I’ MICHIGAN STATE UNIV 1. ll n lljllllllflljlll ll 3193015 .1 31:, iii This is to certify that the dissertation entitled Novel Methodologies to Study Cysteine Status in Proteins by Chemical Cleavage and Mass— Mapping by MALDI-TOF Mass Spectrometry presented by Jiang Wu has been accepted towards fulfillment of the requirements for Ph. D. degreein Chemistry We“ V Major professor \ Date May 16, 1997 MSUicnnAfI-Irmnliim‘ ‘ '1, m” ‘ I ' ‘ 0-12771 ) A 'v—v '_v ‘ fivf‘vv 4 4'. d“ 0' * «8 04‘444—~ LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. ‘ TO AVOID FINES return on or before date due. \.-.- DAME DATE DUE DATE DUE '\ DEC 2 81998i MSU Is An Affirmative Action/Equal Opportunity Institution clclmmpma-DJ NOVEL METHODOLOGIES TO STUDY CYSTEINE STATUS IN PROTEINS BY CHEMICAL CLEAVAGE AND MASS—MAPPING BY MALDI-TOF MASS SPECTROMETRY By Jiang Wu ‘Ii ”‘1 A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1997 I ABSTRACT NOVEL METHODOLOGIES TO STUDY CYSTEINE STATUS IN PROTEINS BY CHEMICAL CLEAVAGE AND MAS S-MAPPING BY MALDI-TOF MASS SPECTROMETRY By J iang Wu Current methodology for characterizing free sulfhydryls and disulfide bonds in proteins involves the modification of cysteine residues, enzymatic digestion of protein chain for cleavage between half cystinyl residues, I-IPLC fractionation, and sequence determination of peptide fragments. This strategy is tedious, cumbersome, and can promote disulfide exchange. A simple methodology has been developed to localize free cysteine groups in peptides and proteins. This new approach employs a specific reaction between free sulfliydryls and 2-nitro-5-thiocyanobenzoic acid (NTCB) or 1-cyano-4-dimethy1amino- pyridinium tetrafluoroborate (CDAP) to specifically cyanylate cysteine thiols. The N- terminal peptide bond of the cyanylated cysteinyl residue can then be cleaved under alkaline conditions to form an amino-terminal peptide and a series of 2-iminothiazolidine- 4-carboxylyl peptides which can be mapped to the sequence by MALDI-MS. The cleavage conditions have been systematically investigated using a number of peptides containing different amino acids adjacent to the N-terminal side of cysteine residues. While the cleavage reaction was traditionally performed in pH 8.5-9.5 buffer at 37-65°C for 12-80 hours, optimal results have been obtained in 1M ammonium solution in which the cleavage is complete within an hour at ambient temperature. This improvement also minimizes side reactions resulting from prolonged exposure to the alkaline medium. A novel strategy is described for assignment of disulfide pairings in proteins. A denatured protein is subjected to limited reduction in acidic solution to produce a mixture of partially reduced protein isomers; the nascent sulfhydryls are immediately cyanylated by CDAP under the same buffered conditions. The cyanylated protein isomers, separated by and collected from reversed-phase HPLC, are subjected to the cleavage reaction to form truncated peptides, which, after further reduction of the remaining disulfide bonds, can be mass-mapped by MALDI-MS. This simple, fast, and sensitive strategy avoids disulfide scrambling and is applicable to disulfide characterization in proteins containing adjacent cysteines. Several proteins with various disulfide structures have been studied to demonstrate the feasibility of the methodology. Two techniques are introduced for the study of disulfide structures of protein folding intermediates. The cyanylation of thiol groups by the CDAP under acidic conditions has been applied to trapping folding intermediates of recombinant human epidermal grth factor (hEGF). Disulfide structures of seven CDAP-trapped intermediates have been identified by the approach described above. Both native and non- native intermediates were found in the folding process. The analysis of the disulfide intermediates provides new insight into the folding pathway of hEGF. ACI Iwould like to express my 51 Watson for his guidance. encourage study at MSL’. Iwould also like to II and support, and for sening as a sec Eugene LeGoii for sening on my I COl tram to acknowledge Drs l knowledge in protein chemistry an members, past and present. thank yo mani'lhanks to other members of the Bti’, Mel. Without their technical sup And last, but not the least, I daUShier, Mengyni, for their love, sup; 80 very gratefiil to my parents moth: their ' calmg, encouragement, and und State University. ACKNOWLEDGIVIENTS Iwould like to express my sincere gratitude to my research advisor, Dr. J. Throck Watson, for his guidance, encouragement, and financial support throughout my graduate study at MSU. I would also like to thank Dr. Doug Gage for his continual encouragement and support, and for serving as a second reader of my dissertation; Drs. John Allison and Eugene LeGoff for serving on my committee. I want to acknowledge Drs. Eugene Zaluzec, and Zhi H. Huang for sharing their knowledge in protein chemistry and organic chemistry. To all the Watson group members, past and present, thank you for your friendship and cooperation. I also owe many thanks to other members of the Mass Spectrometry Facility: Melinda, Paochi, Gary, Bev, Mel. Without their technical support, my graduate study would have been difficult. And last, but not the least, I would like to thank my wife, Ping Gong, and my daughter, Mengyu, for their love, support, and patience that make my life enjoyable. I am also very grateful to my parents, mother-in—law, brothers and sister, and sister-in-laws, for their caring, encouragement, and understanding during my graduate study at Michigan State University. ll Chemj TAE List of Tables. . . List ofFigureS , . - Chapter 1. Introduction 1. introduction _. . ll. Characterization of Cysteine Sun A Quantitative Determination of B. Localization of Sulihydryl Grc C. Disulfide Bond Assignment Ill. innit-Assisted laser Desorption (MALDI-MS) n The Matrix ....... _ B. Sample Preparation. C Instrumentation..... D. Characterization Features of M E. New Techniques in MALDl-TC F. Application of MALDl-TOF M W. References .......... Cha t Per 2. A Strategy to Locate Cystei Specific Chemical Cleavage byMALDi-TOF MS .. 1. Introduction... TABLE OF CONTENTS List of Tables ......................................................................................... viii List of Figures .......................................................................................... x Chapter I. Introduction .............................................................................. l I. Introduction ........................................................................................ 1 II. Characterization of Cysteine Status in Proteins ............................................... 3 A. Quantitative Determination of SH and 8-8 Groups in Proteins ........................ 5 B. Localization of Sulfhydryl Groups in Proteins ........................................... 16 C. Disulfide Bond Assignment ................................................................ 21 III. Matrix-Assisted Laser Desorption/Ionization Mass Spectrometry (MALDI-MS) ................................................................................... 33 A. The Matrix ................................................................................... 35 B. Sample Preparation ......................................................................... 36 C. Instrumentation ............................................................................. 38 D. Characterization Features of MALDI-TOF MS ........................................ 40 E. New Techniques in MALDI-TOF MS ................................................... 41 F. Application of MALDI-TOF MS ......................................................... 43 IV. References ........................................................................................ 46 Chapter 2. A Strategy to Locate Cysteine Residues in Proteins by Specific Chemical Cleavage Followed by Mass Mapping by MALDI-TOF MS ................................................................. 57 I. Introduction ..................................................................................... 57 II. Chemical Cleavage at Cysteine Residues .................................................... 58 III. Experimental Section ........................................................................... 65 IV. Results and Discussion ........................................................................ 70 A. Cyanylation and Cleavag B. Effects of pH and Tempe C. Identification of Free SH D. Choice ofReducrng Ager E. Stability of Disulfide Bon F. Analysis by MALDI-h-lS G. Reactions of Proteins imn l’. Conclusions ._, l’l References. . . . Chapteri. Further Study on the Lee Residues by Chemical C it Optimization and Applica i. introduction ..... . H. Materials and Methods . ill. Results and Discussion. A. Cyanylation of SH groups l B. Optimization of Cleavage F C. Effects of Amino Acid Stru B-elimination Reactions ._ D. Application of O W. Conclusions... . . . .. . V. Future Work ........ VI. References ......... ptimized N Chiliter 4. I u ..... .... ..... ----- A. Cyanylation and Cleavage of SH Groups by NTCB ............................... 70 B. Effects of pH and Temperature on Cleavage Kinetics... . ........................ 75 C. Identification of Free SH and Total Cysteines in Proteins ........................ 76 D. Choice of Reducing Agent ............................................................ 89 E. Stability of Disulfide Bonds during NTCB Reaction .............................. 90 F. Analysis by MALDI-MS .............................................................. 92 G. Reactions of Proteins immobilized on a Zetabind Membrane .................... 94 V. Conclusions ....................................................................................... 99 VI. References ...................................................................................... 100 Chapter 3. Further Study on the Localization of Protein Cysteine Residues by Chemical Cleavage and MALDI—MS: Optimization and Application ...... . .................................... , .......... 105 I. Introduction ..................................................................................... 105 II. Materials and Methods ................................. . ...................................... 107 III. Results and Discussion ........................................................................ 110 A. Cyanylation of SH groups by CDAP ............................................... 110 B. Optimization of Cleavage Reactions ................................................ 113 C. Effects of Amino Acid Structures on Cleavage and tit-elimination Reactions .............................................................. 125 D. Application of Optimized Methodology ............................................ 127 IV. Conclusions ..................................................................................... 134 V. Future Work .................................................................................... 135 VI. References ................ . ..................................................................... 139 Chapter 4. A Novel Strategy for Assignment of Disulfide Bond Pairings in Proteins ................................................................... 141 I. Introduction .................................................................................... 141 11, Our Novel Strategy for Disuiftde Assignment ............................................. 144 III. Experimental Section ......................................................................... 148 vi IV. Results and Discussion A. Partial Reduction of Prot B. Cyanylation of Nascent S C. HPLC Separation of Part D. Cleavage of Peptide C hai E. Interpretation of MALDI F. Application to Proteins C Adjacent Cysteine Residu V. Pioomole Scale Reaction VT. Characteristics of the MCLDOdOlOl Vii. Conclusions, .. .. ill]. References. . . . . . . Chapter 5. Disrlfide Mapping of Folc Epidermal Growth Factor Cleavage, and Mass Spect I Introduction ..... . ii. Experimental Section. ,. ill. Results and Discussion. . . A Refolding of hEGF under C B. Trapping of Folding Intenn C. Separation of Folding Inten D. Disulfide Mapping of Well I - Aspects for Further I - Conclusions ................ Vi. References IV. Results and Discussion ........................................................................ 151 A. Partial Reduction of Proteins ....................................................... 15.1 B. Cyanylation of Nascent Sulthydryls ................................................ 155 C. HPLC Separation of Partially Reduced and Cyanylated Protein Isomers. . . ..156 D. Cleavage of Peptide Chains .......................................................... 158 E. Interpretation of MALDI Data ...................................................... 159 F. Application to Proteins Containing Closely Spaced and Adjacent Cysteine Residues ......................................................... 176 V. Picomole Scale Reaction ...................................................................... 194 VI. Characteristics of the Methodology ................................ . ........................ 196 VII. Conclusions ................................................................................... 198 VIII. References .................................................................................... 199 Chapter 5. Disulfide Mapping of Folding Intermediates of Human Epidermal Growth Factor by CDAP Trapping, Chemical Cleavage, and Mass Spectrometry ................................................. 201 I. Introduction ..................................................................................... 201 11. Experimental Section ......................................................................... 21 0 III. Results and Discussion ........................................................................ 214 A. Refolding of hEGF under Controlled Conditions ................................ 214 B. Trapping of Folding Intermediates ................................................ 216 C, Separation of Folding Intermediates by HPLC ................................... 217 D. Disulfide Mapping of Well Populated Intermediates ............................. 222 IV, Aspects for Further Improvement ........................................................... 240 V. Conclusions .................................................................................... 242 VI. References ...................................................................................... 244 vii TableZl Calculated "”2 “11”“ °f peptides after the NTCB Table2.2 Mass assignment of pepti Tablcll Calculated m/z values of l peptides afier the cleavat latest Relative yields (%) of B-t conditions... . .. _ Tablet] Calculated and observed n resulting from the cleavag at designated cysteine pai; Table42 Calculated and observed m resulting from the cleavag at deSlgnated cysteine pair tablet] Calculated and observed ml resulting from the cleavage inhibitor chains at designatt able 4-4 Calculated and observed ml; resulting from the cleavage at sites of designated Cystei Table - Calculated and observed m/z LIST OF TABLES Table 2.1 Calculated m/z values of fragments for sulfhydryl—containing peptides after the NTCB reaction .................................................. 71 Table 2.2 Mass assignment of peptide fragments after different treatments ............... 78 Table 3.1 Calculated m/z values of fragments for sulfhydryl-containing peptides after the cleavage reaction .............................................. 112 Table 3.2 Relative yields (%) of B-elimination under different cleavage conditions ........................................ . ................................... 1 15 Table 4.1 Calculated and observed m/z values for possible fragments resulting from the cleavage reaction of ribonuclease A chains at designated cysteine pairs ....................................................... 161 Table 4.2 Calculated and observed m/z values for possible fragments resulting from the cleavage reaction of or—lactalbumin chains at designated cysteine pairs ................. . ..................................... 168 Table 4.3 Calculated and observed m/z values for possible fragments resulting from the cleavage reaction of soybean trypsin inhibitor chains at designated cysteine pairs ................................... 173 Table 4.4 Calculated and observed m/z values for possible fragments resulting from the cleavage of recombinant hEGF chains at sites of designated cysteine pairs ............................................. 178 Table 4.5 Calculated and observed m/z values for possible fragments viii resulting from the cleav sites of designated cyst: Table 4.6 Calculated and observed Table 5.1 Table 5.2 fragments resulting from chains at designated cvst Expecred mfz values for 1 the cleavage of 15 singlv Fragments of cleavage pr. cyanylated l-disulfideinte resulting from the cleavage reaction of LRslGF-I chains at sites of designated cysteine pairs .................................................. 187 Table 4.6 Calculated and observed m/z values [M-H]' for possible fragments resulting from the cleavage reaction of insulin chains at designated cysteine pairs ................................................ 191 Table 5.1 Expected m/z values for possible fragments resulting from the cleavage of 15 singly reduced/cyanylated hEGF isomers ................. 224 Table 5.2 Fragments of cleavage products corresponding to 15 cyanylated l-disulfideintermediates of hEGF .................................. 237 ix Figure 1.1 Intra- and Inter-chain d figure 1.2 The modification of cys S-carboxymethvlcvsteil Figure 1.3 Chemical reaction betwr Figure 1.4 The reaction of Ellman‘: acid), with cysteinyl res Figure 15 Modification of cysteine fobenzoate (37331 Figure 1.6 Derivatization of the cyst Figure] 7 S-[Z-((iodoacety1)amin0) (IAEDANS) Tigue 1.8 Illustration of sulfhydryV< lg11te1.9 Classical strategy for disu F191161.10 Current protocol for disul 1' litre 1.11. Scheme of matrix-assistei Figure 1.1 Figure 1.2 Figure 1.3 Figure 1.4 Figure 1.5 Figure 1.6 Figure 1.7 Figure 1.8 Figure 1.9 Figure 1.10 Figure 1.11. Figure 2.1 Figure 2.2 Figure 2.3 Figure 2.4 LIST OF FIGURES Intra- and Inter-chain disulfide bonds in proteins .................................. 4 The modification of cysteine with iodoacetic acid to form S-carboxyrnethylcysteine ............................................................. 6 Chemical reaction between p-mercun'benzoate and sulfliydryl group ........... 8 The reaction of Ellman’s reagent, 5,5’-dithio-bis-(2-nitrobenzoic acid), with cysteinyl residues in proteins ........................................... 