5 Western blot detection of p65 (total, N-tenninal epitope) from p65- immunoprecipated from untreated (IPT) and TNF-u-treated (lP‘) THP-1 nuclear extracts. The immunoprecipitation of p65 from the untreated and treated samples were completed on separate occasions. For this reason the experiments are differentiated here. Control experiments involving immunoprecipitation of p65 from phosphate-buffered saline are also shown (*). Higher amounts of nuclear p65 is present in the treated than the untreated samples after p65-immunoprecipitation, consistent with the results shown in Figure 5.4 for the samples prior to p65-immunoprecipitation. The unaltered image is shown in Figure 5.16 ........................................................................... 192 Figure 5.6 Representative MALDI-mass spectrum of the samples resulting from phosphopeptide enrichment of the trypsin digested p65-lmmunoprecipitated nuclear extracts from TNF-a-treated THP-1 cells, where * = identified xiii phosphorylated peptides; o = identified nonphosphorylated peptides; ? = unidentified peptides; and m = matrix ions. ....................................................... 194 Figure 5.7 CID-MS/MS product ion spectrum of m/z 2061 (from the MS spectrum shown in Figure 5.6). This peptide was identified as KEEpSEEpSDDDM(cam- 105)GFGLFD from the P1 or P2 isoforms of the 608 acidic ribosomal protein, and resulting from the gas-phase neutral loss of 2-(methylthio)acetamide (105 Da) from the S-carbamidomethylmethionine (M(cam))-containing species (m/z 2166). A = —98 Da (—H3PO4 or —(H20+HP03)); CI = —80 Da (—HP03); ° = —18 Da (—H20). .............................................................................................................. 196 Figure 5.8 ClD-MS/MS product ion spectra of m/z 2109 and 2125 (from the MS spectrum shown in Figure 5.6). These peptides were identified as the unmodified methionine— and oxidized methionine-containing species of KEEpSEEpSDDDMGFGLFD from the P1 or P2 isoforms of the 608 acidic ribosomal protein, respectively. A = -98 Da (—H3PO4 or —(HZO+HP03)); CI = —80 Da (—HP03); ° = -18 Da (-H20); :1: = —64 Da (—CH3$OH or methane sulfenic acid). ................................................................................................................. 199 Figure 5.9 CID-MS/MS product ion spectrum of m/z 2166 (from the MS spectrum shown in Figure 5.6). This peptide was identified as the S- carbamidomethylmethionine-containing species of KEEpSEEpSDDDM(cam)- GFGLF D from the P1 or P2 isoforms of the 608 acidic ribosomal protein. ....... 203 Figure 5.10 Residue structures of methionine, oxidized methionine (M(ox)). and S-carbamidomethylmethionine (M(cam)). ......................................................... 204 Figure 5.11 CID-MS/MS product ion spectra of m/z (A) 1677, (B) 1693, (C) 1629, and (D) 1734 (from the MS spectrum shown in Figure 5.6). These peptides were identified as various isoforms of KEEpSEEpSDDDMGF, including the unmodified methionine (m/z 1677), oxidized methionine (m/z 1693), product ion resulting from the gas-phase neutral loss of 2-(methylthio)acetamide (105 Da) from the S- carbamidomethylmethionine (m/z 2166), and S-carbamidomethylmethionine— containing species (m/z 1734) from the P1 or P2 isoforms of the 608 acidic ribosomal protein. A = —98 Da (—H3PO4 or —(H20+HP03)); CI = -80 Da (-HP03); ° = -18 Da (—H20); :I: = —64 Da (—CH3$OH or methane sulfenic acid). ............. 207 Figure 5.12 CID-MS/MS product ion spectrum of m/z 1367 (from the MS spectrum shown in Figure 5.6). This peptide was identified as pSRpSFDYNYR from the translocation liposarcoma (TLS)-associated serine-arginine protein (TASR). A = —98 Da (-H3PO4 or —(H20+HP03)); I] = —80 Da (—HP03); ° = -18 Da (—H20). ........................................................................................................ 212 Figure 5.13 Representative MALDI-mass spectrum of the samples resulting from phosphopeptide enrichment of the trypsin digested p65-immunoprecipitated nuclear extracts from untreated THP-1 cells, where * = identified phosphorylated xiv peptides; o = identified nonphosphorylated peptides; ? = unidentified peptides; and m = matrix ions. .......................................................................................... 217 Figure 5.14 MALDI-mass spectra of the (A) TNF-d-treated THP-1 cells and (B) untreated THP-1 samples resulting from p65 immunoprecipitation, trypsin digestion, phosphopeptide enrichment, and addition of 100 amol of an internal standard C03C02CO-LFTGHPEpSLEK (“05”, m/z 1398), where * = identified phosphorylated peptides; o = identified nonphosphorylated peptides; ? = unidentified peptides; and m = matrix ions. ....................................................... 219 Figure 5.15 Unedited Western blot of p65 (total, C-terminal epitope) and actin from replicate untreated and TNF-a-treated THP-1 nuclear extracts. Replicate samples labeled with * were prepared previously (1 year prior to the preparation of replicates 1 and 2) by another laboratory member and had been stored at -80 °C. Due to their similar amount of p65 in the untreated and treated previously prepared samples, they were not used for any of the experiments described in Chapter Five nor were they included in Figure 5.4. ........................................... 226 Figure 5.16 Unedited Western blot of p65 (total, N-terrninal epitope) from untreated and TNF-q-treated THP-1 nuclear extracts (THP—1 NE); and p65- immunoprecipated untreated (IP*) and TNF-a-treated (IPI) THP-1 nuclear extracts. The immunoprecipitation of p65 from the untreated and treated samples were completed on separate occasions. For this reason they are differentiated here. Control experiments including immunoprecipitation of p65 from phosphate-buffered saline are also shown (*). An additional protein band (lPa) is present in the samples resulting from immunoprecipitation of the treated sample (IP‘). However, because this band is also present in the PBS control sample, it lPal results from the immunoprecipitation process. ............................ 227 Figure 5.17 ClD-MS/MS product ion spectrum of m/z 1191 (from the MS spectrum shown in Figure 5.6). This peptide was identified as HMYHSLYLK from ribosomal protein L19. ° = —18 Da (—HzO) ................................................ 228 Figure 5.18 ClD-MS/MS product ion spectrum of m/z 1022 (from the MS spectrum shown in Figure 5.6). This peptide was identified as AAIDWFDGK from translocation liposarcoma (TLS) protein. ° = -18 Da (—HzO); * = —17 Da (— NH3). ................................................................................................................. 230 Figure 5.19 CID-MS/MS product ion spectrum of m/z 1515 (from the MS spectrum shown in Figure 5.6). This peptide was identified as IWHHTFYNELR from actin (B—isoform). ° = -18 Da (—H20). ....................................................... 232 XV LIST OF SCHEMES Scheme 1.1 Previously proposed mechanism for cleavage of the P-O phosphoester bond of negatively charged phosphorylated peptides forming [M- nH-79]<""* and P03‘(m/z 79). .............................................................................. 15 Scheme 1.2 Fragmentation pathways involving the phosphate group upon CID- MS", where dS and dT represent dehydrated serine and threonine residues, respectively. ........................................................................................................ 23 Scheme 1.3 Analysis of napthyl chromophore-derivatized phosphorylated serine and threonine peptides by UVPD at 266 nm. B-elimination of phosphoserine and phosphothreonine peptides followed by Michael addition to form the napthyl- derivatized labeled peptides. UVPD of the derivatized peptides results in selective cleavage of the chromophore and radical-induced cleavage allows for the selective formation of sequence ions adjacent to the formerly phosphorylated residue. ............................................................................................................... 32 Scheme 1.4 Proposed site-specific an+1-97 ion formation of singly protonated, phosphorylated peptides that contain N-terminal basic residues upon UVPD at 193 nm, where R1 is the side chain for any amino acid ....................................... 35 Scheme 1.5 Previously proposed fragmentation mechanism for the production of c- and z-type product sequence ions for the reaction between a low-energy electron with a multiply protonated peptide ion. .................................................. 39 Scheme 1.6 Penning ionization of a protonated peptide (M) ion, where n 2 1, by interaction with a metastable noble gas atom (m*) .............................................. 47 Scheme 2.1 Potential mechanisms for the neutral loss of H3PO4 from phosphoserine- and phosphothreonine-containing peptides via (A) a charge- remote B—elimination reaction, (B) a charge-directed E2 reaction, and (C) a charge-directed 8N2 neighboring group participation reaction. For clarity, the mechanisms are presented for phosphoserine-containing peptides only. ........... 58 Scheme 2.2 Proposed mechanisms for the dominant loss of H3PO4 via a charge- directed SN2 neighboring group participation reaction (Pathway 1) and for the loss of formaldehyde (CHzO, 30 Da) or acetaldehyde (CH3CI‘IO, 44 Da) from the [M+H-H3PO4]" product ions (Pathway 2) of protonated phosphoserine- and phosphothreonine-containing peptide ions .......................................................... 89 Scheme 3.1 Previously proposed mechanisms for the “charge directed” neutral loss of H3PO4 and RCHO from protonated phosphoserine- or phosphothreonine- (where R = H or CH3, respectively) containing peptides during ClD-MS/MS.206 (A) Intramolecular E2 mechanism and (B) intramolecular SN2 neighboring group participation mechanism. .................................................................................. 100 xvi Scheme 3.2 Competing pathways for the formation of isobaric product ions via the loss of 98 Da (—H3PO4 or —(HPO3+H20)) from a protonated phosphopeptide containing multiple potential phosphorylation sites. R1 = H or CH3 for phosphorylated serine or phosphorylated threonine, respectively; R2 = any amino acid side chain; R3 = H or CH3 for non-phosphorylated serine or non- phosphorylated threonine, respectively. ............................................................ 101 Scheme 3.3 Proposed mechanism for the intramolecular gas-phase phosphate group transfer facilitated by hydrogen bonding to a protonated basic residue. .144 Scheme 3.4 Proposed mechanism for intramolecular phosphate group transfer for deprotonated peptide ions. .......................................................................... 159 xvii CHAPTER ONE Mass Spectrometry Methods for Phosphoproteome Analysis 1.1 Introduction Protein phosphorylation is perhaps the most pervasive post translational modification, playing an essential role in the coordination and regulation of . 1, 2 . . vrrtually all cellular processes. SInce the genome does not encode for protein phosphorylation directly, phosphorylation events are dynamic and reflect the current physiological needs of a cell. Accordingly, protein phosphorylation is reversible, whereby phosphorylation by kinases and dephosphorylation by phosphatases serves to regulate protein activity, expression levels, sub-cellular localization, and protein—protein interactions. Due to its broad biological importance, it is not surprising that deregulation of phosphorylation events has been associated with the onset and progression of many pathological conditions}8 Therefore, understanding protein phosphorylation on a molecular level has been of great interest towards understanding the difference between “healthy" and “abnormal” cellular states, as well as for opening new avenues of The concepts discussed in Chapter One have been submitted for publication in Mass Spectrometry Reviews: Palumbo, A. M.; Smith, S. A.; Stemmer, P. M.; Kalcic, C. L.; Dantus, M.; Reid, G. E. Tandem Mass Spectrometry Strategies for Phosphoproteome Analysis. disease therapy.8.12 Due to the elaborate nature of protein phosphorylation, its analysis presents a formidable challenge. For example, since phosphorylation is a transient modification, the lifetime of a phosphorylated site may be short and therefore the associated structure may be difficult to “capture” for analysis. Furthermore, a particular phosphorylated form of a protein may be substoichiometric, meaning that only a small fraction of the population of a particular protein may be phosphorylated at any given time. Moreover, differentially phosphorylated isoforms (i.e., different sites of phosphorylation) of a given protein may also exist simultaneously in this substoichiometric population. The sheer number of potential phosphorylation sites also complicates their analysis. In eukaryotes, phosphorylation predominantly occurs on serine, threonine, and tyrosine residues (Figure 1.1), which represent approximately 17% of the total amino acid content in an average protein.13 Based on this fact it has been calculated that in an average eukaryotic cell, there exists nearly 700,000 different potential phosphorylation sites.14 OH OH ,‘p’OH I I O \\ HO-P=O HO-P=O o o o a,” /['( KE'IIK H. N o o H o Phosphoserine, pS Phosphothreonine, pT Phosphotyrosine, pY Figure 1.1 Structures of phosphorylated serine, threonine and tyrosine residues. To overcome these inherent biological challenges, numerous mass spectrometry (MS) based approaches have been developed, and have now become the methods of choice for the identification and characterization of an individual phosphoprotein or for the comprehensive examination of an entire “phosphoproteome” from a given tissue, cell or sub-cellular fraction under a given set of physiological or pathological conditions.15'21 Phosphoproteome analysis methods are generally based on the interrogation of peptides derived from enzymatic digestion of intact phosphoproteins, i.e., the “bottom-up approach”. For complex biological mixtures, peptide digests are typically fractioned off-line by chromatographic separation techniques (e.g., Reversed Phase High Performance Liquid Chromatography, RP-HPLC).22 To address challenges associated with the analysis of low abundance phosphoproteins or substoichiometric phosphorylation, phosphopeptides are typically enriched prior to on-line RP-HPLC and introduction to the mass spectrometer. Many different enrichment techniques are available for this purpose, including immunoaffinity chromatography,23 ion exchange chromatography,24'27 Immobilized Metal ion Affinity Chromatography (IMAC),28'31 Metal Oxide Affinity Chromatography (MOAC),32'36 and chemical derivatization.37'46 Furthermore, enrichment on the phosphoprotein level may be employed to decrease sample complexity prior to enzymatic digestionfw.49 Comprehensive reviews describing the development and application of these phosphopeptide and phosphoprotein enrichment methods have recently appeared elsewhere in the Iiterature,5o' 51 and so will not be further described here. Although these sample preparation methods have undoubtedly improved phosphoproteomic analysis strategies, significant challenges remain associated with the mass spectrometry-based aspects of the phosphoproteome characterization process. In particular, the presence of the phosphate group can significantly alter the gas-phase chemistries of peptide ions, particularly with regard to phosphopeptide structural analysis using tandem mass spectrometry (MS/MS). This ultimately impacts the effectiveness of MS/MS strategies used for phosphopeptide identification and characterization (i.e., phosphorylation site assignment). For this reason, this chapter focuses on providing an overview of both traditional and newly emerging tandem mass spectrometry strategies employed for phosphopeptide identification and characterization. Particular attention is given to highlighting the underlying chemical principles, mechanisms and other factors that influence the gas-phase fragmentation reactions associated with each of these strategies, as these issues directly influence their utility. Furthermore, several of the analytical challenges that remain to be overcome in order for this important sub-discipline of the maturing field of proteomics to be used more efficiently and effectively are discussed. 1.2 Mass Spectrometry Methods Employed for Phosphoproteome Analysis The major goals of mass spectrometry-based phosphoproteomic analysis are: (i) identify the presence of phosphorylated peptides, (ii) determine the identity of phosphopeptides by their amino acid sequences, (iii) localize the specific sites of phosphorylation, and (iv) quantitatively determine any temporal dependence to the phosphorylation site occupancies, under the specific biological conditions from which the phosphoprotein was isolated. These aims may be achieved by “soft" ionization of phosphopeptides followed by the use of mass spectrometry (MS), tandem mass spectrometry (MS/MS), and multistage tandem mass spectrometry (MSn) techniques. Each of these mass spectrometry approaches are described in detail below for phosphoproteome characterization 1.2.1 Ionization of Phosphorylated Peptides Phosphopeptide ionization is typically achieved by electrospray ionization (ESI) or matrix-assisted laser desorption/ionization (MALDI). The use of these two sources often allow complementary MS and MS/MS information to be obtained, as they produce different types of ions, i.e., multiply charged ions are typically formed by ESI whereas singly charged ions are formed by MALDI. The ionization, subsequent detection and/or structural analysis of phosphopeptides by these sources may, however, be limited by the presence of extraneous peptides. It has previously been proposed that the ionization efficiency of phosphorylated peptides may be lower than their nonmodified peptide analogs 52, 53 using MALDI. For these reasons, the use of alternative MALDI matrices such as 2’,4’,6’-trihydroxyacetophenone, and a range of matrix additives such as ammonium salts and phosphoric/phosphonic acids, have been employed to enhance the phosphopeptide ionization yields.54'57 The potential for decreased ionization efficiency of phosphorylated peptides versus their nonphosphorylated analogs has also been examined for ESI. However, the results of this study did not indicate selective suppression.58 Rather, it was concluded that phosphopeptide ions produced by ESI may be nonspecifically suppressed owing to their low stoichiometry.58 These results further indicate a necessity for phosphopeptide enrichment prior to MS analysis. 1.2.2 MS Analysis of Phosphorylated Peptide Ions Perhaps the simplest method for identification of the presence of a phosphopeptide by MS is via analysis of a sample before and after phosphatase treatment.53’ 59 The observed reduction in mass by n multiples of 80 Da of a peptide is indicative of the presence of n phosphorylation sites. Additionally, derivatization of phosphopeptides with mass tags, either in solution or in the gas- 37-41, 46, 60 phase (via ion-molecule reactions), and subsequent MS analysis may be used to determine the presence of phosphopeptides from a mixture. Derivatization techniques have also been used for the incorporation of differential labeling tags, allowing quantitative analysis of the occupancies of 38, 46, 61 phosphorylation sites. All of these techniques are, however, limited to relatively simple mixtures, where the mass shift is readily observable. Thus, the majority of methods for phosphoproteome analysis rely on the use of tandem mass spectrometry techniques. 1.2.3 Tandem Mass Spectrometry and Gas-phase Ion Chemistry of Phosphorylated Peptides Tandem mass spectrometry (MS/MS) involves isolation of a selected precursor (reactant) ion, followed by a gas-phase reaction (typically fragmentation) and analysis of the resultant product ions. In certain types of mass spectrometry instrumentation, product ions formed from a previous stage of reaction can be subsequently isolated and then subjected to further reaction and mass analysis. This is termed MSn analysis, where n represents the total number of reaction stages. A plethora of tandem mass spectrometry methods have been employed for the characterization of peptide ions, including collision- )(reviewed in 62) (reviewed in induced dissociation (CID post-source decay (PSD), 63) (reviewed in 64) infrared multiple photon dissociation (IRMPD), ultraviolet in 65) photodissociation (UVPD),(reViewed femtosecond laser induced ionization/dissociation (stID),66 electron capture dissociation (ECD),(rev'ewed m 7 67) (reviewed in 68) electron transfer dissociation (ETD), electron ionization dissociation (EID),69 electronic excitation dissociation (EED),(revlewed m 67) (reviewed in 67) electron-detachment dissociation (EDD), and metastable atom- activated dissociation (MAD).70 Many of these tandem mass spectrometry methods have been applied to phosphoproteome analysis, and the demonstrated strategies are summarized in Figure 1.2. These tandem mass spectrometry strategies have allowed for the identification of the presence of phosphorylated peptides, the identification of the phosphopeptide by its amino acid sequence, the determination of phosphorylation sites, and quantitative analyses of phosphorylation site occupancies. Figure 1.2 Flow chart of previously demonstrated tandem mass spectrometry strategies for phosphopeptide analysis. The phosphopeptide precursor ion charge state (n) requirement for each technique is specified. Some techniques, namely ClD-, PSD-, and lRMPD-MS/MS, may involve formation of “non- sequence” product ions that diagnostically indicate the presence of phosphorylated peptide ions. The observation of these diagnostic product ions has been used to trigger the acquisition of CID-MS3 spectra of the neutral loss product ion, or to trigger the acquisition of MS/MS spectra of the original precursor ion by other methods (e.g., ECD-MS/MS). *Phosphopeptide analysis by UVPD of deprotonated precursor ions has been performed on phosphate- derivatized peptides. Figure 1.2 (cont'd) mm: ems. 258990 $5.06 :38 2an 82 ozwotmmfi c6 .2 .3 ca Fa .me Buzzes: scofv no 8&2 5982c $55-12: 1015 no ow-co_ .62:on Ace E:_coEE_ >3 can 68.95 29558. n a. an .aw on 95.08 $0.: NE 2 60930 20263: n 95 A3. .mm 68 $5 9.2 6235 osmocmmfi cocoa—poms: 8920 9.2 8:85 unmocmflu mucoscomco: confine 3o>=mmoz zeszaoo wwc TN: TM: In: 8m 9m now 9.3 § a: E 2: I. N c r... N c F+ N : ESE. owe 90 E 5 S AI :2 .8585 ouwaoaocamocn. 10 The ability to achieve such a wealth of phosphoproteome information from these MS/MS techniques is ultimately dependent on the types of product ions formed. The possible product ions resulting from the dissociation of a peptide ion along its backbone have been defined previously and are summarized in Figure 71, 72 1.3 Briefly, fragmentation can occur at the Ca-Ccarbonyl. Ccarbony|-N (i.e., amide), or N-Ca bonds to give rise to a-, b-, or c-ions if the charge is retained on the N-terminal segment and x-, y-, or z-ions if the charge is retained on the C- tenninal segment. Therefore, if a population of precursor peptide ions undergo fragmentation to produce a series of consecutive a-, b-, c-, x-, y-, or z-type “sequence” ions, the amino acid sequence of the peptide could be determined from the difference between these consecutive ions. Importantly, since the residue masses of phosphorylated serine (167 Da), phosphorylated threonine (181 Da), and phosphorylated tyrosine (243 Da) are unique, the product sequence ions of phosphorylated peptides are indicative of the presence of phosphopeptides and the location of the phosphorylation sites within the peptide. 11 ...... OOOOOOOOOOOOOOOOOO OOOOOOOOOOOOOOOOOO r r . .. O R' O R'" Figure 1.3 Nomenclature of peptide backbone fragment ions or sequence ions. In addition to backbone fragmentation reactions, fragmentation of certain amino acid side chains may occur. Although these ions are not indicative of the amino acid sequence or the location of the phosphorylation site within the peptide sequence, and are therefore termed “nonsequence” ions, they may provide information regarding the amino acid composition. In particular, some ion dissociation techniques induce selective cleavage of the phosphoester bond of phosphopeptides, resulting in the observation of nonsequence product ions which indicate the presence of a phosphorylated peptide (Figure 1.2-Diagnostic Ions). This is in contrast to ion dissociation techniques that give rise only to sequence ions, where interpretation of the sequence is required to determine whether or not a peptide is phosphorylated Importantly, the specific method by which the ion is subjected to fragmentation dictates the type (i.e., sequence vs. nonsequence) and abundance 12 of the product ions that are observed following an MS/MS experiment. The observed sequence and/or non-sequence ions formed are also heavily dependant on the properties of the precursor ion (e.g., precursor ion polarity and charge state). As a result, the success of any given tandem mass spectrometry strategy employed for phosphoproteome identification and characterization is critically dependant on the underlying chemistry of these techniques.21 The remainder of Chapter One therefore focuses on describing the current state of knowledge regarding the gas-phase ion chemistry associated with the various MS/MS methods employed to date for phosphopeptide identification and phosphorylation site assignment (i.e., characterization), and highlights the influence of gas-phase ion chemistry on the applicability of the analytical strategies used. Reviews describing the development and application of quantitative tandem mass spectrometry based phosphoproteomic analysis strategies have recently been discussed elsewhere, and therefore will not be discussed further herezo’ 73 1.2.3.1 Collision-Induced Dissociation (CID) Collision-induced dissociation, CID, (also known as collision activated dissociation, CAD) is the most widely used tandem mass spectrometry method available for peptide sequencing. Many different mass analyzer platforms are compatible with CID, including quadrupole ion trap (QIT) and Fourier transform- ion cyclotron resonance (FT-ICR) mass spectrometers. Additionally, utilization of quadrupole-based collision cells allow for CID-tandem mass spectrometry in 13 triple quadrupole (QqQ), quadrupole time of flight (qTOF), and TOF/T OF instruments. Ion activation by CID occurs by conversion of ion translational energy into internal (vibrational) energy upon collision with a neutral inert target gas (e.g., He or Ar). The acquired vibrational energy is then quickly redistributed throughout the entire ion (510'12 s) and ergodic fragmentation of a particular bond occurs once the vibrational energy exceeds the activation barrier of that bond.74 Since CID is under kinetic control and is generally a “slow heating” process (i.e., the vibrational energy deposition occurs over many collision events), the most labile bonds fragment preferentially.74 CID of nonphosphorylated peptides generally leads to non-selective sequence ion fragmentation, forming b- and y-sequence ions relating to fragmentation of amide bonds (activation energy barrier of amide bond cleavage ~ 40 kcal/mol cleavage).75' 76 Phosphorylated peptides, however, often undergo preferential nonsequence ion fragmentation of the 77, 78 phosphoester bond (activation energy barrier ~12 kcal/mol). Selective cleavage of labile phosphoester bonds may give rise to neutral losses of 80 Da (HPO3) and 98 Da (H3PO4 or H20+HPO3) or charged losses of 63 Da (P02) 79 Da (P03) and 97 Da (H2PO4') forming either nonsequence product ions (e.g., [M+nH-981n'1) or sequence product ions resulting from sequential fragmentation reactions (e.g., bn-98),24' 26' 79‘89 14 ClD-MS/MS has been used for the identification of phosphopeptides and localization of phosphorylation sites in both negative and positive ion modes (Figure 1.2-1 and 1.2-2, respectively). The gas-phase ion chemistry governing these ClD-MS/MS techniques, as well as its impact on the utility of these methods, is described in detail below. 1.2.3.1.1 Formation of 79 mlz (P03) and [M+nH-79](n'1)' Ions CID of negatively charged phospho-serine, -threonine, and -tyrosine peptide precursor ions commonly result in the cleavage of the phosphoester bond giving rise to P03' ions (mlz 79). When the precursor ion is multiply de- 101-1)- protonated, a charge reduced counterpart ion ([M-nH-79 ) may also be observed. A mechanism for the formation of these ions has been previously proposed, and is shown in Scheme 1.1 for a doubly deprotonated precursor . 85-88 IOD. (190 0Q:— o' o H ClD-MS/MS_ i H + PO' FJLN N.“ r u N.“ 3 H o o [M-nH-79]("'1" m/z 79 Scheme 1.1 Previously proposed mechanism for cleavage of the P-O phosphoester bond of negatively charged phosphorylated peptides forming [M- nit-791""1 * and Poe'(m/z 79). 15 Observation of the PO3' nonsequence product ion (mlz 79) by CID in negative ion mode has been used to indicate the presence of phosphorylated 85. 86 peptides (Figure 1.2-1 a). Since m/z 79 may be too low to be observed in ion trap platforms under typical peptide analysis conditions, the [M+nH-79](n'1)' product ion may alternatively be used for phosphopeptide detection (Figure 1.2- 1a).88 Additionally, the observation of these nonsequence product ions has been used to trigger the acquisition of a positive ion mode ClD-MS/MS spectrum of the corresponding protonated precursor ion for phosphopeptide sequence analysis (Figure 1.2-1a followed by Figure 1.2-2). This has been achieved by either switching the polarity during sample analysis or performing a second round of . . . . . 83, 84, 90 analySIs In posrtive Ion mode. 1.2.3.1.2 Formation of Phosphotyrosine-Specific Immonium Ions at mlz 216.043 In positive ion mode, CID-MS/MS of protonated phosphotyrosine- containing peptide ions can give rise to nonsequence phosphotyrosine-specific 91, 92 immonium ions at m/z 216.043 (Figure 1.4). Immonium ions provide composition-specific information and may result from cleavage of the N-terminal residue or from two backbone cleavages surrounding a particular internal residue.93 It has been previously shown that the abundance of the 16 phosphotyrosine (pY) immonium ion is dependent upon the sequence of the peptide,94 and the immonium ion abundance is highest when the phosphotyrosine residue is located at the first or second residue from the N- ten'ninus.94 This is in accordance with the observation that N-tenninal residues generally favor the formation of immonium ions.93 Additionally, since immonium ions may result via secondary fragmentation of bn sequence ions (bn-)an+CO) and b2 ion formation is favored, localization of the phosphotyrosine residue at the second position from the N-terminus may also preferentially yield these ions?3 Furthermore, a fragmentation study on a series of synthetic phosphopeptides with the general sequence DQQDFFPK where pY was placed in positions 19 demonstrated that the closer the pY residue was to the C-terminal lysine, the less likely the corresponding immonium ion was to form.94 OH I HN=CH-R HN=CH-CH2©—O-fi-OH O Figure 1.4 General structure of an immonium ion, where R is the amino acid side-chain and the phosphotyrosine-specific immonium ion, m/z 216.043. Precursor ion scanning of the pY immonium ion has been used for selective detection of phosphotyrosine-containing peptides from within complex 91, 92, 95 mixtures (Figure 1.2-2a). Since this diagnostic ion may have the same 17 nominal mass as other ions (216 Da), this technique has proven most effective when using high resolution/high mass accuracy instruments such as a qTOF.91' 92' 95 However, considering that this method is reliant upon the formation of this ion, this technique may not allow identification of the presence of phosphotyrosine-containing peptides when the residue is located near the C- terminus. 1.2.3.1.3 Product Ions Formed From 80 or 98 Da Neutral Losses The analysis of peptides containing phospho-serine, -threonine, and/or - tyrosine residues by ClD-MS/MS in either positive or negative ion mode may give rise to product ions corresponding to the neutral losses of 80 Da (HP03) and 98 24, 26, 80-83, 87, 89 Da (commonly assigned as H3PO4). These neutral loss processes yield nonsequence product ions (i.e., [M-nH—80]n', [M-nH-981n', [M+nH-80]n+, and [M+nH-80]n+) that diagnostically indicate the presence of phosphorylated peptide ion (Figure 1.2-1b, 1.2-2b). Furthermore, product ions formed via sequential nonsequence and sequence type fragmentation reactions (e.g., bn-80, bn-98, yn-80, and yn-98), may also be observed.96 Competition between the formation of nonsequence ions, which may be used to determine the presence of phosphopeptides, and the formation of sequence ions,97 which may be used for peptide identification and localization of 18 the phosphorylation site, critically affects the utility of a particular ClD-MS/MS approach for phosphoproteome characterization. When the formation of b- and y- sequence product ions is suppressed by the dominant formation of a 98 Da neutral loss nonsequence product ion, alternative methods have been used to obtain the required sequence information. These include automatically subjecting the initial 98 Da neutral loss product ion to a further stage of ClD-MS/MS (i.e., M83) (Figure 1.2-2b.i) or “pseudo-M83” (Figure 1.2-2b.ii) in ion trap mass spectrometers in a data dependent mode of 24, 98-101 operation. CID-MS3 entails isolation and further fragmentation of the . 3 . . . . neutral loss speCIes, whereas pseudo-MS Involves Simultaneous activation of the precursor ion and the resultant 98 Da neutral loss product ion during a single CID-MS/MS event.101 Therefore, the pseudo-MS3 product ion spectrum contains a “composite” of the product ions generated from fragmentation of both the precursor peptide and the initial neutral loss product ion. Alternatively, observation of these 98 Da neutral loss product ions has been used to trigger ECD-MS/MS analysis of the original precursor ion (Figure 1.2-2b followed by Figure 1.2-7), under which conditions the phosphate group remains intact 102, 103 (described further in its respective section below). These data-dependent strategies have been used for peptide identification and phosphorylation site determination. The utility of these methods are, however, highly dependent upon the formation of these neutral loss ions as abundant species. Additionally, the appearance of the product ion spectrum resulting from CID-MS3 of the 98 Da 19 neutral loss product ion is dependent on the structure of the initial neutral loss ion and the mechanism(s) by which it is formed. For these reasons, it is important that the factors that govern the _ abundance and structures of these neutral loss ions, as well as the mechanisms by which they are formed, are determined. Previous studies have noted that these neutral loss product ions are formed at higher abundance at higher . . . 8 4 . colhsron energies, and In mass spectrometers that employ more energy transferred in short activation times (microsecond), like triple quadrupole instruments.105 The structure of the phosphorylated amino acid also affects the extent of the formation of these neutral loss product ions .by CID, whereby phosphotyrosine undergoes the neutral loss of 80 Da more frequently than 98 87. 89 a. D while phosphoserine and phosphothreonine undergo the neutral loss of 87. 89 98 Da more frequently than the loss of 80 Da. Woods and coworkers have also shown that the neutral loss of 98 Da from phosphotyrosine-comaining peptides is highly dependent upon the presence of basic residues, where the neutral loss of 98 Da was observed only when arginine and lysine residues were present within the peptide};2 This has been hypothesized as requiring the presence of strong hydrogen-bonding interactions between the phosphate group . . . . . 82.106-110 and a protonated arginine or lysrne Side chain. It is generally agreed that the precursor ion charge state also influences the abundance of the 98 Da neutral loss ion, where higher charge states result in 20 87,89,111 lower phosphate neutral loss product ion abundance. Another factor that may influence the abundance of these nonsequence neutral loss ions is the combined effect of precursor ion charge state and basic residue content (i.e., proton mobility). The “mobile proton” model is a central concept to rationalize the appearance of CID product ion spectra with the features of the protonated precursor peptide ion.112'114 In particular, this model considers the fact that the ionizing proton may be directly involved in the fragmentation reaction (charge- directed), or sequestered to a region of high proton affinity (e.g., side chains of basic residues) and therefore unavailable for involvement in fragmentation (charge-remote). Therefore, the proton mobility of a precursor ion, which is influenced by both the precursor ion charge state and the amino acid composition, rather than the charge state alone, is expected to dictate the type and abundance of product ions observed and the mechanisms by which they are formed. A body of evidence exists in the literature that demonstrates the important role of proton mobility on the gas-phase fragmentation reactions of protonated peptide ions containing other post-translational and process-induced 115,116 modifications. Consequently, it is critical to obtain detailed insights into the role of proton mobility on the multistage gas-phase fragmentation reactions of protonated phosphopeptides, as this will ultimately dictate the utility of CID tandem mass spectrometry methods for phosphoproteome analysis. The work described in Chapter Two aims to elucidate the relationship between proton 21 mobility and the gas-phase fragmentation of phosphorylated peptides under CID conditions. Many mechanisms for fragmentation of the phosphate group have been previously proposed. These include the direct loss of H3PO4 from the 78, 87, 89, 117 phosphorylated residues, the losses of HPO3 and H20 from a phosphorylated and a nonphosphorylated hydroxyl (or carboxyl)-containing 89, 118 amino acid residue, and the potential for intramolecular gas-phase phosphate group transfer upon CID (Scheme 1.2).89 It is clear from Scheme 1.2 that these specific fragmentation pathways involving the phosphate group will dictate the utility of CID tandem mass spectrometry methods for accurate and unambiguous phosphorylation site assignment, since differentially phosphorylated isomers may give rise to the same product ion structures. In particular, the presence of sequence ions that have undergone a phosphate neutral loss (e.g., bn-80 or bn-98) complicate phosphate site interpretation since they are isostructural to nonphosphorylated ions; while sequence ions resulting from the neutral loss of H20 from a nonphosphorylated residue may be confused with sequence ions resulting from the losses of H3PO4 from a phosphorylated residue. Additionally, sequence ions resulting from neutral losses of HPO3 are indistinguishable from nonphosphorylated residues, and as a result, the reliability of phosphate-intact sequence ions may be compromised. Consequently, the work described in Chapters Two and Three aims to reveal the potential for these 22 fragmentation pathways, elucidate the specific mechanisms involved, determine the resulting product ion structures, and establish the factors that govern these processes. Importantly, discussion of the implications of these processes on CID strategies for phosphoproteome analyses is also included in these chapters. 13 999009990 I “20(3). -H3P0.(pT) C'Dlv's" -HP03 I i I OGGODOGOO Pfisfisfle GGOODOGOB t I I 41200): CIDIMS" 'H3P04IPS) -HP03(pS) eeeoooege Scheme 1.2 Fragmentation pathways involving the phosphate group upon CID- MS", where dS and dT represent dehydrated serine and threonine residues, respectively. 1.2.3.1.4 Chemical Derivatization and CID-MSIMS To circumvent the potential limitations associated with the Iability of the phosphate group upon CID-MS/MS, other phosphopeptide characterization strategies involving chemical derivatization or conversion of the phosphopeptides prior to CID-MS/MS have been developed. One of the simplest derivatization techniques involves B-elimination of phosphate groups to form the dehydrated species (dehydroalanine for phosphoserine and dehydroaminobutyric acid for phosphothreonine) under alkaline conditions followed by Michael addition with a 23 nucleophile.37 This method has been adapted such that dithiol nucleophiles may act as crosslinkers between modified peptides and affinity materials such as biotin or solid phase resins allowing for both phosphopeptide enrichment and altered CID gas-phase fragmentation behavior.38.41 This approach has also been used to transform phosphorylated serine to aminoethyl cysteine, a lysine mimic.42 By this method, the formerly phosphorylated sites can be targeted for proteolysis by trypsin then subsequently identified by analysis of their CID- MS/MS spectra. Another method involves conversion of the phosphate group to 2-dimethylaminoethanesulfoxide followed by CID to give rise to the characteristic neutral loss of sulfenic acid (mlz 122.06).119 Alternatively, others have converted phosphate groups to S-pyridylethyl groups, which fragment to produce characteristic ions at m/z 106 upon CID.120 The abundance of each of these characteristic product ions is, however, dependent upon the proton mobility of the peptide precursor ion and as a result, they may not be universally detected. Furthermore, due to their low m/z, these methods may not be applicable for detection in ion traps under typical CID conditions. Drawbacks for all of these techniques also exist since (i) O-linked sugars and nonphosphorylated serine and threonine residues may undergo elimination, (ii) phosphotyrosine does not undergo elimination, and (iii) precautions that may be taken to reduce elimination of nonphosphorylated residues may cause limited conversion efficiency of the phosphorylated residues.40 24 Since the gas-phase interactions between phosphate groups and side 107,108,121 chains of basic residues has been demonstrated, other derivatization approaches that block these interactions may also be used to decrease the Iability of the phosphate group upon CID. In fact, malondialdehyde has been used to decrease the basicity of arginine in phosphorylated peptides and CID- MS/MS of these derivatized phosphopeptides has been shown to result in an increase of phosphate-containing b- and y-fragment ions and a decrease in the 98 Da neutral loss species.122 Gronert et al. have shown that protonated phosphopeptides may react in the gas-phase with alkoxy-boron species in an 78,123 ion-molecule reaction. ClD-MS/MS of the resulting product ions that correspond to alkoxy moiety loss gives rise to the boron derivatized peptide ions (e.g., [M+B-H]+). Further CID activation of the boron derivatized ions results in phosphopeptide backbone fragmentation. Production of phosphate-containing sequence ions from these boron derivatized ions is greater in abundance than for their underivatized counterparts-’8’ 123 decreasing the ambiguity of phosphorylation site assignment by CID. 1.2.3.2 MALDI Post-Source Decay (PSD) Post-source decay (PSD) has also been used for phosphopeptide characterization (Figure 1.2-3). PSD involves metastable decomposition of a precursor ion after leaving the MALDI source in a field free region of a TOF instrument.124 In essence, this process relies on ions that are stable enough to 25 exit the source, but are not stable enough to survive the flight to the detector, and decompose. Ion activation for PSD is thought to originate from different mechanisms, where those of major influence occur in the source (photon- . . . . . 63 molecule Interactions, matnx-molecule Interactions, temperature effects, etc.). The resulting product ions are then resolved by their difference in kinetic energy (as a result of their difference in mass) in a reflector TOF.125 Similar to CID, PSD of peptides gives rise to sequence information as b- and y-product ions.126 PSD product ion spectra of phosphorylated peptides are also similar to that of CID, where the phosphotyrosine-specific immonium ion has been observed and may be used to indicate the presence of phosphotyrosine- containing peptides (Figure 1.2-3a)127 and where nonsequence product ions resulting from the neutral losses of 98 Da (H3PO4 or HP03+H20) and 80 Da (HPO3) may be used as diagnostic indicators of the presence of phosphopeptides (Figure 1.2-3b).127'129 The effects of the identity of the phosphorylated amino acid on the abundance of the phosphate neutral loss product ions formed by PSD have been previously reported. Similar to CID, the loss of 98 Da is more prevalent for PSD of phosphoserine- and phosphothreonine-containing peptides compared to phosphotyrosine-containing peptides.127 It has also been determined that the loss of 80 Da occurs more readily for phosphorylated tyrosine-containing peptides than for those containing phosphoserine or phosphothreonine.127 26 Furthermore, the relative abundance of phosphate-intact product sequence ions vs. their “dephosphorylated” counterparts (e.g., bn-80 and bn-98) is high for phosphotyrosine-containing peptides and varies for phosphoserine- and phosphothreonine-containing peptides.127'129 Due to the similarities in CID and PSD product ion spectra, it is likely that the mechanisms by which the phosphate neutral loss product ions are formed are similar. Accordingly, it has been previously hypothesized that the neutral loss of 98 Da upon PSD of phosphorylated serine- and threonine-containing peptides may result from the direct loss of H3PO4 (98 Da) or from the consecutive losses of HPO3 from a phosphorylated residue and H20 from a nonphosphorylated residue (combined mass of 98 Da) (Scheme 1.2).127' 130 Additionally, intramolecular phosphate group transfer from phosphorylated tyrosine to a formerly unmodified residue, followed by the loss of H3PO4, has also been proposed for PSD (Scheme 1.2).128 In that study, it was demonstrated that the neutral loss of H3PO4 depended on the presence of aspartic acid and arginine residues. By this observation it was proposed that the HPO3 group was transferred from tyrosine to aspartic acid followed by cleavage to yield a succinamide-containing product ion.128 Consequently, the same possible ambiguities regarding phosphorylation site determination by CID can occur with PSD. 27 1.2.3.3 Photodissociation Ions may also be excited and subsequently dissociated by absorption of photons. Photodissociation (PD) techniques may be selective in comparison to the aforementioned techniques because only ions that absorb at the wavelength of the light used are activated. These techniques are often used with ion trapping mass spectrometers, because they allow for ions to be confined to the area irradiated by the light. 1.2.3.3.1 Infrared Multiple Photon Dissociation (IRMPD) 64, 131 Ion activation may be achieved using infrared lasers. Due to its relatively low energy (0.1 eV/photon), the absorption of multiple IR photons (tens to hundreds) are required for ion dissociation. Ions are typically irradiated for tens to hundreds of milliseconds for a continuous-wave 002 (10.6 um) laser. Like CID, IRMPD is a slow heating method and allows for intramolecular energy redistribution over all of the vibrational degrees of freedom prior to the next photon absorption event. As a result, ergodic dissociation of low-energy pathways predominates and the resulting spectra are often comparable with those of CID. Since the P-O stretch (9.6-11 pm) is in resonance with 10.6 um light, the phosphate groups of phosphorylated peptides provide strong chromophores for 77, 132 efficient dissociation by IRMPD; hence, IRMPD has been used for phosphopeptide characterization (Figure 1.2-4). Similar sequence coverage, 28 types of product ions formed (b- and y-sequence ions; nonsequence and sequence product ions resulting from phosphate neutral loss), product ion abundance (including the predominant formation of the nonsequence phosphate neutral loss ions) are formed by IRMPD as those formed by CID. Accordingly, selective detection of phosphorylated peptides from mixtures has been achieved by the observation of 80 and 98 Da neutral loss species upon IRMPD (Figure 1.2-4b).133‘136 However, due to the various mechanisms that may be involved in the cleavage of the phosphate group (Scheme 1.2), the same complications for phosphorylation site determination by CID and PSD, exist for IRMPD. 1.2.3.3.2 Ultraviolet Photodissociation (UVPD) Photodissociation of peptides in the UV range has targeted common chromophores such as amide bonds (193 and 157 nm) as well as residue- specific chromophores such as aromatic amino acids (220, 266, and 280 reviewed in 65 nm) Phosphopeptide analysis using ultraviolet photodissociation 1 (UVPD) has been performed using 220,137 157, 38 266 (after derivatization with a chromophore),139 and 193 nm light (Figure 1.2-5).140 One study by Dugourd and coworkers demonstrated that UVPD of peptides containing unmodified tyrosine (chromophore) and phosphorylated serine residues at 220 nm (5.64 eV/photon) result in the predominant production of radical product ions formed by homolytic cleavage and subsequent loss of twosine side chains.137 The formation of radical ions at this wavelength was to 29 some extent expected since nonphosphorylated tyrosine-containing peptides . . . . 141 . . . also give nse to Similar cleavages. When this Side chain cleavage was suppressed, the 98 Da neutral loss product ion was also formed predominantly.137 Kim and Reilly have recently shown that UVPD of singly protonated phosphoserine- and phosphothreonine-containing peptide ions at 157 nm (7.90 eV/photon) results in dominant formation of the neutral loss of 98 Da from the precursor ion, while phosphotyrosine-containing peptides largely retain the phosphate group.138 Interestingly, time-resolved detection of the UVPD product ions demonstrated that the nonsequence product ion resulting from the neutral loss of 98 Da was formed within 300 ns (the measurable limit) post photoexcitation, while product sequence ions (a-, b-, x-, y-, and z-ions) retained the phosphate group up to 1 us post photoexcitation.138 Additionally, a complete series of phosphate-intact sequence ions was formed for all of the model phosphopeptides examined under these conditions.138 Collectively, these observations demonstrate the potential of this strategy for phosphoproteome analysis since the observed 98 Da neutral loss nonsequence product ion may be used as a diagnostic indicator of the presence of phosphorylated peptides while the complete series of sequence ions retaining the intact phosphate group may be used to unambiguously assign the site of phosphorylation. Diedrich and Julian have developed an alternative UVPD approach involving conversion of the phosphate group to a UV-absorbing chromophore.139 30 This method involves B-elimination of the phosphate group followed by introduction of sulfhydryl-containing napthyl chromophores by Michael addition (Scheme 1.3).139 Upon UVPD at 266 nm of singly deprotonated or multiply protonated ions, the chromophore allows for enhanced absorbance of the formerly phosphorylated, labeled peptides. Furthermore, the carbon-sulfur bond connecting the label to the peptide is susceptible to direct photodissociation resulting in homolytic cleavage and subsequent loss of a napthylsulfide radical (Scheme 1.3).139 This resulting ion could be used as a diagnostic indicator of the presence of phosphopeptides. Sequence ions resulting from consecutive cleavage of the napthylsulfide radical and the peptide backbone adjacent to the former position of the label were also observed.139 Interestingly these ions were the only sequence ions observed in the product ion Spectrum. The production of these site-specific sequence ions was proposed to occur by the mechanism shown in Scheme 1.3, where the radical in the B-position (post napthylsulfide cleavage) induces backbone fragmentation adjacent to the labeled residue.139 Such Site-specific sequence ions could allow for unambiguous assignment of the phosphorylation Site within a grouping of potential phosphorylation sites. However, this method would incur the drawbacks of other previously described 8- elimination strategies. 31 Scheme 1.3 Analysis of napthyl chromophore-derivatized phosphorylated serine and threonine peptides by UVPD at 266 nm. B-elimination of phosphoserine and phosphothreonine peptides followed by Michael addition to form the napthyl- derivatized labeled peptides. UVPD of the derivatized peptides results in selective cleavage of the chromophore and radical-induced cleavage allows for the selective formation of sequence ions adjacent to the formerly phosphorylated residue. 32 Scheme 1.3 (can’t) a: $6 u E com .I u E v cozflcoEmm: 05053-96 6835-32an 0‘:- 621 5‘ I2 \=O + ‘__._ (1’. £21 8.. + ncon wO omm>mo_o nooznc_-_mo__omm Ec mom .0124 O I wI O I O I J) 2.9.. ‘ + 5) Z...) + J... 2.8. H coszzmzcoo H _ co=m£E=m R we» Fm rm 0 Fm onooz DO I. 33 Site-specific dissociation of phosphopeptides (non-derivatized) at the phosphorylated residues has also been recently demonstrated by Shin et al. using UVPD at 193 nm.140 In that study, phosphorylated peptide ions containing a basic residue (arginine, lysine, and histidine) at the N-terminus of the peptide resulted in the selective formation of intense an-97 ions C-terrninal to the location of the phosphorylated residues.140 This was demonstrated for singly protonated ions containing one, two, or three phosphorylations. Hydrogen—deuterium exchange experiments were completed and demonstrated that the an-97 ions formed from H3PO4 loss from the an+1 radical cation, which, in turn, is formed by Cot-Ccarbonyl cleavage of the peptide backbone (Scheme 1.4).140 Importantly, since these ions form selectively at the phosphorylation Sites, they could be used to localize the phosphorylation sites of peptides with multiple serine, threonine, and tyrosine residues. Additional product ions were also formed upon UVPD at 193 nm of these phosphorylated peptides, including nonsequence product ions resulting from the neutral loss of 98, 80, and 18 Da, and combinations thereof, as well as other sequence ions. Despite the potential for this technique for phosphorylation Site determination, this strategy may be limited by the choice of proteolytic enzyme available that would produce N-terminal basic residues. 34 NH O:é—OH NH Ozé—OH ’ O UVPD 193 nm / ‘~..O (CI-bk“ O Homolytic cleavage (CHZI3N O J‘\ H2N \(IkN .. F H2N j/IKN ' 0 R1 H o 0 R1 H Protonated phosphopeptide an+1 Scheme 1.4 Proposed site-specific an+1-97 ion formation of singly protonated, phosphorylated peptides that contain N-temiinal basic residues upon UVPD at 193 nm, where R1 is the side chain for any amino acid. 1.2.3.3.3 Femtosecond Laser-Induced Ionization/Dissociation (stID) Another photodissociation technique which has been applied to the analysis of phosphopeptides is femtosecond laser-induCed ionization/dissociation (stID; Figure 1.2-6).66 In stlD, the photons of a near-IR laser (A = 798i30 nm (~1.55 eV/photon)) are concentrated into ultra-short pulses (~35 fS duration) to enable very high photon intensities (~1014 W/cmz) for ion activation. With such a high-power femtosecond laser, novel activation processes may be accessible by stID. Additionally, since fSLID is capable of overcoming the threshold energy for ion oxidation (electron removal) on a timescale similar to or faster than that of bond vibrations (commonly 10-100 fs),142 it is just'rfiable to hypothesize that the technique can achieve non-ergodic dissociation enabling nonselective fragmentation and retention of labile functionalities, such as phosphate groups. In a recent study, Kalcic et al. have shown that stlD-MS/MS of singly and multiply protonated phosphorylated peptides (and nonmodified tryptic peptides) 35 results in the formation of ample product sequence ions, including those of the a- , b-, c-, x-, y-, and z-type.66 It was determined that many of these sequence ions were formed as a result of radical-driven dissociation of the photo-oxidized precursor.66 Importantly, this study showed that stlD-MS/MS analysis of phosphorylated peptides produces sufficient sequence information to enable unambiguous phosphorylation site localization. Although the mechanisms associated with both ionization and dissociation have yet to be elucidated, plausible ionization mechanisms for stID include multi-photon ionization, or field ionization where the electric field produced by the photon packet warps the ion potential energy surface to an extent that the ions may incur the loss of an electron. 1.2.3.4 Electron-Driven Dissociation Tandem mass spectrometry methods that involve dissociation by electron- ion reactions (ECD, EDD) or ion-ion reactions (ETD) as the “activation” component of the MS/MS experiment have also been used for phosphoproteomic analyses. These techniques largely maintain the phosphate group upon dissociation of the precursor ion and therefore overcome the potential problems associated with CID, PSD, and IRMPD. However, because most of these techniques involve charge state reduction by the addition or loss of an electron, multiply charged precursor ions are typically required. 36 1.2.3.4.1 Electron Capture Dissociation (ECD) Electron capture dissociation (ECD) involves exothermic capture of a low energy electron (<0.2 eV) by a positively charged species followed by dissociation of the resultant odd electron ion.143 Since the electron capture process requires low energy electrons (<10 eV) and long interaction times, ECD has traditionally been confined to instruments that employ static electromagnetic fields that avoid energizing or heating electrons, such as FT-ICRs. In particular, ECD was initially precluded for use in quadrupole ion traps since they utilize radio frequency (rf) electric fields with amplitudes of hundreds of volts that can either accelerate electrons greater than 10 eV or repel them from the region where the trapped ions are located. However, recently, the addition of magnetic fields to ion traps have allowed for ECD in such electrodynamic trapping instruments.144’146 Additionally, the use of ECD in a digital ion trap mass spectrometer has also been reported, where the application of a magnetic field was not necessary since this instrument employs a relatively constant electric field by rapidly switching between two dc voltage levels rather than using an rf tieiti.147 Protonated precursor ions (i.e., M+nH]n+) where n 2 2 are required for ECD-MS/MS because charge reduction of a singly protonated precursor ion would result in an undetectable neutral species. Furthermore, multiply protonated precursor ions formed by ESI are favored for ECD since the electron capture cross section is proportional to the square of the ion charge.148 37 Additionally, there have been reports of limited dissociation of doubly protonated precursor ions despite the expected high efficiency of electron capture.149’ 150 This has been rationalized by the potential for noncovalent interactions (e.g., salt bridges) that prevent the separation of the product ion pairs.149 To overcome such noncovalent interactions and increase the extent of dissociation, McLafferty and coworkers have introduced activated ion-ECD (Al-ECD) which involves ion heating before, during, or after the ECD process.151 Additionally, CID supplemental activation of the newly formed charge-reduced product ion may be performed simultaneous to the ECD experiment, thereby allowing for increased fragmentation efficiency and a reduction in the extent of multiple electron capture events that may otherwise be observed while performing ECD.152 The mechanism for bond dissociation by ECD is dictated by radical reaction chemistry. For peptides, this largely results in the formation of c- and 2- type product ions. A mechanism for this N-Ca bond cleavage has been proposed and is shown in Scheme 1.5,153 where a H. from the initial electron capture site is transferred to a backbone carbonyl group followed by cleavage to form c- and z-product ions. Importantly, ECD is a nonergodic process, occurring in a timescale faster than that of internal energy redistribution. Therefore, nonselective fragmentation is achieved rather than kinetically controlled fragmentation (like in CID, PSD, and IRMPD). For this reason, ECD of multiply protonated phosphopeptide ions results in retention of the phosphate group upon 38 154, 155 fragmentation. As a result, ECD-MS/MS has been widely used for phosphopeptide identification and unambiguous phosphorylation site assignment (Figure 1.2-7).156’ 157 + o H2N H250 H e- HfW , o __.. . (o H3N\'/II\N HaNj/IKN H R1 Fl C) R1 C) H2N H2N + OH + 0H m H3N\(§NH + H . ‘ H3N ,NV R1 0 1 H o c 2' Scheme 1.5 Previously proposed fragmentation mechanism for the production of c- and z-type product sequence ions for the reaction between a low-energy electron with a multiply protonated peptide ion. Since the phosphate group is maintained upon performing ECD and is lost upon performing CID, Cooper and coworkers have proposed and demonstrated the combined utility of these processes through automatic triggering of an ECD- MS/MS spectrum of a protonated precursor ion following the observation of an initial neutral loss of 98 Da upon performing CID (Figure 1.2-2b followed by 39 102,157 Figure 1.2-7). The success of this method is highly dependent on the abundant formation of the phosphate neutral loss product ion formed by CID to initiate the triggering event.157 Consequently, the factors that govern the abundance of the nonsequence 98 Da neutral loss product ion will greatly affect the utility of this method. Therefore, the work described in Chapter Two elucidates these factors. ECD of peptide polycations may also be achieved using high energy electrons (3-13 eV) in a method termed “hot” ECD (HECD).158 Similar to conventional ECD, capture of a hot electron by a peptide polycation induces N- CO. bond cleavage, producing c- and z-type ions (similar to that shown in Scheme 1.5). However, due to the excess energy of the captured electron, the resultant product ions may undergo secondary fragmentation. This method has been successfully used to distinguish leucine (Leu) and isoleucine (Ile), due to the abundant secondary fragmentation of their side chains from He or Leu-terminated 158.159 sequence product ions. An initial study demonstrated that HECD (11 eV electrons) of a phosphopeptide resulted in phosphate-intact c- and z-type product ions. In fact, the only product ions resulting from secondary fragmentation were product ions resulting from the neutral loss of 98 Da from amide bond cleavage ions (i.e., b and y-type ions).158 It was hypothesized that those product ions resulting from secondary neutral loss 98 Da formed via inelastic ion-electron collisions, Similar to collisions with neutral molecules for CID. 40 1.2.3.4.2 Electron Transfer Dissociation (ETD) Electron transfer dissociation (ETD) is similar to ECD in that it also induces nonergodic cleavage of the N-Ce bond on a peptide’s backbone to produce c- and z-type product ions, while maintaining phosphate groups and other potentially labile modifications.160 However, rather than the direct capture of an electron, ETD involves transfer of an electron from a radical anion reagent to a multiply protonated precursor ion. The use of reagent anions as electron donors makes ETD amenable for use in quadrupole ion trap mass spectrometers. Various anions with relatively low electron affinities (<60-70 kcal/mol), such as anthracene, fluoranthene, and azobenzene, have been successfully used as electron donors for ETD.16O'163 After electron transfer to a protonated peptide, the mechanism for dissociation of the N-Ce bond is thought to be the same as that for electron capture (Scheme 1.5).160 For these reasons, ETD has increasingly been used for phosphoproteome analysis (Figure 1.2-8), enabling phosphoprotein identification and unambiguous phosphorylation site . . 161,164-166 localization. The performance factors associated with ETD have been previously discussed by Good et al.162 In that study, the most Significant factor governing ETD peptide fragmentation efficiency was determined to be the charge density or the charge/residue ratio of a precursor peptide ion, where low precursor ion charge density gave rise to low fragmentation efficiency. Accordingly, less than 41 1% of the unmodified peptides identified using ETD were derived from doubly . 1 . protonated precursor ions. 62 In another large-scale analySIs, Swaney et al. similarly demonstrated that ETD is most effective for precursor ions with low m/z (high charge density).167 This apparent trend of decreased ETD fragmentation efficiency for low charge density precursor ions has been attributed to the formation of the non-dissociated (intact) charge-reduced radical cation (e.g., [M+2H]2+ -> [M+2H]+.).162’ 168 In order to increase the ETD efficiency of doubly protonated precursor ions and other peptide ions with low charge density, application of supplemental low-energy CID of these radical cations has been performed.168 By using this hybrid technique termed ETcaD, Molina et al. demonstrated that 60% of identified tryptic peptides were from doubly protonated precursor ions.169 Furthermore, it has been shown that ETcaD provides an improvement for nonmodified peptide identification (89%) over CID (77%) or ETD (63%) alone.168 In that study, ETcaD of doubly protonated phosphorylated peptide ions was also completed. Not surprisingly, the product ion formed from the neutral loss of 98 Da was the most abundant ion present and neutral losses of 98 Da from c- and z-product ions were also observed.168 As a result, ETcaD may incur the same limitations for phosphorylation site assignment as for conventional CID. Since both the charge and length of the peptide ion affects ETD efficiency, Molina et al. have studied how the chosen proteolytic enzyme affects ETD for 42 phosphoproteome analyses.164 In particular, they compared trypsin (cleaves C- terrninal to arginine and lysine), Lys-C (cleaves C-terminal to lysine), and Glu-C (cleaves C-tenninal to glutamic acid and to a lesser extent aspartic acid). Due to the cleavage specificity, it is expected that Lys-C would produce peptides with more basic residues (and therefore higher charge) than their trypsin counterparts, increasing the number of phosphopeptides identified using ETD. However, Molina et al. showed that Lys-C and trypsin performed similarly.164 This was attributed to inefficient trypsin digestion near phosphorylated residues, resulting in phosphopeptides with similar peptide lengths and protonation states as Lys-C peptides.164 Molina et al. also studied the effects of Glu-C for ETD analysis of phosphopeptides, since the average peptide length produced by Glu- C digestion is thought to be between that of LyS-C and trypsin and therefore produce peptides suitable for ETD.164 However, due to poor cleavage specificity and efficiency, an order of magnitude less phosphopeptides were identified using Glu-C as compared with LyS-C and trypsin.164 Heck and coworkers have shown that Lys-N may also be an appropriate protease for use in conjunction with ETD and ETcaD analysis of phosphorylated 170, 171 peptides. As the name suggests, Lys-N cleaves N-terminal to lysine residues and therefore produces peptides that are similar to Lys-C, but differs in the position of the lysine residue. Taouatas, et al. have showed that ETcaD of phosphopeptides produced by Lys-N digestion readily produce (N-terminal) c-ion 43 product ion series, corresponding to the retention of the ionizing protons on the N-tenninal basic residue,170'173 enabling facile phosphopeptide 170, 171 identification. Additionally, a recent study by Gauci, et al. showed that Lys-N and trypsin may provide complementary identification of phosphopeptides (albeit using CID), since Lys—N produces lysine-containing phosphopeptides, which were poorly represented in the phosphopeptides identified by the analogous trypsin digest.174 One limitation of ETD (as well as ECD) is that it is not amenable to the analysis of peptide anions. Moreover, phosphopeptides tend to ionize better in negative ion mode than in positive ion mode, due to their relatively low isoelectric points.175 Considering these seemingly opposing points, Gunawardena et al. have demonstrated charge inversion of singly deprotonated phosphopeptide anions with multiply protonated amino-terminated dendrimerS, to form polycationic precursors in the gas-phase for ETD analysis.176 This strategy was particularly beneficial because the phosphopeptides analyzed were either not present or present in low abundance in the corresponding positive ion mode MS 1 76 spectra. 44 1.2.3.4.3 Electron Detachment Dissociation (EDD) Electron detachment dissociation (EDD), another technique for electron- induced ion activation/dissociation, is particularly notable in that it allows for fragmentation of polyanions directly. EDD involves reaction of a polyanion with an electron (>10 eV), inducing electron detachment and resulting in a charge- reduced radical anion.177 Fragmentation of the radical anion is governed by radical reaction chemistry and produces a- and x-type product ions resulting from 150, 177 cleavage of the Ca-Ccarbonyl bond. To date, and Similar to conventional ECD, the application of EDD has been confined to FT-ICR instruments, where magnetic rather than electrodynamic fields are used. Like the other electron- driven methods, EDD is a nonergodic process, and phosphorylated peptides largely retain their phosphate group upon dissociation, allowing for sequence determination and accurate phosphorylation site assignment (Figure 1.2-9).150' 178 Kweon and Hakansson have demonstrated that despite its lower fragmentation efficiency, EDD may give rise to complementary sequence information to ECD for singly phosphorylated peptides.150 1.2.3.5 Metastable Atom-Activated Dissociation (MAD) _ Metastable atoms produced by either fast atom bombardment or a glow discharge source have also been used to impart energy to precursor peptide ions for fragmentation in a technique termed metastable atom-activated dissociation 70, 179-181 (MAD) or metastable-induced dissociation of ions (MIDI). Since the 45 beam of metastable atoms is neutral, MAD may be completed in quadrupole ion traps without causing problems relating to charge capacity. Misharin et al. showed that fragmentation of polycationic peptide ions with metastable helium or argon atoms in an ion trap gives rise to c- and z-sequence ions, similar to ECD and ETD, while fragmentation of polyanionic peptide ions by similar methods gives rise to a- and x-sequence ions, similar to EDD.7O As a result, two mechanisms for MAD were proposed: for cations, electron transfer from the metastable atoms; and for anions, electron detachment and de-excitation of the metastable atom (i.e., Penning ionization; Scheme 1.6), each giving rise to radical-induced non-ergodic peptide backbone fragmentation.70 Berkout and Doroshenko have demonstrated that singly protonated peptides also undergo fragmentation by interaction with metastable atoms, substantiating the contribution of a Penning ionization mechanism.180 Similarly, Cook et al. have recently demonstrated extensive backbone fragmentation may be achieved using MAD of both singly and doubly protonated peptides.181 Notably, product ions including a-, b-, c-, x-, y-, z-types, consecutive side-chain and backbone cleavage ions, and product ions resulting from through-ring cleavage of proline were observed, despite its relatively low fragmentation efficiency.181 Importantly, fragmentation of phosphorylated peptides by MAD results in the formation of sequence ions that maintain the phosphate groups at their native sites, demonstrating that MAD could be employed for phosphoproteome analysis (Figure 1.2-10).18°' ‘81 46 [M+nH]’.’* + m' -> [M+nH]("*1)"' + m +e' Scheme 1.6 Penning ionization of a protonated peptide (M) ion, where n 2 1, by interaction with a metastable noble gas atom (m*). 1.2.3.6 Comparison of Tandem Mass Spectrometry Techniques for Phosphoproteome Analysis Each of the tandem mass spectrometry methods described above differs in their effectiveness for the analysis of phosphopeptides. In particular, they differ by the instrument platform which may be used, the precursor ions for which they are amenable, and the types of product ions that are formed (sequence vs. nonsequence, phosphate neutral loss, phosphate group transfer, etc.). Consequently, not one technique has emerged as a universal method for phosphopeptide characterization. In fact, many studies comparing these tandem mass spectrometry methods have been completed and have shown that many of these techniques produce complementary information. For example, studies of nonphosphorylated peptides have reported that CID and ECD provide complementary sequence information, and that the application of both provides a higher confidence in sequence assignment than 154, 155 can be attained with either technique alone. Moreover, a recent large- scale comparison of ECD and CID of tryptic phosphorylated peptides by Sweet et al. Showed that although ECD gives rise to more sequence coverage, it is complementary to the information made available by CID.157 However, for the purpose of unique phosphopeptide identification, it was also found that CID 47 (34%) outperformed ECD (13%).157 This was attributed to predominant production of doubly protonated precursor ions from the analysis of tryptic peptides, which are poor targets for electron dissociation techniques but are generally considered to be optimal for CID. The ECD Spectra, however, gave rise to more confident phosphorylation site localization than the CID spectra.157 The results of a small scale study on synthetic “tryptic” phosphopeptides comparing CID and ECD using a hybrid LTQ-FT-ICR instrument also agreed with the study by Sweet, et al.182 Accordingly, combination techniques that use the strengths of both ECD and CID have proven successful for phosphoproteomic . . 25. 102. 156, 162 applications. Other efforts toward comprehensive analyses have employed the combination of IRMPD and ECD since, as with the CID/ECD pairing, they also give rise to complementary structural information and hence allow for higher confidence in peptide identification.183 Additionally, similar to the Al-ECD, CID+ECD, and ETcaD methods, IRMPD of the non-dissociated charge-reduced radical cation species formed by ECD can also be performed in order to improve sequence coverage.183 This technique has further been applied to a tryptic phosphorylated peptide from protein kinase A.149 Interestingly, no product ions resulting from the neutral loss of 98 Da were observed and, as a result, the PhOSphorylation site was easily identified.149 48 The complementary nature of CID and ETD has also been demonstrated in a large-scale study on unmodified peptides by Good et al., where there was a 12% overlap in peptide identification between CID and ETD spectra. Swaney et al. have Shown that more peptides are identified using CID, but with lower sequence coverage as compared with ETD.167 Additionally, a large-scale comparison of ETD and CID for phosphoproteomic profiling of human embryonic kidney 293T cells completed by Molina et al. resulted in the identification of 60% of the 312 phosphopeptides by ETD and only 27% by Cit).164 It should also be noted that although the product ions corresponding to the non-dissociated charge-reduced cations were typically much more intense than the c— and z-ions in the resultant ETD spectra, ETD was still better for phosphopeptide identification than CID. Interestingly, this same study resulted in the identification of more nonphosphorylated peptides by CID than ETD. Importantly, due to the low amount of overlap between the two methods (13%), the combination of ETD and CID data resulted in a greater number of phosphopeptides identified than could be determined from the individual techniques and 80% of the known phosphorylation sites were observed from the sample.164 Other complementary sequence information that may be achieved from CID, ECD/ETD, and MAD relates to peptide backbone cleavage adjacent to proline residues. For CID of protonated peptide ions, amide bond cleavage N- terminal to Pro may be a preferred cleavage site, while cleavage C-terminal to 184. 185 Pro is often not observed. For ECD/ETD, cleavage to produce the 49 corresponding c- and z-ions on the N-terminal Side of Pro requires scission of two bonds and its observation has, therefore, remained rare.186 This double cleavage has, however, been observed in other strategies such as MAD.181 Spectral information available from newly emerging techniques for phosphopeptide analysis such as UVPD, stlD, EDD, and MAD may also be complementary to that obtained by the more established techniques, ClD/PSD/IRMPD and ECD/ETD. This is particularly true when considering the analysis of precursor ions with low protonation states (e.g., 1 or 2). In particular, ClD/PSD/IRMPD of precursor ions with low charge states exhibit the predominant neutral loss of the phosphate group, limiting the amount of sequence information that may be observed; and limited dissociation efficiency of precursor ions with low charge state is achieved using ECD/ETD. UVPD, stlD, and MAD, on the other hand, are capable of ionizing low charge state peptides. Additionally, UVPD, MAD, and EDD enable the analysis of anionic phosphopeptides, which are not accessible by ETD or ECD due to a low reaction cross-section, resulting in sequence information and/or phosphorylation-site specific information. 1.3 Specific Aims A hypothetical ideal technique for phosphoproteome analysis could be described as: (i) globally applicable with regard to precursor ion composition, sequence or charge state, (ii) enables facile differentiation between phosphorylated and nonphosphorylated peptides, and (iii) provides unambiguous 50 sequence information for peptide identification and phosphorylation site assignment. Unfortunately, no existing technique has been demonstrated to fulfill all of these criteria Simultaneously. As a result, efforts to provide an improved understanding of the gas-phase ion chemistry involved in these existing methods, would enhance the utility of these techniques. For this reason, the major goals of the research described in this dissertation were: (i) to provide insights into the gas-phase ion chemistry of phosphorylated peptides, (ii) to consider the effect of these chemistries on the applicability of tandem mass spectrometry methods for phosphopeptide identification and characterization, and (iii) to apply those methods towards the phosphoproteome analysis of biologically relevant systems. These goals are discussed in the remaining chapters with regard to these specific aims: 1. Obtain insights into the role of proton mobility on the multistage gas-phase fragmentation reactions of protonated phosphoserine- and phosphothreonine-containing peptides under CID conditions. 