9 Modification of cysteine sulfhydryl by 2-nitro-5-thiosul- fobenzoate (NT SB) ................................................................. 15 Derivatization of the cysteinyl group by SBD-F or ABD-F .................... 18 5-[2-((iodoacetyl)amino)-ethyl]naphthalene—1-sulfonic acid (IAEDAN S) ........................................................................... 19 Illustration of sulthydryl/disulfide exchange in proteins ......................... 22 Classical strategy for disulfide bond assignment of a protein ................... 23 Current protocol for disulfide bond assignment by mass spectrometry ....... 28 Scheme of matrix-assisted laser desorption/ionization time- of-flight mass spectrometer (MALDI-TOF MS) ................................ 39 Reaction between cysteine residue and 2-nitro-5-thiocyanobenzoic acid (N TCB), (A) Cyanylation and (B) Cleavage reaction ..................... 59 A mechanism for base catalyzed cleavage reaction .............................. 61 Chemical modification to identify free sulthydryls and total cysteines ........ 64 Conceptional scheme for microscale analysis of proteins immobilized on an inert membrane ............................................... 69 figure 25 MALDI mass spectra c after reaction with a 5- peptide and (B) cleava to doubly charged spec calculated and observe Figurezt’; MALDI Wm Ofcleav TCVEWLRRYLKN, (E (‘31 WIRQEAVDCLKr with 5-fold molar excess identities of the peaks , Figure 2.7 strum feature; of me Farm MALDl mass spectrum ( reaction With 500 Pmol ‘ “131er was applied to :1 Peaks corresponding to 1 for the identities of mark and (B) TCEP/NTCB trea of other marked peaks anc observed m/z values ..... Flam I i .10 Imass spectra of or (B) TCEP/NTCB treatment Cyanylated/uncleaved peptit Products. See Tab1e2.2 for Figure 2.5 MALDI mass spectra of peptide RYVVLPRPVCFBKGMNYTVR after reaction with a 5-fold molar excess of NTCB, (A) cyanylated peptide and (B) cleavage products. ++ indicates peaks corresponding to doubly charged species. See Table 2.1 for the correlation of calculated and observed m/z values ............................................... 72 Figure 2.6 MALDI spectra of cleavage products of peptides (A) TCVEWLRRYLKN, (B) DRVYIHPCHLLYYS, and (C) MHRQEAVDCLKKFNARRKLKGA, after reaction with 5-fold molar excess of NTCB. See Table 2-1 for the identities of the peaks ............................................................. 73 Figure 2.7 Structural features of model proteins ........................................... 77 Figure 2.8 MALDI mass spectrum of 100 pmol of spinach ferredoxin after reaction with 500 pmol of NTCB. Approximately 1.5 pmol of analyte was applied to the probe tip. Asterisks (*) indicates peaks corresponding to B-elimination products. See Table 2.2 for the identities of marked peaks and the correlation of calculated and observed m/z values ............................................ 80 Figure 2.9 MALDI mass spectra of papain after (A) NTCB/TCEP treatment and (B) TCEP/NTCB treatment. See Table 2.2 for the identities of other marked peaks and the correlation of calculated and observed m/z values ............................... . ................................ 82 Figure 2.10 MALDI mass spectra of ovalbumin after (A) NTCB/TCEP and (B) TCEP/NTCB treatment. * indicates peaks corresponding to cyanylated/uncleaved peptides and/or their B-elimination products. See Table 2.2 for the identities of marked peaks and the correlation of calcu Pigurele MALDI mass spectra C (B) NTCB/Dialysis/l C l Table 2.2 for the correl figure 2.12 MALDI mass spectrum reaction with NTCB 1 calculated and observed Figure 2.13 Stability study of somau molar excesses of MC] excess of NTCB reagen‘ somatostatin decompose ngre2l4 MALDI mass spectrum ‘ NTCBflCEP treatment Figural-15 MALDI mass spectrum c TCEP/NTCB treatment l Em“ (A) Cyanylation of sulfhy bond cleavage catalde Figure 3.2 HPLC chromatograms of in(A) pH 9.0, 37°C, 18h, marked with I, II, B, C/U, B-elimination product cya Figure 2.11 Figure 2.12 Figure 2.13 Figure 2.14 Figure 2.15 Figure 3.1 Figure 3.2 Figure 3.3 the correlation of calculated and observed m/z values ........................... 84 MALDI mass spectra of B-lactoglobulin after (A) NTCB/TCEP, (B) NTCB/Dialysis/TCEP, and (C) TCEP/NTCB treatment. See Table 2.2 for the correlation of calculated and observed m/z values .......... 87 MALDI mass spectrum of or- and B-hemoglobin mixtures alter reaction with NTCB. See table 2.2 for the correlation of calculated and observed m/z values ................................................ 88 Stability study of somatostatin in the presence of different molar excesses of NTCB reagent. After exposed to 50-fold excess of NTCB reagent in pH 9.0 buffer at 37°C for 16 hours, somatostatin decomposes completely .............................................. 91 MALDI mass spectrum of 10 pmol of papain after in situ NTCB/TCEP treatment on a Zetabind membrane ............................... 96 MALDI mass spectrum of 15 pmol of ovalbumin after in situ TCEP/NTCB treatment on a Zetabind membrane ...................... . ........ 97 (A) Cyanylation of sulfhydryl group by CDAP and (B) peptide bond cleavage catalyzed by ammonia ............................................ 116 HPLC chromatograms of cleavage products of SLRRSSCFGGR in (A) pH 9.0, 37°C, 18h, and (B) 1 NINE-14011, rt, 1h. The peaks marked with I, II, B, C/U, and D are fragment 1, fragment 11, B-elimination product, cyanylated/uncleaved species, and peptide dimer, respectively ................................................................. \\9 HPLC chromatograms of cleavage products of DRVYIHPCHLLYYS in (A) pH 9.0, 3'7"C, 1811, and (B) 1 M NH40H, rt, 1h, respectively ............... . ................................ \ZQ xii Figure 3.4 HPLC chromatograms EKPLQNFFLCFR in ( l M NlLOll rt, In re Figure 3.5 HPLC chromatograms MHRQEM’DCLKKFT 18h, and (B) l M MM 53111816 MALDI spectra of the c alter cleavage in (A) pH SOlUIiOrL rt llt respecti represent fragment 1. fra uncleaved peptide, and i 5311163 7 MALDI spectra of the c‘ DRVYIHPCHLLYYS, 2 18h and (B) l M NlL0l Symbol ++ represents a < Figure3.8 MALDI spectra of the cy after cleavage in (A) pH solution, rt, lh, respectiv Flgure3 9 ' MALDI Spectra of the cy: Figure 3.4 HPLC chromatograms of cleavage products of EKPLQNFTLCFR in (A) pH 9.0, 37°C, 18h, and (B) lMNH40H, rt, 1h, respectively....................................... ..........121 Figure 3.5 HPLC chromatograms of cleavage products of peptide MHRQEAVDCLKKFNARRKLKGA in (A) pH 9.0, 37°C, 18h, and (B) 1 M NH40H, rt, 1h, respectively ................................ 122 Figure 3.6 MALDI spectra of the cyanylated peptide SLRRSSCFGGR, after cleavage in (A) pH 9.0, 37°C, 18h and (B) 1 M NH40H solution, rt, 1h, respectively. Symbols 1, 11, D, C/U, and (3 represent fragment 1, fragment 11, peptide dimer, cyanylated/ uncleaved peptide, and its [ii-elimination product, respectively... ... .......l23 Figure 3.7 MALDI spectra of the cyanylated peptide DRVYIHPCHLLYYS, after cleavage in (A) pH 9.0, 37"C, 18h and (B) l M NH40H solution, rt, 111, respectively. Symbol ++ represents a doubly charged species ............................. 123 Figure 3.8 MALDI spectra of the cyanylated peptide EKPLQNFTLCFR, after cleavage in (A) pH 9.0, 37°C, 18h and (B) l MNH40H solution, rt, lh, respectively. [P represents the intact peptide ........... 12.4 Figure 3.9 MALDI spectra of the cyanylated MHRQEAVDCLKKFN ARRKLKGA, after cleavage in (A) p11 9 .0, 37°C, 18h and (B) l M NH40H solution, rt, 1h, respectively ............ 12.4 Figure 3.10 MALDI spectra of spinach ferredoxin (W. 10483) after cyanylation by the CDAP and subsequent cleavage in (A) pH 9.0 bufi‘er for 18 hours at 37°C and (B) 1 MNH4011 solution for one hour at room temperature, respective1y ................ 12% Figure 3.11 MALDI mass SW” OfO‘ cyanylation by CDAP and buffer at 37C for 18 bout c and d represent cyanylal B-elimination products. rt ligire312 MALDI mass spectra of r: after cyanylation by C DA] pH 9.0 at 37C for 18 hos respectively. The peaks " impurity in the sample T related to two unidentifie Fl11116.13 HPLC chromatograms of phosphokinase afier react for 18 hours, and (B) 1 hr The main HPLC peaks cc at cysteine residues are rr Figure314 Chemical modification of and thiocyanoPYIidine, re Flgure 4. 1 Descriptive overview of c assignment of disulfide b1 11 ' gure42 Chemical overview of on ii i gu 64.3 Conceptional scheme to i Flgure44 - C separation of dena partially reduced/ cyanyl‘ protein were separated b Figure 3.11 MALDI mass spectra of ovalbumin (MW. 42699) after cyanylation by CDAP and subsequent cleavage in (A) pH 9.0 buffer at 37°C for 18 hours and (B) 1M NH40H, rt, 1h , respectively. c and d represent cyanylated/uncleaved peptides and its 1.3-elimination products, respectively ............................................ 129 Figure 3.12 MALDI mass spectra of rabbit muscle creatine phosphokinase after cyanylation by CDAP and subsequent cleavage in (A) pH 9.0 at 37°C for 18 hours and (B) 1 M NH40H, rt, 1h, respectively. The peaks "*" are due to "carry-over" from an impurity in the sample. The question marks "7" in (A) are related to two unidentified species ............................................... 131 Figure 3.13 HPLC chromatograms of cleavage products of creatine phosphokinase after reaction in (A) pH 9.0 buffer at 37°C for 18 hours, and (B) 1 M NH40H solution, rt, 1h respectively. The main HPLC peaks corresponding to the specific cleavage at cysteine residues are marked ................................................. 133 Figure 3.14 Chemical modification of SH groups by 2,2'-dipyridyldisulfide and thiocyanopyridine, respectively ............................................ 137 Figure 4.1 Descriptive overview of our proposed methodology for assignment of disulfide bond pairings in proteins ............................ 145 Figure 4.2 Chemical overview of our methodology ...................................... 146 Figure 4.3 Conceptional scheme to illustrate "Partial Reduction" ..................... 153 Figure 4.4 HPLC separation of denatured ribonuclease A and its partially reduced/ cyanylated isomers. Ten-nmol of the protein were separated by reversed-phase HPLC on a xiv Figure 4.5 Figure4.6 FlSure 4.7 Fiturets Vydac C18 column at a fit gradient 204076 B in 90 ll watetand3=0 l%TFA singly reduced/Cyanylated MALDI-TOP analysis Disulfide structure of ribor The MALDI mass spectra the cleavage of the four 5 A isomers, corresponding respectively The symbo species and protonated B Thepeaks 1-124 in ( A ) Behrnination only occuri do not provide any speci calculated and observed The MALDI mass spectn the cleavage of a lelUlt H and " represent the d Products, respectively .. HPLC separation of dena reduoed/ cyanylated isom Vydac C4 column at a fit gradient 40.50% Bin 90 ”dB=0.1% TFA in C1 cyanylated species, as de' Figure 4.5 Figure 4.6 Figure 4.7 Figure 4.8 Vydac C18 column at a flow rate of 1.5 ml/min with a linear gradient 20-40% B in 90 minutes, where A = 0.1% TFA in water and B = 0.1% TFA in CH3CN. Peaks 14 represent singly reduced/cyanylated species, as determined by MALDI-TOF analysis .............................................................. 157 Disulfide structure of ribonuclease A ............................................ 160 The MALDI mass spectra of peptide mixtures resulting from the cleavage of the four singly reduced/cyanylated ribonuclease . A isomers, corresponding to the HPLC peaks 1-4 in Figure 4.4, respectively. The symbols ++ and * represent the doubly-charged species and protonated B-elimination products, respectively. The peaks 1-124 in ( A) and (B) represent intact proteins in which B-elimination only occurred at one cysteinyl residue. These products do not provide any specific information. See also Table 4.1 for calculated and observed m/z values ............................................. 162 The MALDI mass spectrum of peptide mixtures resulting from the cleavage of a mixture of HPLC peaks 1 and 2. The symbols ++ and * represent the doubly-charged species and B-elimination products, respectively ............................................................ 165 HPLC separation of denatured oe-lactalbumin and its partially reduced/ cyanylated isomers. Separation was carried out on a Vydac C4 column at a flow rate of 1.5 ml/min with a linear gradient 40-50% B in 90 minutes, where A = 0.1% TFA in water and B = 0.1% TF A in CH3CN. Peaks 1-4 represent singly reduced/ cyanylated species, as determined by MALDI-TOP analysis ............... 167 Figure” The MALDI mass spectra cleavage of the four 51an isomers, corresponding tr respectively The symbol species and B-elimination 4.2 for the calculated and Figure 4.10 HPLC separation of dena its partially reduced" cya carried out on a Vydac ( “1111 a linear gradient 15 TFA in water and B = 8‘ 0.1% TFA. Peaks 1 and species, as determined b Figure 4.11 The MALDI mass spectr the cleavage of the two prsin inhibitor (STl) is peaks 1 and 2 in Figure peaks are due to impunt Figure 4.12 The MALDI mass spectr the cleavage of unsepari soybean trypsin inhibitor the mixture was dialyset excess reagents. See tal m/z values ...... Figure4 ' .13 may structure and dis Figure 4.9 The MALDI mass spectra of peptide mixtures resulting from the cleavage of the four singly reduced/cyanylated or-lactalbumin isomers, corresponding to the HPLC peaks 1-4 in Figure 4.8., respectively. The symbols ++ and "‘ represent the doubly-charged species and B-elimination products, respectively. See also Table 4.2 for the calculated and observed m/z values ................................. 169 Figure 4.10 HPLC separation of denatured soybean trypsin inhibitor and 'its partially reduced/ cyanylated isomers. Separation was carried out on a Vydac C4 column at a flow rate of 1.2 ml/min with a linear gradient 15-50% B in 30 minutes, where A = 0.1% TFA in water and B = 80% 1-propanol/20% H20 containing 0.1% TFA. Peaks 1 and 2 represent singly reduced/cyanylated species, as determined by MALDI-TOP analysis ............................. 172 Figure 4.11 The MALDI mass spectra of peptide mixtures resulting from the cleavage of the two singly reduced/cyanylated soybean trypsin inhibitor (STI) isomers, corresponding to the HPLC peaks 1 and 2 in Figure 4.10, respectively. The unidentified peaks are due to impurities in the sample ...................................... 174 Figure 4.12 The MALDI mass spectrum of peptide mixtures resulting fi'om the cleavage of unseparated, partially reduced/cyanylated soybean trypsin inhibitor (STI) isomers. Before the cleavage, the mixture was dialysed against 1% HAc solution to remove excess reagents. See table 4-3 for possible fragments and their m/z values .......................................................................... 175 Figure 4.13 Primary structure and disulfide bond linkage of recombinant human epidermal grOMh Figure 4.14 HPLC separation of denat cyanylated isomers Sepa column at a flow rate of 1 B in 15 min and 35-55% in water and B = 0 1% Tl reduced cyanylated specie iigue415 The MALDI mass spectra cleavage of the three sing HPLC peaks 1-3 in Fig and u represent the dOU products, and cyanylated also Table 4.4 for the cal Figure416 The MALDI mass spectra the cleavage of the three isomers and a completeli pending to the HPLC pe The symbols ++ , ‘, and B-elimination products a respectively. See also T m/z values ..... ”BM 1 - 7 C separation of dena reduced/ cyanylated ison Vydac C18 column at a 8 gradient 30-50% B in 45 human epidermal grth factor (hEGF) (MW. 6347.2) ...................... 177 Figure 4.14 HPLC separation of denatured hEGF and its partially reduced/ cyanylated isomers. Separation was carried out on a Vydac C18 column at a flow rate of 1.0 ml/min with a linear gradient 15-3 5% B in 15 min, and 35-55% B from 15—50 min, where A = 0.1% TFA in water and B = 0.1% TF A in CH3CN. Peaks 1-3 represent singly reduced/cyanylated species, as determined by MALDI-TOP analysis ....... 179 Figure 4.15 The MALDI mass spectra of peptide mixtures resulting from the cleavage of the three singly reduced/cyanylated hEGF isomers, HPLC peaks 1-3 in Fig. 4.14, respectively. The symbols ++ , *, and ** represent the doubly-charged species, B-elimination products, and cyanylated/uncleaved products, respectively. See also Table 4.4 for the calculated and observed m/z values ................... 181 Figure 4.16 The MALDI mass spectra of peptide mixtures resulting from the cleavage of the three doubly reduced/cyanylated hEGF isomers and a completely reduced/cyanylated hEGF, corres- ponding to the HPLC peaks 4—7 in Figure 4.14, respectively. The symbols -H- , *, and ** represent doubly-charged species, B-elimination products, and cyanylated/uncleaved products, respectively. See also Table 4.4 for the calculated and observed m/z values .......................................................................... 183 Figure 4.17 HPLC separation of denatured LR3IGF-I and its partially reduced/ cyanylated isomers. Separation was carried out on a Vydac C18 column at a flow rate of 1.0 ml/min with a linear gradient 30-50% B in 45 minutes, where A = 0.1% TFA in xvii water and B = 0 1% TFA singl.v reduwdicyanylaled 10F anaii'siS iigUre 4.18 The MALDI mass 513%"2 the cleavage of the two 5 isomers. conesponding l 4.17, respeciiVel." The 5 text. See also Table 4 5 Figure 4.19 HPLC separation ofdena cyanylated isomers Sep column at a flow rate of B in 40 minutes. where .1 TFA in CH3CN Peaks species, as determined b' A and B are insulin A-cl FigureiZO The negative ion MALDI resulting from the cleave cyanylated insulin isome 1-3 in Figure 4.19, respe the B-elimination produc reSPCCthely. See also T observ Figure421 Separ ~ ed mil values. H ation of nbonucleasi Species by microbore C 1: conditions are the same a water and B = 0.1% TF A in CH3CN. Peaks 1 and 2 represent singly reduced/cyanylated species, as determined by MALDI- TOF analysis ........................................................................ 185 Figure 4.18 The MALDI mass spectra of peptide mixtures resulting from the cleavage of the two singly reduced/cyanylated LR3IGF-I isomers, corresponding to the HPLC peaks 1 and 2 in Figure 4.17, respectively. The symbol * represent the B-elimination the text. See also Table 4.5 for the calculated and observed m/z values ....... 188 Figure 4.19 HPLC separation of denatured insulin and its partially reduced/ cyanylated isomers. Separation was carried out on a Vydac C18 column at a flow rate of 1.0 m1/min with a linear gradient 20-50% B in 40 minutes, where A = 0.1% TFA in water and B = 0.1% TFA in CH3CN. Peaks 1-3 represent singly reduced/cyanylated species, as determined by MALDI-TOP analysis. The peaks marked A and B are insulin A-chain and B-chain, respectively ....................... 