2. Determine the mechanisms and product ion structures associated with the formation of nonsequence neutral loss product ions resulting from phosphate bond cleavage and to examine the influence of these structures on the appearance of the Spectra obtained by CID-MS3 of the neutral loss product ions formed by CID-MS/MS. 3. Evaluate the potential competing fragmentation pathways involving the phosphate group upon ClD—MS/MS, including the direct loss of H3PO4; 51 II consecutive losses of HPO3 and H20; and intramolecular phosphate group transfer. Determine the effect of these potential competing fragmentation pathways on phosphoproteome analysis by tandem mass spectrometry strategies. Determine the phosphorylation status of proteins involved in the NF-KB pathway, specifically p65 and its interacting proteins, in a differentially treated leukemia cell line. 52 CHAPTER TWO Mechanistic Insights into the Multistage Gas-Phase Fragmentation Behavior of Phosphoserine- and Phosphothreonine-Containing Peptides 2.1 Introduction The most common tandem mass spectrometry method used for phosphopeptide analysis is collision-induced dissociation (CID) tandem mass 15, 26, 80 Spectrometry (MS/MS). Under these conditions, phosphoserine- and phosphothreonine-containing peptides often undergo the facile neutral loss of Phosphoric acid (H3PO4, 98 Da) from their protonated precursor ions.81’ 83’ 127 The formation of these “nonsequence” ions can be particularly useful in providing diagnostic information regarding the presence of a phosphorylated peptide within a complex peptide mixture and may therefore be attractive for use as a 24. 98-102 diagnostic “target” for selective phosphoproteome analysis, as well as to improve the specificity of subsequent database search analysis strategies.187 188 However, when these ions are observed as the dominant species in the product ion spectrum, the formation of b- and y-type “sequence" ions may be The results described in Chapter Two have been published in: Palumbo, A.M.; Tepe, J. J. Reid, G.E. Mechanistic Insights into the Multistage Gas-Phase Fragmentation Behavior of Phosphoserine- and Phosphothreonine-Containing Peptides. J. Proteome Res. 2008, 7, 771-779. 53 suppressed. This can make it difficult to unambiguously identify the sequence of the protein from which the phosphopeptide originated or to localize the Site of phosphorylation to a Specific amino acid residue within the modified peptide. In such cases, alternate analysis strategies have been employed to obtain the required sequence information, involving automatically subjecting the initial neutral loss product ion to multistage CID-MS/MS (M83) or “pseudo M83“ in a 24. 98-101 data-dependent mode of operation. Alternatively, the loss of H3PO4 during ClD-MS/MS can be used to trigger MS/MS analysis of the original precursor ion by electron capture dissociation (ECD) or electron transfer dissociation (ETD), under which conditions the phosphate group remains . 102.160.161.164 intact. The utility of these data-dependent acquisition methods is highly dependent on the ability to form the initial [M+nH-H3PO4]n+ neutral loss product ion as an abundant species.24 Therefore, studies aimed at the development of an improved understanding of the mechanisms and other factors that influence the relative abundance of these ions, as well as determination of the structure(s) of the initial [M+nH-H3PO4jm product ions subjected to further dissociation by CID-M83, would enable the development of improved analytical mass spectrometry approaches for comprehensive phosphoproteome analyses. The results from these studies would also enhance ongoing efforts to develop in silico 54 spectral intensity prediction algorithms for use in enhanced database search . . 189-192 analySIS strategies. To date, it has been widely accepted that under low-energy CID-MS/MS conditions, the loss of H3PO4 from protonated phosphoserine- and phosphothreonine-containing peptides occurs via a B-elimination mechanism 87, 89 (Pathway A, Scheme 2.1). In this mechanism, the ionizing proton is distal from the site of fragmentation and is therefore referred to here as “charge- remote”. This reaction pathway would involve transfer of the hydrogen atom from the a-carbon on the phosphoamino acid to the phosphate oxygen on the modified amino acid Side chain via a six-centered transition state, resulting in the formation of dehydroalanine- or dehydroaminobutyric acid-containing product ions from phosphoserine- or phosphothreonine- containing precursor ions, respectively. Support for this mechanism, including the formation of product ions corresponding to cleavage on either side of the dehydroalanine (69 Da) or dehydroaminobutyric acid (83 Da) residues during MS3 of the initial [M+nH- H3PO4]n+ neutral loss products.89 Additionally, observation of the loss of HD2PO4 (100 Da) following CID-MS/MS of a deuterated phosphopeptide precursor ion,87 has been presented by several groups. However, the former observation could be accounted for by an alternate mechanism (see below), while the latter result must be considered inconclusive due to incomplete 55 hydrogen/deuterium exchange of the precursor ion and the relatively low resolution of the mass analyzer employed at that time. The former study also reported that the loss of H3PO4 from phosphothreonine-containing peptide ions was highly dependent on the Charge state of the precursor ion (i.e., that low charge states underwent a dominant loss of H3PO4, while the highest observed charge state tended to retain the phosphate moiety and instead undergo extensive fragmentation along the peptide backbone), while the same loss from phosphoserine-containing peptides was independent of the precursor ion charge state.89 However, this conclusion was reached based on consideration of the effect of charge state alone and therefore did not account for the combined effect of precursor ion charge state and the number and composition of basic amino acids in the peptide sequence (i.e., proton mobility) on the observed product ion abundances.112' 193 Two alternate mechanisms have Since been proposed for the side chain neutral loss of H3PO4, based on results obtained by molecular orbital calculations and from the gas-phase fragmentation reactions of uniformly deuterated phosphoserine and model phosphoserine peptide analogues.78’ 117 The first of these involves an E2 elimination reaction, via intramolecular transfer of the hydrogen atom from the a—carbon on the phosphoamino acid to an adjacent nucleophile, resulting in the formation of either dehydroalanine- or dehydroaminobutyric acid-containing product ions (Pathway B, Scheme 2.1). 56 (Note that the site of protonation for the product ions formed by this pathway would differ from that for the B-elimination mechanism shown in Pathway A of Scheme 2.1.) The second mechanism involves an SN2 neighboring group participation reaction, involving intramolecular nucleophilic attack by the N- terminal amide carbonyl group on the [3-0arbon of the phosphoamino acid side chain, resulting in formation of a cyclic five-membered oxazoline product ion (Pathway C, Scheme 2.1). In both of these mechanisms, the ionizing proton is directly involved in the fragmentation reaction, and they are therefore referred to here as “charge-directed". 57 (A) HO\ ,OH H- (B) +OH +9” HO-P-OH R (‘o 0‘, H 1,.an N DI “h H O (C) H+ CID MS/MS ‘ o > H -H3PO4 pJLNJk’rN.“ H o R CID MS/Ms> 0“ I H -I'I3PO4 ' rpkfi NH"! H o CID MS/MS R > O -H3PO4 R = H (p8). CH3 (pTl Scheme 2.1 Potential mechanisms for the neutral loss of H3PO4 from phosphoserine- and phosphothreonine-containing peptides via (A) a charge- remote B-elimination reaction, (B) a charge-directed E2 reaction, and (C) a charge-directed SN2 neighboring group participation reaction. mechanisms are presented for phosphoserine-containing peptides only. Several reports in the literature indicate that proton mobility plays an important role in the gas-phase fragmentation reactions of protonated peptide ions containing posttranslational or process-induced modifications. Furthermore, the appearance of product ion spectra obtained by MS3 58 For clarity, the 15, 116 dissociation of the neutral loss product ions formed by MS/MS is expected to be highly dependent on both the proton mobility and the structure(s) of the neutral loss product ion. The purpose of the work described herein, therefore, was to obtain insights into the role of proton mobility on the multistage gas-phase fragmentation reactions of protonated phosphoserine- and phosphothreonine- containing peptide ions, to determine the mechanisms and product ion structures associated with their formation and to examine the influence of these structures on the appearance of the Spectra obtained by MS3 of the neutral loss product ions formed by MS/MS. 2.2 Results and Discussion 2.2.1 Evaluation of Charge-Remote and Charge-Directed Fragmentation Mechanisms for the Loss of H3PO4 from Phosphoserine- and Phosphothreonine-Containing Peptides via Multistage Tandem Mass Spectrometry Analysis of Their Regioselectively and Uniformly Deuterated Precursor Ions Differentiation between charge-remote and charge-directed fragmentation pathways for neutral loss of the side chains from modified peptide ions can be readily achieved via examination of the CID-MS/MS product ion spectra of regioselectively or uniformly deuterium-labeled peptide ions.115'117 For example, the product ion spectra obtained by dissociation of the singly and doubly protonated precursor ions of the regioselectively Side chain deuterated 59 phosphoserine-containing peptide, GAIL(d3-D,L)pSGAILK, are shown in Figure 2.1A and B, respectively. The product ion spectra obtained from the nondeuterated form of this peptide, containing all L-amino acids, are shown in Figure 2.2. Loss of the phosphorylated side chain from the regioselectively deuterated peptide via the previously proposed charge-remote B-elimination mechanism shown in Scheme 2.1, Pathway A, would result in the loss of HzDPO4 (99 Da), whereas the loss of H3PO4 (98 Da) would result from the charge-directed mechanisms (Scheme 2.1, Pathways B and C). Clearly, it can be seen from Figure 2.1A and B that the ion corresponding to the neutral loss of H3PO4 (98 Da) was the major product observed from dissociation of both the singly and doubly protonated precursor ions, indicating that the fragmentation process had occurred predominantly via a charge-directed mechanism in each case (approximately 97% and 98.5% of the neutral loss product ion abundance, respectively). An expanded region of the product ion spectrum is shown in the inset Figure 2.1A and reveals that a product ion resulting from fragmentation via the charge-remote pathway (loss of HzDPO4, 99 Da) was also observed, albeit at a level representing only 3% of the neutral loss product ion abundance. Although the presence of a similar charge-remote product ion could not be observed by direct inspection of the doubly protonated product ion spectrum (presumably due to limited mass analyzer resolution caused by local space charge effects), evidence for the formation of this product at a level of approximately 1.5% of the neutral loss product ion abundance was obtained by MS3 (see below). 60 Figure 2.1 Multistage tandem mass spectrometry of regioselectively deuterated GAIL(d3-D,L)pSGAILK. CID-MS/MS of the (A) [M+H]+ and (B) [M+2H]2+ precursor ions. CID-MS3 of the (C) singly and (D) doubly protonated [M+nH- H3PO4]I|+ neutral loss product ions from panels A and B, respectively. CID-MS3 of the (E) singly and (F) doubly protonated [M+nH-H2DPO4]n+ neutral loss product ions from panels A and B, respectively. A = -H3PO4, o = -H20, * = -NH3. 61 Figure 2.1 (cont’d) NE 89 _ BS _ owe _ 2.8 _ owe _ omm _ _ ._ _. .2 4. 2:1 _m. .. .. .. . . a a... £3 Erma .. can? Mmswnmwfi .. ...4 a 4a .. ma Q q ®>< x...» ..... N> a. +Nw> . a - m. +N od I awe? loo? NE 8m. 82 8m 80 84 _ _ _ . ._. m _. _ a W _ N - as. i... ka m.» o I < ma < .w 4. F < 4 n Ao~z+eodmzv - e8 ewe m8 _ _ ,- J, U - <. _ \ ....... m i z A v m r . or on. I aonawz. fie amend eouepunqv eAlieleu % C ‘- 62 NE 63 Figure 2.1 (cont’d) ooN _. ooo F oom ooo oov ooN ooN _. ooo_. _ _ — _ I— _ F — _ _ _ _ _ _ _ . i —tl at I I q a i a a . 0 _ e ___ x m _ / ma... ea .. § .. of- P m ma W. mi. . r A n.» .. E» e W+N on W . I n U a w +NO I l . T u If. 0 a... a o> .I._. -2: 2 mm. ..o _. ooNF ooo r ooo ooo oov ooN ooN r ooow _ _ _ h _ t. i w... «p i. t. _ a JP. _ _ p .0 ...lm. mm.» a“... max. Wm o/o a...s.......s..a. a o . n .. a . >1. ...r n o> .. ...+N . B. as .... no ma .. m .. F I. N VD > i A V a. r n u D. B I r w S o 0 a [CO—s [00F vmomé VMmoN Figure 2.2 Multistage tandem mass spectrometry of GAILpSGAILK. CID- MS/MS product ion spectra of the (A) [M+H] and (B) [M+2H]2 precursor ions. CID-MS3 product ion spectra of the (C) singly and (D) doubly protonated [M+nH- H3PO4]'1+ neutral loss product ions from panels A and B, respectively. A =- H3PO4, ° = -H20, * = -NH3. 64 Figure 2.2 (cont'd) NF: N\E com com 8.. 8m 83 89 com _ _ — _ _ . _ _ _ _ _ _ . _ . 7.... . _.: .. m _ . O m w m E. fl o/ n n 3 m - . - a w> n ma M> .<. m w> Wu- +~ m n. +NONI1 l 0 I1... I m 3.8090- “now 1 1 u p B n> 1 u m m w an» O. -02 n n -2: mm? F r 8m 80 8.. 8m 8m. 82 08 _ _ _ _ _ _ — _ _ ._ . «_w — p. .2 _. __fi 5. ‘ ___ N- a; ..__ % O I .. m < Mm.» m9 .1 o o P r» N» F w .9. .. < < a n AoNIEonIV- v < - - w. w o w> u a m b» > +N m 33.0%: o v m - -oS . - roe 2.5+... on_ I $3.0 Ex: 65 Importantly, regioselective stable isotope labeling allows for the gas-phase separation of isobaric product ions formed from different reaction mechanisms, thereby allowing differences in the structures of these MS/MS product ions to be independently interrogated using M83. Isolation and MS3 dissociation of the [M+nH-H3PO4]n+ product ions from both the singly and doubly protonated product ion spectra in Figure 2.1A and B resulted in the product ion spectra shown in Figure 2.10 and D, while the product ion spectra obtained by M83 dissociation of the [M+nH-HzDPO4]n+ ions are shown in Figure 2.1E and F. Clearly, the product ions formed in each case, and their relative abundances, are quite different, indicative of the different structures of their precursor ions. Notably, the product ion spectra shown in Figure 2.1E and F are essentially identical to those previously obtained by dissociation of the [M+nH-CH3SOH]n+ product ions of an S-alkyl cysteine sulfoxide containing peptide GAILC(Sme(ox))GAILK, where the charge-remote B-elimination mechanism was shown to be the dominant process.112 Interestingly, MS/MS dissociation of the singly and doubly protonated dehydroalanine (dA)-containing precursor ions of GAIL(dA)GAlLK, produced in solution via B—elimination from its phosphoserine analogue (data not shown), gave rise to significantly different product ion spectra than those observed by MS3 of the [M+nH-H2DPO4]|n+ product ions formed in the gas-phase. As the sites of protonation in these ions are expected to be similar, the differences in 66 fragmentation may be due to differences in the conformations of the precursor ions formed in solution versus in the gas-phase. This result therefore highlights the potential limitations associated with obtaining evidence for gas-phase fragmentation mechanisms by comparison of MS3 fragmentation behaviors with the MS/MS fragmentation behavior of proposed product ions formed from solution. The differences in the MS3 data obtained from dissociation of the [M+nH-H3PO4]n+ and [M+nH-HZDPO4]n+ product ions from the regioselectively deuterated peptide also rule out the possibility that the charge-remote fragmentation pathway was the dominant mechanistic process occurring, but that the resultant [M+nH-HzDPO4]n+ product ion and the neutral HzDPO4 molecule had undergone deuterium exchange prior to their separation from the ion— molecule complex initially formed upon MS/MS, thereby resulting in formation of the [M+nH-H3PC)4]n+ ion. Furthermore, ClD-MS/MS of the singly protonated GAIL(d3-D,L)pSGAlLK peptide ion using a range of activation times (30 ms to 30 us at an activation Q of 0.75) was observed to result in no significant change in the ratio of the [M+H-H3PO4]+ and [M+H-H20PO4]+ neutral loss products (data not shown), again suggesting that no deuterium exchange had taken place. Further evidence for the neutral loss of H3PO4 predominantly via a charge-directed mechanism was obtained by ClD-MS/MS of various uniformly deuterated phosphoserine- and phosphothreonine-containing peptides, including some that had previously been examined in mechanistic studies of 67 phosphopeptide fragmentation161’ 164 (Table 2.1). Under these conditions, the losses of D3PO4 (101 Da) and HD2PO4 (100 Da) would be expected for the charge-directed and charge-remote mechanisms, respectively. The product ion spectra obtained by ClD-MS/MS of the singly and doubly protonated precursor ions of the phosphoserine-containing Kemptide peptide, LRRApSLG, and the phosphothreonine-containing Kemptide analogue, LRRApTLG, are shown in Figure 2.3 and Figure 2.4, respectively. Monoisotopic isolation and dissociation of the uniformly deuterated singly [M(D19)+D]+ and doubly [M(|;)19)+ZD]2+ charged precursor ions of LRRApSLG both resulted in the dominant loss of D3PO4 (80% and 82% neutral loss product ion abundance, respectively). This is indicative of fragmentation occurring via the charge-directed fragmentation mechanisms, whereas the abundances of product ions resulting from the charge- remote fragmentation mechanism, via the loss of HDzPO4, were observed at only 20% and 18%, respectively. Importantly, ClD-MS/MS of the singly and doubly charged precursor ions from all the uniformly deuterium labeled peptide ions studied to date have resulted in the dominant loss of 03PO4, unequivocally demonstrating that the loss of H3PO4 from protonated peptide ions occurs predominantly (>80%) via charge-directed mechanisms (Table 2.1). As mentioned in the Introduction section (2.1) above, previous molecular orbital calculations have predicted that the activation energy of the transition state structure associated with the charge-remote B—elimi'nation mechanism is higher 68 (i.e., less thermodynamically favored) compared to those for the charge-directed 78, 11 8N2 or E2 processes, 7 thereby supporting the experimental results observed here. Finally, no significant differences in the ratio of charge-directed versus charge-remote fragmentation product ions were observed between phosphoserine- and phosphothreonine-containing peptides or as a function of the identity or number of basic amino acid residues within the peptide sequence (arginine, lysine, and histidine). 69 Table 2.1 Summary of the MS/MS and MS3 neutral loss product ion abundances from the gas-phase fragmentation reactions of protonated phosphoserine- and phosphothreonine-containing peptides. a “Partially mobile” b c d proton.193 “Mobile” proton.193 “Nonmobile” proton.193 Results obtained by ClD-MS/MS of uniformly deuterium labeled peptide ions. e Note that these values differ from those shown in Figure 2.1 for the regioselectively deuterated GAIL(d3-D,L)pSGAILK peptide and are likely due to either their acquisition on a different instrument platform (LCQ-DECA versus LTQ), the effect of the D,L- versus L-phosphoamino acid, or deuterium isotope effects on the charge-remote versus charge-directed fragmentation mechanisms. Results obtained by ClD- MS/MS of protonated peptide ions. 9 Calculated as the ratio of % total product ion abundances of the [M+nH-H3PO4]n+ ions from isomer phosphopeptide ions (e.g., GAILpSGAILK versus GAILpTGAILK). h Results obtained by CID-MS3 of [M+nH-H3PO4-CHZO]n+ (pSer) or [M+nH-H3PO4-CH30HO]n+ (pThr) product ions initially formed by MS/MS. 70 Table 2.1 (cont’d) UoEEQoD - E 2 3 ~+ o n . 2 529320 8 - ow 2 5 or. 855.98 8.. 2 n 8 ~+ 62 a a «:28 .56 R :4 8 2 mm o: m m Em... o - . . e .N. 529320 e - 8 2 am «I nogfiggoo hm. F F F oNF owm nw+ £28320 m X! mu m2 05 a? 810 u m EB E._.&._oma _eOamI _vOamoI _vOamo EV 39w moccacom maroon. . +: +c +: GENSO .I u 1 $an -155: -153: .155: c o. 255.. Mm”... flows .83.... -..onIiSZ no so coocmcc3< moocmbcznlq. hoocmucsnd. co. co. :2 boocmocan< Hosted Eaten. : o._ so go: 9 So o .38 e. .93 .s We“ om. _ o n. _ e \o .6 25m 0 71 Table 2.1 (cont’d) m: chm m _o _uoz. o - 9. 3 8 a? xmfiamazotu F - w... m. 8 a: 855.28 o: em 9 mm m~+ “oz xmumamazotu to t; mm a 2 m: N - ...m E 2 om+ 925.5 N - E u. 8 o: F 8.. 2 2 mm om.- 925.5 Nd mo; co om cm 05 810 u 1 EB Ekatowa _vOamI _vOdNQI _voamo A5 99m 8533 onion. . +c +c +c 09650 .1 u 1 5mg -155: -155: -155: co. E819. emcee 8.8% 5282a sax-:55: -m .20 -0920 coocmnczn< moocmuczn< mocmvcana. co. co. ._ co. mocmcczn< “eaten «035.5 0: 0.. :0 o: 0.. m o o .985 .9055 a can. : a 2:5 .93 5 .0 8mm 72 N5: .mmm — 0mm _ “mm — om” - mm” b . 5 0 +~ 0%0: 9)sz pm”.Av _wmv —hm—.v mom-V pmm—v_ : —mm—.v —wm—.v —Nm-v —©m-v —mm—¢_ c0330."... -1<:< ms. 0 NE. com. owe own on» one cam one can cow - . _- m Ill/m .12.. ..1... o +~< n §E+~v0mm¥ m Em: SW. 2.5 ow» 2.: 2.0m 8m 2.: own 0N:- m». m an . ma > H/ < < f. Amzz+v0amIY 3.810%..- < . 88.... .09 oo_. N5: 0.55 . wmh . th . or: . mm: .me. V N - Cd 01 £692 2% p... a... .51.... 1 530.00. www. 0mm . Vmw. finw. 0mm. awn 4. ms. GOUBPUDQV GARB|GH °/o Figure 2.3 ClD—MS/MS product ion spectra of protonated and deuterated 2+ , (o) [M(019)+D]+, and (o) [M(D19)+ZD] 2+ LRRApSLG. (A) [M+H]+, (B) [M+2H] ions. A = -H3PO4. 73 Figure 2.4 Multistage tandem mass spectrometry of LRRApTLG. CID-MS/MS product ion spectra of the (A) [M+H]+ and (B) [M+2H]2+ precursor ions. CID-M83 product ion spectra of the (C) singly and (D) doubly protonated [M+nH-H3PO4]n+ neutral loss product ions from panels A and B, respectively. A = -H3PO4, * = - NH3. 74 Figure 2.4 (cont’d) N5: NE. 88 08 o8 2: 88 8o 8e 88 8N 82 8o 08 2K 08 88 8e 8” _ _ _ _ _ _ _ _ I. _ _ _ _ a .4- 11m4 .u u a - . 4. _ m.» % s. , .. W... - m .1 ........ f... M m ..x... on N. we .. +9. - - e .. t... +Noo $30585- W m .. r n 0st 0I0 I0- .. .. u +N 00 x D. +Ns mg 9 +N - r U m m - D. 82 Iz 0. -2: mm? o $83 82 o8 08 8s 88 8m 8e 88 8N 82 8o 08 08 88 8m 8e 28 _ F P _ _ _ {I _ _ _ _ _ b _ _ _ _ e. we .- 1-... a - -.e_. mm /o e N - o < - 0 I q o - a .. If}. B J.......... (1..-r... . far... my ..... ... N. +N on AoNI+e0omIY A: 1 A58: Ao\om: +N¢OamIr m 1 00—. VOQmIr < r 00.. 88.. BEN SOUBpUl'qu GAIIBBH °/o l O O F 79 eouepunqv eAnelea % Figure 2.6 Multistage tandem mass spectrometry of GAllipTGAILR. ClD- + + MS/MS product ion spectra of the (A) [M+H] and (B) [M+2H] precursor ions. CID-MS3 product ion spectra of the (C) singly and (D) doubly protonated [M+nH- H3PO4]n+ neutral loss product ions from panels A and B, respectively. A =- H3PO4, * = -NH3. 80 Figure 2.6 (cont’d) N5: N5: cow coo com com P coo? com com oov _ _ _ p _ _ _ _ _ L _ _ 1L.||_ . . . - . _- :- .l. 4.- 1111:1141...- .11.. 0% GD GD... ND .... N1 MD M _ ‘ d... 0/0 . OD > ”at an mn- 1 H 0> an MwIZt 02 ND 5 e N - - N. O I- e :4» +N 5.» V O. - - n U D. B - r U m +N D .- OOV Ao\ommV 0100—1—01 o .- 00—. mmeod omwod com com com com _- ooo F com com oov _ III-1L||[LI_I|LI. _ _ _ . _ _ _ _ _ - N lat—.1 no N.- .......-.- _ % <5 o5 eeo J - 0 I- AowI+e0oNIY m» - w _ < e No 5 M m5 - r w n ma < r - U we < N.» w b» . N r r U q es e m A$oFV+Ne0oNI- m .22 A580 0.. I- < -o2 mumvd nwnmd 81 Figure 2.7 Multistage tandem mass spectrometry of GAILpTGAILK. CID- MS/MS product ion spectra of the (A) [M+H]+ and (B) [M+2H]2+ precursor ions. CID-MS3 product ion spectra of the (C) singly and (D) doubly protonated [M+nH- H3PO4]'1+ neutral loss product ions from panels A and B, respectively. A =- H3PO4, ° = -H20, * = -NH3. 82 Figure 2.7 (cont’d) mmwwd NE. N5: 88 08 8e 08 8N2 82 88 C8 8... _ -. .ILl: _d J...¢ -—qw — _ . __ _ 5... . I /.. m. m . w J. .. 9...... m... e...- f I .3...) . ...e o w . o .1 - . B - ...o ... ! +N0NI- +NN» .. o - 0 I m .9... o - m §80I0 I0- so we v o o O. N - - m . w - no - U 0 9 o +NN.» D .22 c we 0. -o2 $8.. $80 o8 o8 8e 8N 82 82 o8 8e 8e — _ _ _ _ _ L — — p _ F _ _ _ _ 1 d .1 u 1 J 4 4 4 _ _ 0NI-_ _. m» N.» e»... _ % N.» quo u m O ONI+mn o m V o 1 m EH 0- o - w m 09.8; o m 0.» ONI- Io/m > 0 roow $8.: 8m: 8m: 8: o8 8N 8m 08 8m: 8m: 88 _l _ _ _ _ _ _ _ _ . E .1 . _. _. d . l . . . o; 8N an .._ Q. m, «a _ % 9 a < a M; - of- , w 0.» N... .. +N 9 Pl. +NO II . +N< l I m V “a I N v m - r U m +~6N1+e0mm$ 6 :+ on. Iv w m - - w .6. 9 w §va+mvoam¥ m -8: Agomvvommz- < r8: m Nmad 885 b F 85 Figure 2.9 Multistage tandem mass spectrometry of LFTGHPEpTLEK. CID- MS/MS product ion spectra of the (A) [M+H]+ and (B) [M+2H]2+ precursor ions. CID-MS3 product ion spectra of the (C) singly and (D) doubly protonated [M+nH- H3PO4]r|+ neutral loss product ions from panels A and B, respectively. A =- H3PO4, * = -NH3. 86 NNE NNE oom F oom F 00 F F com com com oomF oomF 09 F oom con com — _ P _ . _ _ _ _ h .. ..I. ._ _.. ..1 .J ..1: .1 _ .. . 4. _ . _. w _...._..Nn A a. _ IN % N o: o . ...n... .0 Sn... N - e .. 8N \H, a o m. .. Z 9» r e + GD 0- m OFD I n U 8 D. O B N H l u o I- m 0N m 0> D rooF ONIN- ONI+o Fe O rooF 88 N 3° 3 0:090. 88 : com F oom F oo F F com com com com com F oom F 00 F F com 005 com — _ _ _ _ _ _ F _ _ _ _ _ . .._..... a. n _ n _u a . 4 _ ....x m”. _ _ A d 1‘ — 0 > < o I mo .. n r < < u a N oh Q OmI- P < +N 3 m +N< N. N. - - a +NO I- W \I/ .- ..,.. I n HO .. .x. W: m N v m . w +N6 1+ 0m IY- SNIEOINIY - w W e Z v m - v. m - e Aamsa cm I m. .8F Isms on. I <. .8F m swan F owFN N .mu F 87 (Li . _‘ r 1 Sc Considering that the loss of H3PO4 proceeds predominantly via charge- directed mechanisms (where an ionizing proton is directly involved in the fragmentation reaction), we propose that the neutral loss of H3PO4 as an increasingly dominant product ion under nonmobile proton conditions is due to the formation of a strong hydrogen bonding interaction between the phosphate group and the guanidine side chain of a protonated arginine residue (shown in Scheme 2.2, Pathway 1 for the 8N2 neighboring group participation reaction mechanism). The presence of this hydrogen bonding interaction would be expected to lower the barrier for transfer of the ionizing proton from its initial site of protonation (the side chain guanidine group of the arginine residue, proton affinity 251.2 kcal/mol)194 to the side chain of the phosphoserine or phosphothreonine residues (proton affinity 224.2 and 224.8 kcal/mol, respectively)195 either prior to or following collisional activation of the ion, thereby facilitating cleavage of the B—methylene-phosphate oxygen side chain bond. This hypothesis is consistent with a recent report describing the important role of phosphate-arginine side chain interactions in the stabilization and gas-phase . . . 107, 108, 196 fragmentation reactions of noncovalent peptide complexes. 88 H H 11+“ A. H H’ (TN H’ Hc‘)“ f‘,- H’ ,N H (Pathway 1) NH HO- p: 0' OH CID MS/MS_ OH RG3 HN -H po 7 R HN 3 4 o A O H A O i'; NM‘PEPTIDE o HAW Nmpspnoe o L U 0 HF (Pathway 2)lH:n’N Transfer MT: 0 < -RCHO o ”N 00 \J ANN N-MPEPTIDE o #1:th 7” NMPEPTIDE'JL " OH R = H (pS), CH3 (pT) Scheme 2.2 Proposed mechanisms for the dominant loss of H3PO4 via a charge-directed 8N2 neighboring group participation reaction (Pathway 1) and for the loss of formaldehyde (CH20, 30 Da) or acetaldehyde (CH30HO, 44 Da) from the [M+H-H3PO4]+ product ions (Pathway 2) of protonated phosphoserine- and phosphothreonine-containing peptide ions. Although the peptides described here represent only a very limited data set, the results described above indicate that the magnitudes of the H3PO4 neutral loss from phosphoserine-containing peptide ions are consistently higher compared to those from their analogous phosphothreonine—containing peptide ions (with an average ratio of 1.28) (Table 2.1). These results therefore highlight the subtle but appreciable effect of the fi-methyl group on the phosphothreonine residue, compared to the B—hydrogen atom on the phosphoserine side chain, on the specificity of the gas-phase fragmentation reactions. This is presumably due 89 to steric hindrance effects on the transition state activation barriers for the respective fragmentation reactions. 2.2.3 Experimental Evidence for the Neutral Loss of H3PO4 from Phosphoserine- and Phosphothreonine-Containing Peptides via an 8N2 Neighboring Group Participation Reaction Mechanism. Experimental evidence for loss of the phosphoserine and phosphothreonine side chains via the 8N2 neighboring group participation reaction mechanism (Scheme 2.1, Pathway C) was obtained by inspection of the product ion spectra resulting from CID-MS3 of the [M+nH-H3PO4]n+ ions initially formed by MS/MS. From these spectra (Figure 2.1 to Figure 2.10), product ions corresponding to the loss of 30 Da from the phosphoserine-containing peptide ions or the loss of 44 Da from the phosphothreonine-containing peptide ions were observed. As the pairs of analogous phosphoserine and phosphothreonine peptides examined in this study (e.g., GAILpSGAILK and GAILpTGAILK) differ only by the side chain of the phosphorylated amino acid (a 14 Da difference), the losses of 30 and 44 Da (a 14 Da difference) are likely to involve the side chains of the serine and threonine amino acids. Thus, we have assigned the loss of 30 Da from the [M+H-H3PO4]+ ions of the phosphoserine-containing peptides as corresponding to formaldehyde (CH20) and have assigned the loss of 44 Da from the [M+H-H3PO4]Jr ions of the phosphothreonine-containing peptides as 90 corresponding to acetaldehyde (CH3CHO). Note that the loss of 32 Da (CD20) was observed upon MS3 dissociation of the [M+H-H3PO4]+ product ion from the regioselectively deuterated peptide GAIL(d3-D,L)pSGAlLK (Figure 2.1C), consistent with this proposal. A proposed mechanism for these aldehyde losses, via intramolecular proton transfer and subsequent fragmentation resulting in the formation of a nitrilium product ion, is shown in Scheme 2.2, Pathway 2. Evidence can be found in the literature indicating that nitrilium product ions can be formed as stable product ions in the gas-phase, via dehydration involving the carbonyl oxygen atoms of amide bonds contained within protonated peptide . 197 . . . . Ions. Despite extensrve consuderatlon, we were unable to propose a reasonable mechanism for these losses from the [M+H-H3PO4]+ ions initially formed via the E2 elimination pathway. We note that the ion trap instrumentation employed in this study does not have the requisite mass resolution and mass accuracy to distinguish between the losses of CH3CHO (44.05256 Da) and 002 (44.0095 Da), so we can not rule out that the loss of 002 does not also occur to some extent from the phosphothreonine-containing peptides. However, as this loss would also be expected from the analogous phosphoserine-containing peptides, we consider that this process was unlikely. Interestingly, the proposed 8N2 mechanism for the loss of H3PO4 that results in the subsequent loss of the aldehyde upon M83 (Scheme 2.2) may also be used to rationalize the consecutive side-chain losses of H20 (18 Da) and CzH4O (44 Da) recently 91 observed for the fragmentation of protonated tryptic peptides containing N- terrninal Thr-Thr or Thr-Ser residues under mobile proton conditions in a triple quadrupole mass spectrometer.198 Additional fragmentation processes observed upon MS3 dissociation of the [M+nH-H3PO4]n+ product ions from the LFTGHPEpSLEK and LFTGHPEpTLEK peptides (Figure 2.8 and Figure 2.9) included the formation of unexpected product ions corresponding to bn+H20. Notably, these “hydrolysis"-type product ions have also been observed previously from the M33 fragmentation reactions of both nonphosphorylated and phosphorylated peptide ions.4o’ 99’ 199' 200 92 _s C C % Relative Abundance _L O O % Relative Abundance Figure 2.10 Similar to that described above for the neutral loss of H3PO4 upon MS/MS 1.63E6 -NH3 ' -CH20(0._2%) _ 5 b6 I l . I '1; 1‘]. I l I I I I I I 300 400 500 600 700 800 900 1000 1.44E6 *2+ _ b6 2+ b 2+ - 5. $9.6 b433, -CH202*<1%> b3 ‘ 1 Al A “‘3'; if]. _- —l—_l— I I I I I I I 200 300 400 500 600 700 800 900 1000 m/z CID-MS3 product ion spectra of the (A) singly and (B) doubly protonated [M+nH-H3PO4]n+ neutral loss product ions of LRRApSLG from Figure 2.3A and B, respectively. * = -NH3. of the protonated phosphoserine- and phosphothreonine-containing precursor ions, the abundance of the CH20 or CH3CHO neutral loss product ions upon 93 MS3 of the [M+nH-H3PO4]n+ ions was observed to increase with decreasing proton mobility (Table 2.1). However, the abundance of the CH3CHO neutral loss product ions from the [M+nH-H3PO4]n+ ions of phosphothreonine-containing peptide ions was consistently higher than that for the loss of CH20 from the analogous phosphoserine-containing ions (Table 2.1). Note that the observation of low-abundance product ions corresponding to fragmentation of the amide bond on the N-terminal side of the phosphoamino acid side chain upon MS3 of the [M+nH-H3PO4]n+ product ions from the GAIL(d3-D,L)pSGAlLK peptide (e.g., the ya and b4 product ions in Figure 2.1C and D, respectively) is suggestive of the presence of an acyclic dehydroalanine residue and could be used as evidence for formation of the [M+nH—H3PO4]n+ neutral loss product ions during MS/MS (at least to some extent) via a competitive E2 elimination pathway. However, the observation of these ions could also be rationalized by ring opening of the 8N2 directed oxazoline ring- containing product by intramolecular proton transfer, yielding the acyclic dehydroalanine-containing product, prior to its further dissociation. Unfortunately, we can not experimentally differentiate between these mechanistic possibilities from the results obtained here. 94 2.3 Conclusions Multiple competing charge-directed and charge-remote mechanisms can contribute to the overall appearance of the resultant MS/MS and MS3 product ion spectra formed via gas-phase fragmentation of protonated peptide ions. However, in contrast to the widely accepted charge-remote B-elimination mechanism for the gas-phase neutral loss of H3PO4 from phosphoserine- and phosphothreonine-containing peptide ions, the results obtained from this study unequivocally demonstrate that this loss occurs predominantly via charge- directed mechanisms. Experimental evidence for formation of the neutral loss product ion via a charge-directed 8N2 neighboring group participation reaction mechanism under conditions of low proton mobility was obtained by the observation of novel diagnostic product ions corresponding to the loss of formaldehyde (CH20, 30 Da) or acetaldehyde (CH3CHO, 44 Da) upon CID-MS3 of the [M+nH-H3PO4]n+ product ions from phosphoserine- and phosphothreonine-containing peptide ions, respectively. Further studies are undenNay to examine the potential value of these diagnostic MS3 neutral loss product ions for enhancing the utility of data-dependent MSn methods for the characterization of phosphopeptides containing multiple potential phosphorylation sites. 95 CHAPTER THREE Evaluation of Gas-Phase Rearrangement and Competing Fragmentation Reactions on Protein Phosphorylation Site Assignment Using Collision- induced-dissociation-MS/MS and M83 3.1 Introduction Due to the vital role that reversible protein phosphorylation plays in the regulation of a large number of cellular processes and in the development of various pathological conditions,1’ 3' 23 significant effort has been directed in recent years toward the development and application of a range of mass spectrometry (MS)—based analytical strategies for comprehensive phosphoproteome analysis.15'18 In a typical approach, phosphopeptides produced by enzymatic digestion are first enriched by ion exchange chromatography, immobilized metal affinity chromatography (IMAC) or metal 17, 24, 79 oxide (e.g., TiOz and Zr02) resins, then introduced to the mass spectrometer by either electrospray ionization (ESI) or matrix assisted laser desorption/ionization (MALDI). The resultant protonated ions are then isolated The results described in Chapter Three have been published in: Palumbo, A.M.; Reid, G.E. Evaluation of Gas-Phase Rearrangement and Competing Fragmentation Reactions on Protein Phosphorylation Site Assignment using ClD—MS/MS and M83. Anal. Chem. 2008, 80, 9735-9747. 96 and subjected to collision-induced dissociation (CID)- tandem mass spectrometry (MS/MS), where the presence of phosphorylated tyrosine, serine or threonine residues within these peptides are commonly recognized by the formation of characteristic “non-sequence” product ions corresponding to the losses of 80 Da 24. 26, 79-82. 127 (HPO3) or 98 Da (H3PO4) from the initial precursor ions. When these product ions are observed as abundant fragmentation pathways, subsequent isolation and M83 dissociation is often employed to obtain additional 24, 79, 98—100. 201 structural information. “Pseudo M33” methods may also be employed, whereby activation of the precursor ion and the initial neutral loss product ions are simultaneously carried out in a single step during CID- MS/MS.101 Alternatively, observation of the characteristic neutral loss product ions may be employed to trigger electron capture dissociation (ECD)- or electron transfer dissociation (ETD)-MS/MS of the original precursor ion, under which 102,160,161,164 conditions the phosphate group remains intact. Identification of the phosphopeptide sequence and assignment of the site of phosphorylation are then performed via algorithm based interpretation or manual interrogation of the 24, 187, 202-204 MS/MS or MS3 spectra. Despite significant advances made in each of these areas and the subsequent emergence of experimental datasets containing thousands of protein phosphorylation site assignments,24’ 79' 205 various analytical obstacles remain that may hinder efforts to unambiguously localize the sites of phosphorylation to specific residues within an identified 97 phosphopeptide of interest. For example, when multiple potential sites of phosphorylation exist within a given peptide, the utility of ClD-MS/MS or MS3 analysis strategies for phosphorylation site assignment are ultimately dependent on the observation of sufficient “sequence” ions (i.e., b— and y-type product ions) from which the sequence of the peptide may be determined, and from which the site of phosphorylation may be unambiguously assigned. It is rare however, that complete “sequence” ion information is obtained upon ClD-MS/MS or MS3 of protonated peptides, thereby often limiting the ability to localize the modification to within only a region of the peptide sequence. The observation of “non-sequence” fragmentation pathways following CID- MS/MS or MS3 may also complicate phosphorylation site assignment. For example, CID-MS/MS product ions formed via the non-sequence loss of 98 Da from protonated phosphopeptides are typically assigned as corresponding to the 81, 83 loss of H3PO4. Due to the potentially limiting effects on phosphorylation site assignment when this product ion is observed at high abundance, we have recently evaluated the mechanisms and other factors that influence the formation and abundance of H3PO4 neutral loss product ions.206 In that study, it was determined that the loss is observed as a dominant process under conditions of limited proton mobility due to the formation of a strong hydrogen bonding interaction between the phosphate group and the side chain of a protonated arginine or lysine residue. Furthermore, it was proposed that the 98 Da loss primarily occurs via either a charge directed intramolecular E2 reaction, resulting 98 in the formation of either dehydroalanine- or dehydroaminobutyric acid-containing product ions from phosphoserine- or phosphothreonine, respectively (Scheme 3.1A), or via an 8N2 neighboring group participation reaction mechanism to yield cyclic five-membered oxazoline containing product ions (Scheme 3.18).206 Experimental evidence for the 8N2 mechanism was provided by the observation of characteristic products formed via the losses of 30 Da (CHzO) and 44 Da (CH3CHO) upon CID-MS3 dissociation of the initial [M+H-98]+ products formed from phosphoserine- and phosphothreonine-containing model peptides, respectively.206 However, it has also been previously recognized that the product ions formed by the loss of 98 Da could also be formed via the sequential non-sequence losses of HPO3 and H20 from a phosphorylated amino acid residue and a non-phosphorylated hydroxyl (or carboxyl) containing amino acid 89, 118 residue, respectively (Scheme 3.2). Thus, CID-MS3 and subsequent analysis of the product ions formed from this competing fragmentation process could result in erroneous assignment of the phosphorylation site. 99 H N+N ,‘H HO’I’ €Nr Ho pfo‘ H H " OH (‘5 HN 0 H J H %j— O O I + {Mfr H ,NH N OJ H O ’R‘KNMPEPTIDEAOOH H DU 6‘0) I HN AoOH N-M PEPTIDE O OH NN\ N ll H’ ‘1’ NH CID MS/MS_ OH -H3PO4 R HN 0” l H A \fi NMPEPTIDE o H o H' T NH CID MS/Ms_ OH -H3PO4 O R HN H A o \+ NMPEPTIDE 0 LII] r V H o lH+ Transfer H’ ‘11” NH 3 < MS R HN OH -RCHO O H A O 0 AN 3 erPEPTlDE o 4.))