191 Figure 4.20 The negative ion MALDI mass spectra of peptide mixtures resulting from the cleavage of the three singly reduced/ cyanylated insulin isomers, corresponding to the HPLC peaks 1-3 in Figure 4.19, respectively. The symbols * and ** represent the B-elimination products and cyanylated/uncleaved peptides, respectively. See also Table 4.6 for the calculated and observed m/z values .......................................................... 193 Figure 4.21 separation of ribonuclease A and its partially reduced/cyanylated species by microbore C18 HPLC column (1 x100mm). HPLC conditions are the same as those in Figure 4.4, except the flow xviii rate is 005 ml/min Figure 51 Conventional approach fc intermediates Figrre 5.2 New approach for disulfrr Figure53 Primary structure and dis factor (hEGF)Figure 5 4 folding intermediates of l Fl8111e55 Comparison of HPLC prr and CDAP-trapped foldi iigure5.6 HPLC separation of 111-! 11gure57 The MALDI mass spectr the cleavage of the three of non-native hEGF 111-; peaks 1-3 in Figure 56 1 represent the doubly-cha products, respectively “8116.8 Comparison of HPLC prr singly redUCed/cyanylate represented by peaks 1 a Film-9 HPLC separation of “H Fl 81610 The MALDI mass spectr; of the 3 singly reduced/c A~C corresponding to 11 respectively. . . . rims ........ . . . l 1 C separation of TM Figure 5.1 Figure 5.2 Figure 5.3 Figure 5.5 Figure 5.6 Figure 5.7 Figure 5.8 Figure 5.9 rate is 0.05 ml/min ............................................................. Conventional approach for disulfide mapping of protein folding intermediates .................................................................. New approach for disulfide mapping of protein folding intermediates... . . .. Primary structure and disulfide linkage of human epidermal growth factor (hEGF)Figure 5.4. HPLC separation of CDAP-trapped folding intermediates of hEGF at different time courses .................. Comparison of HPLC profiles of iodoacetate-trapped, acid-trapped, and CDAP-trapped folding intermediates of hEGF ...................... HPLC separation of III-A and its partially reduced/cyanylated species The MALDI mass spectra of peptide mixtures resulting from the cleavage of the three singly reduced/cyanylated species of non-native hEGF III-A, A-C corresponding to the HPLC peaks 1-3 in Figure 5-6, respectively. The symbols 4+ and * represent the doubly-charged species and B-elimination ..... 195 ..... 203 208 ..... 218 ..... 220 ..... 225 products, respectively ............................................................. 226 Comparison of HPLC profiles of cleavage products of two singly reduced/cyanylated HI-A isomers. (A) and (B) are represented by peaks 1 and 2 in Figure 5.6, respectively ................ HPLC separation of III-B and its partially reduced/cyanylated species. . . .. Figure 5.10 The MALDI mass spectra of peptide mixtures from the cleavage ..... 228 .230 of the 3 singly reduced/cyanylated species of non-native hEGF III-B, A-C corresponding to the HPLC peaks 1-3 in Figure 5.9, respectively ................................................................... ..... 231 Figure 511 HPLC separation of II-A and its partially reduced/cyanylated species ...... 233 xix — Figure 5.12 The MALDI mass spectra of (A) the cyanylated hE reduced/cyanylated speci peaks 1 and 2 in Figure ligureSl3 Disulfide mapping ofl Figure 5.14 Disulfide structures of ofhEGF............ . Figure 5.12 The MALDI mass spectra of peptide mixtures from the cleavage of (A) the cyanylated hEGF H-A, (B) and (C) the partially reduced/cyanylated species of II-A, corresponding to the HPLC peaks 1 and 2 in Figure 5.11, respectively ..................................... 235 Figure 5.13 Disulfide mapping of l-disulfide folding intermediates of hEGF ........... 238 Figure 5.14 Disulfide structures of seven well populated folding intermediates of hEGF ........................................................................... 241 1 Introduction Since its discovery in 1912, broad and diverse disciplines. Howcv of biomolecules stems in large techniques in the early 19803 (1, 2). sciences have led to a dramatic ' inforination of biologitu active undiscipline-biological mass spec Structural isms of biomedically impor ins accelerated dramatically, due to ionization techniques: matrix—mister electrospiay ionization (EST) (4). Con: routine the analysis of high mass com]: mural products, and drug metabolites MS has been revolutionary in spurrin and biotechnology, and most recently ir Protein analytical techniques a Siructrrre as well as detecting and defir thatthe complete structure of a proteir CHAPTER 1 INTRODUCTION 1. Introduction Since its discovery in 1912, mass spectrometry (MS) has provided key insight for broad and diverse disciplines. However, the involvement of mass spectrometry in studies of biomolecules stems in large part from the discovery of desorption ionization techniques in the early 1980s (1, 2). At the same time, rapid advances in biological sciences have led to a dramatic increase in the demand for chemical and structural information of biologically active materials. The combination led to the creation of a new discipline-biological mass spectrometry—which addresses the challenging unsolved structural issues of biomedically important molecules. In the past decade, this expansion has accelerated dramatically, due to the discovery of two new mass spectrometry ionization techniques: matrix-assisted laser desorption/ionization (MALDI) (3) and electrospray ionization (ESI) (4). Commercial availability of these instruments has made routine the analysis of high mass compounds including proteins, peptides, carbohydrates, natural products, and drug metabolites with picomole to femtomole sensitivity. Today, MS has been revolutionary in spurring research in protein biochemistry, glycobiology, and biotechnology, and most recently in DNA sequencing (5). Protein analytical techniques are essential for determining a protein’s primary structure as well as detecting and defining posttranslational modifications of proteins so that the complete structure of a protein can be obtained (6, 7). Nucleotide sequences of l chnnically modified in viva after Defining postnanslational modificati are destined for use as pharmaceuti same chemical structure as their Umceptable structtm could pharmacological or immunological Defining the locations and isone of the mom important contribu tomake in the field of protein bi evident when it is uwd in conjunctii protein sequence and amino acid an which identification is hm entirely 20 commonly occurring amino acids molecular mass, an intrinsic physical of choice for characterization of postr welldeveloped, accurate, and semi available. Disulfide bond formation at Postttanslational modifications frequr Mass spectrometry, combined with extensively applied to the recognitii genes embody the information required to deduce the primary structure (amino acid sequence) of proteins but they do nOt reveal whether the side chains of amino acids are chemically modified in vivo after translation of genes or the extent of any modifications. Defining posttranslational modifications is essential in the case of expressed proteins that are destined for use as pharmaceuticals. This is required to ensure that they have the same chemical structure as their natural counterparts or an acceptable alternate structure. Unacceptable structures could be detrimental to recipients due to unwanted pharmacological or immunological activities. Defining the locations and structures of posttranslational modifications in proteins is one of the most important contributions that mass spectrometry has made and continues to make in the field of protein biochemistry (8). The unique potential of MS is most evident when it is used in conjunction with conventional methodologies like automated protein sequence and amino acid analysis. In contrast to the latter two approaches in which identification is based entirely on chromatographic retention relative to one of the 20 commonly occurring amino acids or a derivative thereof, MS relies on measurement of molecular mass, an intrinsic physical pr0perty. Mass spectrometry is clearly the method of choice for characterization of posttranslationally modified proteins, since in most cases well-developed, accurate, and sensitive chemical or biochemical approaches are not available. Disulfide bond formation after gene expression is one of the most important POSttranslational modifications frequently encountered in protein characterization (9, 10). Mass spectrometry, combined with other chemical/biochemical techniques, has been extensively applied to the recognition of both free sulthydryls and disulfide bonds in proteins. In this disserta' tion. novel disuliide bonds, and disulfide struc graterdetail inthechapter 2. 4. an introduce basic aspects of MALDl statisinproteins. ll. Characterization of Cysteine S Among the 20 amino acids Iheredwed form of cysteine contai cystine, is formed by linking two sul wntnhutes to protein biological fun active site for enzyme catalysis such metals (ll). Sulthydryl groups c physiological and biochemical procc: division, oxidative phosphorylation at hand comprises the major covalent 1 Chain in nature (Figure 1.1). Intra-ch inribonuclease A, serve to confer Chain Additionally, by limiting or the correct orientation of the amino antibodies and other biologically functional in maintatmn' ' g the quate only linkage between subunits or pro proteins. In this dissertation, novel methodologies for recognizing free sulfliydryl groups, disulfide bonds, and disulfide structures of folding intermediates will be discussed in greater detail in the chapter 2, 4, and 5, respectively. The objective of this chapter is to introduce basic aspects of MALDI and current methodologies for characterizing cysteine status in proteins. H. Characterization of Cysteine Status in proteins Among the 20 amino acids that compose proteins, cysteine has unique properties. The reduced form of cysteine contains a sulihydryl (thiol) group while its oxidized form, cystine, is formed by linking two sulfltydryls together to form a disulfide bond. Cysteine contributes to protein biological functions by using its free sulthydryl (-SH) group as the active site for enzyme catalysis such as in cysteine proteases, and as the chelating site for metals (11). Sulfhydryl groups can also play an important role in a variety of physiological and biochemical processes such as muscular contraction, nerve activity, cell division, oxidative phosphorylation and photosynthesis (11). Disulfide bond, on the other hand, comprises the major covalent cross-linkage in proteins and may be intra- or inter- chain in nature (Figure 1.1). Intra-chain disulfide bOnds, such as the four disulfide bonds in ribonuclease A, serve to confer conformational stability on the folded polypeptide chain. Additionally, by limiting or directing the folding these bonds may contribute to the correct orientation of the amino acid residues that form the active sites of enzymes, antibodies and other biologically active proteins. Inter-chain disulfide bonds are functional in maintaining the quaternary structure of multi-chain proteins, serving as the only linkage between subunits or providing covalent stability to structures otherwise Inter Figure 1.1. Inna-and mannined by non-covalent forces proteins play a unique role in its act themechanism of activity in glutathi that cysteine sulfliydryls and cysti hmctions, the recognition of cyst: characterization (9, ii). For a new protein with cysteine content is very beneficial t the amino acid sequence of the pr CDNA sequence, difi'erentiating be m— U)..— Intra—chain disulfide bonds l s l s J Inter-chain disulfide bonds Figure 1.1. Intra- and Inter-chain disulfide bonds in proteins. maintained by non-covalent forces (e. g., insulin, IgG). Disulfide bonds in particular proteins play a unique role in its activity. For example, disulfide bonds are involved in the mechanism of activity in glutathione reductase and thioredoxin. Due to diverse roles that cysteine sulfhydryls and cystine disulfide bonds play in protein structures and functions, the recognition of cysteine status is a very important aspect of protein characterization (9, 11). For a new protein with unknown sequence, the quantitative determination of cysteine content is very beneficial to further characterization of the protein. Even though the amino acid sequence of the protein can be easily deduced from the corresponding cDNA sequence, differentiating between free cysteine residues and those involved in disulfide bonding (cystines) is also the cysteine residues after translati pimarystructure. Characterization bonds in proteins has become a cruci hearse molecules having inco biological activity than that of th cysteine status in protein chemistry cystine; (2) localization of free sul linkage. A Quantitative Determination of 1 The cysteine sulfhydryl grou acids by its high reactivity and by it they participate, such as alkylatio exchange, charge-transfer complexe (ll). Sulihydryls can dissociate into Sulfltydryl groups is attributable to ions, which exist at reasonable con (pK, 8—105). Thiolate-dependent r at moderate speed in the pH range 6- disulfide bonding (cystines) is also very essential because the state and connectivity of the cysteine residues after translation can not be predicted readily from the protein’s primary structure. Characterization of the location of sulfltydryl groups and disulfide bonds in proteins has become a crucial part of the analysis of recombinant DNA products, because molecules having incorrect disulfide bond linkages may have much lower biological activity than that of the desired product. The problems associated with . cysteine status in protein chemistry include: (1) quantitative evaluation of cysteine and cystine; (2) localization of free sulflrydryl groups; and (3) assignment of disulfide bond linkage. A. Quantitative Determination of SH and S-S groups in proteins The cysteine sulfhydryl group is distinguished from other side chains of amino acids by its high reactivity and by the exceptionally diverse chemical reactions in which they participate, such as alkylation, acylation, oxidation, sulthydryl-disulfide bond exchange, charge-transfer complexes, reactions with organic heavy metal compounds (11). Sulthydryls can dissociate into a proton and a thiolate anion. The high reactivity of sulfhydryl groups is attributable to the high nucelophilicity of the correSponding thiolate ions, which exist at reasonable concentrations at neutral to weakly alkaline pH-values (pK, 8-10.5). Thiolate-dependent reactions thus proceed readily at pH 8 and above, and at moderate speed in the pH range 6-8. RSH —-> RS" + H’“ A variety of methods, disulfidc bonds participate, have disulfidc bond content in proteins. chaptetsUl-B). L Methods Based on Alkyhtion 0 II + I -NH*~CH -C — Figure 1.2. The modifica S-carboxymethylcysteine Alkylation (S-carboxymcthyl isthe most frequently used reaction be accomplished with compounds icdoacctamide, bromoacetate, etc, and specificity to SH (Figure 1.2 electrochemical method (14) or ab bi "(I-labeled iodoacetate permits radioactivity of the product (16. A variety of methods, based on numerous reactions in which sulfliydryls and disulfide bonds participate, have been studied for the determination of sulfhydryl and disulfide bond content in proteins. Reviews on this subject can be found in several book chapters (1 1-13). 1. Methods Based on Alkylation CIIOOH SH 1 CH2 CH2 0 l I II + ICH2COOH -——> S + H1 —NH—-CH —C -— «. 1 cc c —NH —CH -—C — Figure 1.2. The modification of cysteine with iodoacetic acid to form S-carboxymethylcysteine. Alkylation (S-carboxymethyltion) of SH groups to form S-carboxymethylcysteine is the most frequently used reaction in protein characterization (11:13). Alkylation can be accomplished with compounds containing an alkyl halide group such as iodoacetate, iodoacetamide, bromoacetate, etc., among which, iodoacetamide shows highest reactivity and specificity to SH (Figure 1.2). The iodide liberated can be determined by an electrochemical method (14) or absorbance at 226 nm (15). As an alternative, alkylation by “C-labeled iodoacetate permits the number of thiol groups to be determined from the radioactivity of the product (16, 17). To avoid side reactions connected with the fonnafion of iodine by photooxid: performedindiedark. Based on the same principlt developed in order to improve the : alkylation by iodoacetate is still t1 nilfliydryls because of its simplicity 2. Methods Based on Memptidat Organomercurial compounds Sllgmups (Figure 1.3). In con‘ crumble. Among various organor mmpormds have widely been e mectnibenzoate proposed by Boyer sensitivity, selectivity, and precisi: measurement of the increase in absor binding of p-mercuribenzcate (p~M only the number of SH groups in a of loss of enzymatic activity and rate and the extent of the reaction 0 PH 4.6 than at pH 7-8 because of th result, the reactivity of the reagent i reagents, such as l-(4-chloromercu and p-(hydroxymercuri)-benzoic formation of iodine by photooxidation of iodide ions, the carboxymethylation was performed in the dark. Based on the same principle, a number of other alkylating reagents have been developed in order to irnprove the sensitivity and specificity of the reaction (18). But alkylation by iodoacetate is still the most popular derivatization reaction to modify sulthydryls because of its simplicity and specificity. 2. Methods Based on Mercaptidation Organomercurial compounds are high affinitive and specific reagents for protein SH groups (Figure 1.3). In contrast to alkylation, the mercaptidation reaction is reversible. Among various organomercurial reagents, monofunctional organomercurial compounds have widely been exploited. Spectrophotometric titration with p- mercuribenzoate proposed by Boyer stands in first place because of its simplicity, high sensitivity, selectivity, and precision (19). This classical method is based on the measurement of the increase in absorbance in the 25 0-255 nm region that occurs upon the binding of p-mercuribenzoate (p-MB) to SH groups. It provides a means to measure not only the number of SH groups in a protein, but also the relationship between the degree of loss of enzymatic activity and the number of blocked SH groups. Interestingly, the rate and the extent of the reaction of SH groups of many proteins with p—MB is higher at pH 4.6 than at pH 7-8 because of the high affinity of p—MB for hydroxide ions (20). As a result, the reactivity of the reagent increases as the pH of the solution is decreased. Other reagents, such as 1-(4-chloromercuriphenylazo)—napphthol-2 (CMPN; mercury-orange), and p-(hydroxymercuri)—benzoic acid (21), were also proposed to improve the Spectrophotometric detection of SI carboxyhemoglobin (23). SIH Ct CH2 0 I II + —NH—CH—-c-— @ Figure 1.3. Chemical reaction be 3. Methods Based on Sulthydrylll RISSR, + R: R,SSR + Among oxidants of SH gro reactions with sulfliydryls are ab sclflrydryl/disulfide exchange. As of two steps of nucleophilic substi intermediate stage. Spectrophotometric detection of SH groups in lysine monoxygenase (22) and in carboxyhemoglobin (23). SH COO' CH2 0 Hg l H + "—‘—> I + H+ —NH-—CH —'C -— S Figure 1.3. Chemical reaction between p-mercuribenzoate and sulfliydryl group. 