! 0 OH R = H (pS), CH3 (pT) Scheme 3.1 Previously proposed mechanisms for the “charge directed” neutral loss of H3PO4 and RCHO from protonated phosphoserine- or phosphothreonine- (where R = H or CH3, respectively) containing peptides during ClD-MS/MS. 206 (A) Intramolecular E2 mechanism and (B) intramolecular SN2 neighboring group participation mechanism. 100 OOHO H' JHHEF R10 OH O HN NHL 0 |N| PEPTIDE];k CID MS/MS -HPO3, -H20 (98 Da) (44-IL H \ + E O H H PEPTID O 0 R2 0 Scheme 3.2 Competing pathways for the formation of isobaric product ions via the loss of 98 Da (-H3PO4 or —(HPO3+H20)) from a protonated phosphopeptide containing multiple potential phosphorylation sites. R1 = H or CH3 for phosphorylated serine or phosphorylated threonine, respectively; R2 = any amino acid side chain; R3 = H or CH3 for non-phosphorylated serine or non- phosphorylated threonine, respectively. A further critical requirement for unambiguous phosphorylation site assignment is that the information obtained by interrogation of the MS/MS or MS3 product ion spectra must reflect the structure (i.e., the sequence and 101 phosphorylation site) of the phosphopeptide that was originally introduced to the mass spectrometer. Therefore, any rearrangement of the peptide sequence or the site of phosphorylation in the gas-phase, either prior to, during, or following the dissociation reaction would interfere with both sequence determination and phosphorylation site assignment. It is well established that a range of complex rearrangement reactions can occur in the gas-phase under typical ClD-MS/MS 199, 207-218 conditions involving both protonated precursor and product ions. In particular, the observation of “scrambled” sequence ions resulting from the formation and subsequent fragmentation of an isomeric population of b-type product ion structures following cyclization and subsequent ring opening of an initial oxazolone containing b—type product ion formed upon ClD—MS/MS has 217,218 recently been reported. Notably, another more recent study performed to evaluate the fragmentation reactions of several hundred synthetic peptide sequences of variable composition and length has determined that half the peptides included in the study exhibited evidence for such rearrangement . 219 reactions. Evidence for gas-phase rearrangement reactions have also been reported previously for post-translationally modified peptide ions. For example, it has been shown that formyl and acetyl groups from N-formyl- or N-acetyl- N-hydroxy omithine residues may transfer in the gas-phase under ClD-MS/MS conditions to unmodified lysine residues contained within the peptide sequence.220 Under these circumstances, interpretation of the resultant product ion spectra resulted 102 in erroneous assignment of the known site of modification. The possibility of gas- phase rearrangement reactions have also been hypothesized in order to rationalize the observed loss of 98 Da from phosphotyrosine-containing peptides, 89, 128 since phosphotyrosine may not undergo the direct loss of H3PO4. For example, Metzger and Hoffmann previously proposed that the loss of H3PO4 observed upon post-source decay of a phosphotyrosine-containing peptide may have occurred via initial transfer of the phosphate group from the tyrosine to an aspartic acid residue located within the peptide sequence, resulting in formation of a succinamide containing product.128 DeGnore and Qin hypothesized that the lengthy timescale (millisecond) for activation in an ion trap mass spectrometer may allow for loss of H3PO4 following transfer of the tyrosine-phosphate group to another residue within the peptide.89 Due to the potential limiting effect of gas-phase rearrangement reactions on the assignment of post-translational modifications involving phosphorylation to specific sites within a given phosphopeptide sequence, an improved understanding of the conditions under which rearrangement reactions may be observed is clearly required. Unfortunately, it is not generally possible to extract this information via the interrogation of phosphopeptide datasets of MS/MS or MS3 spectra obtained from proteins isolated from biological sources, as the specific sites of modification (with the exception of those containing known 205 kinase motifs)24’ and the potential for rearrangement in these peptides are 103 not known a priori. However, it is expected that detailed insights may be obtained via analysis of the MS/MS and MS3 spectra obtained from the dissociation of individual synthetic phosphopeptide sequences or synthetic phosphopeptide datasets, where the site of phosphorylation in each case is unequivocally known. 3.2 Results and Discussion 3.2.1 ClD-MSIMS of Protonated Phosphopeptides Results in Gas-Phase Phosphate Group Transfer Reactions. In order to determine whether rearrangement reactions could occur during the ClD-MS/MS of protonated phosphopeptides containing multiple potential phosphorylation sites, resulting in the formation of product ions whose inclusion in subsequent data interrogation strategies could lead to erroneous or ambiguous phosphorylation site assignments, we have examined the gas-phase fragmentation behavior of a series of singly and doubly phosphorylated peptides (Figure 3.1, Figure 3.2 and Table 3.1). The product ion spectra obtained by ESI- ClD-MS/MS of the singly, doubly, and triply protonated precursor ions of the model synthetic singly phosphorylated peptides LFTGHPEpSLER (TpSR) and LFpTGHPESLER (pTSR), using a linear quadrupole ion trap mass spectrometer, are shown in Figure 3.1 and Figure 3.2, respectively. Consistent with previous phosphopeptide fragmentation studies, a range of product ions resulting from both non-sequence and sequence fragmentation pathways were observed in these spectra, the magnitudes of which were dependant on the proton mobility of 104 the precursor ion.206 For example, the % total product ion abundance for the non-sequence product ions corresponding to the loss of 98 Da (typically assigned as -H3PO4 and labeled in Figure 3.1 and Figure 3.2 with a superscript A) from the singly protonated (i.e., non-mobile) precursor ion of the pTSR peptide shown in Figure 3.2A, was found to be 72%, while the same loss from the doubly (i.e., partially mobile) and triply (i.e., mobile) precursor ions (Figure 3.28 and C, respectively) were found to be 43% and 13%, respectively. In addition, non- sequence ions corresponding to the loss of 80 Da (-HP03, labeled in Figure 3.1 and Figure 3.2 with a superscript 0) and 18 Da (-H20, labeled in Figure 3.1 and Figure 3.2 with a superscript 0), as well as various products corresponding to the sequential losses of these species from the protonated precursor ions were also observed. 105 Figure 3.1 lon trap ClD—MS/MS product ion spectra of the (A) singly, (B) doubly, and (C) triply protonated precursor ions of the model synthetic phosphopeptide LFTGHPEpSLER (TpSR). The spectra are labeled according to the known or expected Ser phosphorylation site. Ions labeled with bold text are unambiguously indicative of the expected site of phosphorylation at the Ser residue. A = -98 Da (-H3PO4 or —(H20+HPO3)); [:1 = —80 Da (—HP03); o = —18 Da (—H20); I = +80 Da (+HPO3). 106 Figure 3.1 (cont'd) oo: OONF ooow cow . ._ _. .m ...... _.. a. ... _.. a. 0—2 ... 2 . Q2... N2 m...» m... .0... 2mm. <........ a > 2 <9» h> 4N.» ©> ....... < 0.2.... $220.1- 523%-... Ma 38 T . - wx +N mx . A$Nov+Nfi....§.. 32%.. 2. . . >2”... ...WD .. m > 60209.“-.. < 02>». 0%.» of < t o a w; . .. o< 9. 04 3. o < o . 52.0%- s mmx _. Asmtavwm- < .29 (%) aouepunqv aAnelaH (%) eauepunqv eAnelea 107 Figure 3.1 (cont’d) o_.x 09. i W . m m. 9 . w. .. +m0 n a +~ .2 2 - N . W $25228- . m ..lot. . . Ex . x/w 5.30.0.1- 0 520 108 Figure 3.2 Ion trap ClD—MS/MS product ion spectra of the (A) singly, (B) doubly and (C) triply protonated precursor ions of the model synthetic phosphopeptide LFpTGHPESLER (pTSR). The spectra are labeled according to the known or expected Thr phosphorylation site. Ions labeled with bold text are unambiguously indicative of either the expected or correct site of phosphorylation at the Thr residue, or the unexpected or incorrect site of phosphorylation at the Ser residue. A = —98 Da (—H3PO4 or —(H20+HPO3)); [:1 = —80 Da (—HPO3); ° = —18 Da (-H20); I = +80 Da (+HPO3). 109 Figure 3.2 (cont’d) oov _. CON _. of. an 9 a: .< 9. +N§om¥ $6281-. II o I I I I I on - ODE. OONF o; .. ...m O—fi .. >0 ...o O; < D GENIJ 2859.....- 8.32 7 < Evmoni- 3&2 €28- $.92 $228- cow 3. ow.» %) aouepunqv 9A!IB|GH .oo_.( (%) efluepunqv eAllelea 'o o F 110 Figure 3.2 (cont'd) 2.2m 00 N 8209-..... .. 9.2.. 41 madam- -ooF/t (%) aouepunqv aAnelea 111 Note that the spectra shown in Figure 3.1 and Figure 3.2 are labeled according to the expected (i.e., known) phosphorylation sites at Serg of the TpSR peptide sequence and Thr3 of the pTSR peptide sequence. However, the possibility of phosphorylation occurring at the alternative (i.e., unexpected) site (i.e., Thr3 of the TpSR peptide sequence and Serg of the peptide sequence TpSR) must also be considered prior to unambiguously assigning the site of phosphorylation in this peptide. Figure 3.3 shows the same spectra as those in Figure 3.2, but where the product ions in each spectra are labeled according to the alternative (i.e., unexpected) phosphorylation site at Serg. By comparison of these Figures, it can be seen that the majority of the observed product ions can be assigned to either of the two possible phosphorylated peptide sequences. For example, yn—type product ions assigned to the pTSR peptide in Figure 3.2 that contain the non-phosphorylated serine residue (e.g., y4, y5, W and ya in Figure 3.2A) may also be assigned to the TpSR peptide in Figure 3.3 as corresponding to ynD-type product ions (e.g., y4D. Yea. WU and y3D in Figure 3.3A) formed via the sequential loss of 80 Da (-HPO3) from yn-type ions that contain the phosphorylated serine residue (i.e., sequential sequence and non-sequence fragmentation pathways). Similarly, the yn-type product ions labeled in Figure . . O 0 O O O . . 3.2 With a superscript (e.g., y4 , y6 , w and ya In Figure 3.2A), corresponding to the loss of 18 Da (-H20) from product ions containing the non- 112 phosphorylated serine residue in the expected pTSR sequence may also be labeled as corresponding to the loss of 98 Da (-H3PO4) from the phosphorylated . . . A A A A . senne resrdue In the unexpected TpSR sequence (e.g., y4 , ye , w and ya In Figure 3.3A). b-type product ions labeled in Figure 3.2 with a superscript A corresponding to the loss of 98 Da (-H3PO4,), which contain only a single . . . A A A A. . potential srte of phosphorylation (e.g., b3 , b4 , b5 and by In Figure 3.2B), may also be assigned in Figure 3.3 as corresponding to the loss of 18 Da (-H20) when a non-phosphorylated threonine residue is present (e.g., b3o, b4o, b50 and b7° in Figure 338). Finally, b—type product ions labeled with a superscript D in Figure 3.2 (e.g, b5D, bit3 in Figure 3.2A) corresponding to the loss of 80 Da (HPO3). when a phosphorylated residue is present, may also be assigned in Figure 3.3 as intact b-type ions where the non-phosphorylated residue is present. Without definitively knowing the extent to which the losses of H3PO4, HP03 or H20 occur in these peptides, these ions must all be considered ambiguous, and therefore may not be used to provide conclusive evidence to unambiguously assign the site of phosphorylation. It should also be noted that conclusive evidence to unambiguously localize the site of phosphorylation cannot be obtained from b- or y-type product ions containing a phosphate group (e.g., b9, y9 and b1o in Figure 3.2A), or from bnA- or ynA-type product ions formed via the 113 loss of 98 Da (e.g., bgA, b1oA, ygA and y1oA in Figure 3.2A), when multiple potential phosphorylation sites within these regions of the sequence are present. These results therefore indicate that the only ions in the ClD-MS/MS spectra that can provide conclusive (i.e., unambiguous) evidence for the site of phosphorylation are those in which the phosphate group remains intact and where the product ion contains only a single potential site of phosphorylation. 114 Figure 3.3 lon trap ClD—MS/MS product ion spectra of the (A) singly, (B) doubly and (C) triply protonated precursor ions of the model synthetic phosphopeptide LFpTGHPESLER (pTSR). The spectra are labeled according to the unexpected (i.e., incorrect) site of phosphorylation at the serine residue. A = —98 Da (— H3PO4 or —(H20+HPO3)); «:1 = —80 Da (—HPO3); ° = —18 Da (4120); I = +80 Da (+HPO3). 115 Figure 3.3 (cont'd) cow _. oomw 000—. com 9 ...m D? a; 0N2 <............ amen o a use S .2909. $620.21-. 2°33 $228- oov _. CON _. coo _. cow @o~:.......w Aoovo~zw-m 05...... of 2.9. .. A» 298%- 830:- m.» - mx p $85 228- o_.x NX coo %) eouepunqv GAHBIGH r09.1 %) eouepunqv eAlieIea 221.. 116 Figure 3.3 (cont’d) oow m m + m +N< > 'A 8 . n: 7: +... a. $3209.- . $5: $238- q 0 L NE. oov .+No +0 com ...o i c.» +ND ..u —|<.. 4‘ +N< 2mm oow .oo_. %) eouepunqv eAiieIea ( 117 Inspection of the product ion 'spectrum obtained upon CID-MS/MS of the singly, doubly and triply protonated precursor ions of the TpSR peptide (Figure 3.1) revealed multiple ions in each case (labeled in the Figures in bold text) that were unambiguously indicative of the expected site of phosphorylation at the serine amino acid. In contrast, inspection of the product ion spectrum obtained upon CID-MS/MS of the singly protonated precursor ion of the pTSR peptide (Figure 3.2A) did not reveal any ions that were unambiguously indicative of the expected site of phosphorylation at the threonine amino acid. However, an unexpected series of product ions were observed, whose mlz values corresponded to localization of the phosphate moiety (+80 Da, ') at the serine residue of the peptide (labeled as y4-. ye'. WI, and ya' in Figure 3.2A, and as y4, ye, y7, and ya in Figure 3.3A). CID-MS3 of the ye. ion (sequence PESLER+80 Da) from Figure 3.2A (see Figure 3.4A) revealed the presence of an unmodified y3 ion (LER) as well as a product corresponding to a y4 ion containing the phosphate group (y4') (mlz 584.4), confirming that the phosphate group was located on the serine residue. Importantly, this spectrum was essentially identical (see Figure 3.4B) to that obtained by CID-MS3 of the phosphorylated y5 product ion from the TpSR peptide in Figure 3.1A. These data potentially suggest therefore, that the incorrect peptide (i.e., one containing phosphoserine (TpSR) rather than phosphothreonine (pTSR)) had been introduced to the mass spectrometer, or indicate that transfer of the phosphate 118 residue had occurred either in the interface of the mass spectrometer during the electrospray ionization process, or in the gas-phase during ClD-MS/MS. 119 O O . LO - ID A °° 2 00 O V O V o 9.. a C2. A O I d. A O I ‘ C] N ' C] N ' s, E “,3 s E - ,ug ............... w '-..........,._" 3 III: ..... , q: 3': . 8 A <1 A 3 2 A 5’ 8 ------------ < 75 8 ................ ‘- ........ , O .- ....... O ‘- < ‘1- It) ‘- < LO ' :1" “‘ co ' V co co 1 m ‘- OLD " ‘7 <1“) _1 ' >. ‘1 >. I 1' 4' §-—-« E I o I o 1" '0 LO 1 l0 . LO 1 ‘ N 0 q ‘ \ §——— g 2 E o , o , V- ‘— l o O x ~ to x - :2 <1- <1- >.“’—-‘1 >.‘°—- O 0 l0 - l0 (‘0 0') 9a >‘?‘~ J a a < N m N O O O O V- ‘- (%) eouepunqv eAneIea (%) aouepunqv aAnelaa Figure 3.4 lon trap CID—MS3 product ion spectra of (A) the y5+80 Da (y5+ HPO3) product ion formed from ClD-MS/MS of the singly protonated pTSR precursor ion in Figure 3.2A and (B) the ye product ion formed from ClD—MS/MS of the singly protonated TpSR precursor ion in Figure 3.1A. 120 An unexpected ye. product ion was also observed upon analysis of the ClD-MS/MS spectrum obtained from the doubly protonated precursor ion of the pTSR peptide (Figure 3.2B). However, products corresponding to the correct site of phosphorylation were also observed in this spectrum, (see the b3, a5, b5 and b7 ions), suggesting that a mixture of the threonine and serine phosphorylated peptide sequences were present. In contrast, no evidence was observed for unexpected product ions indicative of the TpSR sequence in the ClD-MS/MS spectrum obtained from the triply protonated precursor ion of the pTSR peptide (Figure 3.20), while various products were observed that could be used to conclusively assign the site of phosphorylation to the expected site of phosphorylation (see the b3, a5, b5, b52+ and b72+ ions). As the presence of the TpSR peptide during introduction of the pTSR peptide to the mass spectrometer would be expected to result in observation of the unexpected product ions in each of the precursor ion charge states subjected to ClD-MS/MS, this data provides evidence for formation of these ions via a gas-phase phosphate transfer process. Further confirmation that the unexpected product ions were not due to the presence of the isomeric phosphoserine (i.e., TpSR) containing peptide being present within the pTSR sample that was introduced to the mass spectrometer, was obtained by repeating the MS/MS analysis following individual RP-HPLC purification of the pTSR and TpSR peptides. Using the conditions outlined in the Materials and Methods section above, the retention times of these peptides were found to be 22.8 min and 20.1 min, respectively. Reinjection of each peptide 121 following their purification did not result in any absorbance at the retention time of the other peptide, indicating that the purification process did not result in rearrangement of the phosphate group. The effects of the ionization method employed (ESI versus MALDI), the type of ion trap mass spectrometer in which the dissociation experiments were carried out (linear versus 30), as well as the time scale for ion activation (ion trap versus quadanole) on the origin of the phosphate group transfer products were also examined. Acquisition of the ClD-MS/MS product ion spectrum from the singly protonated precursor ion of the pTSR peptide following ionization by MALDl on a linear quadrupole ion trap mass spectrometer yielded an identical spectrum to that shown in Figure 3.2A, suggesting that the rearrangement process did not occur during the ESI process. Analysis of the product ion spectra obtained from the singly, doubly, and triply protonated precursor ions of the pTSR peptide following ESI-ClD—MS/MS in a conventional 30 ion trap resulted in essentially identical product ion spectra to those observed using the linear ion trap, including observation of the rearrangement product ions from the singly and doubly protonated precursor ions (data not shown), indicating that the rearrangement process was not limited to a particular ion trap platform. The abundance of the product ions resulting from gas-phase rearrangement reactions upon ClD-MS/MS of the singly protonated precursor ion of the pTSR peptide in the linear ion trap mass spectrometer was found to vary significantly as a function of the activation time (2-2000 msec). In particular, a decrease in the abundance of the phosphate group rearrangement product ions was observed 122 with decreasing activation time (Figure 3.5). However, even at the shortest activation time that could be employed while still observing reasonable activation efficiencies (2 msec); the rearrangement product ions were still present. In contrast, upon analysis of the singly, doubly and triply protonated peptide ions by ESI-ClD-MS/MS in a triple quadrupole mass spectrometer (psec activation timescale), no product ions resulting from the gas-phase phosphate group rearrangement reactions were observed, while the overall appearance of the spectra were relatively unchanged compared to that obtained using the ion trap (Figure 3.6 and Figure 3.7). These results indicate the phosphate group rearrangement reaction occurs in the gas-phase during ClD-MS/MS, in either linear or 3D ion trap mass spectrometers, due to the relatively long (millisecond) timescales associated with ion activation in these instrument platforms, consistent with that previously hypothesized by DeGnore and Qin.89 123 Figure 3.5 lon trap ClD—MS/MS product ion spectra of the singly protonated precursor ion of the model synthetic phosphopeptide LFpTGHPESLER (pTSR) using (A) 10 msec and (B) 2000 msec ion activation times at an activation q value of 0.25. The spectra are labeled according to the known (i.e., expected) site of phosphorylation (i.e, Thr). Ions labeled with bold text are unambiguously indicative of the unexpected site of phosphorylation (i.e, Ser), resulting from the gas-phase rearrangement reaction. A = —98 Da (—H3PO4 or -(H20+HPO3)); El = —80 Da (—HPO3); ° = -18 Da (—H20); I = +80 Da (+HP03); RA = relative abundance. 124 Figure 3.5 (cont’d) 00..V _. 6209...... AoovoNINW 2m 08.: Ammo... I- 228. oov _. 8209...... 809$ 2m $9.8 62.0%- 228- 832 F- OONF mx oov o_.x owx %) aouepunqv GARBIGH m .oS( %) eouepunqv eAiieIea < .oS( 125 Figure 3.6 Triple quadrupole ClD-MS/MS product ion spectra of the (A) singly, (B) doubly, and (C) triply protonated precursor ions of the model synthetic phosphopeptide LFpTGHPESLER (pTSR). The spectra are labeled according to the known (i.e., expected) site of phosphorylation (i.e, Thr). Ions labeled with bold text are unambiguously indicative of the expected site of phosphorylation at the threonine residue. A = —98 Da (—H3PO4 or —(H20+HPO3)); D = —80 Da (— HPO3); ° = —18 Da (—H20). 126 Figure 3.6 (cont’d) OOVF OON_. OOO _. COO m COS. .C._.> < S... < OON_. OOO _. 9 COO > z + Q1. 0 +~ovmo COO OOv OON V838 F- OOV CON b b - 28 3%? 639 F- fine. 11‘ D m limit- 2 .. 2; S. n. 1dr {it“d‘i‘ttitjm .. i. n. %) eouepunqv eAiieieu .OO_.( %) aouepunqv aAgteIaH .OO_.(\ 127 Figure 3.6 (cont’d) CCC _. COO COO CON COO CCC CE.» CCm CON OC 2. i he me +NO _.> Mur .A . ...N 0X ©> W W m .. .. ................. o ; 6 We 00> +N< We O ; ......+N Nm %) eouepunqv eAneIea 0 mini): NS . ....... . 00—. 128 Figure 3.7 Expanded regions of the product ion spectra obtained by (A) ion trap (Figure 3.2A) and (B) triple quadrupole (Figure 3.6A) ClD-MS/MS of the singly protonated precursor ion of the model synthetic phosphopeptide LFpTGHPESLER (pTSR). The spectra are labeled according to the known (i.e., expected) site of phosphorylation (i.e, Thr). Ions labeled with bold text are unambiguously indicative of the unexpected site of phosphorylation (i.e, Ser), resulting from the gas-phase rearrangement reaction. E] = —80 Da (—HPO3); O = -18 Da (—H20); I = +80 Da (+HPO3). 129 Figure 3.7 (cont’d) 1 O to CD $3 :32 O O 0 $2 5 >°.° 'l E e >': co >. . O (O _. (I) (I) a I X 0 > _‘2 o __ O) N — O I!) [x $2 $2 22 O O (D o (O >~. >~ "‘ < P: m C j C O S ,_ (%) eOUEpunqv eAuelea (%) eauepunqv eAiielea 130 950 910 870 830 790 750 710 m/z To determine the generality of these gas-phase phosphate group transfer reactions, 33 singly or doubly phosphorylated peptides, each containing multiple potential phosphorylation sites, were independently synthesized then examined by ESI- and MALDI- ClD-MS/MS using linear quadrupole ion trap mass spectrometers (Table 3.1). Of these peptides, 70% (23 of 33) were singly phosphorylated (12 containing phosphoserine, 7 containing phosphothreonine and 4 containing phosphotyrosine) while the remaining 30% (10 of 33) were doubly phosphorylated (4 containing two phosphoserine, 1 containing two phosphothreonine, 2 containing one phosphoserine and one phosphothreonine, 2 containing one phosphoserine and one phosphotyrosine, and 1 containing one phosphothreonine and one phosphotyrosine). These sequences were taken from a previously published dataset of identified phosphopeptide sequences,24 and were selected in order to provide a reasonable representation of the overall characteristics of the peptides contained within the dataset (i.e., the type and number of phosphorylation sites, peptide mass, complete versus incomplete tryptic cleavages, arginine versus lysine terminated, etc.). All of the precursor ion charge states that were observed by either ESI (primarily doubly and triply protonated precursor ions) or MALDI (singly protonated precursor ions) using the experimental conditions described in Chapter Four and were subjected to CID- MS/MS. Product ion assignment was achieved by manual interrogation of each of the MS/MS spectra. For all of the precursor ion charge states from each of the peptides listed in Table 3.1, the number of observed amide bond cleavages that were used to unambiguously assign either the correct or incorrect site(s) of 131 phosphorylation, as well as the total number of possible amide bond cleavages that could be used to unambiguously assign the site(s) of phosphorylation, have been determined (Table 3.1). Amide bond cleavages that enable unambiguous assignment of the site of phosphorylation to the correct (i.e., expected) residue result from product ion(s) that contain the intact phosphate modification at the expected residue, and (i) where no other potentially phosphorylated residues (i.e., no unmodified serine, threonine or tyrosine) are present within the product ion, and (ii) when the m/z of the product ion is higher than the low mass cutoff and lower than the high mass limit of the instrument, and (iii) where the nominal mlz of the product ion does not overlap with any other product ion. Similarly, amide bond cleavages that enable unambiguous assignment of the phosphorylation site to an incorrect (i.e., unexpected) residue result from product ion(s) that contain a phosphate modification (+80 Da) at an unexpected serine, threonine or tyrosine (i.e., incorrect) residue within the product ion, and (i) where the formerly phosphorylated (i.e., correct) residue is not present within the product ion, and (ii) when the m/z of the product ion is higher than the low mass cutoff and lower than the high mass limit of the instrument, and (iii) where the nominal mlz of the product ion does not overlap with any other product ion. 132 Table 3.1 Summary of the gas-phase phosphate group transfer product ions observed following ESl- and/or MALDI-ClD-MS/MS of independently synthesized tryptic phosphopeptides. a Proton mobility classifications defined by Kapp, et al: Non-mobile: number of ionizing protons 5 number of arginine residues; Partially-mobile: number of arginine residues < number of ionizing protons 5 combined number of arginine, lysine and histidine residues; Mobile: number of ionizing protons > combined number of arginine, lysine and histidine residues. b I indicates +80 Da (i.e., +HP03), II indicates +160 Da (i.e., +2HPO3). c Singly protonated precursor ions observed only by MALDI. 133 Table 3.1 (cont’d) to QC - 0__20E N moez>oomzmme00< 00 ca - 0:858: F :0 E - 0.39: N x00ememem as as - 0:858: F to ca - 258. N 5052808024 2N 2o .0) re. 0.52: 233.8 P so NN - 232: N xodtomeom 30 Na - 252: 23:8 P Na Na - 0.59: 23:8 N $801.3 Na Na - 0:858: F :2... 3... 3.3... w _ ”0.? 200.2002. .020... 2002.50 0020020 :0. 2005005 0:0. 00900.0 000092.202: 134 Table 3.1 (cont’d) $0 08 - 0_.00E v 30 E. - 0.59: m maxamommo0qt>3 $0 5 - 0.50.: 2.03.00 N to E - 0.3950: 0? 08 05 - 0.50.: N m:000:0000u.03> QN 90 It» .I? 0.50800: 9 :0 F: - 0.50:. m :o E - 0.39: N szm_00fimo.u.20n_0._ :o E - 259: 20:8 or 0:0 N\N - 0.30.: m 25 Na - 259: N x0:0:002.En_..>0._0 08 N: - 2.85 23:8 F to ca - 2.8:. 0 «00250020082 "00:00:. 80:00 .000... 0:0. 00:0; 2:50.). :22n. 0.05 00:0:000 .06.. 30:00:. 000.80 000.0 90:00on 090:0 0020000 :0. 50.5005 0:0. 0m0>00.0 000390505. 135 Table 3.1 (cont’d) 0.0 0.0 - 0.50:.. N mqmqquokaawwgw to 0.0 - 0.50:.:0: _. 0.0 0.0 - 0.50:. m 0.0 ...o - 2.8:. N x0<<._0_005_0>: EN 0.0 I9. .lw> 0.50:. 2.00.00 N NE N\N - 0.50:. N _ 030000050 NB NB - 0.50:. 2.05.00 F 3.0 0.0 - 0.50.: 5.05.00 0 EN 05 .0; ..N; ..o; 0.52.8: N m><<30000<0>000<0000< > . . > 0.0 0.0 I: IN; .0. 0.50:.:0: 0F 20 0.0 - 0.50:. m to 05 - 0.50:. N ¥0w0<0000m»..050:00:05<..> N:N N: .0 E ..05 059.. 2.0.0.. N 0 EN N: um E .005 0.50E:0: 0. N ..o ...o - 0.8... 0 N ..o o... - 0.8:. N 0.05000050509005 0.0 0.0 .0 F» .00.. .00.. 0.50.: 2.00.00 0. ...0 E - 0.50:. N 0.50095500000 0.0 F. P - 0.50:. 2.0...00 v 50:00:. 50:00 0:0. 55:0... .0550... :0.0.0 0.0.0 00:00.50 .05... ..00:00:. .05 ... 000:00 000.0 0.0500050 00.050 :0. 0020050 50500.0 0:0. 000>00.0 000005605. 137 Table 3.1 (cont’d) F... v... - 0.50:. 2.050.. N m0>w0mw0omm102020m :0 0. F - 0.50:.:0: OF 0.0 0.0 - 0.50:. m 0.0 0.0 - 0.50:. 2.0...00 N v.m.0.0zo>omu.w012n.0>0wc< 0.0 0.0 - 0.50:. 5.0...00 OF 0.0 F. F - 0.50.: N 500550900.» 0.0 F. F - 0.50.::0: F :0 F. F - 0.50:.. m ..o F.F - 059.. 25:00 N x2000>0>zmmmo> F.0 F.F - 0.50.: >__0...00 F :0 F. F - 0.50:.. N 05.502005 :0 F.F - 0.50.::0: F F :0 0.0 - 0.50:. v 000 F 0N.0 0.0 . F F > . 0 F > . 0.. 0.50E:0: 0 l I I 0 F.F ...o .3 0.50:. 0 80 ...o ...o ..o; 0.59.. 2.00.00 N 0.0000000590000000 0 In 0.0 03.. .IF 5. .00 F.» 0.50:.:0: 0F F... v... - 0.30:. 0 0.5000000001020205 50:00:. .05 ... 50:00 .05 .. 0:0. 55:0... 2:50.... :0.0.0 0.0.0 0800000 .50:00:_ 50.50 000.0 0.0500050 00.050 :0. 0020050 50500.0 0:0. 000>00.0 0000055050 139 Table 3.1 (cont’d) 0N... :0 - 0.50:. 0 «N: ...0 u... 0.00:. 2.00.00 N v.00002550005505....0000. 0N... 0.0 - 0.50:.:0: 0F 0:0 20 - 0.505 2.0...00 0 0:0 5 .00 .0... .00.. 0.39.. 20:00 N v.00..00:>00>00>0: 0 E o... 0.... 0.50... 0.050.. F E... N... - 0.30:. 0 ..F: N: .3. 059.. 2.00.00 N 503005550500515 3.0 N... _...F. .08. ...0. 0.58.8: 0. . :0 ...o - 0.00:. 0 . F: 0.0 .0. 0.02.. 0.00.00 N 500.005.050.05... 0 ..N ...o .0. ..0. 0.50:.:0: . ..F... ...o - 0.30:. 0 $3.055050E0>01<0 50:00:. .05 .. 50:00 0:0. 55:0. ... .5505. :0.0.0 0.0.0 00:00.50 50:00:. .05 .r 50.50 000.0 0.0500050 00.050 :0. 0020000 50500.0 0:0. 000>00.0 000005502. 140 Table 3.1 (cont'd) 00... E - 0.52.. .V on: F.F .an 0.50... 2.00.00 0 00055....00250301550 S I . . > . > 0N: E .00 .NF uFF 0.32.8: N . 0F.. ... 0. I I I 0N: o... .8. 0.58.8... 0. 50555525035550 00:8... 00:00 0:0. 0.2.0: 2.50... :0.0.0 0.0.0 00:00:00 _0.0.F 50:00:. .05 ... .50..00 000.0 0.0500050 00.050 :0. 0020050 50500.0 0:0. 000>00.0 0000055553 141 From the data in Table 3.1, it was determined that 15 of the 33 peptides (45%) exhibited product ions resulting from gas-phase phosphate group transfer reactions. Importantly, the site of phosphorylation could be unambiguously correctly assigned in only 12 of the 33 (36.4%) peptides, while 10 of the 33 (30.3%) peptides gave rise to product ion spectra from which the site of phosphorylation was unambiguously incorrectly assigned. For the remaining 11 peptides (33.3%), the site of phosphorylation could not be definitively assigned to either the correct or incorrect site(s), as either both or neither of the required product ions were observed. When divided according to the ionization method employed, only 10 out of the 33 peptides ionized by MALDI (30.3%) and 12 out of the 33 peptides ionized by ESI (36.4%) resulted in unambiguously correct phosphorylation site assignment. Further analysis of the data in Table 3.1 revealed that 28 of the 94 observed precursor ion charge states (29.8%) gave rise to product ions resulting from gas-phase phosphate group transfer reactions (14 of the 33 precursor ions observed by MALDI (42.4%) and 19 of the 78 precursor ions observed by ESI (24.0%)). From these data, the number of unambiguously correct, unambiguously incorrect and ambiguous phosphorylation site assignments were 31 of 94 (33.0%), 21 of 94 (22.3%) and 42 of 94 (44.7%), respectively. Although this is a relatively limited data set, such that a comprehensive evaluation of the effect of peptide sequence, length, overall basicity/acidity and the relative location of the donor/acceptor residues was not possible, a clear trend was observed based on the effect of the precursor ion proton mobility on 142 the observation of the rearrangement products. The majority of the precursor ions that exhibited evidence for the phosphate group transfer were classified as “non-mobile” (number of ionizing protons 5 number of arginine residues) or “partially-mobile” (number of arginine residues < number of ionizing protons 5 combined number of arginine, lysine and histidine residues).193 Note that the percentage of the precursor ion charge states that could be unambiguously correctly assigned following ClD-MS/MS under “non-mobile”, “partially-mobile” or “mobile” protonation conditions was 23.8% (5 of 21 ), 30.3% (10 of 33) and 42.5% (17 or 40), respectively. Based on this observation, a potential mechanism for the observed rearrangement reaction is shown in Scheme 3.3. This mechanism involves nucleophilic attack by the hydroxyl oxygen of a non-phosphorylated serine or threonine residue to the phosphorous atom of a phosphoserine or phosphothreonine residue, with a concerted proton transfer. Due to the observed effect of proton mobility on the formation of these rearrangement products, it is proposed that this mechanism is facilitated by the initial formation of strong hydrogen-bonding interactions between the phosphate group of the phosphoamino acid and the side-chain of a protonated arginine residue. This is consistent with our previous report indicating the role of such intramolecular interactions in the loss of H3PO4 from protonated phosphoserine and phosphothreonine containing peptides.206 This proposal is also consistent with previous reports with a report describing gas-phase phosphorylation within non- 143 covalent complexes of serine or threonine containing peptide cations and triphosphate anions.221 Moreover, this rationale is in agreement with previously observed intermolecular gas-phase phosphate group transfer within noncovalent heterodimeric peptide complexes. In particular, Woods and coworkers showed that due to the “covalent-like” stability of the phosphate-arginine (and phosphate- quatemary amine) electrostatic interactions, ClD-MS/MS of such non-covalent complexes causes removal of the phosphate group from the serine, threonine, or tyrosine residues by scission of the P-O phosphoester bond, while the phosphate remains noncovalently bound to arginine or quaternary amines of the formerly 106, 107, 109, 110 nonphosphorylated peptide in the complex. Moreover, Jackson et al. have shown that the propensity for noncovalent intermolecular gas-phase phosphate group transfer decreases with increasing charge state.110 ‘71 H N + H .N + N . Ho ————— H I HQ"'H \Nr Ho-eio ----- H' ‘H ClD-MS/M§ HO'P=0'"H’ H OH 0 OWH'O NH OH 0 0H 0 NH 0 ”'JLN N 0 FAN N H o H o H o H 0 Scheme 3.3 Proposed mechanism for the intramolecular gas-phase phosphate group transfer facilitated by hydrogen bonding to a protonated basic residue. Several of the phosphopeptides listed in Table 3.1 were also subjected to MS/MS analysis via electron transfer dissociation (ETD), following esterification of the peptides to increase the relative basicity and promote the formation of 144 multiply protonated ions, and performed with or without the aid of supplemental activation. As expected, no evidence was observed in any of the product ion spectra for the loss of H3PO4. Furthermore, no evidence was observed for the presence of products formed via gas-phase rearrangement reactions, thereby readily allowing the unambiguous determination of the location of the correct site of phosphorylation in these peptides in each case (data not shown). These findings therefore clearly support the increased use of electron transfer dissociation (ETD) (or electron capture dissociation (ECD)). under which the 102, 160, 161, 16 phosphate group remains intact, 4 for large scale or high throughput phosphoproteome analysis studies. 3.2.2 CID-MS3 of [M+nH-98]'1+ Ions May Not Be Used for Unambiguous Phosphorylation Site Localization. The loss of H3PO4 (98 Da) from a phosphoserine or phosphothreonine residue upon ClD-MS/MS would result in the formation of a dehydrated amino acid at the formerly phosphorylated residue (Scheme 3.1), while any non- phosphorylated tyrosine, serine or threonine residues that are present within the peptide would remain intact. Further M83 dissociation of the [M+nH-98]n+ product ions could therefore be use to obtain evidence to assign the site of phosphorylation. However, as mentioned in the Introduction section above (3.1), if the loss of 98 Da occurs via the loss of HP03 from a phosphotyrosine, 145 phosphoserine or phosphothreonine residue combined with the loss of H20 from a non-phosphorylated serine or threonine residue (Scheme 3.2) an isomeric product ion containing an unmodified tyrosine, serine or threonine residue at the formerly phosphorylated amino acid, and a dehydrated amino acid at the formerly non-phosphorylated amino acid would be produced. Analysis of the product ions formed by MS3 of this population would therefore result in assignment of the site of phosphorylation to the incorrect position. Under conditions where gas-phase phosphate group transfer occurs prior to dissociation, an isomen’c mixture of protonated precursor ions with respect to the site of phosphorylation may also be present. For example, the pTSR peptide discussed above was shown to undergo gas-phase phosphate group rearrangement to yield a mixture of itself and the phosphoserine isomer, TpSR. Subsequent MS3 of the product ion corresponding to the neutral loss of H3PO4 from this isomeric mixture would therefore result in the formation of product ions from which the site of phosphorylation could be assigned to either the expected (i.e., correct) or unexpected (i.e., incorrect) phosphorylated residue. Note that the structure of the product ions formed by the combined losses of HP03 and H20, as described above, would be indistinguishable from those formed via the loss of H3PO4 from a precursor ion population that had undergone a gas-phase phosphate group transfer reaction prior to its dissociation. 146 Definitive evidence for the combined losses of HPO3 and H20 was obtained here, however, by analysis of the spectra produced by CID-MS3 of the [M+nH-98]ln+ ions from the singly, doubly and triply protonated precursor ions of the TpSR peptide from Figure 3.1 (see Figure 3.8A-C, respectivelY). where no evidence for the gas-phase phosphate group rearrangement was observed upon MS/MS. In these spectra, a series of product ions (labeled as yn') corresponding to the presence of an unmodified serine residue rather than the expected dehydrated residue at the site of the formerly phosphorylated serine residue, were observed (Figure 3.8). It was also noted that the abundance of the -80 Da (-HPO3) neutral loss product ion in the MS/MS spectra of these ions were increased relative to that of the -98 Da neutral loss product with decreasing activation times in the linear ion trap (Figure 3.5), consistent with the loss of 98 Da being formed, at least to some extent, via the initial loss of HPO3 from the precursor ion followed by the further loss of H20. Further evidence for the presence of a population of ions formed via the combined losses in the -98 Da ClD-MS/MS product ion spectra was also obtained by inspection of the neutral loss product ions observed in the MS3 spectra of the singly and doubly protonated TpSR peptide (Figure 3.8A and B, respectively). As mentioned in the Introduction section above (3.1), a recent study to determine the mechanisms and other factors that influence the formation and abundance of H3PO4 neutral 147 loss product ions resulted in the observation of characteristic product ions formed via the losses of 30 Da (CH20) and 44 Da (CH3CHO) upon CID-M83 dissociation of the initial [M+H-98]+ products formed from phosphoserine- and phosphothreonine-comaining model peptides, respectively.206 The potential utility of these MS3 aldehyde neutral loss product ions were noted in this prior study as diagnostic indicators of the site of phosphorylation (i.e, pSer or pThr) in peptides containing multiple possible sites of phosphorylation. For the TpSR peptide, therefore, if the ClD-MS/MS neutral loss of 98 Da occurred exclusively via loss of H3PO4 from the phosphoserine residue, the loss of CH20 should be observed upon M83, rather than the loss of CH3CHO. It can be seen however, that an abundant ion corresponding to the loss of CH3CHO (44 Da) was observed in these spectra, indicating that the loss of 98 Da had occurred via the combined loss of HPO3 from the phosphoserine residue and the loss of H20 from the threonine residue. This result is consistent with other studies that have shown that the loss of H20 from serine or threonine residues198 can result in the formation of product ions with identical structures to those determined for the loss of H3PO4 from phosphoserine or phosphothreonine residues,117 and can therefore undergo similar subsequent fragmentation reactions. 148 I." 1.4 - . ..“fl 5 '1‘ -!', II. Figure 3.8 lon trap CID—M83 product ion spectra of the (A) singly, (B) doubly, and (C) triply protonated ([M+nH—98]n+) neutral loss product ions from the TpSR peptide in Figure 3.1A—C, respectively. Ions labeled with bold text correspond to the presence of an unmodified serine residue rather than the expected dehydrated residue. The inset in panel B shows the expanded m/z region from 600-630. ° = -18 Da (—H20); ' = +18 Da (+H20). 149 Figure 3.8 (cont’d) 09; CON? oo.o_. cow CON a M... . omo owe owe 2.8 . 6599.0sz omzo- S m .2: 90¢ F CON _. ooo _. com com oov r . 3...... . -- ...: ...- .. 1.. a w .. 11...: . a E T, 1.. m. _4/ 3m m3 an to 033 w on Q . .. M> .. > m .. v .. 9 3.32m“. ......3 mac on o> . ... 0 an m. 0 ma m: o o n N> 9» Q... S” . ......... m o a o . > . ..... Omud OF .. m m H a 659.. 7w o .05 < .2: (%) eouepunqv eAneIea (%) eouepunqv eAneIea 150 Figure 3.8 (cont’d) _ l (I) >. N ........ II > t o D F l x ‘0 >4 é?__— o ......... W (D . > m ....... > .............. 1' Ln .......... N >. 54-0 ............. _.., F ooooooo + o >__~3 ................ b. N S ........ Lo. 93 > .o + + _ m0 + N00 N NN >5 ............... . I. > ................................... 1 l + ....... — N .......... <0 >. N— .o N m— (0/0) eouepUHQV GARBIGH 151 300 400 500 600 700 800 900 1 000 200 m/z Finally, a series of product ions were observed in the M83 spectra from the TpSR peptide in Figure 3.8 (as well as an analogous product ion series in Figure 3.9), that could be assigned as being due to the expected presence of a dehydrated residue at the site of the formerly phosphorylated serine (e.g., y4, y5, ye, w and ya) as well as for the presence of the unmodified threonine residue (b3, b5 and b7), following the loss of H3PO4 from the phosphoserine residue. However, these ions could also be assigned as being due to the loss of water from an unmodified serine residue and the loss of HPO3 from a phosphothreonine residue following the combined loss of H20 and HPO3 from a phosphothreonine containing peptide (i.e., pTSR). Indeed, all of the product ions observed in the MS3 spectra could be assigned as being formed from either of the phosphorylated peptide isomers. 152 Figure 3.9 lon trap CID-M83 product ion spectra of the (A) singly, (B) doubly, and (C) triply protonated ([M+nH—98]n+) neutral loss product ions from the pTSR peptide in Figure 3.2A—C, respectively. lons labeled with bold text correspond to the presence of an unmodified threonine residue rather than the expected dehydrated residue. The inset in panel B shows the expanded m/z region from 600-630. ° = —18 Da (—H20); ' = +18 Da (+H20). 153 Figure 3.9 (cont’d) oov _. omo 0N0 com—- on. ...o; 9...... «.3 3.. omo ooor 9. «00-. $6303..- oov _. J u? AoVONI-3 0mm. +Ao.ovONI-... CON—- +No_._ enro- _3 O 33> Omzo- 3.. 3.0103310- arm-4.. m...o 93.? NE. cow b 3.? own 93 ...-NAOIomzo-oNIV- ... 3......- _ 3.. 9 mg mg . (%) eouepunqv eAueIeu .oor (%) eouepunqv eAneIeu é: o F 154 Figure 3.9 (cont’d) 10x 1001-(3 (%) eouepunqv eAneIea 155 300 400 500 600 700 800 900 1000 200 an Similar results to those described above for the TpSR peptide were also obtained upon CID-MS3 of the [M+nH-98]n+ product ions from the pTSR peptide (Figure 3.9), where a5., b5. and b7. ions indicative of the presence of an unmodified threonine residue rather than the expected dehydrated residue were observed, as well as the loss of CH20. Taken together, these data all indicate that without definitively knowing a priori the extent to which gas-phase phosphate group transfer occurs in a given phosphopeptide, or the extent to which the combined losses of HP03 and H20 contribute to the loss of 98 Da, the product ions formed by CID-MS3 must all be considered ambiguous (note the similarity between the product ions, but not the product ion assignments, in the MS3 spectra of the TpSR peptide in Figure 3.8A and the pTSR peptide in Figure 3.9A), and therefore may not be used to provide conclusive evidence to unambiguously assign the site of phosphorylation within peptides containing multiple potential phosphorylation sites. 3.3 Conclusions Although tandem mass spectrometry based analysis strategies have been widely employed for phosphoproteome analysis, the potential for gas-phase phosphate group transfer rearrangement reactions under typical ion trap CID- MS/MS conditions can significantly impede the ability to unambiguously assign the correct site of phosphorylation in protonated phosphopeptides containing 156 multiple potential sites of phosphorylation. Furthermore, given that competing fragmentation reactions corresponding to the loss of H3PO4 or the combined losses of HPO3 and H20 may be responsible for the neutral loss of 98 Da, the spectra acquired by CID-MS3 of [M+nH-98]n+ ions may not be used for unambiguous phosphorylation site localization. Thus, great care should be taken when exclusively using ion trap ClD-MS/MS methods for phosphorylation site determination. Indeed, the validity of the protein phosphorylation site assignments that have previously been reported using ion trap ClD-MS/MS or MS3 analysis strategies should be re-evaluated prior to making any conclusions regarding their biological significance. It is recommended that this is achieved by either (i) taking into consideration the presence of sequences with known kinase motifs, (ii) by using alternative data acquisition methods such as ETD or ECD, or (iii) by comparison of the experimentally observed CID-MS/MS spectra with those obtained from the dissociation of independently synthesized phosphopeptide standards. 3.4 Other Recent Studies Demonstrating Intramolecular Gas-phase Phosphate Group Transfer Since the publication of this work on the intramolecular gas-phase phosphate group transfer of protonated peptides by CID,222 an analogous phosphate migration has been reported for deprotonated phosphopeptides by Lehmann and coworkers (Scheme 3.4).223 In that study, the phosphate group 157 transfer from phosphotyrosine residues to other functional groups such as hydroxyl and carboxyl followed by H3PO4 neutral loss was investigated. Although no sequence product ions containing the transferred phosphate were observed, ClD-MS/MS of differentially labeled peptides suggested phosphate group transfer from phosphotyrosine to carboxyl groups. In particular, fragmentation of the deprotonated phosphopeptides studied commonly gave rise to the neutral loss of 140 Da, which was attributed to the concurrent losses of 98 Da (H3PO4) from the transferred phosphate and 42 Da (HN=C=NH, 42 Da) from the C-tenninal arginine residue by the mechanism shown in Scheme 3.4.223 Note that this proposed mechanism involves the removal of one of the oxygen atoms from the C-terminal carboxyl as part of the neutral loss of H3PO4. Therefore, upon CID-MS/MS of an 180 C-terrninal carboxyl labeled peptide, the neutral loss would be expected to be 142 Da or H3P1603180 (100 Da) and HN=C=NH (42 Da). In fact, the neutral loss of 142 Da was observed for a deprotonated phosphotyrosine-containing peptide that was 180 labeled on the C- temninal carboxyl upon ClD-MS/MS, providing confirmation of the migration of the phosphate group.223 158 OH _ |+/-\. NH — °Ei"°“ 0 ca - 7 ° )4 ”3 0 A n O NH HO. 1,0 H’VVNH CID-MS/MS ’P\O P > HO L ,NH hosphate CH 0 CH2 Transfer 0 CH2 0&( 23’ H'JL N NH N H 0 H O L— _ - H3PO4(98 Da) - HN=C=NH (42 Da) HN 0 CH2 0 A WW N H 0 Scheme 3.4 Proposed mechanism for intramolecular phosphate group transfer for deprotonated peptide ions. 3.5 Additional Commentary Certainly the results reported in this chapter, which demonstrate that intramolecular gas-phase phosphate group transfer occurs, imposes limitations on the effective use of ClD-MS/MS spectra for accurate and unambiguous assignment of phosphorylation sites in ion trap instruments. However, these results do not imply that CID is not capable of accurate phosphorylation site assignments under these conditions. Moreover, these results do not imply that CID should be replaced by alternative dissociation methods such as electron capture dissociation (ECD) and electron transfer dissociation (ETD), which 159 maintain the phosphorylation site. In fact, the information derived from CID and ECD/ETD spectra may largely overlap, despite the complications originating from the Iability of the phosphate group and the potential for phosphate group transfer by CID. This is because multiply protonated ions have a lower propensity to undergo phosphate group neutral losses and intramolecular phosphate group 87, 89, 111, 206, 222 transfer upon CID, and ECD/ETD are only amenable to the analysis of multiply protonated ions. In brief, ECD/ETD may result in redundant information to that achieved by CID in the analysis of multiply protonated peptides. In light of this discussion, however, it is clear that the analysis of singly protonated peptide ions and peptide ions with low proton mobility by these established mechanisms are limited and the results from such analyses may require validation using alternative dissociation methods of those ions (e.g., using ultraviolet photodissociation, femtosecond laser-induced ionization/dissociation, metastable atom-activated dissociation; discussed in detail Chapter One), the consideration of those suggestions listed in the Conclusions section of Chapter Four, or previously published data on related multiply protonated species. 160 CHAPTER FOUR Experimental Methods for Gas-Phase Fragmentation Studies of Phosphopeptides (Chapters Two and Three) 4.1 Materials All aqueous solutions were prepared using deionized water purified by a Bamstead nanopure diamond purification system (Dubuque, IA). Sequenal grade trifluoroacetic acid (TFA) was purchased from Pierce (Rockford, IL). Peptides were prepared using manual stepwise Fmoc-based solid-phase peptide synthesis (SPPS) on Wang resins pre-loaded with N-q-Fmoc protected amino acids (100-200 mesh) (EMD Biosciences, San Diego, CA). N-a-Fmoc protected acid building blocks with acid-labile orthogonal side-chain protection, where appropriate, were purchased from EMD Biosciences (San Diego, CA) and are listed in Table 4.1. Boc-glycine-OH and dibenzyI-N,N-diisopropylphosphoamidite were also purchased from EMD Biosciences. Anhydrous dimethylforrnamide (DMF) was purchased from Spectrum Chemicals (Gardena, CA) and further dried with a 4 A molecular sieve. Reagent grade piperidine, O-(Benzotriazol-1- yI)-N,N,N’,N’-tetramethyluronium tetrafluoroborate (TBTU), 1- hydroxybenzotriazole hydrate (HOBt), diisopropylethylamine (DIPEA), ninhydrin, phenol, potassium cyanide, pyridine, ethanedithiol (EDT), triisopropylsilane (TIS), and monodeuterated acetic acid (CH3COzD) were purchased from Sigma Aldrich (St. Louis, MO). Deuterium oxide (D20) and deuterated methanol (CD30D) were 161 obtained from Cambridge Isotope Laboratories (Andover, MA). D,L-Serine-2,3,3- d3 (99.4 atom % D) was purchased from CDN Isotopes Inc. (Quebec, Canada). Recrystallized 2,5-dihydroxybenzoic acid (2,5-DHB) was purchased from Laser BioLabs (SOphia-Antipolis Cedex, France). HPLC grade methanol (MeOH) was purchased from Mallinckrodt (Hazelwood, MO) and HPLC grade acetonitrile (ACN) was purchased from EMD (Gibbstown, NJ). All other reagents were commercially available and used without further purification. 162 Table 4.1 Common amino acid building blocks used for solid-phase peptide synthesis. Abbreviation Solid Phase Peptide Synthesis Amino Acid 3 Letter 1 Letter Building Block Alanine Ala A Fmoc-AIa-OH Arginine Arg R Fmoc-Arg(Pmc)-OH Asparagine Asn N Fmoc-Asn(Trt)-OH Aspartic Acid Asp D Fmoc-Asp(OtBu)-OH Cysteine Cys C Fmoc-Cys(Trt)—OH Glutamic Acid Glu E Fmoc-Glu(OtBu)—OH Glutamine Gln Q Fmoc-Gln(Trt)-OH Glycine Gly G Fmoc-Gly-OH Histidine His H Fmoc-His(Trt)-OH Isoleucine lle | Fmoc-IIe-OH Leucine Leu L Fmoc-Leu-OH Lysine Lys K Fmoc-Lys(Boc)—OH Methionine Met M Fmoc-Met-OH Phenylalanine Phe F Fmoc-Phe—OH Phosphoserine pSer pS Fmoc-Ser(PO(Obzl)OH)—OH Phosphothreonine pThr pT Fmoc-Thr(PO(Obzl)OH)-OH Phosphotyrosine pTyr pY Fmoc-Tyr(P03H2)-OH Proline Pro P Fmoc-Pro-OH Serine Ser 8 Fmoc-Ser(tBu)-OH Threonine Thr T Fmoc-Thr(tBu)-OH Tryptophan Trp W Fmoc-Trp(Boc)«OH Tyrosine Tyr Y Fmoc-Tyr(tBu)-OH Valine Val V Fmoc-VaI-OH 4.2 Phosphopeptides The phosphoserine- and phosphothreonine—containing peptides GAILpSGAILK, GAILpTGAILK. GAILpSGAILR, GAILpTGAILR. LFTGHPEpSLEK, LFTGHPEpTLEK, LRRApSLG, LRRApTLG, LFpTGHPESLER (pTSR), and LFTGHPEpSLER (TpSR) were prepared by manual stepwise Fmoc- based solid-phase peptide synthesis, as described in detail below. Detailed experimental methods for the synthesis of the regioselectively deuterated peptide GAIL(d3-D,L)pSGAILK also follow. All other phosphopeptides mentioned in Chapter 3 were synthesized by Sigma-Genosys (The Woodlands, TX) and used without further purification. 4.2.1 Fmoc Solid-Phase Peptide Synthesis (SPPS) 4.2.1.1 Apparatuses Manual solid-phase peptide synthesis was completed in disposable 5 mL Toriq syringes with frits (Fisher Scientific, Pittsburgh, PA). Multiple peptides were synthesized simultaneously using Kontes three-way stopcocks with HDPE plugs (VWR, West Chester, PA) on a Vac-Man laboratory vacuum manifold (Promega, Madison, WI). All steps requiring continuous mixing were completed using a Labquake tube rotator (Thenno Fisher Scientific Bamstead, San Diego, CA). 4.2.1.2 Synthesis Generally, peptide synthesis was completed on a 0.05 mmol scale. The N-d-Fmoc protected C-terminal residue bound to a Wang-resin was swelled in DMF (5 mL) for 15 min with continuous mixing and washed with 2 x 5 mL DMF. C-terminal elongation of the peptide was completed via repetitive Fmoc deprotection and amino acid coupling until the target peptide sequence was 164 achieved. Cleavage of the peptide from the resin and acid-labile orthogonal protecting groups followed. 4.2.1.2.1 Fmoc Deprotection Deprotection of the resin-bound peptide N-terminal Fmoc group to form a free amine N-terminus was achieved by washing with 3 mL 30% piperidine in DMF followed by continuous mixing with 3 mL 30% piperidine in DMF for 25 min. The peptidyl resin was then washed with DMF until pH < 8 was achieved (approximately 8 x 5 mL). 4.2.1.2.2 Amino Acid Coupling N-d-Fmoc amino acids (0.25 mmol) were pre-activated by thorough mixing with DIPEA (65 pl) and a DMF solution consisting of 0.5 M TBTU, 0.5 M HOBt (0.5 mL). To increase coupling efficiencies, monobenzyl-protected N-d-Fmoc- phosphoamino acids were pre-activated using a 3—fold excess of DIPEA (195 pl). Coupling of the pre-activated building blocks to the peptidyl resin was completed by continuous mixing for 15 min. The peptidyl resin was then washed 5 x 5 mL DMF. Directly following coupling and washing, the Kaiser test was completed to determine the coupling efficiency. This color test determines the presence of primary amines, such as the N-terminus of a peptide, by reaction with ninhydrin. A dark blue color is generated upon reaction of ninhydrin with a primary amine . . . . . 224 . and a yellow color IS maintained If no reaction occurs. Therefore, here It was 165 used to indicate the completeness of the coupling of an N-d-Fmoc amino acid to the free amine N—tenninus of the peptidyl resin. Briefly, approximately 2 mg of peptidyl resin was placed in a small glass test tube and washed with 2 x 0.5 mL ethanol. Equal volumes (125 pL) of 0.28 M ninhydrin in ethanol, 42.5 M phenol in ethanol and 20 (M potassium cyanide in water/pyridine (2 mL 0.001 M KCN aq. in 100 mL pyridine) were added to the washed resin and mixed thoroughly. After incubation at 100 °C for 10 min, the color was visually inspected. Upon the formation of any blue resin, the same N-d-Fmoc amino acid was re-coupled to the resin, using the coupling procedure described above. 4.2.1.2.3 Resin Cleavage and Deprotection of Acid-Labile Orthogonal Protecting Groups Upon completion of the target peptide sequence, a final Fmoc deprotection was done. The peptidyl resin was dried by washing with 5 x 5 mL ethanol. Then the peptide was cleaved from the resin and any acid-labile orthogonal side chain protecting groups simultaneously by continuous mixing with a cleavage solution (5 mL) for 3 hr. The composition of the cleavage solution was dependent upon the composition of the peptidyl resin. A flow chart for its determination is summarized in Figure 4.1. Product peptides were precipitated in cold diethyl ether (20 mL), filtered, reconstituted in a minimum volume of 25% acetic acid (aq) solution and Iyophilized. Lyophilized crude peptides were stored at room temperature. 166 Does the peptide ® Does the peptide contain Arg(Mtr)or contain Cys(Trt) or Met? unprotected Trp? Does the peptide Cleave with: contain > 2 95% TFA, 2.5% H20, Arg(Mtr)? 2.5% TIS Cleave with: 94.5% TFA, 2.5% H20, ‘— V 0 0 Cleave with: 2.5 /o EDT, 1 /o TIS 1 M TMSBr/thioanisole in TFA with m-cresol/EDT Cleave with: 81.5% TFA, 2.5% EDT, 5% thioanisole, 5% phenol, 5% H20, 1% TIS figure 4.1 Flowchart to determine cleavage solution composition. Adapted from 4.2.2 Synthesis of GAIL(d3-D,L)pSGAILK The regioselectively deuterated peptide GAIL(d3-D,L)pSGAILK was prepared by the synthesis and incorporation of (d3-D,L)-Fmoc-Ser(O-tert- butyldimethylsilyl (TBDMS))-OH into the protected peptide Boc-GAIL(d3- D,L)S(O-TBDMS)GAILK(Boc). On-resin deprotection of the serine side chain and functionalization of the resin bound hydroxyl group using dibenzyl-N,N- diisopropylphosphoramidite followed. Oxidation, peptide cleavage, and extraction from the resin afforded the product phosphopeptide. Detailed experimental methods for this procedure are below. 167 4.2.2.1 Synthesis of (d3-D,L)-Fmoc-Serine-OH (4.a) A solution of (d3-D,L)-serine (150 mg, 1.39 mmol) in 10% Na2CO3 (aq) (7.5 mL) and 1,4-dioxane (3.75 mL) was treated dropwise with 9-fluorenylmethyl chlorofon'nate (395 mg, 1.53 mmol) in 1,4-dioxane (7.5 mL) at 0 °C. After 1 hr at 0 °C, the mixture was warmed to room temperature, and then stirred overnight. The resulting mixture was diluted with cold water (15 mL) and extracted with ether (2 x 20 mL). The aqueous layer was acidified with 5% HCI (aq) solution followed by extraction with ethyl acetate (2 x 20 mL). The combined organic phases were dried with MgSO4, filtered, and concentrated in vacuo resulting in an off-white solid (388 mg, 85%). The resulting Fmoc-protected amino acid 4.a was used without further purification. 1H NMR (300 MHz), (CD3)2CO: 5 1.29 (s, 1H), 4.25 (t, J = 7.1 Hz, 1H), 4.30-4.39 (m, 2H), 6.57 (s, 1H), 7.32 (ddd, J = 7.5, 1.1 Hz, 2H), 7.41 (t, J = 7.5 Hz, 2H), 7.74 (dd, J = 7.3, 4.0 Hz, 2H), 7.85 (d, J = 7.4 Hz, 2H), 11.01 (s, 1H); 130 NMR (75 MHz), (003)2co: 6 47.93, 56.79, 62.31, 67.32, 120.74, 126.13, 126.16, 127.89, 127.91, 128.48, 142.04, 144.97, 145.04, 157.02, 172.21. Chemical shifts are reported relative to the residue peaks of the solvent (CD3)2CO (2.05 ppm for 1H and 29.84 ppm for 13C). 4.2.2.2 Synthesis of (d3-D,L)-Fmoc-Serine(O-TBDMS)-OH (4.b) A solution of 4.a (250 mg, 0.76 mmol) in 2.5 mL dry DMF was treated with ted-butyldimethylsilyl chloride (148 mg, 0.98 mmol) and imidazole (2.27 mmol) at 0 °C while stirring under N2 (9) atmosphere. After 1 hr at 0 °C, the solution was 168 warmed to room temperature and then stirred overnight. The solution was diluted with 1 N HCI (aq) (5 mL) and extracted with ether (2 x 8 mL). The combined organic layers were concentrated in vacuo and the resulting residue was partitioned between 10% LiBr (aq) solution (5 mL) and ether (2 x 8 mL). The combined organics were dried with MgSO4, filtered, and concentrated in vacuo. The silyl protected product 4.b (210 mg, 62%) was recovered as a yellow oil after silica gel chromatography (EM Science Silica Gel 60 (230-400 mesh)). An initial mobile phase composition of 5% ethyl acetate, 95% hexanes was used to elute impurities followed by elution of the product with 80% ethyl acetate, 20% hexanes. 1H NMR (300 MHZ), (CD3)2CO: 6 0.08 (s, 6H), 0.90 (s, 9H), 4.25 (t, J = 7.0 Hz, 1H), 4.34 (d, J = 7.0 Hz, 2H), 6.45 (s, 1H), 7.32 (t, J = 7.3 Hz, 2H), 7.41 (t, J = 7.4 Hz, 2H), 7.72 (t, J = 7.1 Hz, 2H), 7.85 (d, J = 7.4 Hz, 2H), 11.03 (s, 1H); 13c NMR (75 MHz), (003)2co: 6 -5.34, -5.27, 18.83, 26.18, 47.94, 56.46, 63.54, 67.39, 120.79, 126.12, 126.14, 127.92, 128.52, 142.08, 144.94, 145.07,156.80, 171.90. 4.2.2.3 On-resin SPPS of Boc-GAIL(d3-D,L)pSGAILK(Boc) (4.c) The side chain protected peptide Boc-GAIL(d3-D,L)S(O- TBDMS)GAILK(Boc) was synthesized as described above in section 4.2.1 using 4.b for serine incorporation and Boc-eg-OH as the N-terminal amino acid. The peptidyl resin (0.05 mmol) was dried by washing with ethanol (2 x 5 mL) and dichloromethane (2 x 5 mL). Selective deprotection of the serine silyl protecing 169 group was completed by incubation of the dried peptidyl resin with a solution of tetrabutylammonium fluoride trihydrate (TBAF-3H20, 31.6 mg, 0.10 mmol) in dimethylacetamide (DMA) (2 mL) while rotating overnight. The peptidyl resin was washed with DMA (3 x 5 mL) and dichloromethane (3 x 5 mL) and dried overnight in vacuo. A 0.45 M tetrazole solution in acetonitrile (8.9 mL, 4.0 mmol) was mixed with dry DMF (5 mL) and dibenzyI-N,N-diisopropylphosphoramidite (1.0 g, 2.9 mmol). The solution was then added to the peptidyl resin and the mixture was stirred for 2.5 hr at room temperature under a N2 (9) atmosphere. The resin was washed with DMF followed by oxidation with 6 M t—butyl hydroperoxide (0.36 mL, 2 mmol) for 0.5 hr, resulting in 4.c. The resin was then washed with DMF and cleaved and extracted as described in section 4.2.1.2.3 affording GAIL(d3-D,L)pSGAILK. 4.2.3 Peptide Purification The product peptides GAILpSGAILK, GAILpTGAILK, GAILpSGAILR, GAILpTGAILR. LFTGHPEpSLEK, LFTGHPEpTLEK, LRRApSLG, LRRApTLG, and GAIL(d3-D,L)pSGAILK were purified by reversed phase-high performance liquid chromatography (RP-HPLC) using an Aquapore RP-300 column (4.6 mm; Perkin Elmer, Wellesley, MA). Peptides were eluted at 1 mL/min with monitoring at 215 nm using a linear gradient elution from 0-100% B. Solvent A was 0.1% TFA in water and solvent B was 0.089% TFA, 60% acetonitrile in water. The product peptides LFpTGHPESLER (pTSR), and LFTGHPEpSLER (TpSR) were purified by RP-HPLC (Shimadzu, model LC-20AB, Columbia, MD) 170 using a Vydac Protein and Peptide C18 column (150 mm x 10 mm) (Grace, Hesperia, CA). Peptides were eluted at 5 mL/min with monitoring at 215 nm using the following gradient elution conditions: 045 min, 0-25% B; 15-35 min, 25- 35% B. Solvent A was 0.1% TFA in water and solvent B was 0.089% TFA, 60% acetonitrile in water. This method afforded separation of LFpTGHPESLER and LFTGHPEpSLER and their retention times were 22.8 min and 20.1 min, respectively. 4.3 Multistage Tandem Mass Spectrometry Analysis of Phosphopeptides 4.3.1 Phosphopeptide Sample Preparation All peptide samples were prepared in 50% methanol, 49% water, 1% acetic acid (v/v/v) for subsequent mass spectrometry analyses and stored at -20 °C. For H/D exchange experiments, peptides were prepared in 50% deuterated methanol (CD3OD), 49% deuterated water (D20), and 1% deuterated acetic acid (CH3002D) (v/v/v), Iyophilized, and then redissolved in the same deuterated solvent system to maximize exchange. Mass spectrometry analysis of the deuterated samples immediately followed sample preparation, purging of the source with N2 (9), and equilibration of the sample injection lines and syringe with deuterated methanol. For matrix-assisted laser desorption/ionization (MALDI) analysis, 5 pmol of each sample was mixed on-plate with 0.25 pL of a 2,5-DHB matrix solution (10 mg/mL in 50% acetonitrile, 50% water (v/v)). 171 Samples subjected to triple quadrupole mass spectrometry analysis were dissolved in 50% methanol, 50% water (v/v). 4.3.2 Mass Spectrometry Mass spectrometry analysis was performed using (i) an LCQ DECA 3D quadrupole ion trap mass spectrometer or (ii) an LTQ linear quadrupole ion trap mass spectrometer, each equipped with electrospray (ESI) or nano electrospray ionization (nESI) sources, (iii) an LTQ XL linear quadrupole ion trap mass spectrometer equipped with a vMALDl source, or (iv) a TSQ Quantum Ultra triple quadrupole mass spectrometer equipped with a chip-based nESI source (Advion NanoMate, Ithaca, NY). All mass spectrometers used were manufactured by Therrno Fisher Scientific (San Jose, CA). Samples subjected to ion trap ESI analysis (100 11M) (Chapter Two) were introduced by infusion at a flow rate of 5 pL/min. The spray voltage was maintained at 4.5 kV. Nitrogen sheath gas and auxiliary gas (H/D exchange only) were supplied at 25 (arbitrary unit) and 10 (arbitrary unit), respectively. The heated capillary temperature was set to 180 °C. Samples subjected to ion trap nESI analysis (20 pM for purified peptides and 50 pM for the crude peptides) (Chapter Three) were introduced by infusion at a flow rate of 0.5 pL/min. GAIL(d3-D,L)pSGAlLK (Chapter Two), however, was infused at a flow rate of 1.0 pL/min. The spray voltage was maintained at 2.1 kV and the heated capillary temperature was set at 200 °C. 172 ' r "."-".!'_'i Ion trap collision-induced dissociation (ClD)—MS/MS and -MS3 experiments were performed on mass selected precursor ions. Typical isolation widths were 5.0 m/z for singly protonated ions, 3.5 m/z for doubly protonated ions and 1.5 m/z for triply or higher protonated charge states. The automatic gain control (AGC) was set to 3.0 x 104 for full MS and 1.0 x 104 for MS". The activation time was maintained at 30 msec using an activation q value of 0.25, unless otherwise stated. For triple quadrupole mass spectrometry analysis, 13 uL of sample (20 pM) was loaded into a Whatman Multichem 96-well plate (Fisher Scientific, Pittsburgh, PA) and sealed with Teflon Ultra-Thin Sealing Tape (Analytical Sales and Services, Pompton Plains, NJ). Infusion mode operation was achieved using an ESI HD-A chip, a spray voltage of 1.4 W, a gas pressure of 0.3 psi and an air gap of 2 11L. The ion transfer tube of the mass spectrometer was maintained at 150 °C. MS and MS/MS spectra were acquired using scan rates of 500 mlz-second'1, while maintaining Q1 and QS peak widths at 1.0 Th. The Q2 argon collision gas pressure was set at 0.7 mTorr. All MS/MS and MS3 experiments were performed by collision induced dissociation (CID) on mass selected ions using standard isolation and excitation procedures. The spectra shown were typically the average of 20—200 scans (3 microscans / scan) and were collected in profile mode. Repeated analysis of individual samples resulted in less than 5% variation in relative product ion abundances. 173 4.3 Quantitative Analysis of Percent Product Ion Abundances Quantitative analysis of non-sequence percent product ion abundances was performed by initially filtering the spectra to remove any ions present below the empirically determined level of chemical noise (5-200 counts). Then, the % total product ion abundances were determined by expressing the summed abundances of product ions formed by the loss of H3PO4 (98 Da) or HPO3 and H20 (98 Da), CH20 (30 Da), CD20 (32 Da), or CH3CHO (44 Da) as a fraction of the total product ion abundance. 174 CHAPTER FIVE Application of Phosphoproteomic Analysis Strategy to Elucidate p65- Associated Phosphoproteins in the NF-KB Signaling Pathway From Treated and Untreated THP-1 Cell Nuclear Extracts 5.1 Introduction Since its first discovery in 1986,226 and more than 30,000 scientific publications later,227 the protein nuclear factor (NF)—KB has proven central in the coordination of innate and adaptive immunity and inflammatory responses in nearly all cell types. NF-KB exerts its activity by functioning as a transcription factor, regulating over 150 genes that are involved in the development of immune cell function, promotion of inflammatory mediators, inhibition of apoptosis, and regulation of cellular growth.(rew'med m 228-230) These genes include those that produce pro-inflammatory cytokines and chemokines (e.g., lL-6, TNF-a, and IL- 8)231'234, cell adhesion molecules (e.g., lCAM-1)235, growth factors (e.g., FGF- 236 8) : and apoptosis regulators (e.g., Bcl-2 and caspase11)237' 238 . Due to its biological importance, it is not surprising that aberrant regulation of NF-KB and the associated biological pathways that control its activity have been implicated in many different diseases including arthritis, asthma, diabetes, and AIDSWViewed in 239’ 240) In addition, the NF-KB pathway has been implicated in a variety of processes leading to the development and progression of cancer, 175 including tumor proliferation, prevention of apoptosis, metastasis, and . . ' d' 231, 241-24 . angrogenesrs.(remewe m 3) For these reasons, the mechanisms by which NF-KB exerts its influence, specifically pertaining to its activation and modulation of its transcriptional activity, have been of great interest for the - . 44- 4 treatment and prevention of cancer and other diseases.2 2 6 NF-kB consists of different homo— and hetero-dimers comprised of five proteins: NF-kB1 (p50 and its precursor p105), NF-KB2 (p52 and its precursor p100), RelA (p65), RelB, and c-Rel. Each of these proteins has an N-terminal Rel homology domain (RHD), which contains a nuclear localization signal (NLS) and is involved in dimerization, sequence-specific DNA binding, and interaction (structures reviewed in 247) with inhibitory proteins. In addition to the RHD, p65, RelB, and c-Rel each contain non-homologous C-terminal transactivation domains, which are responsible for the recruitment of co-activators and basal (structures reviewed in 247) transcriptional machinery. In mammalian cells, the p50/p65 heterodimer is ubiquitous and is largely responsible for regulating inflammation as well as the life cycle of lymphoid cells during the immune response.248 For these reasons, the remainder of this introduction and the research described in Chapter Five will focus on the p50/p65 isoform, specifically highlighting the p65 subunit. (From this point on the discussion of “NF-KB” will refer to the p50/p65 isoforrn.) In unstimulated cells, NF-KB is negatively regulated in the cytoplasm by its interactions with inhibitory (I) KB proteins (illustrated in Figure 5.1). Activation of 176 the NF-KB pathway may be initiated by a host of cellular stimuli, including pro- inflammatory cytokines, such as tumor necrosis factor (TNF)-d (as shown in Figure 5.1)249; microbial components, such as Iipopolysaccharide (LPS)250; and ' . 2 1 . . ' d' 252 cytotoxrc agents, such as chemotherapeutics, 5 ox1dative stress,(rev'ewe '" ) and UV light.253 In response to such inducers, the upstream IKB kinase (IKK) complex is activated by phosphorylation, which in turn phosphorylates IKB, subsequently targeting it for ubiquitination and proteasomal degradation (Figure 5.1). This liberates the p50/p65 dimer, allowing nuclear translocation and transcriptional regulation of target genes, resulting in variety of cellular responses. 177 Extracellular matrix mm mm writ? :ff “ ‘ j" Cytoplasm i_ Serine ‘ kinases IKK kinases %® L» 8%" Inhibited NF-KB J‘— 9 6 9° .36 _ f—“K IKB degradation Gene transcription Cell survival > Inflammation Immunity \ \ \ \ F \ | I Figure 5.1 The NF-KB pathway. Adapted from 254. 178 Given the host of cellular stimuli that may induce the NF-KB pathway and the variety of NF-kB gene targets, many regulatory mechanisms are likely involved that relate a specific stimulus to a specific NF-KB-induced cellular response (i.e., NF-KB gene specificity and the resulting translational profile) in a specific cell line. Regulatory mechanisms of the NF-KB pathway may be divided into two distinct parts: (i) those involving translocation of NF-KB into the nucleus and (ii) those involving mediation/promotion of transcriptional activity. It is generally accepted that regulation of NF-KB nuclear transport, regardless of stimulus, originates from activation of the IKK complex by phosphorylation, which subsequently phosphorylates IkB, leading to dissociation of the IKB-NF-KB complex in the cytosol (as shown in Figure 5.1).(mwev‘led in 227’ 255) Although it is known that phosphorylation is the key regulatory mechanism for IKK activation (reviewed in 255) and subsequent NF-KB nuclear localization, the kinases that are directly responsible are largely unknown. Additionally, the mechanisms that ultimately converge to enable phosphorylation/activation of IKK by such diverse cellular stimuli also remains to be answered. Numerous studies have demonstrated that NF-KB transcriptional activity, including DNA binding, gene specificity, and resulting translational profile, is complex, and depends on phosphorylation of NF-KB and involves the interactions KB (reviewed in 255-258;259) of numerous proteins with NF- In fact, for the p65 subunit alone, ten phosphorylation sites have been elucidated (summarized in Figure 5.2, including the responsible kinases and phosphatases involved), many 179 of which mediate NF-kB transcriptional activity.(reVIeWGd m 256’ 257) For example, phosphorylation of p65 at Ser536 has been shown to enhance transcription of interleukin-8 (IL-8), which is responsible for production of pro-inflammatory chemokines.260 Another example includes phosphorylation of Ser468, which, interestingly, has been described to both enhance and suppress p65 transcriptional activity under different conditions.261'263 The differential function of Ser468 phosphorylation may arise from a difference in phosphorylation status of other p65 sites (e.g., phosphorylation of both Ser468 and Ser536 under one condition and phosphorylation of only Ser468 in another condition). Or, perhaps, the difference in p65 activity despite constant phosphorylation of Ser468 corresponds to a difference in the transcriptional co-activators or other p65- interacting proteins involved in p65—transcriptional regulation. In fact, greater than 130 proteins have shown to directly interact with the p65 subunit alone and have demonstrated involvement in NF-KB-mediated transcriptional activity.257 For example, increased lL-4 gene expression, has been directly linked to the interaction between p65 and STAT6 (signal-transducer and activator of transcription 6) in cells stimulated with lL-4.264 Another example includes increased NF-kB transcriptional activity as a result of the nuclear association between p65 and the protein UXT (ubiquitously expressed transcript) in comparison with systems that have reduced UXT expression.265 180 [IKKa] [IKKB] [lKKs] [PKAc] [NAK] [MSK1/2] [2] [Chk2][RSK1] S276 8281 3529 S536 ['KKB] [?l [?l w lPKCfJ [Fig] [$515313] [Chk1] w 5205 T254 S311 T435 S468 T505 9 9 \/ 9 9 9 9 \/ | | RHD I l TAD I 1 19 301 428 551 Figure 5.2 Summary of the known phosphorylation sites of p65 and the corresponding [kinases] and [phosphatases*]. The kinases responsible for Ser205, T254, and Thr435 phosphorylation are25a6t Egg/sent unknown, indicated with [?]. This information was compiled from ; and adapted from 2 6 RHD is the Rel homology domain and TAD is the transactivation domain Given that phosphorylation of p65 is highly involved in the regulation NF- KB transcriptional activity, it is expected that the phosphorylation status of p65 influences the recruitment of p65 transcriptional co-activators and other NF-KB regulatory proteins. Additionally, the post-translational modifications of the p65- interacting proteins may also play a role in p65-interactions and subsequent regulation of p65-mediated transcriptional activity. In fact, a recent study has shown that tyrosine phosphorylation of STAT6 is required for both p65-binding to STAT6 and the resulting p65-STAT6-mediated IL-4 gene expression.264 Despite the significance of this result and the results of other individual studies described above that associate a specific phosphorylation event or a specific NF-KB- interacting protein with particular NF-KB cellular responses, many of these studies may not be comparable or may not be generally applicable. This is because some of these studies have involved overexpression of either p65, the 181 interacting protein, or both and, therefore, may not be truly representative of “real” systems. Moreover, each study varies in the stimulating agents, concentrations of the stimulating agent, incubation times, and cell lines used, resulting in different phosphorylation events and different p65 interactions with different proteins. Accordingly, a true representation of the NF-KB regulatory mechanisms and the relationship between stimulus and NF-KB activity would require global examination of the stimulus-specific interacting proteins, their post- translational modifications, and the stimulus-specific NF-KB transcriptional response. Such a global study would involve a variety of stimuli, treatment regimens, and cell types. Additionally, the analysis technique used in such a study should not be limited by the post-translational modifications that may be examined and should be able to unambiguously and accurately localize those modifications. To date, however, such a comprehensive study that would elucidate context-specific regulatory mechanisms and NF-KB activity has yet to be completed. To aid in this achievement, here, a phosphoproteome analytical approach was developed and applied to determine the p65-associated phosphoproteins from the nuclear fractions of untreated and TNF-d-treated THP-1 acute monocytic leukemia cells. A schematic of the analytical strategy employed is shown in Figure 5.3. Briefly, p65-associated proteins were co- immunoprecipitated with nuclear p65 and the resulting protein mixture was proteolytically digested with trypsin. The phosphorylated peptides were then enriched using a titanium-based method followed by collision-induced 182 dissociation tandem mass spectrometry analysis. The resultant product ion spectra were interpreted by database searching and manual analysis for protein identification and phosphorylation site localization. Importantly, for phosphorylation site assignment of the identified phosphorylated peptides, the effects of gas-phase ion chemistry of phosphorylated peptides (as discussed in Chapters Two and Three) on the formation of analytically relevant information was considered during the downstream data interpretation. Using this process, five p65-associated proteins were identified, two of which have not been previously reported to interact with p65. 