3. Methods Based on SulfhydryllDisulfide Exchange RISSR1 + RSH ——-> RISH + RlSSR RISSR + RSH ——> RISH + RSSR Among oxidants of SH groups, disulfides occupy a special position since their reactions with sulfhydryls are absolutely specific. This reaction is referred to as sulfhydryl/disulfide exchange. As can be seen fiom the equations, this reaction consists of two steps of nucleophilic substitution with the formation of a mixed disulfide at the intermediate stage. Among the various disulfi nitrobemic acid), DTNB, has thiolate anion with excess E11 stoichiometric formation of TNB l disolfide (RS-TNB) (Figure 1.4). 1 his bufl’er, pH 8, at a concentratit ccntaininglmMEDTAinafivet groups (EDTA is included not only metal ions can affect the developme =13,600 Worn"), very fast and st: ofSHgmupsinboth native and (let Figure 1.4. The reaction of Eli with cysteinyl residues in prot Among the various disulfides proposed, Ellman’s reagent, 5,5’-dithio—bis-(2- nitrobenzoic acid), DTNB, has received the widest application (24). The reaction of thiolate anion with excess Ellman’s reagent at pH 7-8 is favored toward the. stoichiometric formation of TNB thiolate (5-thio-2-nitrobenzoate, TNB') and a mixed disulfide (RS-TNB) (Figure 1.4). Ellman’s reagent usually is dissolved in phosphate or tris bufi‘er, pH 8, at a concentration of 10 mM. It is added to a solution of protein containing lmM EDTA in a five to ten-fold excess with respect to the number of SH groups (EDTA is included not only to protect SH groups from oxidation but also because metal ions can affect the development of the color). The method is highly sensitive (a412nm ' = 13,600 M'1cm"), very fast and strictly specific, and may be used for the determination of SH groups in both native and denatured proteins (25-27). N02 N02 s. 9 SI. e + 9 COOH 1| —NH-—CH --C - ‘et ' | COOH —NH—CH —C — NO2 —————> + —m——m— _m_m _ Figure 1.4. The reaction of Ellman’s reagent, 5,5’-dithio-bis-(2-nitrobenzoic acid), with cysteinyl residues in proteins. Grassetti and Murray (28, dithiodipyridineweregood substi stronger reactivity to SH groups ' widely md as the binding bed 01 proteins(3l). 4. Methods Based on Eledrochem One of the important pmpert each other though oxidation/reductit Undermild conditions (e.g., in the bonds Finther oxidation results in reducing agent, disulfide bonds can Those specific oxidation/red thebasis of amperometric titration c the first portion of AgNO, binds t small current to increase. As soo appear in solution and are redu Current which is proportional to the 10 Grassetti and Murray (28, 29) demonstrated 2, 2’-dithiodipyridine or 4, 4’- dithiodipyridine were good substituents for Ellman’s reagent. These compounds showed stronger reactivity to SH groups in a wide pH range (pH 1-8) (30) and are therefore widely used as the binding bed of affinity chromatography for sulfhydryl-containing proteins (3 1). 4. Methods Based on Electrochemical Titration One of the important properties of sulfhydryl/disulfide is that they can convert to each other though oxidation/reduction reactions. Cysteine is very sensitive to oxidation. Under mild conditions (e.g., in the air), SH groups undergo oxidation to form disulfide bonds. Further oxidation results in the formation of cysteic acid. Under the presence of reducing agent, disulfide bonds can be reduced to form sulfhydryls. [O] RSSR + 2H+ + 2e' [R] 2RSH Those Specific oxidation/reduction properties of a sulfllydryl/disulfide pair form the basis of amperometric titration of sulfliydryls with silver nitrate (32). In this method, the first portion of AgNO3 binds to the SH groups of the protein and do not cause the small current to increase. As soon as all the SH groups are blocked, free metal ions appear in solution and are reduced on the platinum electrode, resulting in a diffusion current which is proportional to the concentration of the metal ions released to solution. The amperometn'c titration sulfhydryl groups. With prior re enmded to quantitation of disulfid (33) niece-shiny measured the con unified pig kidney Na‘. K'-ATPa: ammonium nitrate buffer (pH 7.6). Recently Sun et a1 (34) imultaneously both sulfliydryls an detectorsystem actually has five ele steel amtiliary electrode, and three I used to distinguish between sulfhy chromatographic analysis, as illustrt mdpartially reduwd bovine insulin. Although the advantage r spectrophotometry lies in the possi? solutions, in enzymological studies, image as the Spectrophotometric sensitivity, the necessity of special a 5. Methods based on Oxidation Shipton et a1 (35) found th selectively and quantitatively oxid 11 The amperometric titration method was first proposed for the measurement of sulfhydryl groups. With prior reduction of disulfide bonds, this method was also extended to quantitation of disulfide bonds in proteins. For example, Gevondyan et al (33) successfully measured the content of free SH groups and disulfide bonds in the purified pig kidney Na“, K+-ATPase by amperometric titration with silver nitrate in ammonium nitrate buffer (pH 7.6). Recently Sun et al (34) developed an electrochemical detector to detect simultaneously both sulfliydryls and disulfides in peptides. The three-electrode EC detector system actually has five electrodes, one Ag/AgCl reference electrode, a stainless steel auxiliary electrode, and three Hg/Au amalgam working electrodes, and thus can be used to distinguish between sulfliydryl- and disulfide—containing peptides in a single chromatographic analysis, as illustrated with proteolytic digests of bovine OLA-crystallin and partially reduced bovine insulin. Although the advantage of electrochemical titration in comparision with spectrophotometry lies in the possibility of carrying out analysis in cloudy or colored solutions, in'enzymological studies, the electrochemical method has not received as wide a usage as the spectrophotometric method. This is evidently because of its lower sensitivity, the necessity of special apparatus. 5. Methods based on Oxidation Shipton et a! (35) found that benzofuroxan (benzo-Z-cxa-l, 3-diazole N-oxide) selectively and quantitatively oxidizes the SH groups of proteins, the reagent being reduced to o—benzoquinone dioxime. absorbance at wavelengths far remo Alternatively, proteins are treated amiable residues of cysteine into cy thenbereliably determined by amin perfonnic acid leads to destruction quantitative conversion of methioni is therefore limited h Methods Based on Enzyme Ass RSH + Papain-S-SCP Singh et a1 (37) developed group determination using an inset llle sulthydryl/disulfide exchange n SCH, results in the stoichiometri motivated papain catalyzes the hy nnplified spectrophotometn'c signal is about 100-fold more sensitive sulfhydryls, e. g., glutathione, cys determined by this approach. Howe f_——— 12 reduced to o-benzoquinone dioxime. The reduction is accompanied by a large increase in absorbance at wavelengths far removed from the region where proteins absorb strongly. Alternatively, proteins are treated with performic acid (3 6). This is to convert the unstable residues of cysteine into cysteic acid, which withstands acid hydrolysis and may then be reliably determined by amino acid analysis. However, treatment of a protein with performic acid leads to destruction of tryptophan, partial destruction of tyrosine, and quantitative conversion of methionine into methionine sulfone. The practical application is therefore limited. 6. Methods Based on Enzyme Assay RSH + Papain-S-SCH3 -——> RS-SCH3 + Papain-SH Singh et al (37) developed a sensitive Spectrophotometric assay for sulfliydryl group determination using an inactive disulfide derivative of papain (papain-S-SCH3). The sulfliydryl/disulfide exchange reaction of a protein cysteinyl residue with papain-S- SCH3 results in the stoichiometric formation of active papain (papain-SH). The reactivated papain catalyzes the hydrolysis of a chromogenic substrate, resulting in an amplified Spectrophotometric signal proportional to the initial amount of SH. The assay is about 100-fold more sensitive than that using Ellman’s reagent. A variety of sulfliydryls, e. g., glutathione, cysteamine, penicillamine, etc. have been successfully determined by this approach. However, the feasibility of the method for the measurement of thiols in large proteins remain 1 proteins, the accessibility of protein 7. Methods Based on Mass Spec Mass spectrometry is determining protein cysteine/cysti limitations of sensitivity, multiple many classical methods (38-40). Feng er al (38) demonstrat PSI-MS for rapidly counting cyste proteins. In this analysis, a peptid iodmceticacidintheabsence ofa ndfliydryl group(s). Parallel analys datafiom which the number of fr: molecular weight of the alkylated d usefor iodoacetic acid, a mass shifi protein. A complementary experimi protein to a disulfide reducing r differentiation of the total numbe Recently Sun et a1 applied this pr littoglobulin B (39). However, the Shift from alkylation by iodoacetic 13 of thiols in large proteins remain to be tested. Because of the steric hindrance of large proteins, the accessibility of protein sulfliydryls to papain might be a problem. 7. Methods Based on Mass Spectrometry Mass spectrometry is becoming an alternative means of quantitatively determining protein cysteine/cystine content in a way that promises to overcome limitations of sensitivity, multiple derivatization, and protein contamination that burden many classical methods (3 8-40). Feng et al (3 8) demonstrated a method that combines reduction/alkylation and ESI-MS for rapidly counting cysteines, free sulfhydryl groups, and disulfide bonds in proteins. In this analysis, a peptide or denatured protein is first allowed to react with iodoacetic acid in the absence of a disulfide reducing agent to selectively label any free sulfhydryl group(s). Parallel analyses using underivatized protein provide mass spectral data from which the number of free sulfhydryls can be determined by the shift in the molecular weight of the alkylated derivative relative to that of the native protein. In the case for iodoacetic acid, a mass shift of 59 u is observed for each derivatized cysteine of a protein. A complementary experiment involving the same reaction, after exposure of the protein to a disulfide reducing reagent, provides mass spectral data that allows for differentiation of the total number of cysteine(s) and/or cystine(s) originally present. Recently Sun et al applied this procedure to ribonuclease A, lysozyme, insulin, and [3- lactoglobulin B (3 9). However, the above method is limited to small proteins as the mass shift from alkylation by iodoacetic acid is small and the limited mass resolution available and potentially prevent small in drawback, Zaluzec er a1 (40) iydroxymercuribenzoate (pHMB) derivatizationthemass shifl of32l masshitt minimizes the error whic topoor resolution of the instrumen protein samples and provide an e cysteine content even in a protein in & Determination of S-S Content Conventional methodologie reductive cleavage of the S-S bonds the methods mentioned above. Th deduced A sensitive and quantitative thiosulfobenzoate (NTSB) has dra‘ actually composed of two sequentir dimlfide bond with sodium sulfite: RSSR' + so; 14 could potentially prevent small mass shift from being detected. To circumvent the drawback, Zaluzec et al (40) used monofunctional organomercurial reagent (p- hydroxymercuribenzoate (pHMB) for selective derivatization of thiol groups. After derivatization, the mass shift of 321 u is obtained for each derivatized cysteine. The large mass shift minimizes the error which might occur during mass spectrometric analysis due to poor resolution of the instrument. Both MALDI and E81 can detect low picomole of protein samples and provide an excellent method for the quantitative assessment of cysteine content even in a protein mixture. 8. Determination of S-S Content Conventional methodologies for S—S counting in proteins are based on the reductive cleavage of the S-S bonds to form cysteinyl residues that can be determined by the methods mentioned above. The content of disulfide bonds can then indirectly be deduced. A sensitive and quantitative method involving the use of the reagent 2-nitro-5- thiosulfobenzoate (NTSB) has drawn great attention (41, 42). The NTSB assay is actually composed of two sequential reactions. The first reaction is the cleavage of a disulfide bond with sodium sulfite: RSSR' + 8032' ——> RSSO3' + R'S' producedinreactiononN'lSBto du'cbcnzoate (NTB), the latter can the reaction of NTSB with of both dissolved oxygen and the inconvenience associated with the wellastheneedtoremovethered SH sso, I c 0 iHi n + —NH—CH-—c—— (01 Figiue 1.5. Modification of cysteir Himse er a! (43) descri wlyacrylamide gel electrophoresi quantitative determination of irn electrophoresis is one of the most j biological systems, since it requii Visualized by selective methods of Withers of cleaved disulfide bonds 15 The second reaction (Figure 1.5) involves nucleophilic attack of the thiolate produced in reaction on NTSB to yield 1 mol each of a thiosulfonate and 2-nitro-5- thiobenzoate (NTB), the latter can be measured by the absorbance at 412 nm. The reaction of NTSB with thiols and disulfides can be carried out in the presence of both dissolved oxygen and the reducing agent which eliminates the inaccuracy and inconvenience associated with the necessity to work under an oxygen-free atmosphere as well as the need to remove the reducing agent. SH S.‘ | CH2 0 s ‘ ~NH—éH—t': + T ('3 + No2 —NH—CH —-C — No2 Figure 1.5. Modification of cysteine sulfhydryl by 2-nitro-5-thiosu1fobenzoate (N TSB). Hirose et al (43) described another novel method that makes use of polyacrylamide gel electrophoresis (PAGE) following two-step alkylation for the quantitative determination of intramolecular disulfide bonds in proteins. Gel electrophoresis is one of the most powerful techniques for protein analysis in complex biological systems, since it requires very small amounts of proteins which can be visualized by selective methods of gel staining. In this method, proteins with different numbers of cleaved disulfide bonds are alkylated with iodoacetic acid or iodoacetamide asthefirst step. The remaining di and the newly generated free sulfli (iodoacetamide, iodoacetic acid. 0 acid-urea PAGE, different intermed hands, depending on difierences alkylation used in combination analysis of disulfide bonds in protei B. localization of Sulfltydryl G localization of cysteine res bonds (see section C) in protei manipulations in protein chemistry been tested and uwd both in the re available SH groups, and there is nt Yd. 1- Classical Approach The general strategy for conventional methods involves seve trollps, usually, by an irreversible r intent or minimize sulfliydryl/dis enzymes or chemical reagents betw lSseparated by chromatography. F 16 as the first step. The remaining disulfide bonds were reduced by excess dithiothreitol, and the newly generated free sulfliydryls were alkylated with the reagent not yet used (iodoacetamide, iodoacetic acid, or vinylpyridine) as the second step. By subsequent acid-urea PAGE, different intermediates formed in two steps can be resolved into distinct bands, depending on differences in their net charge and conformation. Two-step alkylation used in combination with autoradiography was especially useful for the analysis of disulfide bonds in proteins synthesized in complex biological systems. B. Localization of Sulfhydryl Groups Localization of cysteine residues and their possible pairings in forming disulfide bonds (see section C) in proteins involves some of the oldest and best studied manipulations in protein chemistry. A large number of reagents and procedures have been tested and used both in the reduction of disulfide bonds and in the reaction of the available SH groups, and there is no indication that the search for better methods is over yet. 1. Classical Approach The general strategy for locating free sulfhydryl groups in proteins by conventional methods involves several steps. The first step is to modify free sulfhydryl groups, usually, by an irreversible reaction such as alkylation, under conditions that can prevent or minimize sulfhydryl/disulfide exchange. Second, the protein is cleaved by enzymes or chemical reagents between cysteine residues. Third, the mixture of the digest is separated by chromatography. Finally, the derivatized sulfhydryl-containing peptides — are recognized, mapped to sequem andrelated to specific segments of A key to the success of thi reaction and reagent for sulfhydryl medium and label sulfhydryl gror conditions (ideally under weak a exchange which is minimized at 1 strong UV or fluorescent absorp absorption of proteins, or easily an of derivatimd peptides afier HPLC bedistinguishable from other min the identification of amino acids derivatives. While iodoacetate and pyriv hzsbeenfargreater interestinthe u larger molecule which can serve a 2,1,3—benzoxadiazole (47), 7-fluor (aminosulfonyl)-7-fluoro~2, 1 ,3 —barn assay of sulfliydryls by reason of The adducts exhibit fluorescence a shill corresponding to environment under borate buffer, pH 9.5, on oxidation of the sulthydryls. For 1 17 are recognized, mapped to sequence by the Edman technique and/or mass spectrometry, and related to specific segments of the protein. A key to the success of this approach is to choose an appropriate derivatization reaction and reagent for sulfliydryl groups. The reagent should be soluble in the reaction medium and label sulfhydryl groups selectively, rapidly, and irreversibly under mild conditions (ideally under weak acidic conditions to avoid sulthydryl/disulfide bond exchange which is minimized at pH 2-6.5). Furthermore, the reagent should possess strong UV or fluorescent absorption which does not overlap with the maximum absorption of proteins, or easily attach a radioactive element to facilitate the recognition of derivatized peptides after HPLC separation. Finally, the derivative of cysteine should be distinguishable from other amino acids by the Edman degradation technique in which the identification of amino acids exclusively relies on the retention time of PTH- derivatives. While iodoacetate and pyridylethylation are still extensively used (44-46), there has been far greater interest in the use of this chemistry as a mechanism for introducing a larger molecule which can serve as a sulfhydryl probe. Two derivatives of 7-fluoro- 2,1,3-benzoxadiazole (47), 7—fluoro-2,1,3-benzoxadiazole 4-sulfonate (SBD-F) and 4- (aminosulfonyl)-7-fluoro-2,1,3-banzoxadiazole (ABD-F), have proven usefiil for the assay of sulfhydryls by reason of their formation of fluorescent products (Figure 1.6). The adducts exhibit fluorescence at a long wavelength (~515 nm) and show a spectral , shift corresponding to environmental hydrophobicity (48, 49). The reaction is performed under borate buffer, pH 9.5, containing 1 mM EDTA to prevent metal-catalyzed oxidation of the sulflrydryls. For the derivatization of sulfltydryls in large proteins, the — proteinshavetobedenann'edtom 51). l ‘i N col . / N/ l | sown,+ 502NH2 SBD-F ABD-F Figure 1.6. Derivatization Chin and Wold (52) emplc reduce disulfide bonds and ABD-l sulfhydryls and disulfide bonds i apparently do not react with each c general method for the complete Uting the similar procedure, Kirk: bonds and one free sulfltydryl group Another reagent, 5-[2-((i (IAEDANS) (Figure 1.7) has been 1 Recently, Stnrroek et a1 (58) used ‘ testicular angiotensin-eonverting er from enzymatic digests by HPLC, 18 proteins have to be denatured to ensure that every SH can be attacked by the reagents (50, 51). F. N F. N ‘ r N a / \ / \ / \ / ‘N/O ‘N/O sod::rate> ‘N /O 6 NE) SIO3'NH4+ SIOzNHz $O3'Na+ SIOZNHQ SBD-F ABD-F . 3313-311 ABD-SR Figure 1.6. Derivatization of the cysteinyl group by SBD-F or ABD-F. Chin and Wold (52) employed the combinatiOn of tributylphosphine (Bu3P) to reduce disulfide bonds and ABD-F. to block free sulfliydryls to characterize both free sulfhydryls and disulfide bonds in a. number of proteins. } Since the two reagents apparently do not react with each other, their combination offers a convenient and quite general method for the complete characterization of free and cross-linked cysteines. Using the similar procedure, Kirley (53, 54) determined the location of three disulfide bonds and one free sulfhydryl group in the B-subunit of (Na+,K*)-ATPase. Another reagent, S-[2-((iodoacetyl)amino)-ethyl]naphthalene-1 -sulfonic acid (IAEDANS) (Figure 1.7) has been proven powerful for sulfhydryl derivatization (55-57). Recently, Sturrock et al (58) used this reagent to label a free cysteine residue in human testicular angiotensin-converting enzyme (tACE). After isolating the fluorescent peptide from enzymatic digests by HPLC, the sequence of the fluorescent peptides was mapped by MALDI-PSD. The sequence LACE. r rnzcc- Figure 1.7. 5-[2-((iodoacetyl)arr 2. Affinity Chromatography The only functional group enable covalent bond which can The sulflrydryl-containing protei interaction (59). After enzymati peptides do not bind to the colum containing fragments are then or structural characterization by Ednu ' interactions include (60): a) bind binding to affinity media with tea exchange, and (iii) binding to af methods were originally used to pt 19 by MALDI-PSD. The sequence data established that Cys496 is a free thiol form in tACE. u IHZCC—NH-CHZCHZ-NH ©© SO3H Figure 1.7. 5-[2-((iodoacetyl)amino)-ethyl]naphthalene-l~sulfonic acid (IAEDANS). 2. Affmity Chromatography The only functional group for which conditions are available for the formation of a stable covalent bond which can be Split under mild conditions is the sulfhydryl group. The sulfhydryl-containing protein is attached to an' affinity column by covalent interaction (59). After enzymatic digestion on column, the nonsulfhydryl-cbntaining peptides do not bind to the column and can be washed away. The attached sulfliydryl- containing fragments are then removed from the column and - subjected to further structural characterization by Edman degradation and/or mass spectrometry. The affinity interactions include (60): (i) binding to affinity media that contain heavy metals, (ii) binding to affinity media with reactive disulfide that undergo facile sulfhydryl/disulfide exchange, and (iii) binding to affinity media that contain chelated zinc. While those methods were originally used to purify sulflrydryl-containing proteins, this approach was recently used to locate cysteine snlfliydryl-containing peptides (61 groups in high mass proteins bwa But conditions have to be establ' grontpseanbeattachedtothecol 3. Cleavage at Cysteine Residues The third method utilizes a l ofanSH group into an SCN gn ream ofa protein with Ellman showed that 2-nitro-5-thiocyanobe sulfliydryls, which subsequently cle residues under mildly alkaline conc of 2-iminothiazolidine-4—carboxyly residues, the cleavage reaction rest alignment of which indicates the potentially a very useful method, 1‘ because all but the N-terminal p cuboxylyl group, which is not am convenient method to remove the Because cysteines are relatively sc ‘ Produces large fragments which 4 Wllacrylamide gel electrophoresi: 20 recently used to locate cysteine residues in proteins by mass mapping of bound sulfhydryl-containing peptides (61). This approach is powerful for locating sulfhydryl groups in high mass proteins because the HPLC separation of the digests is unnecessary. But conditions have to be established and optimized under which all the sulfhydryl groups can be attached to the column. 3. Cleavage at Cysteine Residues The third method utilizes a cleavage reaction at cysteine residues. The conversion of an SH group into an SCN group was first achieved in two stages by successive treatment of a protein with Ellman’s reagent and then with cyanide (62, 63). Stark (64) showed that 2-nitro—5-thiocyanobenzoic acid (NTCB) specifically cyanylates cysteine sulfhydryls, which subsequently cleave at the N-terminal side of the cyanylated cysteinyl - residues under mildly alkaline conditions to form an amino-terminal peptide and a series of 2-iminothiazolidine-4-carboxylyl (ITC) peptides. If a protein contains 11 cysteine residues, the cleavage reaction results in the formation of n+1 peptide fragments, mass alignment of which indicates the number and location of cysteine residues. While potentially a very useful method, it has seldom been used for sequence determination because all but the N-terminal peptide become blocked by the iminothiazolidine- carboxylyl group, which is not amenable to Edman degradation. Thus far, there is no convenient method to remove the ITC, group from the cleavage products (65, 66). Because cysteines are relatively scarce in proteins, cleavage at these residues usually produces large fragments which can be mass mapped by sodium dodecyl sulfate- polyacrylamide gel electrophoresis (SDS-PAGE) (63, 64, 67). However, peptide fi assignments using this approach an 5%) (64, 67). Papayannopoulos spectrometry to sequence the NTC isolated from Sarcophaga buIlala. mass assignment and can be applier tlie NTCB approach has primarily proteins although selective identifi bonds was also reported (68-73). The cleavage reaction men mas spectrometry are the theme 0 theclwvage reaction will be discus C Disulfide Bond Assignment Although the amino acid 5 corresponding cDNA, modification; predicted accurately. Disulfide l POSttranslaticnal modifications, is e: As a result, most disulfide-containi bonding is correct. Improved anal 0f S-S are thus important to m However, although there is 800d m 21 assignments using this approach are often complicated by its poor mass accuracy ( error > 5%) (64, 67). PapayannOpoulos and Biemann (68) first used CID tandem mass spectrometry to sequence the NTCB cleavage reaction products of a protease inhibitor isolated from Sarcophaga bullata. Mass spectrometry provides much more accuracy for mass assignment and can be applied to the sequencing of ITC blocked peptides. To date, the NTCB approach has primarily been employed to locate total cysteines in various proteins although selective identificatiOn of free cysteine in the presence of disulfide bonds was also reported (68-73). The cleavage reaction mentioned above and mass mapping of the products by mass spectrometry are the theme of this dissertation. The chemistry and mechanism of the cleavage reaction will be discussed in detail in chapter 2. C. Disulfide Bond Assignment Although the amino acid sequences of proteins are readily deduced from the corresponding cDNA, modifications occurring to the protein after translation can not be predicted accurately. Disulfide bonding, one of the most frequently encountered posttranslational modifications, is essential for stabilization of protein’s tertiary structure. As a result, most disulfide-containing enzymes have maximum activity only if disulfide bonding is correct. Improved analytical methods for the rapid and definitive assignment of SS are thus important to many areas of biochemical research and technology. However, although there is good methods for quantifying the number of disulfide bonds i inproteins (41, 42), the unambiguo taskwhich is sometimes only resoli Two frequently used ten exchange and disulfide bond sen sulflli’dfltl/disulfide exchange (sue both sulfliydryl/disulfide exchange mildly alkaline media when stru d'uulfide bond surnames. Figun protein Figure 1.8. Illustration 0 Sanger and his coworkers 1 insulin, recognized that an interc Protein was subjected to partial formation of mixed disulfides. S disulfide exchange, disulfide bond was minimal over the range of pH 22 in proteins (41, 42), the unambiguous determination of disulfide bonds remains a difficult task which is sometimes only resolved by a main commitment of time and sample. Two frequently used terms in cysteine chemistry are sulfliydryl/disulfide exchange and disulfide bond scrambling (or disulfide bond interchange). Although sulfhydryl/disulfide exchange (such as Ellman’s reagent) is the most specific reaction, both sulfhydryl/disulfide exchange and disulfide bond scrambling in proteins occur in mildly alkaline media when structures permit and result in mismatched nonnative disulfide bond structures. Figure 1.8 shows the sulfhydryl/disulfide exchange in a protein. Figure 1.8. Illustration of sulfliydryl/disulfide exchange in proteins. Sanger and his coworkers (74-76), in their pioneering work on the structure of insulin, recognized that an interchange reaction took place with disulfide bonds when the - protein was subjected to partial hydrolysis in cold concentrated HCl, leading to the formation of mixed disulfides. Subsequent studies demonstrated that , like sulfhydryl/ disulfide exchange, disulfide bond scrambling also occurs in slightly alkaline buffers but was minimal over the range of pH 2 to pH 6.5 (77, 78). R'S-SR' + R"S In alkaline and neutral met from the hydrolytic cleavage of dis [menu of alkylating agents. such 1. Classical Strategy Multiple step en Fractn Modific R'S-SR' + R"S-SR" ——-—-> 2R'S—SR" In alkaline and neutral media, the reaction is catalyzed by thiols which can arise from the hydrolytic cleavage of disulfides. Disulfide bond interchange is inhibited by the I presence of alkylating agents, such as iodoacetate (79). 1. Classical Strategy Multiple step enzymatic/chemical cleavage of protein l Fractionate peptides by HPLC l Reduce disulfide bonds l Modification of nascent SH groups l Separation of the resulting peptides l Sequencing by Edman degradation Figure 1.9. Classical strategy for disulfide bond assignment of a protein. As illustrated in Figure remgnin'ng disulfide bond strucu similar to the assignment of fre< enzymes or chemical reagents bet avoid or minimize disulfide scram reversed-phase HPLC. Third, the peptides are determined using Ed identified peptides are related to sp the methodology focused on the specially various mass spectrom claracteristic aspects of the strategj 2. Cleavage of Cystine Peptides The choice of enzymes for specificity of the enzyme to pro identification process; the condi capability of cleaving between ev no more than one disulfide bond. factors. The identification of di Specific cleavage reagent is used ( with a variety of enzymes, such as or less specifically, chymotrypsin. highly specific for Arg and Lys resi 24 As illustrated in Figure 1.9, a well established approach (65, 80, 81) for recognizing disulfide bond structure of proteins involves several steps which are very similar to the assignment of free sulfliydryl groups. First, a protein is cleaved by enzymes or chemical reagents between half-cystinyl residues under conditions that can avoid or minimize disulfide scrambling. Second, the mixture of digests is separated by reversed-phase HPLC. Third, the amino acid sequence or molecular masses of these peptides are determined using Edman degradation or mass spectrometry. Finally, the identified peptides are related to specific segments of the protein. Recent development of the methodology focused on the identification of the disulfide-containing peptides, especially various mass spectrometric techniques are more and more involved. The characteristic aspects of the strategy will be discussed below. 2. Cleavage of Cystine Peptides The choice of enzymes for cleavage is largely dictated by three requirements: the specificity of the enzyme to produce well-defined fragments that can simplify the identification process; the conditions to minimize disulfide interchange; and the capability of cleaving between every half—cystinyl residue to obtain peptides containing no more than one disulfide bond. Usually compromise has to be made among these three factors. The identification of disulfide-containing peptides is facilitated greatly if a specific cleavage reagent is used (10, 39, 82, 83). Proteins may be cleaved selectively with a variety of enzymes, such as trypsin, staphyococcus aureus V8, cyanogen bromide, or less specifically, chymotrypsin. Trypsin has been used most frequently because it is highly specific for Arg and Lys residues. Unfortunately, trypsin has maximum activity at pll 8.3, and is not active in acid. enzymolysis. However, in spite of published papers on the disulfide cleavage reagent, among which Chyrnotrypsin is less specific conditions. staphyococcw aureus 4.0 and 7.8, cleaving on the GI Cyanogen bromide is also used fn Me. This reaction is very attractir of side reactions, and because th volatile. In addition, disulfide br proteins in their native state are of isnsed to open up or unfold the prr isnot a particularly abundant cons large peptides. Subsequent digesti peptides (82). Unfortunately, if proteins c must be cleaved by non-specific r lhemrolysin, and partial acid hydro sllt‘tcific proteases it is difficult SCtlmcttts of the protein simply by case,multiple-step Edman degradat 0fthe disulfide-bridged peptides. 25 pH 8.3, and is not active in acid. As a result, disulfide scrambling may occur during enzymolysis. However, in spite of the jeopardy, our survey indicated that among over 60 published papers on the disulfide assignment in various proteins, 70% used trypsin as a cleavage reagent, among which 80% of the digestion was performed at pH>8. Chymotrypsin is less specific than trypsin, and likewise active only under alkaline conditions. staphyococcus aureus V8 is useful because it has maximum activity at pH 4.0 and 7.8, cleaving on the C-terminal side of Glu or Glu and Asp, respectively. Cyanogen bromide is also used frequently to cleave proteins on the C-terminal side of Met. This reaction is very attractive as it is highly specific for Met and is generally free of side reactions, and because the reagents (cyanogen bromide and formic acid) are volatile. In addition, disulfide bonds are stable during treatment with CNBr. Since proteins in their native state are often resistant to enzymatic attack, cleavage with CNBr is used to open up or unfold the protein, rendering it susceptible to enzymolysis. As Met is not a particularly abundant constituent of proteins, cleavage with CNBr usually gives large peptides. Subsequent digestion by other enzyme(s) is required to produce smaller peptides (82). Unfortunately, if proteins can not be cleaved by specific cleavage reagents, they must be cleaved by non-specific reagents. Pepsin (maximum activity around pH 3.0), thermolysin, and partial acid hydrolysis are useful for this purpose. However, using non- specific proteases it is difficult to relate the disulfide-bridged peptides to specific segments of the protein simply by determining their molecular masses by MS. In such a case, multiple-step Edman degradation and/or MS/MS is a useful tool for characterization of the disulfide-bridged peptides. 3. Purification of Cystine-Contai Following digestion of a 1 original disulfide bonds remain inta toidentify, by N-terminal amino : dinilfide bonded peptide. Reverse acetonitrile in 0.1% trifluomacetie : rapid and employing conditions nsuallydetected byUV at215 nm‘ tinm high-mass proteins is still 8 HPLC columns are used, because hydrolysis. An alternative method I identification of disulfide-bonded based on the chemical modification Orpaper electrophoresis is carried performic acid vapour. After th repeated in the perpendicular di migrate along a line at 45° to ea formed cysteic acid peptides migra then eluted and analyzed by amin method is diagonal electrophoresis 3. Purification of Cystine-Containing Peptides Following digestion of a protein to give a mixture of peptides in which the original disulfide bonds remain intact, the next step is to pry each of these peptides and to identify, by N-terminal amino acid and if necessary, partial sequence analysis, each disulfide bonded peptide. Reversed-phase HPLC under acidic conditions ( a gradient of acetonitrile in 0.1% trifluoroacetic acid, for example), is ideal for this purpose, being both rapid and employing conditions favoring stability of disulfide bonds. Peptides are usually detected by UV at 215 nm with acceptable sensitivity. The purification of digests from high-mass proteins is still a challenging task even though microbore or capillary HPLC columns are used, because so many peptide fragments might be produced after hydrolysis. An alternative method to reversed-phase HPLC for the separation and identification of disulfide-bonded peptides is diagonal electrophoresis. This method is based on the chemical modification of cysteinyl residues to cysteic acids. The thin-layer or paper electrophoresis is carried out in the horizontal direction, and then exposed to the performic acid vapour. After the performic acid is removed, the electrophoresis is repeated in the perpendicular direction. All peptides unaffected by performic acid migrate along a line at 45° to each direction of the electrophoresis, while the newly formed cysteic acid peptides migrate off the diagnal. Peptides containing cysteic acid are then eluted and analyzed by amino acid composition or sequences. An example of this method is diagonal electrophoresis map of a peptic digests of or—chymotrypsin. 