183 gt..- hp- 0330 0030505000330 n 03 033x38 0H=M0d ... .. fl...” , bu EOEcotc0 i o a 9gb 3 WM 0030003005 00000-253290 0.6300303 0.303 Edie-:0 5085. 0:0 coszocamoca HOV 0503030 02300035600 0030:0_0050::EE_-00 w mom 0:0 cozflbocamoca "9 00050000500050 :0_30c_E..0300 0:0 cozflbocawoca w cozmoEE0E E03030 20.00.303.035 023 0:00am ...... «0 %9...M99 ...... @033...» - 5858-000 5:. 20.330333360323553 .33 c0200: 300.032 . ll- 000E E00009 0:32-004 m. 02-5.33 200300330: ..0 0.0230 05:30 __00 Tax... - ) -5590 fl. kgb§00 Figure 5.3 Schematic of the phosphoproteome analytical strategy employed for identification and phosphorylation site determination of p65-interacting proteins from the nuclear fractions of untreated and TNF-d-treated THP-1 cells. 184 5.2 Results and Discussion 5.2.1 Development of the Phosphoproteome Analysis Strategy Employed for Identification of p65-Associated Phosphoproteins From Differentially Treated THP-1 Cells To determine the identity and phosphorylation sites of the p65-associated phosphoproteins from the nuclear fractions of TNF-d-treated and untreated THP- 1 cells, immunoprecipitation of these samples was completed using a p65 polyclonal antibody (p65 C-terminal epitope) that had been covalently cross- linked to agarose beads. The co-immunoprecipitated, p65-associated proteins were then analyzed using the phosphoproteome approach described in detail below. 5.2.1.1 Proteolytic DigestiOn The immunoprecipitated proteins were digested directly on the immunoprecipitation resin, using trypsin. On-resin digestion was used here because it enables decreased risk of sample loss as compared to in-gel digestion or in-solution digestion post protein elution, which may incur losses from inefficient peptide extraction from the gel or inefficient protein elution, respectively. Additionally, utilizing the covalently cross-linked IP resin decreases the presence of extraneous p65 antibody peptides, by decreasing the number of tryptic sites available through modification of the lysine residues on the p65 anfibody. 185 5.2.1.2 Phosphopeptide Enrichment Phosphopeptide enrichment was completed by Mr. Weihan Wang (Professor Merlin Bruening’s group, Department of Chemistry, Michigan State University) on a matrix-assisted laser desorption/ionization (MALDI) sample plate, utilizing a titanium-based metal-oxide affinity chromatography (MOAC) method developed by Mr. Weihan Wang and Professor Merlin Bruening. On- plate phosphopeptide enrichment methods have exhibited high recovery of phosphopeptides (70%) from mixtures containing a few proteins.(reViewed in 50) Additionally, these on-plate methods enable decreased sample handling and, therefore, decreased sample loss, and are convenient for samples with small volumes (“Lionewewe‘d in 50) As a result, on-plate phosphopeptide enrichment strategies are optimal for immunoprecipitation samples and hence were used here. 5.2.1.3 Mass Spectrometry Analysis and Database Searching MALDI-MS and ClD-MS/MS analyses of the enriched phosphopeptides were performed using a linear ion trap mass spectrometer. MALDI-MS allows for decreased complexity in the MS spectrum because only singly protonated ions are formed, as opposed to electrospray ionization (ESI), where multiple protonated charge states for each peptide would be present. Also, MALDI is compatible with the upstream on-plate enrichment strategy. The work described in Chapter Two demonstrate that the precursor ions targeted here (singly protonated phosphorylated tryptic peptides i.e., precursor 186 ions with low proton mobility) under ClD—MS/MS conditions acquired on the MALDI-ion trap instrument platform promote predominant formation of the nonsequence product ion resulting from the neutral loss of 98 Da. As a result, these nonsequence product ions were used for this work as diagnostic markers for the presence of phosphorylated peptides, enabling targeted mass spectrometry analyses. However, as discussed in Chapter Two, when the product ion resulting from the neutral loss of 98 Da is dominant, the formation of sequence ions, which may be used for peptide identification and localization of the phosphorylation site, may be limited. To overcome this potential limitation, strategies involving CID-MS3 of “the initial 98 Da neutral loss precursor ion has been used in other studies to obtain the sequence information and 24. 98-101 phosphorylation site localization. However, the work described in Chapter Three, demonstrate that the sites of phosphorylation should not be determined from product ion spectra acquired using such CID-MS3 strategies, since the initial neutral loss product ion may represent a mixture of ions. For these reasons, CID-MS3 activation was not used here for that purpose. Instead, high quality ClD-MS/MS product ion spectra were achieved by averaging between 500-3000 scans (with 3 microscans), such that sequence ions could be clearly observed above the level of noise for peptide identification and phosphorylation site determination. Additionally, isolation and activation of the protonated precursor ion was limited to either the monoisotopic ion or the monoisotopic ion and the 130 isotope (the first heavy isotope), to decrease the 187 amount of signal present in the product ion spectrum resulting from activation of extraneous ions with similar m/z. The work described in Chapter Three also demonstrate that the propensity for intramolecular gas-phase phosphate group migration is highest for precursor peptide ions with low proton mobility (i.e., singly protonated tryptic peptides), and using ClD-MS/MS in ion trap instruments, which are the conditions used in this study. As a result, unambiguous phosphorylation site analysis may be hindered. This may be particularly limiting when using database search algorithms for phosphorylation site determination, since they do not account for the possibility of such rearrangement reactions. However, this potentially limiting case does not apply for peptides that contain equivalent numbers of phosphate groups and potential phosphorylation sites (serine, threonine, and tyrosine residues). This was true for two of the phosphopeptides identified in this study. In cases where the number of phosphate groups was greater than the number of potential phosphorylation sites, the ambiguities of phosphorylation site assignment were decreased by considering phosphorylation sites that have been previously identified from product ion spectra obtained using alternative dissociation techniques to CID. Additionally, all ClD-MS/MS product ion spectra were manually interpreted to obtain partial sequence information, information pertaining to the amino acid composition, and other modifications (e.g, methionine oxidation) to enhance the specificity of the database search as well as enable validation of those search results. In many cases, manual interpretation also enabled elucidation of the phosphate-intact sequence ions. 188 Importantly, manual interpretation was completed considering the potential for the formation of sequence product ions resulting from (i) the neutral losses 98 Da (H3PO4 or HPO3+H20), 80 Da (HPO3), 18 Da (H20) and combinations thereof; and (ii) intramolecular gas-phase phosphate group transfer. For this study, peptide identities and the proteins from which they are derived were determined using the Mascot database search engine. The definable search parameters for each Mascot query include: taxonomy; proteolytic enzyme; number of proteolytic missed cleavages; constant and variable modifications (either post-translational or process-induced); and precursor and product ion mass tolerances. To determine an initial set of search parameters as well as to empirically determine the tolerances of those parameters for accurate identification of phosphorylated peptides, Mascot was used to examine the ClD-MS/MS spectra of singly protonated synthetic “tryptic” phosphorylated peptides (data not shown). From this study it was found that trypsin (cleavage C-terminal to KR, but not when P is the subsequent residue) could be chosen as the proteolytic enzyme allowing for up to three missed cleavages. However, for the real samples, if the search results did not return a manually validated peptide, then “semi-trypsin” was chosen as the proteolytic enzyme to account for peptides that have tryptic specificity at one terminus, but where the other terminus may be a non-tryptic cleavage. No constant modifications were chosen. From the standard peptides, it was determined that if the modifications that were present in the peptide were not chosen as a variable modification for the search, then the correct peptide could not be identified. For 189 this reason, an initial set of variable modifications were chosen including: oxidation of methionine; S-carboxyamidomethylation of cysteine; and phosphorylation of serine, threonine, and tyrosine residues. Additionally, if no result could be manually verified, then other modifications such as lysine acetylation were also included in the search. Based on the analysis of the synthetic peptides, mass tolerances of 10.5 Da for both the precursor and product ions were used to optimize the search specificity, while decreasing the resulting number of false positives. 5.2.2 Differentially Treated THP-1 Cells and Resultant Differential Levels in Nuclear p65 THP-1 cells, which activate the NF-KB pathway upon stimulation with a variety of known stimulating agents including TNF-a,266'268 were used in this study. The nuclear fraction from cells treated with TNF-o (10 ng/mL for 0.5 hr) and untreated cells were extracted, and the amount of nuclear p65 was determined by Western blot (Figure 5.4). Similar to the results from previously published studies,268-270 the TNF-o-treated cells exhibited significantly higher levels of nuclear p65 than the untreated cells. 190 Untreated TN Fa Treated I j I l 1 2 1 _ 2 _ReplicatelD M ”M ”r Actin Figure 5.4 Western blot detection of p65 (total, C-terrninal epitope) and actin from replicate untreated and TNF-a-treated THP-1 nuclear extracts. Western blot of the actin loading control indicates that the untreated and TNF-o-treated THP-1 nuclear extracts contain comparable amounts of protein; and therefore, that the amount of nuclear p65 is greater for the TNF-a-treated nuclear extract than the untreated nuclear extract. This gel was edited to enable facile comparison between the untreated and treated samples. The unaltered image is shown in Figure 5.15. ' Since the TNF-o stimulated cell nuclear extracts contained significantly more p65 than their untreated counterparts, immunoprecipitation of p65 and its interacting proteins was initially completed for the treated cell samples only. At a later time (~2 months later), the untreated cell samples (prepared on the same day as the treated cellular extracts) were also immunoprecipitated. Western blot analysis of p65 from these immunoprecipitated samples is shown in Figure 5.5. As expected, the amount of p65 immunoprecipitated from the nuclear fraction of the TNF-o-treated cells resulted in significantly higher nuclear p65 than that of the untreated cells. Identification of the p65-associated proteins was completed for both the trypsin digested, phosphopeptide-enriched nuclear fractions of the treated and untreated THP-1 cells. Discussion involving the protein identification 191 process, comparison of the results from the treated and untreated samples, and the relevance of the identified proteins is given in detail below. IPT |P* * - f + TNFd *1" . _.. . 'l’i'. .' . “1-3131. .... Figure 5.5 Western blot detection of p65 (total, N-terminal epitope) from p65- immunoprecipated from untreated (IPT) and TNF-o-treated (IP ) THP-1 nuclear extracts. The immunoprecipitation of p65 from the untreated and treated samples were completed on separate occasions. For this reason the experiments are differentiated here. Control experiments involving immunoprecipitation of p65 from phosphate-buffered saline are also shown (*). Higher amounts of nuclear p65 is present in the treated than the untreated samples after p65-immunoprecipitation, consistent with the results shown in Figure 5.4 for the samples prior to p65-immunoprecipitation. The unaltered image is shown in Figure 5.16. 5.2.3 Identification of p65-Associated Phosphoproteins From TNF-q- Treated THP-1 Cells A representative mass spectrum of the phosphopeptide-enriched tryptic digests of the nuclear extracts from TNF-d-treated THP-1 cells is shown in Figure 5.6. Nearly all of the ions present in the mass spectrum were determined to be peptides, as indicated by the appearance of their ClD-MS/MS product ion spectra. The only exceptions included those ions corresponding to matrix clusters (labeled as “m” in Figure 5.6), which were also present in the product ion spectra of standard peptides (data not shown). The observed predominant neutral loss of 98 Da in nine ClD-MS/MS product ion spectra indicated that those 192 precursor ions were phosphorylated peptides (labeled as “*” in Figure 5.6). This diagnostic ion was, however, not present in the product ion spectra of other peptide ions. Upon further investigation of these ions, three were identified as nonphosphorylated peptides (labeled as “o” in Figure 5.6) and some others were unfortunately not identified from their ClD-MS/MS product ion spectra (labeled as “7" in Figure 5.6). However, the unidentified peptides did not give rise to predominant 98 Da neutral product ions in their spectra, and therefore, may not be phosphorylated peptides. 193 3500 4000 3000 o 8 NE E ,_ ‘8 N-k o o o N ‘o 0) q—a s a. ._ co *7 ‘- 8 LE I!) N ‘— Z a '— ‘- (I) 5. a o —I 8 < ‘- E 0 o ‘— (%) eouepunqv eAueIea Figure 5.6 Representative MALDI-mass spectrum of the samples resulting from phosphopeptide enrichment of the trypsin digested p65-immunoprecipitated nuclear extracts from TNF-o-treated THP-1 cells, where * = identified phosphorylated peptides; O = identified nonphosphorylated peptides; ? = unidentified peptides; and m = matrix ions. ' 194 The ClD-MS/MS product ion spectra were used to identify the peptides, and, when present, their phosphorylation sites. The ClD-MS/MS product ion spectrum of the most abundant ion in the mass spectrum, m/z 2061, is shown in Figure 5.7. The spectrum contains product ions resulting from single and double neutral losses of 98 Da from both the precursor (labeled as “A” and AA, respectively, in Figure 5.7) and some sequence product ions (i.e., sequence ions that have differences in 98 Da), demonstrating that the peptide is multiply phosphorylated. Despite the significant amount of sequence product ions present, and the quality of the product ion spectrum, the identity of the peptide was initially not determined by either manual interpretation or by Mascot searching (while allowing for carboxyamidomethylation of cysteine; oxidation of methionine; and phosphorylation of serine, threonine, and tyrosine residues). However, upon further investigation of the ClD-MS/MS product ion spectra from other peptide ions in the mass spectrum (Figure 5.6), the identity of m/z 2061 was elucidated and the limitations that initially interfered with its identification were made clear, vide infra. 195 Figure 5.7 ClD-MS/MS product ion spectrum of m/z 2061 (from the MS spectrum shown in Figure 5.6). This peptide was identified as KEEpSEEpSDDDM(cam-105)GFGLFD from the P1 or P2 isoforms of the 608 acidic ribosomal protein, and resulting from the gas-phase neutral loss of 2- (methylthio)acetamide (105 Da) from the S-carbamidomethylmethionine (M(cam))-containing species (m/z 2166). A = —98 Da (—H3PO4 or - (H20+HPO3)); [:1 = —80 Da (-HP03); ° = -18 Da (—H20). 196 Figure 5.7 (cont'd) III N38 ooom com? com: 8.: OONF 00.9 com com -1...” 41. A .-1..- __4. 3.1.... _... . . _ . . . . , ....fl 84W . .3.» em a. 0n .. .o...o<<.....< cw > mi or... < < 8 <4 9 m; < 2 < n .. < o<< €8.20..- 32.583 0 o A........— .— 3...... 3.3.131 _ . ...... .Mrw M1 .. 333. 033.33.. 0M3333333 033 0.3.3.033 03.3 033 0333 03 333 0033 33 0 33 - 3030.00-33. . 330.33 . 0030 333...... 033 . 00033 00 3330.0- ......0%000 3 333.300-3000- 30303.- . z .383 "30333032033003.0me0? 0003 0333303330330 m cot. 0003 00.3 00.303 00.3 oo.N_. so: 083 com com com com com . . . ..H . M1 1. .. 3i... ._ . .... ...... N 0 Q m M. M 0.33 0033 0.... 0 33 33 3.33 0 33 3.333 0 33 - 30309- 3333 00 33 00 .33 .... - ..d, .3 3 m. m M ..n 333300.300- 5. 30300. . 3.. . 383 m w. MOEDQDmamwwawwv. tow £452-90 < .331. (%) a9UBPUI’ICI‘17' eAneleH (%) eouepunqv eAnelea 208 Figure 5.11 (cont’d) N33: 833 0003 0003 803 0003 803 833 803 000 000 83 80 80 I h! 1 up 4‘ 1‘. .. 1 . .1 u g - .u - n . L! p - a 4? MM M...— M 033 W .... - ... - W 8300:- 00 3U” 30003535 - m 003- - m .3./)0 3030. M83( .03emovzooomamwmfimx 003302.02d6 D 0003 803 0003 0003 803 833 0003 000 80 80 M1. . . .0 w 0 33 w. W. 30300:- 3.033 3 m . v q n U . D. e U .. 0 .. :D I 9 3300-300? o \/w 25m- _ r33 .-OOF“ “.030 TEmoMEODQmawmmammx omow m5:w.2-0_0 O 209 doub\y translc CID-h labeli‘ to the this s the 1 is m. poter phos Mas. site pro diff Dho Dh pm the Dh ha\ as The last phosphopeptideidentified from the TNF-o-treated sample was the doubly phosphorylated peptide (R)SRSFDYNYR(R) (m/z 1367) from the human translocation liposarcoma (TLS)—associated serine-arginine protein (TASR). The ClD-MS/MS product ion spectrum of m/z 1367 is shown in Figure 5.12, where the labeling represents the pSRpSFDYNYR isofon'n. The product ion corresponding to the 98 Da neutral loss is the most abundant ion in the spectrum, indicating that this species is phosphorylated. Other neutral loss product ions observed include the 196 Da (labeled as “AA") and 178 Da (labeled as “150”) indicating that the ion is multiply phosphorylated. Since there are two phosphate groups and four potential phosphorylation sites (two serine residues and two tyrosine residues), phosphorylation site assignment required consideration of many factors. The Mascot results indicated that the two serine residues were the phosphorylation sites and that assignment is reflected in Figure 5.12. Inspection of the assigned product ion spectrum demonstrate that there exists a y5/y7 ion pair with a difference of 163 03, indicating Ser3 phosphorylation. Additionally, no phosphate-intact sequence ions unambiguously corresponding to tyrosine phosphorylation were present upon manual interpretation. However, the potential for intramolecular gas-phase phosphate group transfer remains under these conditions, and gives rise to some ambiguity to the origin of any phosphate-intact sequence ions. Previous mass spectrometry-based studies have shown that Sen and Ser3 are phosphorylation sites, further justifying these 164, 276-279 assignments. Notably, electron transfer dissociation tandem mass 210 spectrometry (ETD-MS/MS) has been used to unambiguously assign these phosphorylation sites of the identified quadruply protonated RRpSRpSRpSFDYNYR and triply protonated pSRpSRpSFDYNYR peptides, which were derived from trypsin digestion and Ti02 phosphopeptide enrichment of human embryonic 293T whole cell lysates.164 Note that under ETD-MS/MS conditions, the phosphate group remains intact and therefore the potential limitations associated with CID (i.e., phosphate Iability and phosphate migration) do not apply. Additionally, phosphorylation of Wm and Wm of this peptide has not been described previously. Considering all of this information, it is likely that m/z 1367 predominantly represents the doubly serine-phosphorylated peptide pSRpSFDYNYR. However, using an alternative tandem mass spectrometry method for activation of this ion from this sample would further validate this hypothesis. 211 Figure 5.12 ClD-MS/MS product ion spectrum of m/z 1367 (from the MS spectrum shown in Figure 5.6). This peptide was identified as pSRpSFDYNYR from the translocation liposarcoma (TLS)—associated serine-arginine protein (TASR). A = —98 Da (-H3PO4 or —(H20+HPO3)); E] = —80 Da (—HPO3); O = —18 Da (-H20). 212 N33: 091. com M. oow _. 00.. 3 ooor com com com com com oov ri'lllifldiilllulliuflflm w. 3 3.3 M . 0 D> . V 03 m. 3.0 3 m 30 m m 3 m w. m m > 2M>MQTM00 m 03. . ..u. 0 0 .3 0 .m 33 . \I ( << d 0/0 3300- _ mx 3. .oow(\ 03323333003.me 3003 02.02.06 Figure 5.12 (cont’d) 213 Since the proteins identified in this sample have other known phosphorylation sites, the corresponding precursor ion masses for those modified peptides were searched for in the MS spectrum. Additionally, phosphorylation of p65 may also occur, so the presence of those phosphopeptides was also examined for the TNF-o-treated sample. This was completed by allowing for tryptic cleavages only, with up to three missed cleavage sites for the 608 acidic ribosomal protein (P1 and P2 isoforms) and p65 and up to five missed cleavage sites for the TASR protein; as well as variable modifications including the known phosphorylation sites, 8- carbamidomethylcysteine, S-carbamidomethylmethionine, and other known post- translational modifications of those proteins (e.g., lysine acetylation). However, using these parameters, no other peptides from 608 acidic ribosomal protein (P1 and P2 isoforms), TASR, or p65 were identified from this sample. One plausible explanation for the lack of their presence includes that the peptides were not phosphorylated in this sample. Alternatively, if these phosphorylated peptides were present in the sample, they may not have been effectively enriched; they may not have been eluted from the enrichment material; they may not have ionized above the noise; or their masses were above or below the user defined mass range (m/z 800-4000). Nonphosphorylated peptides were also identified from the TNF-d-treated sample. Those peptides included HMYHSLYLK from ribosomal protein L19 (m/z 1191; product ion spectrum shown in Supplemental of Chapter Five Figure 5.12), AAIDWFDGK from the translocation liposarcoma (TLS) protein (mlz 1022; 214 pi product ion spectrum shown in Supplemental of Chapter Five Figure 5.18), and IWHHTFYNELR from actin (B isoforrn) (m/z 1515; product ion spectrum shown in Supplemental of Chapter Five Figure 5.19). Since TLS and actin are known to interact with p65 (as discussed in detail below),280'283 these nonphosphorylated peptides were likely non-specifically enriched during MOAC, but are likely specifically associated with p65. 5.2.4 Identification of p65-Associated Phosphoproteins From Untreated THP-1 Cells For comparison to the TNF-d-treated cell sample, immunoprecipitation, trypsin digestion, and phosphopeptide enrichment of p65 and its associated proteins was also performed for untreated THP-1 cells. Since p65 was present in significantly lower abundance in the untreated sample as compared to the TNF- a-treated sample (Figure 5.4 and Figure 5.5), it was expected that any protein that has specific interactions with p65 (rather than nonspecific interactions with the p65 antibody) would be present in lower abundance in the untreated sample than in the treated sample, assuming similar enrichment efficiencies. A representative MS spectrum of the untreated sample is shown in Figure 5.13. Similar to the treated sample, the KEEpSEEpSDDDMGFGLFD peptide (mlz 2061, 2109, 2125, and 2166) from the P1 and P2 isoforms of the 608 acidic ribosomal protein, the pSRpSFDYNYR peptide (mlz 1367) from TASR, and the IWHHTFYNELR peptide (m/z 1515) from actin were present. The most predominant ions in the spectrum, however, were those relating to the matrix 215 (labeled as “m” in Figure 5.13). In fact, many more matrix ions were present in the untreated sample (Figure 5.13) than in the treated sample (Figure 5.6), potentially due to a lower abundance of peptides in the untreated sample. 216 4000 o _o to co 0 _o o co 0 8 NQ‘ E o o o N 'o 9 'o <0 o 9 LD H _r- c D (D 5. D 8 -l o < 1- 2 (%) eouepunqv eAneIea Figure 5.13 Representative MALDI-mass spectrum of the samples resulting from phosphopeptide enrichment of the trypsin digested p65-immunoprecipitated nuclear extracts from untreated THP-1 cells, where * = identified phosphorylated peptides; O = identified nonphosphorylated peptides; ? = unidentified peptides; and m = matrix ions. 217 To compare the relative quantities of the peptides in the treated and untreated samples, the phosphopeptides from these samples were enriched in parallel and after elution, 100 amol of an internal standard phosphorylated peptide CD3CD2CO-LFTGHPEpSLEK (“053284 was added followed by addition of matrix. The resulting MS spectra for these samples is shown in Figure 5.14. By comparison, it is clear that the abundance of the ions identified from the TNF- d-treated sample are significantly lower than that in the untreated sample. These results rationalize the high abundance of matrix ions in the MS spectrum of the untreated sample (Figure 5.13). As a result, the corresponding proteins from which these peptides are derived are likely to be specifically associated with p65 in the cell. 218 4000 <1: 0'” E m do : 3 6 — E r E (U _ m - o O O ‘ '- o - o + + 3 U - U — 9 _ 9 _ m (U 5'9 4 _ g F _ C F 52 _ D h g z m m I— 2' E m 5 2' a] C‘. 5 E 0 MI co 8 < _x 8 N E 2 N — E LO . — o) O) * :3 ~ . -)< e .0 _ 'k to N _ O " N 1500 1 000 O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O (O V" N O co (.0 VI' N (D V N O CO (0 <1- N 1— ‘— v- ‘— 1— ‘— 1— ‘— Airsueiui Figure 5.14 MALDI-mass spectra of the (A) TNF-d-treated THP-1 cells and (B) untreated THP-1 samples resulting from p65 immunoprecipitation, trypsin digestion, phosphopeptide enrichment, and addition of 100 amol of an internal standard CD3CD200-LFTGHPEpSLEK (“D5", m/z 1398), where * = identified phosphorylated peptides; O = identified nonphosphorylated peptides; ? = unidentified peptides; and m = matrix ions. 219 5.2.5 Relevance of Identified Proteins Co-lmmunoprecipitated With p65 5.2.5.1 Ribosomal Proteins: 60$ Acidic Ribosomal Protein from the P1 or P2 Isoforms and Ribosomal Protein L19 Ribosomal proteins are largely thought of as pertaining to cytoplasmic translational processes, however, many studies have demonstrated that a number of ribosomal proteins have “extra-ribosomal” functions, including regulation of gene transcription, mRNA translation outside of the ribosome, and (reviewed in 285, 286) DNA repair. In fact, a study that aimed to determine a physical and functional map of the NF-KB pathway using tandem affinity purification (TAP) and key NF-KB proteins demonstrated that a variety of ribosomal proteins are involved in its regulation. 59 In fact, nine ribosomal proteins and four other proteins involved in ribosomal function were co-purified with p65.259 Additionally, a recent study has showed that the ribosomal protein S3 (RPS3) is a DNA-binding partner for both the p65 homodimer and the p50/p65 heterodimer, and enhances DNA-binding of those dimers to specific genes.287 These results are consistent with the identification of ribosomal proteins originating from co-immunoprecipitation with p65 demonstrated in this study. A brief description of these identified ribosomal proteins as well as the implications regarding their associations with p65 are discussed below. The large subunit (608) of eukaryotic ribosomes contain the P0, P1, and P2 isoforms of the 608 acidic ribosomal protein (collectively known as the “P- proteins”), whose sequences are highly conserved across species.288 Both the 220 P1 and P2 proteins have demonstrated involvement in peptide synthesis and interact with elongation factors (EF) 1 and 2, which facilitate translational elongation.289'291 A recent study has also demonstrated that the P1 and P2 proteins may modulate cytoplasmic translation by influencing the interaction between the two ribosomal subunits.292 Importantly, the P2 isoform of the 608 acidic ribosomal protein has been previously co-immunoprecipitated with key players of the NF-KB pathway, including IKK (y subunit), p52, NF—kB-inducing kinase (NIK), and RelB, indicating that it too plays a potentially important role in this pathway.259 However, no reports have been described that associate the P- proteins directly with p65. Consequently, the association between the P1 or P2 isoform of the 608 acidic ribosomal protein and p65 demonstrated here substantiates a function for these proteins in the NF-KB pathway. Since, at present, there is no evidence for a non-cytoplasmic role for the P-proteins, this identified association may result from a cytosolic contaminant or may indicate a potentially novel nuclear function for these proteins. Ribosomal protein L19 (RPL19) is part of the eukaryotic 60$ ribosomal subunit and its exact function has yet to be elucidated. Interestingly, over- expression of RPL19 has been associated with malignant prostate cells and . . . . . . 294 . tissues,293 IS differentially expressed in colorectal cancer tissue, and is up- regulated in human breast tumors.295 No known interactions between RPL19 and p65 have been reported to date. As a result, the demonstrated association 221 here may be novel. Similar to the association between the ribosomal P-protein and p65, this association could be cytoplasmic or the nuclear. 5.2.5.2 TLS and TLS-Associated Serine-Arginine Protein (TASR) Translocation liposarcoma (TLS), is a protein that has shown cellular involvement with the formation of the transcription factor (TF)IID complex and association with RNA polymerase II, implicating TLS in transcriptional activation.296'299 Recently, it has been shown that TLS enhances NF-KB mediated transactivation induced by TNF-d, interleukin-1B, and overexpression of NIK by its interaction with the p65 subunit.280 These results suggest that TLS acts as a co-activator of NF-KB and are consistent with its identification in this study resulting from its co-immunoprecipitation with nuclear p65. TLS-associated serine-arginine protein (TASR) serves as a splicing factor, processing pre-mRNA, and as the name implies, interacts with TLS.298' 300 Specifically, it has been shown that the C-temiinal region (residues 266 to the C- tenninal residue 526) of TLS links gene transcription and RNA splicing by 298, 300 recruitment of TASR to RNA polymerase II. However, when TLS constitutes part of an oncogenic fusion protein, where its C-tenninal region is replaced with another protein, it is unable to recruit TASR to RNA-polymerase 298, 301 II In the study discussed here, the TLS peptide identified constitutes residues 349-357, indicating that TLS is likely present in its wild-type form. Collectively these results suggest that the TASR protein co-immunoprecipitated 222 with p65 here may result from a higher order complex of p65, TLS, TASR, and RNA polymerase II. 5.2.5.3 Actin Actin is a highly conserved major component of the cytoskeleton and plays a critical role in a variety of essential cellular processes. These include traditional mechanical functions, such as cell motility, contractility, mitosis, cytokinesis, intracellular transport, endocytosis, and secretion.(rev'ewed m 302’ 303) And, within the last five years, it has been generally accepted that actin is also involved in the regulation of gene transcription, through sub—cellular localization of transcription factors and in the assembly of transcriptional ' ' 4- 7 . . (rev'ewed m 30 30 ) Importantly, interaction regulatory complexes in the nucleus. between p65 and actin has been previously demonstratedzm'283 In particular, it has been shown that the state of the actin cytoskeleton (e.g., filamentous) dictates the cytoplasmic localization of p65 via the p65-actin interaction in rat embryo fibroblastas.281 In addition, recent evidence has suggested that the cytoplasmic p65-actin interaction regulates p65 nuclear transport, and is sensitive to actin filament stabilization/destabilization in endothelial cells treated 282. 283 with thrombin. These studies, therefore, are consistent with the co- immunoprecipitation of actin with p65 demonstrated here. However, since actin is also a nuclear protein that is known to interact with transcriptional machinery 223 (reviewed in 304-307) (e.g., all three RNA polymerases), and the samples prepared here represent a nuclear fraction, the p65-actin interaction proposed in this study may suggest a nuclear role for p65 regulation. 5.3 Conclusions and Future Directions The results from this initial investigation indicate that the phosphoproteomic approach involving co-immunoprecipitation, phosphopeptide enrichment, and collision-induced dissociation tandem mass spectrometry-based methods holds promise for protein interaction discovery and signaling mechanism elucidation. Using this technique five p65-associated proteins were determined, two of which have not been previously reported. This study also confirmed previously published results that implicate a role for three identified proteins in the NF-KB pathway. Additionally, use of internal standards allowed for the comparative analysis of the extent of p65-protein association in treated and untreated cells. Collectively, these results broaden the perspective on the processes involved in the NF-KB pathway. Due to the success of this initial phosphoproteomic strategy, it is clear that a more global study involving a variety of cell types and treatment regimens is well within reach and would enable a greater understanding of the role of protein interactions and signaling mechanisms in biochemical pathways. Such a study would undoubtedly benefit from the maturation of the phosphoproteomic analysis strategy used here. Initial adaptations to this approach should include explorations of alternative phosphopeptide enrichment strategies. This is 224 particularly important considering that only two of the five proteins identified here were derived from their associated phosphorylated peptides. This likely resulted from limitations regarding the capacity of the phosphopeptide enrichment material and the enhanced enrichment efficiency (and adequate elution) of the peptides identified. In fact, all three distinct phosphorylated peptides (and their modified isoforms resulting from sample preparation) were doubly phosphorylated and two of those phosphopeptides were highly acidic, which may suggest that these are optimal peptide conditions for this titanium-based enrichment material. As a result, it is likely that other phosphopeptides were present within these samples, in particular singly phosphorylated peptides, but were either not enriched or not eluted from the enrichment material. 225 5.4 Supplemental Figures Untreated TN Fa Treated * 1 2 * 1 2 Replicate ID 250 150 100 75 p65 50 Actin 37 25 20 Figure 5.15 Unedited Western blot of p65 (total, C-terminal epitope) and actin from replicate untreated and TNF-d-treated THP-1 nuclear extracts. Replicate samples labeled with * were prepared previously (1 year prior to the preparation of replicates 1 and 2) by another laboratory member and had been stored at -80 °C. Due to their similar amount of p65 in the untreated and treated previously prepared samples, they were not used for any of the experiments described in Chapter Five nor were they included in Figure 5.4. 226 THP-1 NE IPT IPi - + * - * + TNFd 250 150 100 75 p65 50 37 IP3 25 20 Figure 5.16 Unedited Western blot of p65 (total, N-terminal epitope) from untreated and TNF-d-treated THP-1 nuclear extracts (THP—1 NE); and p65- immunoprecipated untreated (IP?) and TNF-d-treated (lPt) THP-1 nuclear extracts. The immunoprecipitation of p65 from the untreated and treated samples were completed on separate occasions. For this reason they are differentiated here. Control experiments including immunoprecipitation of p65 from phosphate-buffered saline are also shown (*).. An additional protein band (IPa) is present in the samples resulting from immunoprecipitation of the treated sample (lPi). However, because this band is also present in the PBS control sample, it IPa results from the immunoprecipitation process. 227 Figure 5.17 ClD-MS/MS product ion spectrum of m/z 1191 (from the MS spectrum shown in Figure 5.6). This peptide was identified as HMYHSLYLK from ribosomal protein L19. 0 = —18 Da (—H20). 228 OONM. oo _. —. coo _. com com CON com com Dov JMIM . m .30 ....om m33m .....m 3.33 0033 3 m ...... .... +—I+_>_M 00.... w> ... N> of omn > . Nooi own 0&3 0> 002 333 033 3300:- v_._>._mI>S_I 5:. 9252-90 (%) eouepunqv aAiielea 229 Figure 5.18 ClD-MS/MS product ion spectrum of mlz 1022 (from the MS spectrum shown in Figure 5.6). This peptide was identified as AAIDWFDGK from translocation liposarcoma (TLS) protein. 0 = -18 Da (—H20); * = —17 Da (— NH3). 230 3300:- x00m>>9<< coo Nmow 92392-90 o o _. (%) eouepunqv eAiiEieu 231 Figure 5.19 ClD—MS/MS product ion spectrum of m/z 1515 (from the MS spectrum shown in Figure 5.6). This peptide was identified as IWHHTFYNELR from actin (B-isoform). o = -18 Da (-H20). 232 N33: com? 00.3. com? com? 09. _. Door com com com com com 0.33 .4 30.0.30 03333. 0.33 000. 0090333 3300:- m4m2>uFII>>_ mwmr 92392-90 %) eouepunqv eAiieiea 003.. 233 CHAPTER SIX Experimental Methods for Applications of Mass Spectrometry Methods for Phosphoproteomic Analysis (Chapter Five) 6.1 Materials Primary p65 polyclonal antibodies (C-20, C-20 agarose conjugate, F6), protein A/G PLUS agarose, and normal goat lgG were obtained from Santa Cruz Biotechnology (Santa Cruz, CA). B—actin primary antibody and E2 Blue Coomassie stain were purchased from Sigma-Aldrich (St. Louis, MO). HyGLO HRP Western blotting detection reagents and XAR film were purchased from Denville Scientific (Metuchen, NJ) and secondary antibodies were obtained from Amersham Biosciences (Piscataway, NJ). Bradford reagent dye, materials used to prepare mini gels, and blocking grade non-fat dairy milk (NFDM) were from BioRad (Hercules, CA). 10% SDS-polyacrylamide gels were prepared according to the manufacturer’s protocol and were 0.75 mm in thickness. PBS, cellular growth media, cellular growth media supplements, and TNF-d were purchased from lnvitrogen (Carlsbad, CA). All aqueous solutions were prepared using deionized water purified by a Bamstead nanopure diamond purification system (Dubuque, IA). HPLC grade methanol (MeOH) was purchased from Mallinckrodt (Hazelwood, MO) and HPLC grade acetonitrile (ACN) was purchased from EMD (Gibbstown, NJ). Sequenal grade trifluoroacetic acid (TFA) was purchased from Pierce (Rockford, IL). Dimethyl pimelimidate (DMP) was purchased from Sigma Aldrich (St. Louis, MO). 234 Sequencing grade modified trypsin was purchased from Promega (Madison, WI). Recrystallized 2,5-Dihydroxybenzoic acid (DHB) was purchased from Laser Biolabs (Sophia-Antipolis, Cedex, France). Information regarding the reagents used for synthesis of the “D5 internal standard may be found in Chapter Four. D1o-propionic anhydride was obtained from CDN Isotopes (Pointe-Claire, Quebec, Canada). All other reagents were commercially available of biological grade and used without further purification. 