4. Identification of Disnlfide Fe The conventional approac after HPLC separation is very t made by a comparison between c the disulfide bonds. Peptide peaks SS; for each such peak, two new S,itislikelythat there will be at 1 However, this strategy depends before and after reduction, and i rodwed species are alkylated and Performic acid oxidation (36) co modification has the disadvantagr residues will also have their elutio halfcystine residues may be. identi The advent of modem mass : disulfide bonds in proteins more bombardment (FAB), matrix-as electrospray ionization (ESI) have mass spectrometry, as shown in Fir their unique masses and by comps the signal corresponding to the corresponding to the respective thi 4. Identification of Disulfide Peptides The conventional approach for the identification of disulfide-containing peptides after HPLC separation is very tedious and time consuming. The identification can be made by a comparison between chromatograms of peptides before and after reduction of the disulfide bonds. Peptide peaks that disappear upon reduction are presumed to contain S—S; for each such peak, two new peaks are formed. If a single peptide has an internal S- S, it is likely that there will be at least a small shift in the elution position upon reduction. However, this strategy depends too much on the variation of hydrophobicity of peptides before and after reduction, and is therefore not always reliable. More effectively, the reduced species are alkylated and re-purified by HPLC followed by Edman sequencing. Performic acid oxidation (36) could also be used to identify peptides with S-S, but this modification has the disadvantage that peptides containing methionine and trytophan residues will also have their elution positions altered. Alternatively, peptides containing half-cystine residues may be. identified colorimetrically (52-54). . The advent of modern mass spectrometric techniques has made the task of locating disulfide bonds in proteins more tractable (10, 83). In the last decade, fast atom bombardment (FAB), matrix-assisted laser desorption/ionization (MALDI), and electrospray ionization (ESI) have widely been used for disulfide bond assignment. In mass spectrometry, as shown in Figure 1.10, disulfide linked peptides are identified from their unique masses and by comparison with the spectrum of reduced samples in which the signal corresponding to the S-S linked peptide(s) is replaced by two signals corresponding to the respective thiol peptide components, if inter-bridged, or shifted by Figure 1.10. Current protocol tvt0 mass units (dithiol) if infra-l for identifying peptides because it that becomes important when tl hldrclysis of large proteins wlr 28 SH S *3 Disulfide bonded E protein B s ~s E Enzymic/chemical digestions ir Mixture of peptides SH ' [+Y + M + \’S—S7/ +npeptides l / \ ‘ \ / \ l \ / r \ , l r \ l I , Ml / identification by MS l /// / 4/ / /\ reduction and MS analysis Figure 1.10. Current protocol for disulfide bond assignment by mass spectrometry. two mass units (dithiol) if intra-bridged. Mass spectrometry is a particularly good tool for identifying peptides because it does not require rigorously purified peptides, a feature that becomes important when the amount of protein is small or when investigating hydrolysis of large proteins where purification of peptides to homogeneity may be difficult Furthermore, modem MALDI-PSD, FAB/CIDIMS/MS, provides confirmation of a specific Takao er al (84) first disulfide-containing peptides from be determined Almost at the strategy for the recognition of di became an accepted method an inhibitor of amylase (86) and h group srrbwquently demonstrated ribonuclease A and lysozyme methodologies based on mass 5 containing peptides (39, 83). Peptide—mapping by E81 (1 increasingly in the characterization uscdhereisexactly same as in sensitivity (low picomole to high HPLC has made the purification MALDI-MS provides the capacity Wins and peptides. It is particu mapping without the previous isola interchain disulfide-containing pep excellent tool for the direct diagno 29 difficult. Furthermore, modern mass spectrometric techniques, such as ESI/MS/MS, MALDI-PSD, FAB/CID/MS/MS, permit the sequence determination of peptides which provides confirmation of a specific peptide segment in a protein. Takao et al (84) first recognized the potential of FAB-MS for identifying disulfide-containing peptides from which the location of disulfide bonds in proteins may be determined. Almost at the same time, Morris and Pucci (85) reported the same strategy for the recognition of disulfide bond structure in insulin. Their protocol soon became an accepted method and successfully applied to proteins such as bacterial inhibitor of amylase (86) and human tissue inhibitor of metalloproteinases (87). Smith’s group subsequently demonstrated the feasibility of the FAB/MS for S-S assignment in ribonuclease A and lysozyme, ’ and further developed several supplementary methodologies based on mass spectrometry to facilitate the identification of disulfide- containing peptides (3 9, 83). Peptide-mapping by ESI (88, 89) or MALDI-MS (90, 91) are being used increasingly in the characterization and identification of disulfide bridges. The strategy used here is exactly same as in the F AB-MS. However, they provide much more sensitivity (low picomole to high femtomole). Besides, the easy interface of ESI to HPLC has made the purification of the disulfide-containing peptides an easy task. MALDI-MS provides the capacity of rapidly analyzing small quantities of mixtures of proteins and peptides. It is particularly well suited to the determination of peptide mass mapping Without the previous isolation or fractionation. The in situ disulfide cleavage of interchain disulfide-containing peptide during the analysis by MALDI-MS provides an excellent tool for the direct diagnosis of disulfide bonded peptides (90, 91). The poor resolution of both E31 and MAL identification of intrachain disu] reduction of the S-S bond. Bean and Carr (92) used electronvolt) collisionOactivated di mmes up to 2,000 Da containin cleavage at the disulfide bridge wi fragment, giving a characteristic 32u These peaks are remarkabl products of a disulfide bond cleav 1 Other Methodologies for SS Despite their many successe locating disulfide bond linkages. I inde alkaline solutions. The er because the requirement to find 3 always of priority and most of the ; Another problem is that the cleav the locations of S-S can be deduce separated by the amino acids r reactions. This is especially the example, trypsin is useful only if Even when the half-cystinyl resid 3O resolution of both ESI and MALDI instruments, however, could be a problem for the identification of intrachain disulfide bond which has only two mass units shift after reduction of the S-S bond. Bean and Carr (92) used tandem mass spectrometry with high—energy (kilo- electronvolt) collision-activated dissociation (CAD) to examine a number of peptides of masses up to 2,000 Da containing a single inter-chain disulfide bridge. They observed cleavage at the disulfide bridge with retention of zero, one, or two sulfurs on the charged fragment, giving a characteristic triplet of peaks in the mass spectrum with separation of 32 u. These peaks are remarkably intense, so they can provide ready identification of products of a disulfide bond cleavage. 5. Other Methodologies for S-S Assignment- Despite their many successes, two problems remain in the traditional methods for locating disulfide bond linkages. One problem is the disulfide scrambling which occurs in mild alkaline solutions. The endeavor to minimize the scrambling are often fruitless because the requirement to find a specific enzyme to produce recognizable fragments is always of priority and most of the specific enzymes only work under alkaline conditions. Another problem is that the cleavage reaction does not always give peptides from which the locations of S-S can be deduced. This may be because half-cystinyl residues are not separated by the amino acids required by specific chemical or enzymatic cleavage reactions. This is especially the case for a protein containing adjacent cysteines. For example, trypsin is useful only if the half-cystinyl residues are separated by Lys or Arg. Even when the half-cystinyl residues are separated by amino acids suitable for cleavage by specific enzymes, the reaction highly folded, as it is when disulfi for the reduced and carboxymethy 94). if a protein fails to produ method, partial acid hydrolysis, nixtme of disulfide-containing hdnoed (95). Partial acid hydro ternperatrne, etc.) under which 0 hydrolysis is attractive as disulfid Steric effects play less of a role i under acid conditions, and beca enzymes. However, due to the lo evenamedium-sized protein (e. g. of peptide fragments. Many overlz the cleavage pattern is not predicta identify peptide fragments, com] assisted data interpretation have to Another problem occurs whe any enzyme. lhat is, it is unlike cleave between the two cysteine re: one disulfide bond. The resulting Slmcture and there are two possible 31 by specific enzymes, the reaction may not proceed at an appreciable rate if the protein is highly folded, as it is when disulfide bonds are intact. Hence, cleavage that occur readily for the reduced and carboxymethylated protein may not occur in the native protein (93, 94). If a protein fails to produce desirable fragments after digestion, an alternative method, partial acid hydrolysis, was proposed to replace enzymatic digestion to produce a mixture of disulfide-containing peptides from which the disulfide connections may be deduced (95). Partial acid hydrolysis is performed under controlled conditions (time, temperature, etc.) under which only limited peptide fragments are formed. Partial acid hydrolysis is attractive as disulfide bonds are particularly stable in dilute or weak acids. Steric effects play less of a role in acid hydrolysis because most proteins are denatured under acid conditions, and because the catalyst, H3O+, is much smaller than proteolytic enzymes. However, due to the low specificity of partial acid hydrolysis, the digests of even a medium-sized protein (e. g., hen egg-white lysozyme) are a very complex mixture of peptide fragments. Many overlapped disulfide-containing peptides were produced and the cleavage pattern is not predictable. Therefore, even if mass spectrometry was used to identify peptide fragments, complicated chromatographic separation and computer- assisted data interpretation have to be applied to the definitive identification. Another problem occurs when a protein has a -Cys-Cys- structure that is resistant to any enzyme. That is, it is unlikely, if impossible, to find a‘ cleavage reagent that can cleave between the two cysteine residues and produce peptide fragments that only contain one disulfide bond. The resulting peptide always contains two disulfide bonds in its structure and there are two possible isomers (suppose they are inter-chain disulfides). Different approaches have used the combination of organi containing peptide and high-perf assign the disulfide structure of large investment of time and ma Strategy takes advantage of multi (99) demonstrated that during seqt nophenylthiohydantoin (PTH) can premably remains attached to tht the second Cys, a PTH upper dehydroalanine, which can cast} sequencer. Zhang and Liang (100) Httwentoxin-I that contains two adj Gray (103-105) developed a cRhine-rich small proteins. This disulfide bonds and sequence anan subjected to limited degree of re 32 -——C __ XS 0/ c\ \ I \ / Different approaches have been reported to solve the problem. Stults et al (96) used the combination of organic synthesis of the possible isomers of a disulfide- containing peptide and high-performance tandem mass spectrometry in their attempt to assign the disulfide structure of recombinant relaxin. This method obviously requires ' large investment of time and material, and is therefore not widely accepted. Another strategy takes advantage of multiple step Edman degradation (97-102). Nokihara et al (99) demonstrated that during sequence analysis of a peptide containing a disulfide bond, no phenylthiohydantoin (PTH) can be seen after cleavage of the first Cys residue, since it presumably remains attached to the second Cys by the disulfide bond. After cleavage of the second Cys, a PTH appears as a cysteine and a dithiothreitol adduct of dehydroalanine, which can easily be distinguished from other PTH-AAs on the sequencer. Zhang and Liang (100) used this strategy for studyingthe disulfide linkage of Huwentoxin-I that contains two adjacent cysteine residues. Gray (103-105) developed a new approach for disulfide mapping of tightly folded, cysteine-rich small proteins. This approach utilized a combination of partial reduction of disulfide bonds and sequence analysis of alkylated peptides. In this strategy, a protein is subjected to limited degree of reduction under acidic conditions to form a series of partially opened disulfide bonds sulflrydryls, followed by sequence bond(s) that had been reduced. and has been successfully applied However, one limit on the applica fragments after digestion of the demonstrated by lshibashi e! a1(l In all, disulfide linkage assi Two troublesome problems, the d Cyscontaining protein and the p present There is a considerabl techniques for disulfide bond assi areas in studying protein structur answers to the problems in deter protein chemists. 111. Matrix-Assisted Laser Decor Since the 1960s mass spectrt liltide and protein analysis, but ( c(Impounds real progress in the tie ionization methods; plasma desor bombardment mass spectrometry ( ttthrriques, matrix-assisted laser i 33 partially opened disulfide bonds that are separated by HPLC. Alkylation of free sulfliydryls, followed by sequencer analysis, provided explicit assignment of the disulfide bond(s) that had been reduced. This approach can minimize disulfide bond scrambling and has been successfully applied to a number of proteins containing adjacent cysteines. However, one limit on the application of the method concerns peptide size. Analysis of fragments after digestion of the protein can hopefully overcome the drawback, as demonstrated by Ishibashi et al (106). In all, disulfide linkage assignment in proteins is still challenging protein chemists. Two troublesome problems, the detemiination of disulfide bond arrangement in a -Cys- Cys-containing protein and the prevention of disulfide scrambling, remain bothersome at present. There is a considerable interest in the protein society in developing novel techniques for disulfide bond assignment. Mass spectrometry is one of the most active areas in studying protein structure. It is hoped that mass spectrometry can provide answers to the problems in determination of disulfide linkage that remain bothering protein chemists. III. Matrix-Assisted Laser Desorption/Ionization Mass Spectrometry (MALDI-MS) Since the 19603 mass spectrometry has been considered a promising technique for peptide and protein analysis, but owing to the size, polarity and low volatility of these compounds real progress in the field was only achieved after discovery of the desorption ionization methods; plasma desorption mass spectrometry (PDMS) (l), and fast atom bombardment mass spectrometry (FAB MS) (2). The introduction in 1988 of two new techniques, matrix-assisted laser desorption/ionization (MALDI) (3) and electrospray ionization (ESI) (4), has extended scieneeiarbeyond the scope of P very difiaent ideas and princip difficulties with MS methods. i. biomolecules, and shown consi accurate determination of molec combination with biochemical tec lire first attempts to g desorpnonfronimtion date back to Which can be desorbed and ion molecules are analyzed, more cncr clergy flux leads to pyrolytic ( breakthrough came in the late experimented with the use of a m desorptionfronization of large, non obtained when such samples were large molar excess of an energy-ab: not strongly absorb at the wavelet this technique was conducted with an The ejected ions were analyze demonstrated that the ions generat 67,000 daltons could be formed Despite lack of an established io 34 ionization (ESI) (4), has extended the applicability of mass spectrometry in the biological science far beyond the scope of PDMS and FAB-MS. Although MALDI and E81 utilized very different ideas and principles, both new techniques have overcome the main difficulties with MS methods, i. e., the desorption and ionization of large and labile biomolecules, and shown considerable promise in characterizing biomolecules by accurate determination of molecular mass up to a few hundred thousand daltons. In combination with biochemical techniques, numerous applications have been achieved (5). The first attempts to generate ions of organic molecules by direct laser desorption/ionization date back to early 1970s (107). However, the size of the analytes which can be desorbed and ionized was limited to ~1,000 daltons. When larger molecules are analyzed, more energy is required to desorb them. However, such a high energy flux leads to pyrolytic or photochemical decomposition of analytes. The breakthrough came in the late 19803 when Hillenkamp and Karas successfully experimented with the use of a matrix, nicotinic acid (3, 108). It was discovered that desorption/ionization of large, nonvolatile molecules such as proteins could be similarly obtained when such samples were irradiated by a laser after being codeposited with a large molar excess of an energy-absorbing “matrix” material, even though the analyte did not strongly absorb at the wavelength of the laser radiation. The first experiment with this technique was conducted with frequency-quadrupled Nd-YAG laser operated at 266 nm. The ejected ions were analyzed by a time-of-flight (TOF) mass spectrometer. It was demonstrated that the ions generated from proteins with molecular mass range 10,000- 67,000 daltons could be formed and detected from picomolar amounts of analyte. Despite lack of an established ionization mechanism, the dependence on the matrix material in the desorption/ionized laser desorptionftonization (MAL powerful technique for accurate groaerthan 200,000 Da. With the (PSD) and delayed extraction (DE formicmral elucidation of biopol The majority of the work immanents being available in th (Chapter 2) was conducted on a Biosystem Voyager Elite, which Spectrafor structural analysis; thi: extraction hardware and sofiwan however, are based on the same pri ”he Matrix Matrix is the key to the suc discovery of nicotinic acid as t meounds have been investigated applicable. Among them, sinap tinnamic acid, and a-cyano—4-hyt (111-113). It was considered imp maximum of the matrix compound, 35 material in the desorption/ionization process gave rise to the terminology matrix-assisted laser desorption/ionization (MALDI). Today MALDI has developed into an extremely powerful technique for accurate and sensitive analysis of molecular masses extending to greater than 200,000 Da. With the introduction of two new techniques, post-source decay (PSD) and'delayed extraction (DE), MALDI-TOP MS is also becoming an effective tool for structural elucidation of biopolymers. The majority of the work that will be presented was performed on two MALDI instruments being available in the MSU/NIH mass spectrometry facility. Early work (Chapter 2) was conducted on a Vestec 2000. Others were gained on a PerSeptive Biosystem Voyager Elite, which can be Operated in the reflectron mode to acquire PSD spectra for structural analysis; this instrument has recently been equipped with delayed extraction hardware and software for improving resolution. The two instruments, however, are based on the same principle. A. The Matrix Matrix is the key to the successful analysis of high-mass biomolecules- Since the discovery of nicotinic acid as the first matrix, several hundred different organic compounds have been investigated (109, 110). However, only a few of these are widely applicable. Among them, sinapinic acid (SA), 2,5-dihydroxybenzoic acid (DHB), cinnamic acid, and or-cyano-4-hydroxycinnamic acid (a-CHCA) are extremely useful (111-113). It was considered important that the laser wavelength match the absorption maximum of the matrix compound. All the matrices above have strong UV absorption in — the320-350 nm range and can be tenancy-doubled Nd-YAG lase include miscibility with the anal} required for the dissolution of the chemical composition that promot protons to the analyte, nonreactivi retire low heat of sublimation and The physioochemical even and their ionization in MALDI he believed to serve to minimize sam1 incident laser energy, resulting in 1 gas phase. One model for the me of matrix are induced to undergo l submquent expansion of these n isolated protein molecules into the protein undergoes ionization throu; Protesses that remain to be explain B. Sample Preparation Proper sample preparation i (“’0- Matrix solutions are typi mixtures at a concentration of : 36 the 320-350 nm range and can be used with a much cheaper nitrogen laser (337 nm) or frequency-doubled Nd-YAG lasers (355 nm). Other important matrix characteristics include miscibility with the analyte in the solid phase, solubility in the same solvents required for the dissolution of the analyte, vacuum compatibility (low vapor pressure), a chemical composition that promotes the ionization of matrix substituents that can donate protons to the analyte, nonreactivity with the analyte, and other physical prOperties such as the low heat of sublimation and a capacity to crystallize readily (114). The physicochemical events leading to the transfer of proteins to the gas phase and their ionization in MALDI have not yet been fully elucidated (115). The matrix is believed to serve to minimize sample fragmentation from the laser beam by absorbing the incident laser energy, resulting in the sample and matrix molecules being ejected into the gas phase. One model for the mechanism (116, 117) assumes that the uppermost layers of matrix are induced to undergo a phase transition fiom the solid to the gas phase. The subsequent expansion of these matrix molecules into the vacuum drags the matrix- isolated protein molecules into the gas phase. During the transfer to the gas phase, the protein undergoes ionization through proton transfer reactions with the matrix by reaction processes that remain to be explained. B. Sample Preparation Proper sample preparation is also critical for successful analysis by MALDI-MS (118). Matrix solutions are typically prepared in water/ethanol or water/acetonitrile mixtures at a concentration of 5-10 pig/pl (~40 mM), depending on the solubility properties of the matrix. The solvent that is miscible with the m triflmroacetic acid (IPA) is frequ matrix and analyte solutions are ~5,000:1 to 10,000:l, and final co reportively. An aliquot of ~lul (V12000), and allowed to dry by warm air, or under vacuum. lnstet he sample plate holder used in Va pumit a much higher sample ti codeposits from solution with thi chamber by a probe and a vacuum hand” in the Voyager Elite MALL d1uation(l to 10 ns) pulses ofan L' To date, the chemrsu’y' of the choice of matrix and method 0 difierent matrices and solvents(11 122) afl‘ects the resolution and se analyses under different conditions 37 properties of the matrix. The analyte is prepared at a concentration of ~0.1 pig/p11 in a solvent that is miscible with the matrix solution (for peptides and proteins, aqueous 0.1% trifluoroacetic acid (TFA) is frequently used to maintain the acidity of the solution). The matrix and analyte solutions are mixed to give a final matrixzanalyte molar ratio of ~5,000:1 to 10,00021, and final concentration of ~40 nmol of matrix/ ~1pmol of analyte, respectively. An aliquot of ~1ptl of the mixture is applied to MALDI-MS probe tip (V T2000), and allowed to dry by either ambient evaporation, heating with a stream of warm air, or under vacuum. Instead of the stainless steel probe tips used in the VT2000, the sample plate holder used in Voyager Elite system can accommodate 100 samples and permit a much higher sample throughput. During the drying process, the matrix codeposits from solution with the analytes. The sample is inserted into the vacuum chamber by a probe and a vacuum lock in the VT2000, and by an “automatic mechanical han ” in the Voyager Elite MALDI-TOF mass spectrometer, and irradiated with short duration (1 to 10 ns) pulses of an UV laser beam. To date, the chemistry of “matrix assistance” remains incompletely understood; the choice of matrix and method of application is still empirical. Sample preparation by different matrices and solvents(119), matrix additive (120), and evaporation rate (121, 122) affects the resolution and sensitivity of MALDI. Optimal results require parallel analyses under different conditions. (L instrumentation lnMALDl (Figure 1.”) laser which is directed and fee controlled by an attenuator and is interaction of photons with the m of the cocrystallized sample/man spread of ions generated by Mr moderating voltage or a reflectror resolution Experimentally, a statir sample by application of a two-star nith regard to a closely spaced at the some kinetic energy (assuming towards long (12 m) field-free tin MALDI is typically used in because TOF is simple, cheap, an MALDI. It has virtually no uppe Which can produce very high m/z its capacity to generate the entire information, in contrast to a s traverse the flight tube (T OF) is relationship: TOF = L/l) = L(m/Zz velocity, in is the mass of the ion 38 C. Instrumentation In MALDI (Figure 1.11) the cocrystal of sample/matrix is irradiated by a pulsed laser which is directed and focused by a prism and Optical lens. The irradiance is controlled by an attenuator and is increased gradually until the threshold is reached. The interaction of photons with the matrix and protein results in the desorption and ionization of the cocrystallized sample/matrix from a metal surface. The initial kinetic energy spread of ions generated by MALDI is large, so either a linear TOF with a high accelerating voltage or a reflectron with an ion mirror (or both) is used to improve mass resolution. Experimentally, a static electric field is imposed upon ions generated from the sample by application of a two-stage high voltage (typically i 25 kV) to the sample probe with regard to a closely spaced accelerating electrode. The ions are thus accelerated to the same kinetic energy (assuming the initial kinetic energy is zero) by the electric field toward a long (1-2 m) field-free time-of-flight (TOF) analyzer. MALDI is typically used in conjunction with a time-of-flight mass analyzer (123), because TOF is simple, cheap, and well-suited to the pulsed nature of laser desorption in MALDI. It has virtually no upper mass range and is therefore compatible with MALDI, which can produce very high m/z ions. Another advantage of the TOF mass analyzer. is its capacity to generate the entire spectrum from every single laser shot without losing information, in contrast to a scanning mass analyzer. The time required for ions to traverse the flight tube (TOF) is dependent on their masses and is described by the relationship: TOF = Up = L(m/22 eV)”, where L is the length of flight tube, to is the ion Velocity, m is the mass of the ion, and V is the acceleration potential. Thus, low-mass no. 77—].— Pmbe Ti Figure 1.11. Scheme of matrix-a: Spectrometer (MALDI-TOF MS) 39 Beam Splitter #:2212533} Laser / / \ I I :3' Transient /// - Recorder // Trigger // I] _ Sample // H I I J— Probe \ 6'? G) $ éf—‘érrr 6% ‘9 l n \¥ _1 T‘\ FlightTube U ' \ T \ Detector I Ion \ I Source | \ \ \ \ Probe Tip Figure 1.11. Scheme of matrix-assisted laser desorption/ionization time-of—flight mass spectrometer (MALDI-TOF MS) ions have a shorter flight time spatially discrete individual ion mlz ratio. A detector positioned as each ion packet strikes it. A yields a TOF spectrum. The di pulse and common to all ions, an proportional to (m/z)“" and can b time, tot) into a m/z ratios axis highly efficient because all ions 0 manned; they simply arrive at recorded for a particular ion refle ion source (such as the time/lo distributions). The poor resolving with the mass of the ionized ma isotopic masses are utilized in the detect certain protein modificati masses. In spite of this, it is Pmteins with molecular masses accuracy for proteins above 40 k 1). Characteristic Features of One of the important fea MALDI is a typical soft ionizatio 40 ions have a shorter flight time than heavier ions. They are separated into a series of spatially discrete individual ion packets, each traveling with a velocity characteristic of its m/z ratio. A detector positioned at the end of the field-free flight-tube produces a signal as each ion packet strikes it. A recording of the detector signal as a function of time yields a TOF spectnun. The difference between the start time, triggered by the laser pulse and common to all ions, and the arrival time of an individual ion at the detector is proportional to (m/z)W2 and can be used to convert the x-axis of the spectrttm (ion arrival time, tof) into a m/z ratios axis ( a conventional mass spectrum). The tof analyzer is highly efficient because all ions of different m/z ratio arising from a single laser shot are measured; they simply arrive at the ion detector at different times. However, The tof recorded for a particular ion reflects many different initial conditions experienced in the ion source (such as the time/location of ion formation and initial kinetic energy distributions). The poor resolving power is reflected by peak broadening which increases with the mass of the ionized macromolecules. Therefore, average masses rather than isotopic masses are utilized in the MALDI-TOF measurement. This limits the capacity to detect certain protein modifications and protein sequence variations, especially at high masses. In spite of this, it is possible to achieve a mass accuracy of 0.1-0.01% for proteins with molecular masses between 1 and 40 kDa, and with somewhat poorer accuracy for proteins above 40 kDa ( 123). D. Characteristic Features of MALDI-TOF MS One of the important features of MALDI spectra is its simplicity (116, 118). MALDI is a typical soft ionization technique that yields little fragmentation. Unlike ESI spectra which are dominated by in MALDI is the singly proto also observed to varying degrees or analyte/alkali metal interactio Compmed to other io ' common in biological samples unseparated, heterogeneous corn mixtures or synthetic reaction mi: The most striking feature has been umd to measure protei 500,000 Da (124). MALDI is als Spectra with as little as 1 femt rcported (112). The analysis at I obtained. E. Novel Techniques in MAL MALDI has been a su and peptides, and other biologi because it is a soft ionization te introduction of post-source deca technique (125, 126). is based 0 collision-induced decay taking P 41 spectra which are dominated by multiply charged species, the predominant analyte signal in MALDI is the singly protonated species, though the doubly protonated molecule is also observed to varying degrees. Other adducts arising from analyte/matrix interactions or analyte/alkali metal interactions are of relatively low abundance. Compared to other ionization techniques such as ESI or FAB, MALDI-MS is relatively tolerant to contaminants, such as buffers, salts, and denaturants which are common in biological samples (116). It has therefore the capacity to analyze unseparated, heterogeneous complex mixtures, such as enzymatic or chemical digest mixtures or synthetic reaction mixtures. The most striking feature of MALDI-MS is its very large practical mass range. It has been used to measure proteins and glycoproteins with molecular weight as high as 500,000 Da (124). MALDI is also a very sensitive technique for protein characterization. Spectra with as little as l femtomole of protein applied on the probe tip have been reported (112). The analysis at low picomole to high femtomole range can be routinely obtained. E. Novel Techniques in MALDI-TOF MS MALDI has been a superior method for molecular mass determination of proteins and peptides, and other biological materials, but lacks the capacity for structural analysis, because it is a soft ionization technique. This deficiency has been overcome since the introduction of post-source decay (PSD) analysis of MALDI-generated ions. The PSD technique (125, 126). is based on mass analysis of product ions from unimolecular or collision-induced decay taking place in the field-free region between the ion source and thereflectron. Incontrast to a li energy analyzer to differentiate i velocity in the linear mode. lons accelerated precursors are detect precursor in the linear mode of because the product ions have 1 reflectron mode of the TOF ins PSD, like MS/MS techniques, car sized peptides (@500 Da). Sinc PSD has quickly evolved into a p atthe low picomole scale. Another limitation of distribution of ions in the MALD nearly independent of mass and the analyte. In addition, when presumably lost by collision wi energy dispersion To attain hi ion extraction. In contrast to generated by the laser beam n potential, in delayed extraction, ionization and ion extraction ev isfield-free during the delay. F speller. Application of the ap 42 the reflectron. In contrast to a linear instrument, reflection instruments can be used as an energy analyzer to differentiate ions that are otherwise detected as species with the same velocity in the linear mode. Ions formed as a result of metastable decomposition of fully accelerated precursors are detected at the same arrival time (same apparent mass) as their precursor in the linear mode of TOF because they have the same velocity. However, because the product ions have lower kinetic energy, they can often be resolved in the reflectron mode of the TOF instrument by lowering potential of reflector. Therefore, PSD, like MS/MS techniques, can provide full or partial sequence information of medium sized peptides (<2500 Da). Since its commercialization about three years ago, MALDI- PSD has quickly evolved into a powerful technique for sequence determination of peptide at the low picomole scale. r Another limitation of traditional MALDI-TOF MS is the broad energy distribution of ions in the MALDI source. The initial velocity of desorbed analyte ions is nearly independent of mass and the initial kinetic energy is proportional to the mass of the analyte. In addition, when desorption occurs in a strong electric field, energy is presumably lost by collision with the neutral plume, resulting in further mass-dependent energy dispersion. To attain high resolution, a delayed extraction technique is used for ion extraction. In contrast to conventional MALDI instruments in which the ions generated by the laser beam near the surface of the sample probe are extracted by a dc potential, in delayed extraction, a short time delay (<3 00 ns) is inserted between the laser ionization and ion extraction event. The region between the repeller and extraction grid is field-free during the delay. Following the delay, a pulsed potential is applied to the repeller. Application of the appropriate pulse voltage provides the energy correction necessary to simultaneously det initial energy. The initially less pulseisappliedandtraverseal ions. An energy/spatial correcti reach the detector plane sirnul accuracy, and the quality of reducing chemical noise, and m lhe potential of delayed extrac peptides (1271130) as well as for F. Application of MALDI-TO] Over the past few years, 1 yet available for macromolecula achievements in protein, nucleotir MALDI has been exte structures of proteins derived fro provides routine and reliable m typically have a high degree of extraneous compounds has bee malaria parasite. This remarkab now within the reach of mass 3 MALDI is an excellent m modifications. Covalently mod' 43 necessary to simultaneously detect all ions of the same mass/charge regardless of their initial energy. The initially less energetic ions are closer to the repeller at the time the pulse is applied and traverse a longer segment of the electric field than more energetic ions. Arr energy/spatial correction is thus provided such that all ions of the same m/z reach the detector plane simultaneously. Delayed extraction improves resolution, mass accuracy, and the quality of MALDI mass spectra by suppressing matrix background, reducing chemical noise, and minimizing the effect of laser intensity on performance. The potential of delayed extraction MALDI has been demonstrated for proteins and peptides (127:130) as well as for oligonucleotides (131). F. Application of MALDI-TOF MS Over the past few years, MALDI has become among the most powerful methods yet available for macromolecular characterization of living systems. A wide range of achievements in protein, nucleotide, and glycobiology have been attained in a short time. MALDI has been extensively used for the determination of primary covalent structures of proteins derived from both natural and recombinant sources (132). MALDI provides routine and reliable means to analyzing tryptic digests and glycoproteins which typically have a high degree of heterogeneity (133). The high tolerance of MALDI to extraneous compounds has been used to study the degradation of hemoglobin by a malaria parasite. This remarkable achievement shows direct analysis of cell contents is now within the reach of mass spectrometry (134). MALDI is an excellent method of choice for dealing with protein posttranslational modifications. Covalently modified protein N- and C- termini (135), disulfide bonds (136, 137), phowhorylation (13 DNA interactions (14]) have all A dramatic demonstratio samples is shown with “protein analysis of a mixture of pep degradation (142). Each of the canbeidentified fi'om the mass d Another highly promisin: identification by MALDI of Coc electrophoresis (143-145). This protein spot identification as the protein quantities in 2D gel S] proteins, poor protein recoveries r In additional to the use r analysis of combinatorial librarie etal (146) used MALDI-MS for peptides isolated from support- monoclonal antibody was scree peptides isolated were shown to The capacity for mass lagged behind that of oligopeptid last few years (147, 148), especi 44 (136, 137), phosphorylation (138), glycosylation (139), lipidation (140), and protein- DNA interactions (141) have all been studied by MALDI. A dramatic demonstration of the ability of MALDI to analyze heterogeneous samples is shown with “protein ladder sequencing” which involves the simultaneous analysis of a mixture of peptide/proteins that have undergone a stepwise Edman degradation (142). Each of the fiagments differs from the next by one amino acid and can be identified from the mass difference between successive peaks. Another highly promising technique in the last three years is the unambiguous identification by MALDI of Coomassie-stained protein spots from two dimensional gel electrophoresis (143-145). This technique is superior to Edman microsequencing for protein spot identification as the latter often suffers from problems such as insufficient protein quantities in 2D gel spots, widespread occurrence of N-terminally blocked proteins, poor protein recoveries from gels, and other unknown factors. In additional to the use of MALDI as a sequencing tool, its ability toward the analysis of combinatorial libraries has also been explored. As an early example, Keough et al (146) used MALDI-MS for the rapid sequence determination of biologically active peptides isolated from support-bound combinatorial peptide libraries. An anti-gp120 monoclonal antibody was screened against a hexapeptide library and six of the eight peptides isolated were shown to possess the exact recognition sequence for the antibody. i The capacity for mass spectrometric analysis of oligonucleotides has generally lagged behind that of oligopeptides. However, this position has changed markedly in the last few years (147, 148), especially with the development of new matrices (149, 150). investigation of a 50-mer is now reports of the analysis of much lar It is believed that MAL biological studies. As is often exciting breakthroughs may. in th this suiting. 