6.2 Cells THP-1 cell line (ATCC, Manassas, VA) was cultured in RPMI-1640 medium, complemented with 10% fetal bovine serum, 2 mM L-glutamine, 100 units/ml penicillin, 100 pg/ml streptomycin, and 10 mM HEPES buffer without phenol red. Cells were propagated at 37 °C with 5% 002 and ambient oxygen. 6.3 Nuclear Extraction of THP-1 Cells Cells were seeded at 1.0 x 106 cell/mL in 150 mm x 15 mm plates (30 mL) the night before treatment. On the day of treatment, cells (20 replicate plates) were treated with TNF-o (final concentration 10 ng/mL) for 30 min at 37 °C with 5% C02 and ambient oxygen. Untreated cells (2 replicate plates) were maintained at 37 °C with 5% C02 and ambient oxygen. 235 Unless otherwise stated, all centrifugation steps were completed at 5000 xg for 5 min at 4 °C. Cells were harvested by centrifugation and the cellular pellet was resuspended in ice cold PBS (1 mL/10 mL initial cell volume). After centrifugation, the supernatant was aspirated and the cellular pellet was washed with ice cold PBS (1 mL/10 mL initial cell volume). The cellular pellet was lysed with cytoplasmic extraction buffer (50 pL/10 mL cells) (10 mM HEPES buffer, pH 7.9, 10 mM potassium chloride, 1.5 mM magnesium chloride, 1.0% Nonidet p-40, 0.5 mM dithiothreitol, 0.5 mM phenylmethylsulfonyl fluoride) supplemented with protease inhibitors (1 pg/mL leupeptin and 1 ug/mL aprotinin) and phosphatase inhibitors (1 mM sodium orthovanadate for phosphotyrosine phosphatases and alkaline phosphatase; 10 mM sodium fluoride for acid phosphatases; and 1 mM B-glycerophosphate for serine/threonine phosphatases) for 10 min on ice. The nuclei were pelleted by centrifugation and the cytosolic supernatant was aspirated and stored at -80 °C. Nuclei were lysed with nuclear extraction buffer (60 uL/ 10 mL initial cell volume) (20 pM HEPES buffer, pH 7.9, 0.42 M sodium chloride, 1.5 mM magnesium chloride, 25% glycerol, 0.2 mM EDTA, 0.5 mM dithiothreitol, 0.5 mM phenylmethylsulfonyl fluoride) supplemented with protease inhibitors (1 pg/mL leupeptin and 1 pg/mL aprotinin) and phosphatase inhibitors (1 mM sodium orthovanadate for phosphotyrosine phosphatases and alkaline phosphatase; 10 mM sodium fluoride for acid phosphatases; and 1 mM (3- glycerophosphate for serine/threonine phosphatases) at 4 °C for 20 minutes while shaking at 2000 rpm. Samples were centrifuged at 5000 xg for 15 min at 4 236 °C, and the supernatant of similar treatment was collected, combined, and transferred into chilled centrifuge tubes, and stored at —80 °C until further use. 6.3.1 Bradford Protein Assay: Determination of Protein Concentration in Cellular Extracts The Bradford protein assay utilizes the color change that Coomassie stain undergoes upon binding to protein.308 Here, this change was measured spectroscopically to quantify the concentration of protein in the cellular extract samples. Duplicate IgG standard solutions were made in 40 pL in water (0.000, 0.088, 0.175, 0.350, 0.700, and 1.050 ug/uL). Cellular extract samples to be quantified were prepared in duplicate by diluting 2 pL cellular extract in 38 pL water. Bradford reagent dye (500 pL) was added to each standard and sample and allowed to react for 5 min. Absorbance was measured at 595 nm on a SpectraMax M5e plate reader (Molecular Devices, Sunnyvale, CA). Protein concentration of each sample was then determined by fit to the lgG calibration curve. 6.3.2 Western Blot Analysis of p65 from Nuclear Extracts Nuclear extracts (30 pg total protein) were boiled for 3 min in 6x reducing Laemmli buffer (0.35 M tris, 600 mM dithiothreitol, 0.36 M SDS, 0.18 mM bromophenol blue in 30% glycerol) and were loaded onto 10% polyacrylamide gels, and electrophoresed for 50 min at 200 V in running buffer (2.5 mM tris base, 2 M glycine, and 10% or 35 mM SDS). The gel was wet transferred to a 237 PVDF membrane for 75 min at 60 V in Towbin buffer (25 mM tn's base, 19 mM glycine, 5% methanol). Membranes were blocked with 4% NFDM in tris-buffered saline-tween-20 (TBS-T: 50 mM tris base, 154 mM NaCl, 0.1% tween-20) for 1 hr at room temperature, followed by washing with TBS-T while rocking for 10 min at room temperature. The membranes were then probed with primary p65 antibody (C-20 rabbit polyclonal 1:1000 in 4% NFDM in TBS-T) while rocking overnight at 4 °C. The membranes were then washed 3xTBS-T while rocking for 10 min at room temperature, followed by incubation with horseradish peroxidase conjugated to secondary antibody anti-rabbit (1:3000 in 4% NFDM in TBS-T) while rocking for 1 hr at room temperature. The blots were washed 3 times with TBS-T while rocking for 10 min at room temperature and developed with HyGLO chemiluminescent detection for 1 min at room temperature and exposed on XAR film. 6.4 Sodium Dodecyl Sulfate-Polyacrylamide Gel Electrophoresis (SDS- PAGE) of lmmunoprecipitated Proteins from Nuclear Extracts 6.4.1 Preparation of Immunoprecipitation (IP) Matrix: Covalently Cross- Linked p65 Primary Antibody to Protein AIG-Agarose The p65 antibody was covalently cross-linked at its fragment crystallizable (Fc) domain to protein A/G-agarose using the diimidoester lysine cross-linking reagent dimethylpimelimidate. For the cross-linking process, all centrifugation steps were completed at 5000 xg for 3 min at 4 °C, unless otherwise stated. p65 goat polyclonal primary antibody agarose conjugate (up to 1 mL of resin) was 238 washed with 10 mL with ice-cold PBS and centrifuged. The pelleted matrix was washed with 3x10 mL 0.2 M borate-NaOH solution, pH 9 and centrifuged. Covalent cross-linking was achieved from incubation of dimethylpimelimidate (DMP, 15.5 mg, 0.06 mmol) dissolved in 2 mL 0.2 M triethanolamine-HCI, pH 9 (final concentration of DMP 30 mM) with the drained matrix while rotating for 30 min at room temperature. The suspension was centrifuged and the reaction was quenched by washing the resin with 10 mL 0.2 M ethanolamine-HCI, pH 9 while rotating for 5 min at room temperature. The suspension was centrifuged and the supernatant was aspirated. The resin was then incubated with 10 mL 0.2 M ethanolamine-HCI, pH 9 while rotating for 1 hr at room temperature. The suspension was centrifuged, the supernatant was aspirated, and the resin was washed with 2x10 mL PBS, 2x10 mL 100 mM glycine, pH 3, and 2x10 mL PBS each for 2 min while rotating at room temperature. The resin was resuspended in 0.02% sodium azide (w/v) in PBS to make a 25% slurry (v/v) and stored at 4 °C until further use. Efficiency of the cross-linking reaction was tested each time the reaction was carried out by running samples before and after the cross-linking reaction by SDS-PAGE and Coomassie staining. Equivalent samples (approximately 2 pg p65 antibody) were aliquoted before cross-linking (after the last 0.2 M borate- NaOH washes), after cross-linking (after the second PBS wash), and after cross- linking and stringent 100 mM glycine, pH 3 washes. These samples were microcentrifuged at maximum speed for 30 sec at 4 °C and suspended in 20 pL 2X SDS-PAGE buffer (10 mM tris pH 6.8, 0.2 M sodium dodecyl sulfate, 20% 239 glycerol, 0.01% bromophenol blue, 2% 2-mercaptoethanol). The samples were incubated in boiling water for 3 min and loaded onto a 10% polyacrylamide gel and electrophoresed for 50 min at 200 V in running buffer (25 mM tris base, 19 mM glycine, 3.5 mM SDS pH 8.3). The gel was washed three times with water while rocking for 5 min at room temperature to remove residual SDS. The protein bands of the gel were fixed with 50% methanol, 10% acetic acid (v/v) while rocking for 15 min at room temperature. Removal of the fixing solution was completed by washing the gel 2x5 min with water at room temperature. Protein bands were visualized with Coomassie stain while rocking for 2 hrs at room temperature. Residual Coomassie stain was removed by washing with water for 1-2 hr, changing water as necessary until the background of the gel was clear. Cross-linking was typically achieved with greater than 95% efficiency. 6.4.2 Immunoprecipitation of Cellular Extracts Using Covalently Cross- Linked p65 Primary Antibody to Protein AIG-Agarose All centrifugation steps used for immunoprecipitation methods were completed at maximum speed in a microcentrifuge for 1 min at 4 °C. Samples for immunoprecipitation (IP) were prepared with 1000 pg of total nuclear protein (quantified by Bradford assay; 20 replicates of TNF-d-treated nuclear extracts and 2 replicates of untreated nuclear extracts) and diluted with ice cold PBS to 800 pL. Blank samples (800 pL PBS) were also “immunoprecipitated". IP samples were precleared with 1 pg normal goat lgG 240 and 20 pL Protein A/G PLUS agarose and incubated while rotating for 30 min at 4 °C to ensure complete removal of endogenous lgGs. Samples were centrifuged and the supernatant was collected into chilled microcentrifuge tubes containing 200 pL cytoplasmic extraction buffer supplemented with protease and phosphatase inhibitors (see 6.3 Nuclear Extraction of THP-1 Cells) and covalently cross-linked p65 antibody to agarose (30 pL, 25% v/v). IP samples were incubated overnight while rotating at 4 °C. Samples were centrifuged and the supernatant was aspirated. The pelleted matrix was washed 2x500 pL cytoplasmic extraction buffer and 2x500 pL PBS. Drained samples were stored at -20 °C until further use. 6.4.3 Western Blot Analysis of p65 From p65-lmmunoprecipitated Nuclear Extracts of THP-1 Cells For Western blot analysis, IP samples were resuspended in 20 pL 6x reducing Laemmli buffer (0.35 M tris, 600 mM dithiothreitol, 0.36 M SDS, 0.18 mM bromophenol blue in 30% glycerol) and 5 pL aliquots (or 250 pg of initial total protein) were loaded onto 10% polyacrylamide gels in running buffer (2.5 mM tris base, 2 M glycine, and 10% or 35 mM SDS). For comparison, 30 pg of nuclear extract samples (without immunoprecipitation) were also loaded onto the gels. The gels were electrophoresed for 50 min at 200 V. The gel was wet transferred to a PVDF membrane for 75 min at 60 V in Towbin buffer (25 mM tris base, 19 mM glycine, 5% methanol). Following transfer, membranes were treated as 241 indicated above (6.3.2 Western Blot Analysis of p65 from Nuclear Extracts). The p65 F6 primary antibody was used to probe the samples post-IP. 6.5 ln-Solution Tryptic Digestion and Subsequent Phosphopeptide Enrichment of lmmunoprecipitated Proteins The immunoprecipitated samples were concentrated to dryness by Speed Vacuum. The sample pellets were denatured and reduced by incubation with 20 pL 6 M urea in 50 mM Tris-HCI and 5 pL of 10 mM dithiothreitol at 65 °C for 1 hr. After cooling, 160 pL 50 mM ammonium bicarbonate and 10 pL of 100 mM iodoacetamide were added, followed by incubation at room temperature in the dark for 1 hr. Sequencing grade modified trypsin (10 pL of 0.5 pg/pL solution) was added and protein digestion was performed overnight at 37 °C. The digestion was quenched by addition of 11 pL glacial acetic acid to achieve a 5% solution. The digested samples were then centrifuged for 1 min at maximum speed and the supernatant was transferred to 300 pL PCR tubes. The residual resin was washed with 20 pL water, centrifuged for 1 min at maximum speed, and the wash solution was collected and combined with the initial supernatant. The combined supernatant was concentrated and stored at -20 °C. Phosphopeptide enrichment of the tryptic digest samples was completed on polymer-oxotitanium hybrid materials on gold wafers by Mr. Weihan Wang (Professor Bruening’s group, Michigan State University). Briefly, lP samples were redissolved in 0.1% trifluoroacetic acid (final volume 20 pL) and 2 pL (10% of the total sample) aliquots were enriched using the oxotitanium enrichment 242 wafers prepared by Mr.. Weihan Wang. After enrichment and washing, phosphopeptides were eluted with 1 pL of 1% H3PO4, followed by addition of 0.25 pL of 40 mg/mL DHB (prepared in 1% H3PO4J 50% ACN), and co- crystallization. For incorporation of the 05 internal standard, enriched phosphorylated peptides were first eluted, followed by application of 0.5 pL of a 200 amol/pL solution of D5 and the matrix solution. 6.6 Synthesis of the lntemal Standard “D5” CD3CD2CO-LFTGHPEpSLEK Fmoc-solid phase peptide synthesis of the D5 internal standard was completed as described in Dunn, et al. and using the detailed methods in Chapter Four.284 Briefly, after C-terrninal elongation and Fmoc-deprotection of the N-terrninal leucine residue, acetylation was performed on-resin by addition of d1o-propionic anhydride (3 mmol eq.) and DIPEA (5 mmol eq.) in DMF, while shaking for 15 min. Cleavage of the resin and orthogonal protecting groups was achieved as described by Dunn, et al. (with consideration of Figure 4.1). Product peptides were purified by reversed-phase-high performance liquid chromatography (RP-HPLC) using an Aquapore RP-300 column (4.6 mm; Perkin Elmer, Wellesley, MA) and a linear gradient elution at a flow rate of 1 mL/min from 0-100 % B, where solvent A was 0.1 % TFA in water, and solvent B was 0.089 % TFA/60 % acetonitrile in water. The concentrations of aqueous D5 243 solutions were determined by amino acid analysis completed by the Genomics Technology Support Facility at Michigan State University. 6.7 Matrix Assisted Laser Desorption/Ionization and Multistage Tandem Mass Spectrometry Mass spectrometry analysis was performed using an LTQ XL linear quadrupole ion trap mass spectrometer equipped with a matrix assisted laser desorption/ionization source (vMALDl) (Therrno Fisher Scientific, San Jose, CA). Ion trap ClD-MS/MS and M83 experiments were performed on mass selected precursor ions. For precursor ions with m/z less than 2000, isolation and activation was completed on the monoisotopic precursor ion using an isolation width of approximately 1.1. For those precursor ions with m/z greater than 2000, isolation and activation was achieved using an isolation width of approximately 2.5 centered at the m/z of the monoisotopic precursor ion. The automatic gain control (AGC) was set to 3.0 x 104 for full MS and 1.0 x 104 for MS". The activation time was maintained at 30' ms using an activation q value of 0.25, unless otherwise stated. All spectra shown were typically the average of 500- 3000 scans (3 microscans / scan), were collected in profile mode, and are shown with a 5 point Gaussian smooth. MALDI-MS analysis of technical replicates resulting from individually immunoprecipitated samples and phosphopeptide enrichment of those samples gave rise to reproducible and consistent results similar to what is shown in Figure 5.6 and Figure 5.13 for the TNF-d-treated and untreated samples, respectively. 244 6.8 Peptide Identification by Database Searching The Mascot algorithm (http://www.matrixscience.com/; Matrix Sciences, UK) was used for interpretation of both uninterpreted MS/MS data (using MS/MS Ion Search) and MS/MS data that was manually interpreted to obtain sequence, composition and fragment ion data (using MS Query). DTA files were generated for MS/MS Ion Search from the CID data (RAW files) using Bioworks 3.3 (Therrno Fisher Scientific, San Jose, CA) (parameters: absolute intensity threshold of 1; minimum ion count of 10). All data was searched against the NCBI non-redundant database, compiled from GenBank CDS translations, Protein Information Resource (PIR), Swiss-Prot, Protein Research Foundation, and Protein Data Bank (National Center for Biotechnology Information, National Library of Medicine, National Institutes of Health, Bethesda, MD; 9,680,073 entries). The search parameters used were as follows: human; semi-tryptic; three missed cleavages allowed; carbamidomethylation (C); oxidation (M); phosphorylation (STY) as variable modifications; peptide tolerance of 0.5 Da; MS/MS product ion tolerance 0.5 Da; peptide charge state 1+; experimental mass values monoisotopic; instrument ESI-Trap. The taxonomy filter resulted in 223,846 protein sequences being considered in the database search. All identifications and determined phosphorylation sites were manually validated. 245 10. REFERENCES Hunter, T. Signaling-2000 and Beyond. Cell 2000, 100, 113-127. Cohen, P. The Regulation of Protein Function by Multisite Phosphorylation - a 25 Year Update. Trends Biochem. Sci. 2000, 25, 596-601. Mazanetz, M. P.; Fischer, P. M. Untangling Tau Hyperphosphorylation in Drug Design for Neurodegenerative Diseases. Nat. Rev. Drug Discov. 2007, 6. 464-479. Culmsee, C.; Landshamer, S. Molecular Insights into Mechanisms of the Cell Death ProgramzRole in the Progression of Neurodegenerative Disorders. Curr. Alzheimer Res. 2006, 3, 269-283. Bode, A. M.; Dong, Z. Post-Translational Modification of p53 in Tumorigenesis. Nat. Rev. Cancer 2004, 4, 793-805. Mackay, H. J.; Twelves, C. J. Targeting the Protein Kinase C Family: Are We There Yet? Nat. Rev. Cancer 2007, 7, 554-562. Kewalramani, 6.; An, 0.; Kim, M. S.; Ghosh, S.; Oi, D.; Abrahani, A.; Pulinilkunnil, T.; Sharma, V.; Wambolt, R. B.; Allard, M. F.; Innis, S. M.; Rodrigues, B. AMPK Control of Myocardial Fatty Acid Metabolism Fluctuates with the Intensity of Insulin-Deficient Diabetes. J. Mol. Cell. Cardiol. 2007, 42, 333-342. Frawley Cass, S. M.; Reid, G. E.; Tepe, J. J. Identification of Phosphoproteins and Their Impact as Biomarkers in Cancer Therapeutics. Curr. Signal Transduction Ther. 2009, 4, DOI: 10.1039/papemo. Gong, C. X.; Iqbal, K. Hyperphosphorylation of Microtubule-Associated Protein Tau: A Promising Therapeutic Target for Alzheimer Disease. Curr. Med. Chem. 2008, 1 5, 2321 ~2328. Gong, C.-X.; Liu, F.; Grundke-lqbal, l.; Iqbal, K. Dysregulation of Protein Phosphorylation/Dephosphorylation in Alzheimer's Disease: A Therapeutic Target. Journal of Biomedicine and Biotechnology 2006, 2006, 1-1 1. 246 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. Clark, M. P.; Bookland, R. G. Pyrazolones as Therapeutics for Kinase- Mediated Inflammatory Disorders. Expert Opin. Ther. Pat. 2005, 15, 1617- 1623. Ducruet, A. P.; Vogt, A.; Wipf, P.; Lazo, J. 8. Dual Specificity Protein Phosphatases: Therapeutic Targets for Cancer and Alzheimer‘s Disease. Annual Review of Pharmacology & Toxicology 2005, 45, 725-750. Echols, N.; Harrison, P.; Balasubramanian, S.; Luscombe, N. M.; Bertone, P.; Zhang, 2.; Gerstein, M. Comprehensive Analysis of Amino Acid and Nucleotide Composition in Eukaryotic Genomes, Comparing Genes and Pseudogenes. Nucl. Acids Res. 2002, 30, 2515-2523. Ubersax, J. A.; Ferrell, J. E. Mechanisms of Specificity in Protein Phosphorylation. Nat. Rev. Mol. Cell Biol. 2007, 8, 665-665. Mann, M.; Jensen, 0. N. Proteomic Analysis of Post-Translational Modifications. Nat. Biotech. 2003, 21, 255-261. Goshe, M. B. Characterizing Phosphoproteins and Phosphoproteomes Using Mass Spectrometry. Brief. Funct. Genomics Proteomics 2006, 4, 363- 376. McLachlin, D. T.; Chait, B. T. Analysis of Phosphorylated Proteins and Peptides by Mass Spectrometry. Curr. Opin. Chem. Biol. 2001, 5, 591-602. Paradela, A.; Albar, J. P. Advances in the Analysis of Protein Phosphorylation. J. Proteome Res. 2008, 7, 1809-1818. Carr, 8. A.; Annan, R. S.; Huddleston, M. J. Mapping Posttranslational Modifications of Proteins by MS-Based Selective Detection: Application to Phosphoproteomics. Methods Enzymol. 2005, 405, 82-115. Nita-Lazar, A.; Saito-Benz, H.; White, F. M. Quantitative Phosphoproteomics by Mass Spectrometry: Past, Present, and Future. Proteomics 2008, 8, 4433-4443. Boersema, P. J.; Mohammed, 8.; Heck, A. J. R. Phosphopeptide Fragmentation and Analysis by Mass Spectrometry. J. Mass Spectrom. 2009, 44, 861-878. Zugaro, L. M.; Reid, G. E.; Ji, H.; Eddes, J. S.; Murphy, A. C.; Burgess, A. W.; Simpson, R. J. Characterization of Rat Brain Stathmin Isoforms by Two- 247 23. 24. 25. 26. 27. 28. 29. 30. 31. Dimensional Gel Electrophoresis Matrix Assisted Laser Desorption/Ionization and Electrospray Ionization Ion Trap Mass Spectrometry. Electrophoresis 1998, 19, 867-876. Rush, J.; Moritz, A.; Lee, K. A.; Guo, A.; Goss, V. L.; Spek, E. J.; Zhang, H.; Zha, X.-M.; PoIakiewicz, R. D.; Comb, M. J. lmmunoaffinity Profiling of Tyrosine Phosphorylation in Cancer Cells. Nat. Biotech. 2005,23, 94-101. Olsen, J. V.; Blagoev, B.; Gnad, F.; Macek, B.; Kumar, 0.; Mortensen, P.; Mann, M. Global, In Vivo, and Site-Specific Phosphorylation Dynamics in Signaling Networks. Cell 2006, 127, 635-648. Bergstrom Lind, 8.; Molin, M.; Savitski, M. M.; Emilsson, L.; Astrom, J.; Hedberg, L.; Adams, 0.; Nielsen, M. L.; Engstrom, A.; Elfineh, L.; Andersson, E.; Zubarev, R. A.; Pettersson, U. Immunoaffinity Enrichments Followed by Mass Spectrometric Detection for Studying Global Protein Tyrosine Phosphorylation. J. Proteome Res. 2008, 7, 2897-2910. Macek, B.; Mijakovic, I.; Olsen, J. V.; Gnad, F.; Kumar, 0.; Jensen, P. R.; Mann, M. The Serine/Threoninel'l’yrosine Phosphoproteome of the Model Bacterium Bacillus Subtilis. Mol. Cell. Proteomics 2007, 6, 697-707. Zanivan, S.; Gnad, F.; Wickstrdm, S. A.; Geiger, T.; Macek, B.; Cox, J.; Fassler, R.; Mann, M. Solid Tumor Proteome and Phosphoproteome Analysis by High Resolution Mass Spectrometry. J. Proteome Res. 2008, 7, 5314-5326. Posewitz, M. C.; Tempst, P. Immobilized Gallium(lll) Affinity Chromatography of Phosphopeptides. Anal. Chem. 1999, 71, 2883-2892. Andersson, L.; Porath, J. Isolation of Phosphoproteins by Immobilized Metal (Fe3+) Affinity-Chromatography. Anal. Biochem. 1986, 154, 250-254. Ficarro, S. B.; McCleIand, M. L.; Stukenberg, P. T.; Burke, D. J.; Ross, M. M.; Shabanowitz, J.; Hunt, D. F.; White, F. M. Phosphoproteome Analysis by Mass Spectrometry and Its Application to Saccharomyces Cerevisiae. Nat. Biotechnol. 2002, 20, 301-305. Kokubu, M.; Ishihama, Y.; Sato, T.; Nagasu, T.; Oda, Y. Specificity of Immobilized Metal Affinity-Based lMAC/C18 Tip Enrichment of Phosphopeptides for Protein Phosphorylation Analysis. Anal. Chem. 2005, 77, 5144-5154. 248 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. Kweon, H. K.; Hakansson, K. Selective Zirconium Dioxide-Based Enrichment of Phosphorylated Peptides for Mass Spectrometric Analysis. Anal. Chem. 2006, 78, 1743-1749. Matsuda, H.; Nakamura, H.; Nakajima, T. New Ceramic Titania - Selective Adsorbent for Organic-Phosphates. Anal. Sci. 1990, 6, 911-912. Ikeguchi, Y.; Nakamura, H. Determination of Organic Phosphates by Column-Switching High Performance Anion-Exchange Chromatography Using on-Line Preconcentration on Titania. Anal. Sci. 1997, 13, 479-483. Pinkse, M. W. H.; Uitto, P. M.; Hilhorst, M. J.; Ooms, B.; Heck, A. J. R. Selective Isolation at the Femtomole Level of Phosphopeptides from Proteolytic Digests Using 2D-nanoLC-ESl-MS/MS and Titanium Oxide Precolumns. Anal. Chem. 2004, 76, 3935-3943. Larsen, M. R.; Thingholm, T. E.; Jensen, O. N.; Roepstorff, P.; Jorgensen, T. J. D. Highly Selective Enrichment of Phosphorylated Peptides from Peptide Mixtures Using Titanium Dioxide Microcolumns. Mol. Cell. Proteomics 2005, 4, 873-886. Meyer, H. E.; Hoffmann-Posorske, E.; Korte, H.; Heilmeyer, L. M. G. Sequence-Analysis of Phosphoserine-Containing Peptides - Modification for Picomolar Sensitivity. FEBS Lett. 1986, 204, 61-66. Goshe, M. B.; Conrads, T. P.; Panisko, E. A.; Angell, N. H.; Veenstra, T. D.; Smith, R. D. Phosphoprotein Isotope-Coded Affinity Tag Approach for Isolating and Quantitating Phosphopeptides in Proteome-Wide Analyses. Anal. Chem. 2001, 73, 2578-2586. Oda, Y.; Nagasu, T.; Chait, B. T. Enrichment Analysis of Phosphorylated Proteins as a Tool for Probing the Phosphoproteome. Nat. Biotechnol. 2001, 19, 379-382. McLachlin, D. T.; Chait, B. T. Improved Beta-EIimination-Based Affinity Purification Strategy for Enrichment of Phosphopeptides. Anal. Chem. 2003, 75, 6826-6836. Thaler, F.; Valsasina, B.; Baldi, R.; Jin, X.; Stewart, A.; lsacchi, A.; Kalisz, H. M.; Rusconi, L. A New Approach to Phosphoserine and Phosphothreonine Analysis in Peptides and Proteins: Chemical Modification, Enrichment Via Solid-Phase Reversible Binding, and Analysis by Mass Spectrometry. Anal. Bioanal. Chem. 2003, 376, 366-373. 249 42. 43. 44. 45. 46. 47. 48. 49. 50. Knight, Z. A.; Schilling, B.; Row, R. H.; Kenski, D. M.; Gibson, B. W.; Shokat, K. M. Phosphospecific Proteolysis for Mapping Sites of Protein Phosphorylation. Nature Biotechnology 2003, 21, 1047-1054. Zhou, H.; Watts, J. D.; AebersoId, R. A Systematic Approach to the Analysis of Protein Phosphorylation. Nat. Biotech. 2001, 19, 375-378. Lansdell, T. A.; Tepe, J. J. Isolation of Phosphopeptides Using Solid Phase Enrichment. Tetrahedron Lett. 2004, 45, 91-93. Frawley Cass, S. M.; Reid, G. E.; Tepe, J. J. Synthesis of Diazo Functionalized Solid Supports and Their Application Towards the Enrichment of Phosphorylated Peptides Org. Biomol. Chem. 2009, 7, 3291- 3299. Tao, W. A.; WolIscheid, B.; O'Brien, R.; Eng, J. K.; Li, X.-j.; Bodenmiller, B.; Watts, J. D.; Hood, L.; AebersoId, R. Quantitative Phosphoproteome Analysis Using a Dendrimer Conjugation Chemistry and Tandem Mass Spectrometry. Nat Meth 2005, 2, 591-598. Ficarro, S.; Chertihin, O.; Westbrook, V. A.; White, F.; Jayes, F.; Kalab, P.; Marto, J. A.; Shabanowitz, J.; Herr, J. C.; Hunt, D. F.; Visconti, P. E. Phosphoproteome Analysis of Capacitated Human Sperm. Evidence of Tyrosine Phosphorylation of a Kinase-Anchoring Protein 3 and Valosin- Containing Protein/p97 During Capacitation. J. Biol. Chem. 2003, 278, 11579-11589. Pandey, A.; Podtelejnikov, A. V.; Blagoev, B.; BusteIo, X. R.; Mann, M.; Lodish, H. F. Analysis of Receptor Signaling Pathways by Mass Spectrometry: Identification of Vav-2 as a Substrate of the Epidermal and Platelet-Derived Growth Factor Receptors. Proc. Natl. Acad. Sci. U. S. A. 2000, 97, 179-184. Brill, L. M.; Salomon, A. R.; Ficarro, S. B.; Mukherji, M.; Stettler-Gill, M.; Peters, E. C. Robust Phosphoproteomic Profiling of Tyrosine Phosphorylation Sites from Human T Cells Using Immobilized Metal Affinity Chromatography and Tandem Mass Spectrometry. Anal. Chem. 2004, 76, 2763-2772. ° Dunn, J. D.; Reid, G. E.; Bruening, M. L. Techniques for Phosphopeptide Enrichment Prior to Analysis by Mass Spectrometry. Mass Spectrom. Rev. 2009, 10.1002/mas.20219. 250 51. 52. 53. 54. 55. 56. 57. 58. 59. Thingholm, T. E.; Jensen, O. N.; Larsen, M. R. Analytical Strategies for Phosphoproteomics. Proteomics 2009, 9, 1451 -1 468. Craig, A. G.; Hoeger, C. A.; Miller, C. L.; Goedken, T.; Rivier, J. E.; Fischer, W. H. Monitoring Protein-Kinase and Phosphatase Reactions with Matrix- Assisted Laser-Desorption Ionization Mass-Spectrometry and Capillary Zone Electrophoresis - Comparison of the Detection Efficiency of Peptide- Phosphopeptide Mixtures. Biol. Mass Spectrom. 1994, 23, 519-528. Liao, P. C.; Leykam, J.; Andrews, P. C.; Gage, D. A.; Allison, J. An Approach to Locate Phosphorylation Sites in a Phosphoprotein - Mass Mapping by Combining Specific Enzymatic Degradation with Matrix- Assisted Laser-Desorption Ionization Mass-Spectrometry. Anal. Biochem. 1994, 219, 9-20. Yang, X. F.; Wu, H. P.; Kobayashi, T.; Solaro, R. J.; van Breemen, R. B. Enhanced Ionization of Phosphorylated Peptides During MALDI TOF Mass Spectrometry. Anal. Chem. 2004, 76, 1532-1536. Asara, J. M.; Allison, J. Enhanced Detection of Phosphopeptides in Matrix- Assisted Laser Desorption/Ionization Mass Spectrometry Using Ammonium Salts. J. Am. Soc. Mass Spectrom. 1999, 10, 35-44. Stensballe, A.; Jensen, O. N. Phosphoric Acid Enhances the Performance of Fe(|ll) Affinity Chromatography and Matrix-Assisted Laser Desorption/Ionization Tandem Mass Spectrometry for Recovery, Detection and Sequencing of Phosphopeptides. Rapid Commun. Mass Spectrom. 2004, 18, 1721-1730. Kuyama, H.; Sonomura, K.; Nishimura, O. Sensitive Detection of Phosphopeptides by Matrix-Assisted Laser Desorption/Ionization Mass Spectrometry: Use of Alkylphosphonic Acids as Matrix Additives. Rapid Commun. Mass Spectrom. 2008, 22, 1109-11 16. Steen, H.; Jebanathirajah, J. A.; Rush, J.; Morrice, N.; Kirschner, M. W. Phosphorylation Analysis by Mass Spectrometry - Myths, Facts, and the Consequences for Qualitative and Quantitative Measurements. Mol. Cell. Proteomics 2006, 5, 172-181. Larsen, M. R.; Sorensen, G. L.; Fey, S. J.; Larsen, P. M.; Roepstorff, P. Phospho-Proteomics: Evaluation of the Use of Enzymatic De- Phosphorylation and Differential Mass Spectrometric Peptide Mass Mapping for Site Specific Phosphorylation Assignment in Proteins Separated by Gel Electrophoresis. Proteomics 2001, 1, 223-238. 251 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. Pyatkivskyy, Y.; Ryzhov, V. Coupling of Ion-Molecule Reactions with Liquid Chromatography on a Quadrupole Ion Trap Mass Spectrometer. Rapid Commun. Mass Spectrom. 2008, 22, 1288-1294. Weckwerth, W.; Willmitzer, L.; Fiehn, O. Comparative Quantification and Identification of Phosphoproteins Using Stable Isotope Labeling and Liquid Chromatography/Mass Spectrometry. Rapid Commun. Mass Spectrom. 2000, 14, 1677-1681. Wells, J. M.; McLuckey, S. A., Collision-Induced Dissociation (CID) of Peptides and Proteins. In Methods Enzymol., Academic Press: 2005; Vol. Volume 402, pp 148-185. Spengler, B. Post-Source Decay Analysis in Matrix-Assisted Laser Desorption/Ionization Mass Spectrometry of Biomolecules. J. Mass Spectrom. 1 997, 32, 1019-1036. Brodbelt, J. S.; Wilson, J. J. Infrared Multiphoton Dissociation in Quadrupole Ion Traps. Mass Spectrom. Rev. 2009, 28, 390-424. Reilly, J. P. Ultraviolet Photofragmentation of Biomolecular Ions. Mass Spectrom. Rev. 2009, 28, 425-447. Kalcic, C. L.; Gunaratne, T. C.; Jones, A. D.; Dantus, M.; Reid, G. E. Femtosecond Laser-Induced Ionization/Dissociation of Protonated Peptides. J. Am. Chem. Soc. 2009, 131, 940-942. Cooper, H. J.; Hakansson, K.; Marshall, A. G. The Role of Electron Capture Dissociation in Biomolecular Analysis. Mass Spectrom. Rev. 2005, 24, 201- 222. Wiesner, J.; Premsler, T.; Sickmann, A. Application of Electron Transfer Dissociation (ETD) for the Analysis of Posttranslational Modifications. Proteomics 2008, 8, 4466-4483. Fung, Y. M. E.; Adams, C. M.; Zubarev, R. A. Electron Ionization Dissociation of Singly and Multiply Charged Peptides. J. Am. Chem. Soc. 2009, 131, 9977-9985. Misharin, A. 8.; Silivra, O. A.; Kjeldsen, F.; Zubarev, R. A. Dissociation of Peptide Ions by Fast Atom Bombardment in a Quadrupole Ion Trap. Rapid Commun. Mass Spectrom. 2005, 19, 2163-21 71 . 252 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. Biemann, K. Contributions of Mass-Spectrometry to Peptide and Protein- Structure. Biomed. Environ. Mass Spectrom. 1 988, 16, 99-1 1 1. Roepstorff, P.; Fohlman, J. Proposal for a Common Nomenclature for Sequence Ions in Mass Spectra of Peptides. Biomedical mass spectrometry 1984, 11, 601. Schreiber, T. B.; Mausbacher, N.; Breitkopf, S. B.; Grundner-Culemann, K.; Daub, H. Quantitative Phosphoproteomics - an Emerging Key Technology in Signal-Transduction Research. Proteomics 2008, 8, 4416-4432. McLuckey, S. A.; Goeringer, D. E. Slow Heating Methods in Tandem Mass Spectrometry. J. Mass Spectrom. 1997, 32, 461-474. ' Paizs, B.; Lendvay, G.; Vekey, K.; Suhai, S. Formation of b2+ Ions from Protonated Peptides: An Ab Initio Study. Rapid Commun. Mass Spectrom. 1999, 13, 525-533. Csonka, I. P.; Paizs, B.; Lendvay, G.; Suhai, S. Proton Mobility in Protonated Peptides: A Joint Molecular Orbital and RRKM Study. Rapid Commun. Mass Spectrom. 2000, 14, 41 7-431 . Flora, J. W.; Muddiman, D. C. Determination of the Relative Energies of Activation for the Dissociation of Aromatic Versus Aliphatic Phosphopeptides by ESI-FTICR-MS and IRMPD. J. Am. Soc. Mass Spectrom. 2004, 15, 121-127. Gronert, 8.; Li, K. H.; Horiuchi, M. Manipulating the Fragmentation Patterns of Phosphopeptides Via Gas-Phase Boron Derivatization: Determining Phosphorylation Sites in Peptides with Multiple Serines. J. Am. Soc. Mass Spectrom. 2005, 16, 1905-1914. Ndassa, Y. M.; Orsi, C.; Marto, J. A.; Chen, 8.; Ross, M. M. Improved Immobilized Metal Affinity Chromatography for Large-Scale Phosphoproteomics Applications. J. Proteome Res. 2006, 5, 2789-2799. Li, X.; Gerber, S. A.; Rudner, A. D.; Beausoleil, S. A.; Haas, W.; Villen, J.; Elias, J. E.; Gygi, S. P. Large-Scale Phosphorylation Analysis of alpha- Factor-Arrested Saccharomyces Cerevisiae. J. Proteome Res. 2007, 6, 1190-1197. 253 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. Schlosser, A.; Pipkom, R.; Bossemeyer, D.; Lehmann, W. D. Analysis of Protein Phosphorylation by a Combination of Elastase Digestion and Neutral Loss Tandem Mass Spectrometry. Anal. Chem. 2001, 73, 170-176. Moyer, S. C.; Cotter, R. J.; Woods, A. S. Fragmentation of Phosphopeptides by Atmospheric Pressure MALDI and ESI/Ion Trap Mass Spectrometry. J. Am. Soc. Mass Spectrom. 2002, 13, 274-283. Carr, S. A.; Huddleston, M. J.; Annan, R. S. Selective Detection and Sequencing of Phosphopeptides at the Femtomole Level by Mass Spectrometry. Anal. Biochem. 1996, 239, 180-192. Collins, M. 0.; Yu, L.; Coba, M. P.; Husi, H.; Campuzano, L.; Blackstock, W. P.; Choudhary, J. S.; Grant, 8. G. N. Proteomic Analysis of in Vivo Phosphorylated Synaptic Proteins. J. Biol. Chem. 2005, 280, 5972-5982. Huddleston, M. J.; Annan, R. 8.; Bean, M. F.; Carr, S. A. Selective Detection of Phosphopeptides in Complex Mixtures by Electrospray Liquid Chromatography/Mass Spectrometry. J. Am. Soc. Mass Spectrom. 1993, 4, 710-7. Ding, J.; Burkhart, W.; Kassel, D. B. Identification of Phosphorylated Peptides from Complex Mixtures Using Negative-Ion Orifice-Potential Stepping and Capillary Liquid Chromatography/EIectrospray Ionization Mass Spectrometry. Rapid Commun. Mass Spectrom. 1994, 8, 94-98. Tholey, A.; Reed, J.; Lehmann, W. D. Electrospray Tandem Mass Spectrometric Studies of Phosphopeptides and Phosphopeptide Analogues. J. Mass Spectrom. 1999, 34, 117-123. EdeIson-Averbukh, M.; Pipkom, R.; Lehmann, W. D. Phosphate Group- Driven Fragmentation of Multiply Charged Phosphopeptide Anions. Improved Recognition of Peptides Phosphorylated at Serine, Threonine, or Tyrosine by Negative Ion Electrospray Tandem Mass Spectrometry. Anal. Chem. 2006, 78, 1249-1256. DeGnore, J. P.; Qin, J. Fragmentation of Phosphopeptides in an Ion Trap Mass Spectrometer. J. Am. Soc. Mass Spectrom. 1998, 9, 1175-1188. Annan, R. S.; Huddleston, M. J.; Venna, R.; Deshaies, R. J.; Carr, 8. A. A Multidimensional Electrospray MS-Based Approach to Phosphopeptide Mapping. Anal. Chem. 2001, 73, 393-404. 254 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. Steen, H.; Kuster, B.; Fernandez, M.; Pandey, A.; Mann, M. Detection of Tyrosine Phosphorylated Peptides by Precursor Ion Scanning Quadrupole TOF Mass Spectrometry in Positive Ion Mode. Anal. Chem. 2001, 73, 1440- 1448. Steen, H.; Kuster, B.; Mann, M. Quadrupole Time-of-Flight Versus Triple- Quadrupole Mass Spectrometry for the Determination of Phosphopeptides by Precursor Ion Scanning. J. Mass Spectrom. 2001, 36, 782-790. Ambihapathy, K.; Yalcin, T.; Leung, H.-W.; Harrison, A. G. Pathways to Immonium Ions in the Fragmentation of Protonated Peptides. J. Mass Spectrom. 1997, 32, 209-215. Salek, M.; Alonso, A.; Pipkom, R.; Lehmann, W. D. Analysis of Protein Tyrosine Phosphorylation by Nanoelectrospray Ionization High-Resolution Tandem Mass Spectrometry and Tyrosine-Targeted Product Ion Scanning. Anal. Chem. 2003, 75, 2724-2729. Steen, H.; Kuster, B.; Fernandez, M.; Pandey, A.; Mann, M. Tyrosine Phosphorylation Mapping of the Epidermal Growth Factor Receptor Signaling Pathway. J. Biol. Chem. 2002, 277, 1031-1039. Moon, J. H.; Shin, Y. S.; Kim, M. S. Utility of Reaction Intermediate Monitoring with Photodissociation Multi-Stage (Msn) Time-of-Flight Mass Spectrometry for Mechanistic and Structural Studies: Phosphopeptides. Int. J. Mass Spectrom. 2009, 10.1016/j.ijms.2009.07.008. Froelich, J. M.; Reid, G. E. The Effect of Post-Translational and Process- lnduced Modifications on the Multistage Gas-Phase Fragmentation Reactions of Protonated Peptides. Comb. Chem. High Throughput Screening 2009, 12, 175-184. Lee, J.; Xu, Y.; Chen, Y.; Sprung, R.; Kim, S. C.; Xie, 8.; Zhao, Y. Mitochondrial Phosphoproteome Revealed by an Improved IMAC Method and MS/MS/MS. Mol. Cell. Proteomics 2007, 6, 669-676. ' Chang, E. J.; Archambault, V.; McLachlin, D. T.; Krutchinsky, A. N.; Chait, B. T. Analysis of Protein Phosphorylation by Hypothesis—Driven Multiple- Stage Mass Spectrometry. Anal. Chem. 2004, 76, 4472-4483. Wolschin, F.; Lehmann, U.; Glinski, M.; Weckwerth, W. An Integrated Strategy for Identification and Relative Quantification of Site-Specific Protein 255 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. Phosphorylation Using Liquid Chromatography Coupled to MSlesa. Rapid Commun. Mass Spectrom. 2005, 19, 3626-3632. Schroeder, M. J.; Shabanowitz, J.; Schwartz, J. 0.; Hunt, D. F.; Coon, J. J. A Neutral Loss Activation Method for Improved Phosphopeptide Sequence Analysis by Quadrupole Ion Trap Mass Spectrometry. Anal. Chem. 2004, 76, 3590-3598. Sweet, 8. M. M.; Creese, A. J.; Cooper, H. J. Strategy for the Identification of Sites of Phosphorylation in Proteins: Neutral Loss Triggered Electron Capture Dissociation. Anal. Chem. 2006, 78, 7563-7569. Sweet, S. M. M.; Cooper Helen, J. On-Line Liquid Chromatography Electron Capture Dissociation for the Characterization of Phosphorylation Sites in Proteins. Methods Mol Biol 2009, 527, 191-9. Hoffman, M. D.; Sniatynski, M. J.; Rogalski, J. C.; Le Blanc, J. C. Y.; Kast, J. Multiple Neutral Loss Monitoring (MNM): A Multiplexed Method for Post- Translational Modification Screening. J. Am. Soc. Mass Spectrom. 2006, 17, 307-317. Lehmann, W. D.; Krueger, R.; Salek, M.; Hung, C.-W.; Wolschin, F.; Weckwerth, W. Neutral Loss-Based Phosphopeptide Recognition: A Collection of Caveats. J. Proteome Res. 2007, 6, 2866-2873. Woods, A. S.; Ferre, S. Amazing Stability of the Arginine-Phosphate Electrostatic Interaction. J. Proteome Res. 2005, 4, 1397-1402. Jackson, S. M; Wang, H.-Y. J.; Woods, A. 8. Study of the Fragmentation Patterns of the Phosphate-Arginine Noncovalent Bond. J. Proteome Res. 2005, 4, 2360-2363. Jackson, S. M; Wang, H. Y. J.; Yergey, A.; Woods, A. S. Phosphate Stabilization of Intermolecular Interactions. J. Proteome Res. 2006, 5, 122- 126. Woods, A. S.; Moyer, S. C.; Jackson, S. N. Amazing Stability of Phosphate- Quatemary Amine Interactions. J. Proteome Res. 2008, 7, 3423-3427. Jackson, S. N.; Moyer, S. C.; Woods, A. S. The Role of Phosphorylated Residues in Peptide-Peptide Noncovalent Complexes Formation. J. Am. Soc. Mass Spectrom. 2008, 19, 1 535-1 541 . 256 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. Villen, J.; Beausoleil, S. A.; Gygi, S. P. Evaluation of the Utility of Neutral- Loss—Dependent MS3 Strategies in Large-Scale Phosphorylation Analysis. Proteomics 2008, 8, 4444-4452. Dongre, A. R.; 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. Wysocki, V. H.; Tsaprailis, G.; Smith, L. L.; Breci, L. A. Mobile and Localized Protons: A Framework for Understanding Peptide Dissociation. J. Mass Spectrom. 2000, 35, 1399-1406. Cox, K. A.; Gaskell, S. J.; Morris, M.; Whiting, A. Role of the Site of Protonation in the Low-Energy Decompositions of Gas-Phase Peptide Ions. J. Am. Soc. Mass Spectrom. 1996, 7, 522-531. 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. Froelich, J. M.; Reid, G. E. Mechanisms for the Proton Mobility-Dependent Gas-Phase Fragmentation Reactions of S-Alkyl Cysteine Sulfoxide- Containing Peptide Ions. J. Am. Soc. Mass Spectrom. 2007, 18, 1690-1705. Reid, G. E.; Simpson, R. J.; O'Hair, R. A. J. Leaving Group and Gas Phase Neighboring Group Effects in the Side Chain Losses from Protonated Serine and Its Derivatives. J. Am. Soc. Mass Spectrom. 2000, 11, 1047-1060. Qin, J.; Chait, B. T. Identification and Characterization of Posttranslational Modifications of Proteins by MALDI Ion Trap Mass Spectrometry. Anal. Chem. 1997, 69, 4002-4009. Steen, H.; Mann, M. A New Derivatization Strategy for the Analysis of Phosphopeptides by Precursor Ion Scanning in Positive Ion Mode. J. Am. Soc. Mass Spectrom. 2002, 13, 996-1003. Arrigoni, G.; Resjé, S.; Levander, F.; Nilsson, R.; Degerrnan, E.; Quadroni, M.; Pinna, L. A.; James, P. Chemical Derivatization of Phosphoserine and Phosphothreonine Containing Peptides to Increase Sensitivity for MALDI- Based Analysis and for Selectivity of MS/MS Analysis. Proteomics 2006, 6, 757-766. 257 121. 122. 123. 124. 125. 126. 127. 128. 129. 130. Moyer, S. C.; VonSeggern, C. E.; Cotter, R. J. Fragmentation of Cationized Phosphotyrosine Containing Peptides by Atmospheric Pressure MALDI/Ion Trap Mass Spectrometry. J. Am. Soc. Mass Spectrom. 2003, 14, 581-592. Leitner, A.; Foettinger, A.; Lindner, W. Improving Fragmentation of Poorly Fragmenting Peptides and Phosphopeptides During Collision-Induced Dissociation by ManndiaIdehyde Modification of Arginine Residues. J. Mass Spectrom. 2007, 42, 950-959. Gronert, S.; Huang, R.; Li, K. H. Gas Phase Derivatization in Peptide Analysis I: The Utility of Trimethyl Borate in Identifying Phosphorylation Sites. Int. J. Mass Spectrom. 2004, 231, 179-187. Kaufmann, R.; Spengler, B.; Lutzenkirchen, F. Mass-Spectrometric Sequencing of Linear Peptides by Product-Ion Analysis in a Reflectron Time-of-Flight Mass-Spectrometer Using Matrix-Assisted Laser-Desorption Ionization. Rapid Commun. Mass Spectrom. 1993, 7, 902-910. Tang, X. J.; Ens, W.; Standing, K. G.; Westmore, J. B. Daughter Ion Mass- Spectra from Cationized Molecules of Small Oligopeptides in a Reflecting Time-of-Flight Mass-Spectrometer. Anal. Chem. 1988, 60, 1791-1799. Rouse, J. C.; Yu, W.; Martin, S. A. A Comparison of the Peptide Fragmentation Obtained from a Reflector Matrix-Assisted Laser Desorption- lonization Time-of-Flight and a Tandem 4-Sector Mass-Spectrometer. J. Am. Soc. Mass Spectrom. 1995, 6, 822-835. Annan, R. S.; Carr, S. A. Phosphopeptide Analysis by Matrix-Assisted Laser Desorption Time-of-Flight Mass Spectrometry. Anal. Chem. 1996, 68, 3413- 3421. Metzger, S.; Hoffmann, R. Studies on the Dephosphorylation of Phosphotyrosine-Containing Peptides Dun'ng Post-Source Decay in Matrix- Assisted Laser Desorption/Ionization. J. Mass Spectrom. 2000, 35, 1165- 1177. Hoffmann, R.; Metzger, S.; Spengler, B.; Otvos, L., Jr. Sequencing of Peptides Phosphorylated on Serines and Threonines by Post-Source Decay in Matrix-Assisted Laser Desorption/Ionization Time-of-Flight Mass Spectrometry. J. Mass Spectrom. 1999, 34, 1 1 95-1204. Schnolzer, M.; D. Lehmann, W. Identification of Modified Peptides by Metastable Fragmentation in MALDI Mass Spectrometry. Int. J. Mass Spectrom. Ion Processes 1997, 169/170, 263-271. 258 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. Eyler, J. R. Infrared Multiple Photon Dissociation Spectroscopy of Ions in Penning Traps. Mass Spectrom. Rev. 2009, 28, 448-467. Little, D. P.; Speir, J. P.; Senko, M. W.; Oconnor, P. B.; Mclafferty, F. W. Infrared Multiphoton Dissociation of Large Multiply-Charged Ions for Biomolecule Sequencing. Anal. Chem. 1994, 66, 2809-2815. Flora, J. W.; Muddiman, D. C. Selective, Sensitive, and Rapid Phosphopeptide Identification in Enzymatic Digests Using ESI-FTICR-MS with Infrared Multiphoton Dissociation. Anal. Chem. 2001, 73, 3305-3311. Flora, J. W.; Muddiman, D. C. Gas-Phase Ion Unimolecular Dissociation for Rapid Phosphopeptide Mapping by IRMPD in a Penning Ion Trap: An Energetically Favored Process. J. Am. Chem. Soc. 2002, 124, 6546-6547. Crowe, M. C.; Brodbelt, J. S. Infrared Multiphoton Dissociation (IRMPD) and Collisionally Activated Dissociationof Peptides in a Quadrupole Ion Trap with Selective IRMPD of Phosphopeptides. J. Am. Soc. Mass Spectrom. 2004, 15, 1581-1592. Chalmers, M. J.; Quinn, J. P.; Blakney, G. T.; Emmett, M. R.; Mischak, H.; Gaskell, S. J.; Marshall, A. G. Liquid Chromatography-Fourier Transform Ion Cyclotron Resonance Mass Spectrometric Characterization of Protein Kinase C Phosphorylation. J. Proteome Res. 2003, 2, 373-382. Lemoine, J.; Tabarin, T.; Antoine, R.; Broyer, M.; Dugourd, P. UV Photodissociation of Phospho-Seryl-Containing Peptides: Laser Stabilization of the Phospho-Seryl Bond with Multistage Mass Spectrometry. Rapid Commun. Mass Spectrom. 2006, 20, 507-511. Kim, T.-Y.; Reilly, J. P. Time-Resolved Observation of Product Ions Generated by 157 nm Photodissociation of Singly Protonated Phosphopeptides. J. Am. Soc. Mass Spectrom. 2009, doi: DOI: 10.1016lj.jasms.2009.08.021. Diedrich, J. K.; Julian, R. R. Site-Specific Radical Directed Dissociation of Peptides at Phosphorylated Residues. J. Am. Chem. Soc. 2008, 130, 12212-12213. Shin, Y. 8.; Moon, J. H.; Kim, M. S. Observation of Phosphorylation Site- Specific Dissociation of Singly Protonated Phosphopeptides. J. Am. Soc. Mass Spectrom. 2009, 10.1016/j.jasms.2009.09.003. 259 141. 142. 143. 144. 145. 146. 147. 148. 149. 150. Tabarin, T.; Antoine, R.; Broyer, M.; Dugourd, P. Specific Photodissociation of Peptides with Multi-Stage Mass Spectrometry. Rapid Commun. Mass Spectrom. 2005, 19, 2883-2892. Zewail, A. H. Femtochemistry: Atomic-Scale Dynamics of the Chemical Bond J. Phys. Chem. A 2000, 104, 5660-5694. Zubarev, R. A.; Kelleher, N. L.; McLafferty, F. W. Electron Capture Dissociation of Multiply Charged Protein Cations. A Nonergodic Process. J. Am. Chem. Soc. 1998, 120, 3265-3266. Baba, T.; Hashimoto, Y.; Hasegawa, H.; Hirabayashi, A.; Waki, I. Electron Capture Dissociation in a Radio Frequency Ion Trap. Anal. Chem. 2004, 76, 4263-4266. Silivra, O. A.; Kjeldsen, F.; Ivonin, I. A.; Zubarev, R. A. Electron Capture Dissociation of Polypeptides in a Three-Dimensional Quadrupole Ion Trap: Implementation and First Results. J. Am. Soc. Mass Spectrom. 2005, 16, 22-27. Satake, H.; Hasegawa, H.; Hirabayashi, A.; Hashimoto, Y.; Baba, T.; Masuda, K. Fast Multiple Electron Capture Dissociation in a Linear Radio Frequency Quadrupole Ion Trap. Anal. Chem. 2007, 79, 8755-8761. Ding, L.; Brancia, F. L. Electron Capture Dissociation in a Digital Ion Trap Mass Spectrometer. Anal. Chem. 2006, 78, 1995-2000. Zubarev, R. A.; Horn, D. M.; Fridriksson, E. K.; Kelleher, N. L.; KnJger, N. A.; Lewis, M. A.; Carpenter, B. K.; McLafferty, F. W. Electron Capture Dissociation for Structural Characterization of Multiply Charged Protein Cations. Anal. Chem. 2000, 72, 563-573. Chalmers, M. J.; Hakansson, K.; Johnson, R.; Smith, R.; Shen, J. W.; Emmett, M. R.; Marshall, A. G. Protein Kinase a Phosphorylation Characterized by Tandem Fourier Transform Ion Cyclotron Resonance Mass Spectrometry. Proteomics 2004, 4, 970-981. Kweon, H. K.; Hakansson, K. Metal Oxide-Based Enrichment Combined with Gas-Phase Ion-Electron Reactions for Improved Mass Spectrometric Characterization of Protein Phosphorylation. J. Proteome Res. 2008, 7, 749- 755. 260 151. 152. 153. 154. 155. 156. 157. 158. 159. 160. Horn, D. M.; Ge, Y.; McLafferty, F. W. Activated Ion Electron Capture Dissociation for Mass Spectral Sequencing of Larger (42 kDa) Proteins. Anal. Chem. 2000, 72, 4778-4784. Bushey, J. M.; Baba, T.; Glish, G. L. Simultaneous Collision Induced Dissociation of the Charge Reduced Parent Ion During Electron Capture Dissociation. Anal. Chem. 2009, 81, 6156-6164. McLafferty, F. W.; Horn, D. M.; Breuker, K.; Ge, Y.; Lewis, M. A.; Cerda, B.; Zubarev, R. A.; Carpenter, B. K. Electron Capture Dissociation of Gaseous Multiply Charged Ions by Fourier-Transform Ion Cyclotron Resonance. J. Am. Soc. Mass Spectrom. 2001, 12, 245-249. Stensballe, A.; Jensen, 0. N.; Olsen, J. V.; Haselmann, K. F.; Zubarev, R. A. Electron Capture Dissociation of Singly and Multiply Phosphorylated Peptides. Rapid Commun. Mass Spectrom. 2000, 14, 1793-1800. Shi, S. D. H.; Hemling, M. E.; Carr, S.,A.; Horn, D. M.; Lindh, I.; McLafferty, F. W. Phosphopeptide/Phosphoprotein Mapping by Electron Capture Dissociation Mass Spectrometry. Anal. Chem. 2001, 73, 19-22. Sweet, S. M. M.; Mardakheh, F. K.; Ryan, K. J. P.; Langton, A. J.; Heath, J. K.; Cooper, H. J. Targeted OnIine Liquid Chromatography Electron Capture Dissociation Mass Spectrometry for the Localization of Sites of In Vivo Phosphorylation In Human Sprouty2. Anal. Chem. 2008, 80, 6650-6657. Sweet, S. M. M.; Bailey, C. M.; Cunningham, D. L.; Heath, J. K.; Cooper, H. J. Large Scale Localization of Protein Phosphorylation by Use of Electron Capture Dissociation Mass Spectrometry. Mol. Cell. Proteomics 2009, 8, 904-912. Kjeldsen, F.; Haselmann, K. F.; Budnik, B. A.; Jensen, F.; Zubarev, R. A. Dissociative Capture of Hot (3-13 eV) Electrons by Polypeptide Polycations: An Efficient Process Accompanied by Secondary Fragmentation. Chem. Phys. Lett. 2002, 356, 201-206. Kjeldsen, F.; Haselmann, K. F.; Sorensen, E. S.; Zubarev, R. A. Distinguishing of IIe/Leu Amino Acid Residues in the PP3 Protein by (Hot) Electron Capture Dissociation in Fourier Transform Ion Cyclotron Resonance Mass Spectrometry. Anal. Chem. 2003, 75, 1267-1274. Syka, J. E. P.; Coon, J. J.; Schroeder, M. J.; Shabanowitz, J.; Hunt, D. F. Peptide and Protein Sequence Analysis by Electron Transfer Dissociation Mass Spectrometry. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 9528-9533. 261 161. 162. 163. 164. 165. 166. Chi, A.; Huttenhower, C.; Geer, L. Y.; Coon, J. J.; Syka, J. E. P.; Bai, D. L.; Shabanowitz, J.; Burke, D. J.; Troyanskaya, O. G.; Hunt, D. F. Analysis of Phosphorylation Sites on Proteins from Saccharomyces Cerevisiae by Electron Transfer Dissociation (ETD) Mass Spectrometry. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 2193-2198. Good, D. M.; Wirtala, M.; McAIister, G. C.; Coon, J. J. Performance Characteristics of Electron Transfer Dissociation Mass Spectrometry. Mol. Cell. Proteomics 2007, 6, 1942-1951. Gunawardena, H. P.; He, M.; Chrisman, P. A.; Pitteri, S. J.; Hogan, J. M.; Hodges, B. D. M.; McLuckey, S. A. Electron Transfer Versus Proton Transfer in Gas-Phase Ion/Ion Reactions of Polyprotonated Peptides. J. Am. Chem. Soc. 2005, 127, 12627-12639. Molina, H.; Horn, D. M.; Tang, N.; Mathivanan, S.; Pandey, A. Global Proteomic Profiling of Phosphopeptides Using Electron Transfer Dissociation Tandem Mass Spectrometry. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 2199-2204. Yang, F.; Wu, 8.; Stenoien, D. L.; Zhao, R.; Monroe, M. E.; Gritsenko, M. A.; Purvine, S. O.; Polpitiya, A. D.; ToIic, N.; Zhang, Q.; Norbeck, A. D.; Orton, D. J.; Moore, R. J.; Tang, K.; Anderson, G. A.; Pasa-Tolic, L.; Camp, D. G.; Smith, R. D. Combined Pulsed-Q Dissociation and Electron Transfer Dissociation for Identification and Quantification of iTRAQ-Labeled Phosphopeptides. Anal. Chem. 2009, 81 , 4137-4143. Swaney, D. L.; Wenger, C. D.; Thomson, J. A.; Coon, J. J. Human . Embryonic Stem Cell Phosphoproteome Revealed by Electron Transfer 167. 168. 169. Dissociation Tandem Mass Spectrometry. Proc. Natl. Acad. Sci. U. S. A. 2009, 106, 995-1000. Swaney, D. L.; McAIister, G. C.; Coon, J. J. Decision Tree-Driven Tandem Mass Spectrometry for Shotgun Proteomics. Nat Meth 2008, 5, 959-964. Swaney, D. L.; McAIister, G. C.; Wirtala, M.; Schwartz, J. C.; Syka, J. E. P.; Coon, J. J. Supplemental Activation Method for High-Efficiency Electron- Transfer Dissociation of Doubly Protonated Peptide Precursors. Anal. Chem. 2007, 79, 477-485. Molina, H.; Matthiesen, R.; Kandasamy, K.; Pandey, A. Comprehensive Comparison of Collision Induced Dissociation and Electron Transfer Dissociation. Anal. Chem. 2008, 80, 4825-4835. 262 170. 171. 172. 173. 174. 175. 176. 177. 178. 179. , with Metastable Atoms. Anal. Chem. 2006, 78, 3055-3061. Taouatas, N.; Drugan, M. M.; Heck, A. J. R.; Mohammed, S. Straightforward Ladder Sequencing of Peptides Using a Lys-N Metalloendopeptidase. Nat Meth 2008, 5, 405-407. Taouatas, N.; AIteIaar, A. F. M.; Drugan, M. M.; Helbig, A. O.; Mohammed, 8.; Heck, A. J. R. Strong Cation Exchange-Based Fractionation of Lys-N- Generated Peptides Facilitates the Targeted Analysis of Post-Translational Modifications. Mol. Cell. Proteomics 2009, 8, 190-200. Boersema, P. J.; Taouatas, N.; AIteIaar, A. F. M.; Gouw, J. W.; Ross, P. L.; Pappin, D. J.; Heck, A. J. R.; Mohammed, 8. Straightfonrvard and De Novo Peptide Sequencing by MALDI-MSIMS Using a Lys-N Metalloendopeptidase. Mol. Cell. Proteomics 2009, 8, 650-660. Hennn'ch, M. L.; Boersema, P. J.; van den Toom, H.; Mischen'kow, N.; Heck, A. J. R.; Mohammed, 8. Effect of Chemical Modifications on Peptide Fragmentation Behavior Upon Electron Transfer Induced Dissociation. Anal. Chem. 2009, 81, 7814-7822. Gauci, 8.; Helbig, A. O.; Slijper, M.; Krijgsveld, J.; Heck, A. J. R.; Mohammed, 8. Lys-N and Trypsin Cover Complementary Parts of the Phosphoproteome in a Refined SCX—Based Approach. Anal. Chem. 2009, 81 , 4493-4501 . Janek, K.; Wenschuh, H.; Bienert, M.; Krause, E. Phosphopeptide Analysis by Positive and Negative Ion Matrix-Assisted Laser Desorption/Ionization Mass Spectrometry. Rapid Commun. Mass Spectrom. 2001, 15, 1593-1599. Gunawardena, H. P.; Emory, J. F.; McLuckey, 8. A. Phosphopeptide Anion Characterization Via Sequential Charge Inversion and Electron-Transfer Dissociation. Anal. Chem. 2006, 78, 3788-3793. Kjeldsen, F.; Silivra, O. A.; Ivonin, I. A.; Haselmann, K. F.; Gorshkov, M.; Zubarev, R. A. Calpha -C Backbone Fragmentation Dominates in Electron Detachment Dissociation of Gas-Phase Polypeptide Ponanions. Chem.- Eur. J. 2005, 1 1, 1803-1812. Budnik, B. A.; Haselmann, K. F.; Zubarev, R. A. Electron Detachment Dissociation of Peptide Di-Anions: An Electron-Hole Recombination Phenomenon. Chem. Phys. Lett. 2001, 342, 299-302. Berkout, V. D. Fragmentation of Protonated Peptide Ions Via Interaction 263 180. 181. 182. 183. 184. 185. 186. 187. 188. 189. Berkout, V. D.; Doroshenko, V. M. Fragmentation of Phosphorylated and Singly Charged Peptide Ions Via Interaction with Metastable Atoms. Int. J. Mass Spectrom. 2008, 278, 150-157. Cook, 8. L.; Collin, O. L.; Jackson, G. P. Metastable Atom-Activated Dissociation Mass Spectrometry: Leucine/Isoleucine Differentiation and Ring Cleavage of Proline Residues. J. Mass Spectrom. 2009, 44, 1211- 1223. Stingl, C.; IhIing, C.; Ammerer, G.; Sinz, A.; Mechtler, K. Application of Different Fragmentation Techniques for the Analysis of Phosphopeptides Using a Hybrid Linear Ion Trap-FTICR Mass Spectrometer. BEA-Proteins Proteom. 2006, 1 764, 1842-1852. Hakansson, K.; Chalmers, M. J.; Quinn, J. P.; McFarland, M. A.; Hendrickson, C. L.; Marshall, A. G. Combined Electron Capture and Infrared Multiphoton Dissociation for Multistage MS/MS in a Fourier Transform Ion Cyclotron Resonance Mass Spectrometer. Anal. Chem. 2003, 75, 3256- 3262. Paizs, B.; Suhai, S. Fragmentation Pathways of Protonated Peptides. Mass Spectrom. Rev. 2005, 24, 508-548. Steen, H.; Mann, M. The ABC's (and XYZ's) of Peptide Sequencing. Nat. Rev. Mol. Cell Biol. 2004, 5, 699-711. Cooper, H. J. Hudgins, R. R.; Hakansson, K, Marshall, A. G. Secondary Fragmentation of Linear Peptides in Electron Capture Dissociation. Int. J. Mass Spectrom. 2003, 228, 723-728. Beausoleil, S. A.; Villen, J.; Gerber, S. A.; Rush, J.; Gygi, S. P. A Probability-Based Approach for High-Throughput Protein Phosphorylation Analysis and Site Localization. Nat. Biotech. 2006, 24, 1285-1292. Kocher, T.; Savitski, M. M.; Nielsen, M. L.; Zubarev, R. A. Phostshunter: A Fast and Reliable Tool to Detect Phosphorylated Peptides in Liquid Chromatography Fourier Transform Tandem Mass Spectrometry Data Sets. J. Proteome Res. 2006, 5, 659-668. Elias, J. E.; Gibbons, F. D.; King, O. D.; Roth, F. P.; Gygi, 8. P. Intensity- Based Protein Identification by Machine Learning from a Library of Tandem Mass Spectra. Nat. Biotech. 2004, 22, 214-219. 264 190. 191. 192. 193. 194. 195. 196. 197. 198. Zhang, Z. Prediction of Low-Energy Collision-Induced Dissociation Spectra of Peptides. Anal. Chem. 2004, 76, 3908-3922. Narasimhan, C.; Tabb, D. L.; VerBerkmoes, N. C.; Thompson, M. R.; Hettich, R. L.; Uberbacher, E. C. Maspic: Intensity-Based Tandem Mass Spectrometry Scoring Scheme That Improves Peptide Identification at High Confidence. Anal. Chem. 2005, 77, 7581-7593. Sun, 8.; Meyer-Arendt, K.; Eichelberger, B.; Brown, R.; Yen, C.-Y.; Old, W. M.; Pierce, K.; Cios, K. J.; Ahn, N. G.; Resing, K. A. Improved Validation of Peptide MS/MS Assignments Using Spectral Intensity Prediction. Mol. Cell. Proteomics 2007, 6, 1-17. Kapp, E. A.; Schutz, F.; Reid, G. E.; Eddes, J. 8.; 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. Anal. Chem. 2003, 75, 6251-6264. Hunter, E. P. L.; Lies, 8. G. Evaluated Gas Phase Basicities and Proton Affinities of Molecules: An Update. J. Phys. Chem. Ref. Data 1998, 27, 413- 656. Paladini, A.; Scuderi, D.; Lagana, A.; Giardini, A.; Filippi, A.; Speranza, M. Gas-Phase Basicity and Enantiodiscrimination of Some Phosphorous- Containing alpha-Amino Acid Mimics. Int. J. Mass Spectrom. 2003, 228, 349-358. Jackson, S. N.; Wang, H. Y. J.; Woods, A. 8. Study of the Fragmentation Patterns of the Phosphate-Arginine Noncovalent Bond. J. Proteome Res. 2005, 4, 2360-2363. Reid, 6.; Simpson, R.; O'Hair, R. J. A Mass Spectrometric and Ab Initio Study of the Pathways for Dehydration of Simple Glycine and Cysteine- Containing Peptide [M+H]+ Ions. J. Am. Soc. Mass Spectrom. 1998, 9, 945- 956. Neta, P.; Pu, Q.-L.; Yang, X.; Stein, 8. E. Consecutive Neutral Losses of H20 and C2H4O from N-Terminal Thr-Thr and Thr-Ser in Collision-Induced Dissociation of Protonated Peptides: Position Dependent Water Loss from Single Thr or Ser. Int. J. Mass Spectrom. 2007, 267, 295-301. 265 199. 200. 201. 202. 203. 204. 205. 206. 207. Ballard, K. D.; Gaskell, 8. J. Intramolecular Oxygen-18 Isotopic Exchange in the Gas Phase Observed During the Tandem Mass Spectrometric Analysis of Peptides. J. Am. Chem. Soc. 1992, 114, 64-71. Lee, C. H.; McComb, M. E.; Bromirski, M.; Jilkine, A.; Ens, W.; Standing, K. G.; Perreault, H. On-Membrane Digestion of beta-Casein for Determination of Phosphorylation Sites by Matrix-Assisted Laser Desorption/Ionization Quadrupole/'I'ime-of—Flight Mass Spectrometry. Rapid Commun. Mass Spectrom. 2001 , 1 5, 191-202. Olsen, J. V.; Mann, M. Improved Peptide Identification in Proteomics by Two Consecutive Stages of Mass Spectrometric Fragmentation. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 13417-13422. Lu, B.; Ruse, C.; Xu, T.; Park, S. K.; Yates, J. Automatic Validation of Phosphopeptide Identifications from Tandem Mass Spectra. Anal. Chem. 2007, 79, 1301-1310. Matthiesen, R.; TreIIe, M. B.; Hojrup, P.; Bunkenborg, J.; Jensen, O. N. VEMS 3.0: Algorithms and Computational Tools for Tandem Mass Spectrometry Based Identification of Post-Translational Modifications in Proteins. J. Proteome Res. 2005, 4, 2338-2347. Wan, Y.; Cripps, D.; Thomas, 8.; Campbell, P.; Ambulos, N.; Chen, T.; Yang, A. Phosphoscan: A Probability-Based Method for Phosphorylation Site Prediction Using M82/MS3 Pair lnforrnation. J. Proteome Res. 2008, 7, 2803-2811. DieIIa, F.; Cameron, 8.; Gemund, C.; Linding, R.; Via, A.; Kuster, B.; Sicheritz-Ponten, T.; Blom, N.; Gibson, T. Phospho.EIm: A Database of Experimentally Verified Phosphorylation Sites in Eukaryotic Proteins. BMC Bioinformatics 2004, 5, 79. Palumbo, A. M.; Tepe, J. J.; Reid, G. E. Mechanistic Insights into the Multistage Gas-Phase Fragmentation Behavior of Phosphoserine- and Phosphothreonine-Containing Peptides. J. Proteome Res. 2008, 7, 771- 779. Thome, G. C.; Ballard, K. D.; Gaskell, S. J. Metastable Decomposition of Peptide [M + H]+ Ions Via Rearrangement Involving Loss of the C-Terminal Amino Acid Residue. J. Am. Soc. Mass Spectrom. 1990, 1, 249-257. 266 208. 209. 210. 211. 212. 213. 214. 215. 216. 217. Gonzalez, J.; Besada, V.; Garay, H.; Reyes, O.; Padron, G.; Tambara, Y.; Takao, T.; Shimonishi, Y. Effect of the Position of a Basic Amino Acid on C- Tenninal Rearrangement of Protonated Peptides Upon Collision-Induced Dissociation. J. Mass Spectrom. 1996, 31, 150-158. Farrugia, J. M.; O'Hair, R. A. J. Involvement of Salt Bridges in a Novel Gas Phase Rearrangement of Protonated Arginine-Containing Dipeptides Which Precedes Fragmentation. Int. J. Mass Spectrom. 2003, 222, 229-242. Tang, X. J.; Thibault, P.; Boyd, R. K. Fragmentation Reactions of Multiply- Protonated Peptides and Implications for Sequencing by Tandem Mass Spectrometry with Low-Energy Collision-Induced Dissociation. Anal. Chem. 1993, 65, 2824-2834. Vachet, R. W.; Bishop, B. M.; Erickson, B. W.; Glish, G. L. Novel Peptide Dissociation: Gas-Phase Intramolecular Rearrangement of Internal Amino Acid Residues. J. Am. Chem. Soc. 1997, 119, 5481-5488. Farrugia, J. M.; Taverner, T.; O'Hair, R. A. J. Side-Chain Involvement in the Fragmentation Reactions of the Protonated Methyl Esters of Histidine and Its Peptides. Int. J. Mass Spectrom. 2001, 209, 99-112. Yague, J.; Paradela, A.; Ramos, M.; Ogueta, 8.; Marina, A.; Barahona, F.; Lopez de Castro, J. A.; Vazquez, J. Peptide Rearrangement During Quadrupole Ion Trap Fragmentation: Added Complexity to MS/MS Spectra. Anal. Chem. 2003, 75, 1524-1535. Polfer, N. C.; Oomens, J.; Suhai, 8.; Paizs, B. Spectroscopic and Theoretical Evidence for Oxazolone Ring Formation in Collision-Induced Dissociation of Peptides. J. Am. Chem. Soc. 2005, 127, 17154-17155. Polfer, N. C.; Oomens, J.; Suhai, 8.; Paizs, B. Infrared Spectroscopy and Theoretical Studies on Gas-Phase Protonated Leu-Enkephalin and Its Fragments: Direct Experimental Evidence for the Mobile Proton. J. Am. Chem. Soc. 2007, 129, 5887-5897. Riba-Garcia, I.; Giles, K.; Bateman, R. H.; Gaskell, S. J. Evidence for Structural Variants of a- and b-Type Peptide Fragment Ions Using Combined Ion Mobility/Mass Spectrometry. J. Am. Soc. Mass Spectrom. 2008, 19, 609-613. Harrison, A. G.; Young, A. B.; Bleiholder, C.; Suhai, 8.; Paizs, B. Scrambling of Sequence Information in Collision-Induced Dissociation of Peptides. J. Am. Chem. Soc. 2006, 128, 10364-10365. 267 218. 219. 220. 221. 222. 223. 224. 225. 226. 227. 228. 229. Harrison, A. G. Peptide Sequence Scrambling through Cyclization of b5 Ions. J. Am. Soc. Mass Spectrom. 2008, 19, 1776—1780. Mouls, L.; Aubagnac, J. L.; Martinez, J.; Enjalbal, C. Low Energy Peptide Fragmentations in an ESI-Q-TOF Type Mass Spectrometer. J. Proteome Res. 2007, 6, 1378-1391. Fuchs, R.; Budzikiewicz, H. Rearrangement Reactions in the Electrospray Ionization Mass Spectra of Pyoverdins. Int. J. Mass Spectrom. 2001, 210- 211, 603-612. Cox, H. A.; Hodyss, R.; Beauchamp, J. L. Cluster-Phase Reactions: Gas- Phase Phosphorylation of Peptides and Model Compounds with Triphosphate Anions. J. Am. Chem. Soc. 2005, 127, 4084-4090. Palumbo, A. M.; Reid, G. E. Evaluation of Gas-Phase Rearrangement and Competing Fragmentation Reactions on Protein Phosphorylation Site Assignment Using Collision Induced Dissociation-MSIMS and M83. Anal. Chem. 2008,80, 9735-9747. Edelson-Averbukh, M.; Shevchenko, A.; Pipkom, R.; Lehmann, W. D. Gas- Phase Intramolecular Phosphate Shift in Phosphotyrosine-Containing Peptide Monoanions. Anal. Chem. 2009, 81, 4369-4381. Kaiser, E.; Colescott, R. L.; Bossinger, C. 0.; Cook, P. I. Color Test for Detection of Free Terminal Amino Groups in the Solid-Phase Synthesis of Peptides. Anal. Biochem. 1970, 34, 595-598. Novabiochem 2004/5 Catalog. EMD Biosciences: San Diego, CA, 2004; pp pp 3.15-3.17. Sen, R.; Baltimore, D. Multiple Nuclear Factors Interact with the lmmunoglobulin Enhancer Sequences. Cell 1986, 46, 705-716. Hayden, M. 8.; Ghosh, 8. Shared Principles in NF-kappaB Signaling. Cell 2008, 132, 344-362. Bonizzi, G.; Karin, M. The Two NF-kappaB Activation Pathways and Their Role in Innate and Adaptive Immunity. Trends Immunol. 2004, 25, 280-288. Vallabhapurapu, 8.; Karin, M. Regulation and Function of NF-kappaB Transcription Factors in the Immune System. Annu. Rev. Immunol. 2009, 27, 693-733. 268 230. 231. 232. 233. 234. 235. 236. 237. 238. 239. Karin, M.; Lin, A. NF-kappaB at the Crossroads of Life and Death. Nat lmmunol 2002, 3, 221-227. Richmond, A. NF-kappaB, Chemokine Gene Transcription and Tumour Growth. Nat Rev lmmunol 2002, 2, 664-674. Shimizu, H.; Mitomo, K.; Watanabe, T.; Okamoto, 8.; Yamamoto, K. Involvement of a NF-kappaB-Like Transcription Factor in the Activation of the Interleukin-6 Gene by Inflammatory Lymphokines. Mol. Cell. Biol. 1990, 10, 561-568. Liberrnann, T. A.; Baltimore, D. Activation of Interleukin-6 Gene Expression through the NF-kappaB Transcription Factor. Mol. Cell. Biol. 1990, 10, 2327-2334. Shakhov, A. N.; Collart, M. A.; Vassalli, P.; Nedospasov, S. A.; Jongeneel, C. V. Kappab-Type Enhancers Are Involved in Lipoponsaccharide-Mediated Transcriptional Activation of the Tumor Necrosis Factor alpha Gene in Primary Macrophages. J. Exp. Med. 1990, 171, 35-47. van de Stolpe, A.; Caldenhoven, E.; Stade, B. G.; Koendennan, L.; Raaijmakers, J. A.; Johnson, J. P.; van der Saag, P. T. 12-O- Tetradecanoylphorbol-13-Acetate- and Tumor Necrosis Factor alpha- Mediated Induction of Intercellular Adhesion Molecule-1 Is Inhibited by Dexamethasone. Functional Analysis of the Human Intercellular Adhesion Molecular-1 Promoter. J. Biol. Chem. 1994, 269, 6185-6192. Kelly, A.; Craig, N. R.; Hing, Y. L. NF-kappaB Activation Upregulates Fibroblast Growth Factor 8 Expression in Prostate Cancer Cells. The Prostate 2006, 66, 1223-1234. Catz, S. D.; Johnson, J. L. Transcriptional Regulation of BcI-2 by Nuclear Factor-kappaB and Its Significance in Prostate Cancer. Oncogene 2001, 20, 7342-7351. Schauvliege, R.; Vanrobaeys, J.; Schotte, P.; Beyaert, R. Caspase-11 Gene Expression in Response to Lipopolysaccharide and Interferon-gamma Requires Nuclear Factor-Kappab and Signal Transducer and Activator of Transcription (STAT) 1. J. Biol. Chem. 2002, 277, 41624-41630. Kumar, A.; Takada, Y.; Boriek, A. M.; Agganrval, B. B. Nuclear Factor- kappaB: Its Role in Health and Disease. Journal of Molecular Medicine 2004, 82, 434-448. 269 240. 241. 242. 243. 244. 245. 246. 247. 248. 249. 250. 251. Courtois, G.; Gilmore, T. D. Mutations in the NF-kappaB Signaling Pathway: Implications for Human Disease. Oncogene 2006, 25, 6831-6843. Naugler, W. E.; Karin, M. NF-kappaB and Cancer - Identifying Targets and Mechanisms. Curr. Opin. Genet. Dev. 2008, 18, 19-26. Karin, M.; Cao, Y.; Greten, F. R.; Li, Z.-W. NF-kappaB in Cancer: From Innocent Bystander to Major Culprit. Nat. Rev. Cancer 2002, 2, 301-310. Karin, M. Nuclear Factor-kappaB in Cancer Development and Progression. Nature 2006, 441, 431-436. Lee, C. H.; Jeon, Y.-T.; Kim, S.-H.; Song, Y.-S. NF-kappaB as a Potential Molecular Target for Cancer Therapy. BioFactors 2007, 29, 19-35. Bacher, 8.; Schmitz, M. L. The NF-kappaB Pathway as a Potential Target for Autoimmune Disease Therapy. Curr. Pharm. Des. 2004, 10, 2827-2837. Niederberger, E.; Geisslinger, G. The‘IKK-NF-kappaB Pathway: A Source for Novel Molecular Drug Targets in Pain Therapy? FASEB J. 2008, 22, 3432-3442. Gilmore, T. D. Introduction to NF-kappaB: Players, Pathways, Perspectives. Oncogene 2006, 25, 6680-6684. Hoffmann, A.; Baltimore, D. Circuitry of Nuclear Factor kappaB Signaling. lmmunol Rev 2006, 210, 171-186. Osborn, L.; Kunkel, 8.; Nabel, G. J. Tumor Necrosis Factor-Alpha and Interleukin-1 Stimulate the Human Immunodeficiency Virus Enhancer by Activation of the Nuclear Factor Kappab. Proc. Natl. Acad. Sci. U. S. A. 1989, 86, 2336-2340. Sen, R.; Baltimore, D. lnducibility of kappa lmmunoglobulin Enhancer- Binding Protein NF-kappaB by a Posttranslational Mechanism. Cell 1986, 47, 921-928. Piret, B.; Piette, J. Topoisomerase Poisons Activate the Transcription Factor NF-kappaB in ACH-2 and CEM Cells. Nucl. Acids Res. 1996, 24, 4242- 4248. 270 252. 253. 254. 255. 256. 257. 258. 259. 260. Schreck, R.; Alberrnann, K.; Baeuerle, P. A. Nuclear Factor Kappab - an Oxidative Stress-Responsive Transcription Factor of Eukaryotic Cells. Free Radical Res. Commun. 1992, 1 7, 221-237. Stein, B.; Rahmsdorf, H. J.; Steffen, A.; Litfin, M.; Herrlich, P. UV-lnduced DNA Damage Is an Intermediate Step in UV-lnduced Expression of Human Immunodeficiency Virus Type 1, Collagenase, c-fos, and Metallothionein. Mol. Cell. Biol. 1989, 9, 5169-5181. Sharma, V.; Hupp. C. D.; Tepe, J. J. Enhancement of Chemotherapeutic Efficacy by Small Molecule Inhibition of NF-kappaB and Checkpoint Kinases. Curr. Med. Chem. 2007, 14, 1061-1074 Ghosh, 8.; Karin, M. Missing Pieces in the NF-kappaB Puzzle. Cell 2002, 109, 881-896. Perkins, N. D. Post-Translational Modifications Regulating the Activity and Function of the Nuclear Factor kappaB Pathway. Oncogene 2006, 25, 6717- 6730. O'Shea, J. M.; Perkins, N. D. Regulation of the ReIA (p65) Transactivation Domain. Biochem. Soc. Trans. 2008, 036, 603-608. Mankan, A. K.; Lawless, M. W.; Gray, 8. G.; Kelleher, D.; McManus, R. NF- kappaB Regulation: The Nuclear Response. J Cell Mol Med 2009, 13, 631- 643. Bouwmeester, T.; Bauch, A.; Ruffner, H.; Angrand, P.-O.; Bergamini, G.; Croughton, K.; Cruciat, C.; Eberhard, D.; Gagneur, J.; Ghidelli, 8.; Hopf, C.; Huhse, B.; Mangano, R.; Michon, A.-M.; Schirle, M.; Schlegl, J.; Schwab, M.; Stein, M. A.; Bauer, A.; Casari, G.; Drewes, G.; Gavin, A.-C.; Jackson, D. B.; Joberty, G.; Neubauer, 6.; Rick, J.; Kuster, B.; Superti-Furga, G. A Physical and Functional Map of the Human TNF-alpha/NF-kappaB Signal Transduction Pathway. Nat Cell Biol 2004, 6, 97-105. Buss, H.; Dorrie, A.; Schmitz, M. L.; Hoffmann, E.; Resch, K.; Kracht, M. Constitutive and Interleukin-1-Inducible Phosphorylation of p65 NF-kappaB at Serine 536 Is Mediated by Multiple Protein Kinases Including IkappaB Kinase (IKK)—alpha, beta, gamma, TRAF Family Member-Associated (TANK)-Binding Kinase 1 (TBK1), and an Unknown Kinase and Couples p65 to TATA-Binding Protein-Associated Factor II31-Mediated Interleukin-8 Transcription. J. Biol. Chem. 2004, 279, 55633-55643. 271 261. 262. 263. 264. 265. 266. 267. 268. 269. 270. Buss, H.; Dorrie, A.; Schmitz, M. L.; Frank, R.; Livingstone, M.; Resch, K.; Kracht, M. Phosphorylation of Serine 468 by Gsk—3 beta Negatively Regulates Basal p65 NF-kappaB Activity. J. Biol. Chem. 2004, 279, 49571- 49574. Schwabe, R. F.; Sakurai, H. Ikkbeta Phosphorylates p65 at 8468 in Transactivaton Domain 2. FASEB J. 2005, 19, 1758-1760. Mattioli, |.; Geng, H.; Sebald, A.; HodeI, M.; Bucher, C.; Kracht, M.; Schmitz, M. L. Inducible Phosphorylation of NF-kappaB p65 at Serine 468 by T Cell Costimulation Is Mediated by IKK Epsilon. J. Biol. Chem. 2006, 281, 6175- 6183. Shen, C. H.; Stavnezer, J. Interaction of STAT6 and NF-kappaB: Direct Association and Synergistic Activation of Interleukin-4-Induced Transcription. Mol. Cell. Biol. 1998, 18, 3395-3404. Sun, 8.; Tang, Y.; Lou, X.; Zhu, L.; Yang, K.; Zhang, B.; Shi, H.; Wang, C. Uxt Is a Novel and Essential Cofactor in the NF-kappaB Transcriptional Enhanceosome. J. Cell Biol. 2007, 1 78, 231-244. Verrna, N.; Chaudhury, I.; Kumar, D.; Das, R. H. Silencing of TNF-aIpha Receptors Coordinately Suppresses TNF-alpha Expression through NF- kappaB Activation Blockade in THP-1 Macrophage. FEBS Lett. 2009, 583, 2968-2974. OIIivier, V.; Parry, G. C. N.; Cobb, R. R.; de Prost, D.; Mackman, N. Elevated Cyclic AMP Inhibits NF-kappaB-Mediated Transcription in Human Monocytic Cells and Endothelial Cells. J. Biol. Chem. 1996, 271, 20828- 20835. Moon, D.-O.; Kim, M.-O.; Kang, S.-H.; Choi, Y. H.; Kim, G.-Y. Sulforaphane Suppresses TNF-alpha-Mediated Activation of NF-kappaB and Induces Apoptosis through Activation of Reactive Oxygen Species-Dependent Caspase-3. Cancer Lett 2009, 274, 132-142. Suzuki, 8.; Singhirunnusom, P.; Nakano, H.; Doi, T.; Saiki, I.; Sakurai, H. Identification of TNF-alpha-Responsive NF- -kappaB p65- Binding Element in the Distal Promoter of the Mouse Serine Protease Inhibitor Serpine2. FEBS Lett. 2006, 580, 3257- 3262. Lu, Z.-G.; Liu, H.; Yamaguchi, T.; Miki, Y.; Yoshida, K. Protein Kinase Cdelta Activates ReIA/p65 and Nuclear Factor-kappaB Signaling In 272 271. 272. 273. 274. 275. 276. 277. 278. 279. Response to Tumor Necrosis Factor-alpha. Cancer Res 2009, 69, 5927- 5935. Lagenrverf, F. M.; Van de Weert, M.; Heerrna, W.; Haverkamp, J. Identification of Oxidized Methionine in Peptides. Rapid Commun. Mass Spectrom. 1996, 10, 1905-1910. Lioe, H.; Laskin, J.; Reid, G. E.; O'Hair, R. A. J. Energetics and Dynamics of the Fragmentation Reactions of Protonated Peptides Containing Methionine Sulfoxide or Aspartic Acid Via Energy- and Time-Resolved Surface Induced Dissociation. The Journal of Physical Chemistry A 2007, 111, 10580-10588. Lapko, V. N.; Smith, D. L.; Smith, J. B. Identification of an Artifact in the Mass Spectrometry of Proteins Derivatized with Iodoacetamide. J. Mass Spectrom. 2000, 35, 572-575. KriJger, R.; Hung, C.-W.; Edelson-Averbukh, M.; Lehmann, W. D. Iodoacetamide-Alkylated Methionine Can Mimic Neutral Loss of Phosphoric Acid from Phosphopeptides as Exemplified by Mano-Electrospray Ionization Quadrupole Time-of-Flight Parent Ion Scanning. Rapid Commun. Mass Spectrom. 2005, 1 9, 1 709-1 71 6. 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 -1 150. Beausoleil, 8. A.; Jedrychowski, M.; Schwartz, D.; Elias, J. E.; Villen, J.; Li, J. X.; Cohn, M. A.; Cantley, L. C.; Gygi, S. P. Large-Scale Characterization of HeLa Cell Nuclear Phosphoproteins. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 12130-12135. Macek, B.; Mann, M.; Olsen, J. V. Global and Site-Specific Quantitative Phosphoproteomics: Principles and Applications. Annu. Rev. Pharmacol. Toxicol. 2009, 49, 199-221. Daub, H.; Olsen, J. V.; Bairlein, M.; Gnad, F.; Opperrnann, F. 8.; Komer, R.; Greff, 2.; Keri, G.; Stemmann, 0.; Mann, M. Kinase-Selective Enrichment Enables Quantitative Phosphoproteomics of the Kinome across the Cell Cycle. Mol. Cell 2008, 31, 438-448. Dephoure, N.; Zhou, C.; Villen, J.; Beausoleil, S. A.; Bakalarski, C. E.; EIIedge, S. J.; Gygi, S. P. A Quantitative Atlas of Mitotic Phosphorylation. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 10762-10767. 273 280. 281. 282. 283. 284. 285. 286. 287. 288. 289. Uranishi, H.; Tetsuka, T.; Yamashita, M.; Asamitsu, K.; Shimizu, M.; Itoh, M.; Okamoto, T. Involvement of the Pro-Oncoprotein TLS (Translocated in Liposarcoma) in Nuclear Factor-kappaB p65-Mediated Transcription as a Coactivator. J. Biol. Chem. 2001, 276, 13395-13401. Are, A. F.; Galkin, V. E.; Pospelova, T. V.; Pinaev, G. P. The p65/ReIA Subunit of NF-kappaB Interacts with Actin-Containing Structures. Exp. Cell Res. 2000, 256, 533-544. ' Fazal, F.; Minhajuddin, M.; Bijli, K. M.; McGrath, J. L.; Rahman, A. Evidence for Actin Cytoskeleton-Dependent and -Independent Pathways for ReIAlp65 Nuclear Translocation in Endothelial Cells. J. Biol. Chem. 2007, 282, 3940- 3950. Fazal, F.; Bijli, K. M.; Minhajuddin, M.; Rein, T.; Finkelstein, J. N.; Rahman, A. Essential Role of Cofilin-1 in Regulating Thrombin-Induced ReIA/p65 Nuclear Translocation and Intercellular Adhesion Molecule 1 (ICAM-1) Expression in Endothelial Cells. J. Biol. Chem. 2009, 284, 21047-21056. Dunn, J. D.; Igrisan, E. A.; Palumbo, A. M.; Reid, G. E.; Bruening, M. L. Phosphopeptide Enrichment Using MALDI Plates Modified with High- Capacity Polymer Brushes. Anal. Chem. 2008, 80, 5727-5735. Lindstrém, M. 8. Emerging Functions of Ribosomal Proteins in Gene- Specific Transcription and Translation. Biochem. Biophys. Res. Commun. 2009, 379, 167-170. Wool, I. G. Extraribosomal Functions of Ribosomal Proteins. Trends Biochem. Sci. 1996, 21, 164-165. Wan, F.; Anderson, D. E.; Bamitz, R. A.; Snow, A.; Bidere, N.; Zheng, L.; Hegde, V.; Lam, L. T.; Staudt, L. M.; Levens, D.; Deutsch, W. A.; Lenardo, M. J. Ribosomal Protein 83: A KH Domain Subunit in NF-kappaB Complexes That Mediates Selective Gene Regulation. Cell 2007, 131, 927- 939. Rich, B. E.; Steitz, J. A. Human Acidic Ribosomal Phosphoproteins-P0, Phosphoproteins-P1, and Phosphoproteins-P2 - Analysis of cDNA Clones, Invitro Synthesis, and Assembly. Mol. Cell. Biol. 1987, 7, 4065-4074. MacConneIl, W. P.; KapIan, N. O. The Activity of the Acidic Phosphoproteins from the 80 S Rat Liver Ribosome. J. Biol. Chem. 1982, 257, 5359-5366. 274 290. 291. 292. 293. 294. 295. 296. 297. 298. Sanchez-Madrid, F.; Reyes, R.; Conde, P.; Ballesta, J. P. G. Acidic Ribosomal Proteins from Eukaryotic Cells. Eur. J. Biochem. 1979, 98, 409- 416. Gomez—Lorenzo, M. G.; Spahn, C. M. T.; Agrawal, R. K.; Grassucci, R. A.; Penczek, P.; Chakraburtty, K.; Ballesta, J. P. G.; Lavandera, J. L.; Garcia- Bustos, J. F.; Frank, J. Three-Dimensional Cryo-Electron Microscopy Localization of EF2 in the Saccharomyces Cerevisiae 808 Ribosome at 17.5 a Resolution. EMBO J. 2000, 19, 2710-2718. Martinez-Azorin, F.; Remacha, M.; Ballesta, J. P. G. Functional Characterization of Ribosomal P1/P2 Proteins in Human Cells. Biochem. J. 2008, 413, 527-534. Bee, A.; Ke, Y.; Forootan, 8.; Lin, K.; Beesley, C.; Forrest, S. E.; Foster, C. S. Ribosomal Protein L19 Is a Prognostic Marker for Human Prostate Cancer. Clinical Cancer Research 2006, 12, 2061-2065. Huang, C. J.; Chien, C. C.; Yang, S. H.; Chang, C. C.; Sun, H. L.; Cheng, Y. C.; Liu, C. 0.; Lin, 8. 0.; Lin, C. M. Faecal Ribosomal Protein L19 Is a Genetic Prognostic Factor for Survival in Colorectal Cancer. J Cell Mol Med 2008, 12, 1936-1943. Dressman, M. A.; Baras, A.; Malinowski, R.; Alvis, L. B.; Kwon, I.; Walz, T. M.; Polymeropoulos, M. H. Gene Expression Profiling Detects Gene Amplification and Differentiates Tumor Types in Breast Cancer. Cancer Res 2003, 63, 2194-2199. Bertolotti, A.; Melot, T.; Acker, J.; Vigneron, M.; Delattre, O.; Tora, L. EWS, but Not EWS-FLI-1, Is Associated with Both TFIID and RNA Polymerase II: Interactions between Two Members of the TET Family, EWS and HTAF(II)68, and Subunits of TFIID and RNA Polymerase II Complexes. Mol. Cell. Biol. 1998, 18, 1489-1497. Bertolotti, A.; Lutz, Y.; Heard, D. J.; Chambon, P.; Tora, L. HTAF(II)68, a Novel RNA/ssDNA-Binding Protein with Homology to the Pro-Oncoproteins TLS/FUS and EWS Is Associated with Both TFIID and RNA Polymerase II. EMBO J. 1996, 15, 5022-5031. Yang, L.; Embree, L. J.; Hickstein, D. D. TLS-ERG Leukemia Fusion Protein Inhibits RNA Splicing Mediated by Serine-Arginine Proteins. Mol. Cell. Biol. 2000, 20, 3345-3354. 275 299. 300. 301. 302. 303. 304. 305. 306. 307. 308. Yang, L.; Embree, L. J.; Hickstein, D. D. TLS/ERG and EWS/FLI Fusion Proteins Disrupt RNA Splicing Mediated by Serine-Arginine (Sr) Proteins. Blood 1999, 94, 694a-694a. Yang, L.; Embree, L. J.; Tsai, 8.; Hickstein, D. D. Oncoprotein TLS Interacts with Serine-Arginine Proteins Involved in RNA Splicing. J. Biol. Chem. 1998, 273, 27761 -27764. Yang, L.; Chansky, H. A.; Hickstein, D. D. EWS Center Dot Fli-1 Fusion Protein Interacts with Hyperphosphorylated RNA Polymerase II and Interferes with Serine-Arginine Protein-Mediated RNA Splicing. J. Biol. Chem. 2000, 275, 37612-37618. Chen, H.; Bernstein, B. W.; Bamburg, J. R. Regulating Actin-Filament Dynamics in Vivo. Trends Biochem. Sci. 2000, 25, 19-23. Cartier, M.-F.; Le Clainche, C.; Wiesner, 8.; Pantaloni, D. Actin—Based Motility: From Molecules to Movement. BioEssays 2003,25, 336-345. Grummt, I. Actin and Myosin as Transcription Factors. Curr. Opin. Genet. Dev. 2006, 16, 191-196. Zheng, B.; Han, M.; Bemier, M.; Wen, J.-k. Nuclear Actin and Actin-Binding Proteins in the Regulation of Transcription and Gene Expression. FEBS J. 2009, 276, 2669-2685. Franke, W. W. Actin's Many Actions Start at the Genes. Nat Cell Biol 2004, 6, 1013-1014. Miralles, F.; Visa, N. Actin in Transcription and Transcription Regulation. Curr. Opin. Cell Biol. 2006, 18, 261-266. Bradford, M. M. Rapid and Sensitive Method for Quantitation of Microgram Quantities of Protein Utilizing Principle of Protein-Dye Binding. Anal. Biochem. 1976, 72, 248-254. 276