45 Investigation of a SO—mer is now somewhat routine (151, 152) and there are individual reports of the analysis of much larger RNAs (15 3). It is believed that MALDI-MS will play more and more important roles in biological studies. As is often true in a new and rapidly developing field, the most exciting breakthroughs may, in the end, occur in areas not even anticipated at the time of this writing. IV, References l. Thorgersen, D- F , skowrons Commurr., 60, 616-620(1974§ 2. Barber, M, Bordoli, R S. S (1981). 3 Karas,ir.,andrrrrrenkamp. F 4. Meng C. K, Mann, M». at Spectrometry and Allied TOP‘ 5 Burlingame, A. L, Boyd, R (1996). 6. Carr, S. A, Roberts, G. D, . Materials, C. N. McEwen a (1990), Ch.3, p87. 7- Yarn S. C. B., Grinnell, B W 8- Carr, S. A, and Biemann, K . 9. Creighton, T. E., in Mecham (1992). 10.Hirayama, K., and Akashi, S Edited by Matsuo, T, Caprir Sons Ltd, (1994). 11. Torchinsky, Y. M, Word. (1981). in Surfu. 46 IV. References 1. 10. 11 Thorgersen, D. F ., Skowronski, R. P., and Macfarlane, R. D., Biochem. Biophys. Res. Commun, 60, 616-620(1974). Barber, M., Bordoli, R. S., Sedgwick, R. D., and Tyler, A. N., Chem. Commun., 325 (1981). . Karas, M., and Hillenkamp, F., Anal. Chem, 60, 2299-2301 (1988). Meng, C. K., Mann, M., and Penn, J. B., Proc. 36th ASMS Conference on Mass Spectrometry and Allied Topics, San Francisco, CA, (1988), p771. Burlingame, A. L., Boyd, R. K., and Gaskell, S. J., Anal. Chem, 68, 599R—651R (1996) Carr, S. A., Roberts, G. D., and Hemling, M. E., inMass Spectrometry of Biological Materials, C. N. McEwen and B. S. Larsen (Editors), Marcel Dekker, New York, (1990), Ch.3, p87. Yan, S. C. B., Grinnell, B. W., and Wold, F., Trends Biochem. Sci, 14, 264 (1989). . Carr, S. A., and Biemann, K., Methods Enzymol., 106, 29-58(1984). Creighton, T. E., in Mechanism of Protein Folding, Edited by Pain, R. H., IRL Press, (1992) Hirayama, K., and Akashi, S., in Biological Mass Spectrometry: Present and Future, Edited by Matsuo, T., Caprioli, R. M., Gross, M. L., and Seyama, Y., John Wiley & Sons Ltd, (1994). . Torchinsky, Y. M., in Surfur in Protein; (Metzler, D., Translation Ed), Pergamon, Oxford, (1981). 12. Jocelyn, P. C ., Biochemistry 0_ 13. Jocelyn, P. C, in Methods in W.B.,anf1th. O. W. Inc. (1' 14. Stark G. R.._ and Stein. \‘1'. H. 15. Swoboda G.. and Hasselbac ( 1973). 16. Oliveira, B., and Lamm..M. E 110211111113. Cutmingham B. M., Biochemism'. 7. 1973-191 18. Lundblad, R. L.. in Technique 19.80yer. P. D.. JACS, 76. 4331 30.Kochkina_ V.‘M.. and Torchn (1975). 21- Sakai, H., Anal. Biochem.. 26 22-Ya1nuchi, T., Yamamoto, S.. 23. Gibson, Q. H., J. Biol. Chem. 24' 511111311, G. L., Arch. Biochem 25. Riddles, P. W., Blakeley R (1983). 2 Q 6.G11bert, H. F., Adv. Enzymol 27. H UXtable, R. J ., Biochemistry . 2 - g-Grassetn, D. R. and Mun 9 ay’ 47 12. Jocelyn, P. C., Biochemistry of the SH Group, Academic Press, London, (1974). 13. Jocelyn, P. C., in Methods in Enzymol., Vol. 143, Academic Press, edited by Jakoby, W. B., Griffith, O. W. Inc., (1987). 14. Stark, G. R., and Stein, W. H., J. Biol. Chem, 239, 3755 (1964). 15. Swoboda, G., and Hasselbach, W., Hoppe-Seyler ’s Z. Physio]. Chem, 354, 1611 (1973). 16. Oliveira, B., and Lamm,,M. E., Biochemistry, 10, 26-31 (1971). 17. Gall, W. E., Cunningham, B. A., Waxdal, M. J ., Konigsberg, W. H., and Edelman, G. M., Biochemistry, 7, 1973-1982 (1968). 18. Lundblad, R. L., in Techniques in Protein Modification, CRC Press, (1995). 19. Boyer, P. D., JACS, 76, 4331 (1954). 20. Kochkina, V. M., and Torchinsky Yu, M., Biochem. Biophys. Res. Communs. 63, 392 (1975). 21. Sakai, H., Anal. Biochem., 26, 387 (1968). 22. Yamuchi, T., Yamamoto, S., and Hayaishi, 0., J. Biol. Chem, 250, 7127(1975). 23. Gibson, Q. H., J. Biol. Chem, 248, 1281 (1973). 24. Ellman, G. L., Arch. Biochem. Biophys., 82, 70 (1959). 25. Riddles, P. W., Blakeley, R. L., and Zemer, B., in Methods in Enzymol., 91, 49-60 (1983). 26. Gilbert, H. F., Adv. Enzymol., 63, 69-172 (1990). 27. Huxtable, R. J ., Biochemistry of Sulfitr, Plenum, New York, 1986. ' 28. Grassetti, D. R., and Murray, J. F., Arch. Biochem. Biophys., 119, 41 (1967). 29. Grassetti, 1). R. and Murray-- 30. Stuchbury. T., Shipton. M.. N l, A. L, and Suschitzky. H.. E lllanson, 1.; and Ryden. L.. Prc and Applications. VC H Press. 32. Gruen, L. C.. and Harrap. B. E 33.Gevondyan, .\'.. M., GeVond FEBS Letters. 255(2). 265-26 34.Sun, Y.. Smith. D. L.. and Sh 35. Shipton. M. and Brooklehurs 36. Sun. Y.. and Smith. D. I... An 375318113. Blattler. W. A.. anr ll-Feng, F. Bell. A.. Dumas. Conference on Mass Spectror 3931111, Y., Bauer, M, 1),, Keo Biology, Vol: 61: Protein ar Chapman,J R, Humana Prer 0.ZalllZec,E.J,Gagc,D A a 366(1994) 1Thannhauser, T W. Konish (1984). 42. Thallnhauser, T W McWh ., < 322(1985). 48 29. Grassetti, D. R., and Murray, J. F. J. chromat'ogr. 41, 121 (1969). 30. Stuchbury, T., Shipton, M., Norris, R., Malthouse, J. P. G., Brocklehurst, K., Herbert, J. A. L., and Suschitzky, H., Biochem. J., 151, 417 (1975). 31. Janson, J .; and Ryden, L., Protein Purification: Principles, High Resolution Methods, and Applications, VCH Press, (1993), p252. 32. Gruen, L. C., and Harrap, B. SJ, Anal. Biochem., 42, 377 (1971). 33. Gevondyan, N., M., Gevondyan, V. S., Gavrilyeva, E.,E., and Modyanov, N., N., FEBS Letters, 255(2), 265-268 (1989). 34. Sun, Y., Smith, D. L., and Shoup, R. E., Anal. Biochem., 197, 69-76 (1991). 35. Shipton, M., and Brocklehurst, K., Biochem. J., 167, 799 (1977). 36. Sun, Y., and Smith, D. L., Anal. Biochem., 172, 130-138 (1988). 37. Singh, R., Blattler, W. A., and Collinson, A. R.,, Anal. Biochem., 213, 49-56(1993). 38. Feng, F., Bell, A., Dumas, F., and Konishi, Y., Proceedings of the 38th Annual Conference on Mass Spectrometry and Allied Topics, Tucson, AZ, (1990), p273. 39. Sun, Y., Bauer, M. D., Keough, T. W., and Lacey, M. P., Methods in Molecular Biology, Vol: 61: Protein and Peptide Analysis by Mass Spectrometry, edited by: Chapman, J. R., Humana Press Inc., Totowa, NJ, (1996). 40. Zaluzec, E. J ., Gage, D. A., and Watson, J. T., J. Amer. Soc. Mass Spectrom, 5, 359- 366(1994) 41.'Thannhauser, T. W., Konishi, Y., and Scheraga, H. A., Anal. Biochem., 138, 181 (1984) 42. Thannhauser T. W., McWherter, C. A., and Scheraga, H. A., Anal. Biochem., 149, 322(1985) 43.11irose, M., Takahashi. N" 06 44. Makinen, A. L.. and Nowak. ' 45. Wang, Z. X., Preiss. B.. and 1 46. Male, A. S., and Jones. B. L. 47. Toyo’oka T., and Imai. K.. A 48.103'o’oka T., and Imai. K.. A 49. Sueyoshi. T.. Miyata T.. I' (10166), 97. 1811 (1985). 50. loyo’oka T.. and Imai. K. .4 51Kobayashi, 1.. Rude. 1.. Kar Biochem. (Tokyo). 98. 1017 (? 53.Chin, C. C. Q.. and Wold. F.. 53. Kirley, T. L.,J. Biol. C hem. 54- Kirley. T. L, Anal. Biochem. 55- Pardo, J. P., and Slayman. C. 56' Miyanishj, T., and Borejdo, l 57. Bishop, J. E., Squier, T. C.. (1988). 58.8mmock,E.D.,Yu X C “ 95609566 (1996). 59. J anson, 1.; and Ryden, L. Pr and Applications,VCH Pres s 49 43. Hirose, M., Takahashi, N., Oe, H., and Doi, E., Anal. Biochem., 168, 193-201 (1988). 44. Makinen, A. L., and Nowak, T., J. Biol. Chem, 264, 12148 (1989). 45. Wang, Z. X., Preiss, B., and Tsou, C. L., Biochemistry, 27, 5095 (1988). 46. Mak, A. S., and Jones, B. L., Anal. Biochem., 84, 432 (1978). 47. Toyo’oka, T., and Imai, K., Anal. Chem, 56, 2461-2464 (1984). 48. Toyo’oka, T., and Imai, K., Analyst, 109, 1003 (1984). 49. Sueyoshi, T., Miyata, T., Iwanaga, S., Toyo’oka, T., and Imai, K., J Biochem. (Tokyo), 97, 1811 (1985). 50. Toyo’oka, T., and Imai, K., Anal. Chem, 57, 1931 (1985). 51. Kobayashi, T., Kudo, 1., Karasawa, K., Mizushima, H., Inoue, K., and Nojima, S., J. Biochem. (Tokyo), 98, 1017 (1985). 52. Chin, C. C. Q., and Wold, F ., Anal. Biochem., 214, 128-134 (1993). 53. Kirley, T. L., J. Biol. Chem, 264, 7185-7192 (1989). 54. Kirley, T. L., Anal. Biochem., 180, 231-236 (1989). 55. Pardo, J. P., and Slayman, C. W., J. Biol. Chem, 264, 9373 (1989). 56. Miyanishi, T., and Borejdo, J ., Biochemistry, 28, 1287 (1989). 57. Bishop, J. E., Squier, T. C., Bigelow, D. J., and Inesi, G., Biochemistry, 27, 5233 (1988) 58. Sturrock, E. D., Yu, X. C., Wu, Z., Biemann, K., and Riordan, J. F ., Biochemistry, 35, 9560-9566 (1996). 59. Janson, J .; and Ryden, L., Protein Purification: Principles, High Resolution Methods, and Applications, VCH Press, (1993), p252. 60. Kalef, E, Walfrsh, P. G.‘. and 61Zaluzec, E. J., Gage, D. A. Conference on Mass Spectror 62. Catsimpoolas, -\.. and Wood. 63. Degani, Y., and Patchornik. r‘ 64. Stark, G. R, .lletltotis m 1511:; 65_Biemann, K, in .t-lethods m (1993), p119. 66. Schafier, M. 11.. and Stark, ( (1976). 67. Darbre, A., Practical Protett p149. 68. Papayarmopoulos, 1A., and I 69. Mahboub, 8., Richard, C, Dr (1986). 70. ' Altamirano, M. M., Plumbrid (1992). 71. Tang, C, Yuksel, K. 11. Jac 283. 12.19 (1990). 72. N efsky, B, and Bretscher, A 73. ’ Mmard, P, Desmadril, M B' tochem., 185, 419-423 (198 50 60. Kalef, E.; Walfish, P. G.; and Gitler, 0, Anal. Biochem., 212,325-334( 1993) 61. Zaluzec, E. J .; Gage, D. A., and Watson, J. T., in Proceeding of the 42nd ASMS Conference on Mass Spectrometry and allied Topics, Chicago, (1994), p20. 62. Catsirnpoolas, N., and Wood, J. L., J. Biol. Chem, 241, 1790 (1966). 63. Degani, Y., and Patchornik, A, Biochemistry 13, 1 (1974). 64. Stark, G. R., Methods in Enzymology, vol. 47, 129-132 (1977). 65. Biemann, K., in Methods in Protein Sequence Analysis, Plenum Press, New York, (1993), p119. 66. Schafl'er, M. H., and Stark, G. R., Biochem. Biophys. Res. Commun, 71, 1040-1047 (1976) 67. Darbre, A., Practical Protein Chemistry--A Handbook, John Wiley & Sons, (1985), p149. 68. Papayannopoulos, LA, and Biemann, K., Protein Sci. 1, 278-288 (1992). 69. Mahboub, S., Richard, C., Delacourte, A., and Han, K., Anal. Biochem., 154, 171-182 (1986) 70. Altamirano, M. M., Plumbridge, J. A., and Calcagno, M. L., Biochem., 31, 1153-1158 (1992) 71. Tang, C., Yuksel, K. U., Jacobson, T. M., and Gracy, R. W., Arch. Biochem. Phys, 283, 12-19 (1990). 72. Nefsky, B., and Bretscher, A., Proc. Natl. Acad Sci. USA, 86, 3549-3553 (1989). 73. Minard, P., Desmadril, M., Ballery, N., Perahia, D., and MOuawad, L., Eur. J. Biochem., 185, 419-423 (1989). 74, Sanger, F ., Nature flondonl. l 75. Ryle, A. P., and Sanger. F.. B 76. Ryle, A. P., Sanger, P. Smith 77. Lumper, L., and Zahn. 11.. Ad 78. Spackman. D. 11.. Stein. W. l 79. Marti, T., Rosselet. S. J .. Tit. (1987). 80. Stasnick P. E.. Hermodson. (1984). 31. Haeliner-Gormley. L. Paren‘ 83 (1985). 82' Yazdanpafast R.. Andrews. 1 25072513 (1987). 31. Smith, D. L., and Zhou, Z.. A 84.18180, T., Yoshida, M., Hon Spectrom, 11, 549-556 (198. 8 . 5. Moms, H. R., and Pucci, P. 86I ' Akash1,S.; and Hirayama, K 87. Williamson, R. A Marsto C, n, 88. ' KHShnamurthy, T.; and Haur , , 0 e 51 74. Sanger, F ., Nature (Iondon), 171, 1025-1026 (1953). 75. Ryle, A. P., and Sanger, F., Biochem. J., 60, 535-540 (1955). 76. Ryle, A. P., Sanger, F., Smith, L. F., and Kitai, R., Biochem. J., 60, 541-556 (1955). 77. Lumper, L., and Zahn, H., Adv. Enzymol., 27, 199 (1965). 78. Spackman, D. H., Stein, W. H., and Moore, S., J. Biol. Chem, 235, 648-659 (1960). 79. Marti, T., Rosselet, S. J., Titani, K., and Walsh, K. A., Biochemistry, 26, 8099-8109 (1987). 80. Staswick, P. E., Hermodson, M. A., and Nielsen, N. C., J. Biol. Chem, 259, 13431 (1984). 81. Haeffner-Gormley, L., Parente, L., and Wetlaufer, D. B., Int. J. Pept. Protein Res, 26, 83 (1985). 82. Yazdanparast, R., Andrews, P. C., Smith, D. L., and Dixon, J. E., J. Biol. Chem, 262, 2507-2513 (1987). 83. Smith, D. L., and Zhou, Z., Methods in Enzymol., Vol. 193, 374-389 (1990). 84. Takao, T., Yoshida, M., Hong, Y. M., Aimoto, S., and Shimonishi, Y., Biomed. Mass. Spectrom., 11, 549-556 (1984). 85. Morris, H. R., and Pucci, P., Biochem. Biophys. Res. Commun. 126, 1122 (1985). 86. Akashi, S., and Hirayama, K., Biomed. Environ. Mass Spectrom., 15,541-546 (1988). 87. Williamson, R. A., Marston, F. A.; and Angal, S., Biochem. J., 268, 267-274 (1990). 38. Krishnamurthy, T.; and Hauer, C. C., Biol. Mass Spectrom. 23, 719-726 (1994). 89. Bauer, M.; Sun, Y.; Degenhardt, C.; and Kozikowski, B., J Protein Chem. 12, 759- 764(1993) 90. Patterson. 8.; 811d Katta. V" A 9(Crimmins. D. L., saylor. M" 36111995). 92.Bean,M. P: and Carr. 5. A... 93.Chung, D.. Sairom. M. R.. z (1975). 94.lhannhauser. T. W. Konishi 115-119(1987). 95..Zhou, Z.. and Smith. D. L. J. 96861115.]. 1.. Bourell. J. 11.. C 111-". Griffin. P. R.. Rinder Spectrom., 19, 655-664 (1991 97.1011, J. W., Elzinga M., and 93. Serensen. H. H.. Thomsen. J 11713720 (1990). 99. Nokihara, K, Morita, N., Ya (1992). 1011. Zhang, 1)., and Liang, 8. J. 10 ' 1.Knshnamurthy, T., Hauer Biol. Mass Spectrom., 23 71 102. Tang, Y., and Selsted, M. E 52 90. Patterson, S.; and Katta, V., Anal. Chem. 66, 3727-3732 (1994). 91. Crimmins, D. L., Saylor, M., Rush, J., and Thoma, R. S., Anal. Biochem., 226, 355- 361(1995) 92. Bean, M. F .; and Carr, S. A., Anal. Biochem., 201, 216-226 (1992). 93. Chung, D., Sairom, M. R., and Li, C. H., Int. J. Peptide Protein Res, 7, 487-493 (1975) 94. Thannhauser, T. W., Konishi, Y., and Scheraga, H. A., Methods Enzymol., Vol: 143, 115-119 (1987). 95. Zhou, Z., and Smith, D. L., J. Protein Chem, 9, 523-532 (1990). 96. Stults, J. T., Bourell, J. H., Canova-Davis, E., Ling, V. T., Laramee, G. R., Winslow, J. W., Griffin, P. R., Rinderknecht, E., and Vandlen, R. L., Biomed. Environ. Mass Spectrom., 19, 655-664 (1990). 97. Fox, J. W., Elzinga, M., and Tu, A. T., Biochemistry, 18, 678-684 (1979). 98. Serensen, H. H., Thomsen, J ., and Bayne, S., Biomed. And Environ. Mass Spectrom., 19, 713-720 (1990). 99. Nokihara, K., Morita, N., Yamaguchi, M., and Watanabe, M., Anal. Lett, 25, 513-533 (1992) 100. Zhang, D., and Liang, S., J. Protein Chem, 12, 735-740 (1993). 101. Krishnamurthy, T., Hauer, C. R., Prabhakaran, M., Freedy, J. G., and Hayashi, K., Biol. Mass Spectrom., 23, 719-726 (1994). 102. Tang, Y., and Selsted, M. E., J. Biol. Chem, 268, 6649-6653 (1993). 103.Gray, W. R.. Luque. F. A.. ( Reyes, A., Alford. 1.. Mclnl Biochemistry, 23. 2796 4 1984 104. Gray, W. R., Protein Sci. 2. 105. Gray W. R., Protein Sci, 2. 106.15h1'bashi, J., Kataoka, 11.. Mizoguchi, A., Ishizaki. 11.. a 107. Posthumus. M. A., Kistema Chem, 50. 985-991 (1978 ). 108.Karas. M., Bachmnn. D.. l Processes. 78. 53-68 (1987‘). 109. Ehring, H.. Karas. M.. an (1992). 110. Fitzgerald, M. C., Parr, G. 1 lll.Beavis, R. C., and Chait, B. 112. Strupat, K, Karas, M., ant 111, 89-102 (1991). 113.Beavis, R. C., Chaudhary (1992). 114. Juhasz, P., Costello, C. E.. 409 (1993). 115. Chait,B.T.; and Kent 8 E 53 103. Gray, W. R., Luque, F. A., Galyean, R., Atherton, E., Sheppard, R. 0, Stone, B. L., Reyes, A., Alford, J ., McIntosh, M., Olivera, B. M., Cruz, L. J ., and Rivier, J ., Biochemistry, 23, 2796 (1984). 104. Gray, W. R., Protein Sci., 2, 1732-1748 (1993). 105. Gray W. R., Protein Sci., 2, 1749-1755 (1993). 106.1shibashi, J., Kataoka, H., Isogai, A., Kawakami, A., Saegusa, H., Yagi, Y., Mizoguchi, A., Ishizaki, H., and Suzuki, A., Biochemistry, 33, 5912-5919 (1994). 107. Posthumus, M. A., Kistemaker, P. G., Meuzelaar, H. L. C., and Brauw, M. C., Anal. Chem, 50, 985—991 (1978). 108. Karas, M., Bachmann, D., Bahr, U., and Hillenkamp, F., Int. J. Mass Spectrom. Ion Processes, 78, 53-68 (1987). 109. Ehring, H., Karas, M., and Hillenkamp, F., Org. Mass Spectrom., 27, 472-480 (1992). 110. Fitzgerald, M. C., Parr, G. R., and Smith, L. M., Anal. Chem, 65, 3204-3211 (1993). 111. Beavis, R. C., and Chait, B. T., Rapid Commun. Mass Spectrom., 3, 432-435 (1989). 112. Strupat, K., Karas, M., and Hillenkamp, F., Int. J. Mass Spectrom. Ion Processes, 111, 89-102 (1991). 113. Beavis,.R. C., Chaudhary, T., and Chait, B.T., Org. Mass Spectrom., 27, 156-158 (1992). 114. Juhasz, P., Costello, (3. 13., and Biemann, K., J. Am. Soc. Mass Spectrom, 4. 399- 409 (1993). 115. Chait, B. T.; and Kent. S. B. H., Science, 257, 1885-1894 (1992). 116. Siuedak, G.. Proc. Natl. Acad. 117. Busch, K. L., J. Mass Spectror 118.Zaluzec. E. J.; Gage. D. As. a 6109-123 (1995). 119.Cohen. S. L., and Chait. B. T. 120. Preston, L. M.. Murray. K. 1- 55011993). 121.Vonn, 0.. and Roepstorff. P. 133.1-"mm. 0.. and Mann. M.. .1. . 1218433115. R. C.. and Chait. B. 124.11111enkamp. F.. Karas. M.. 1198A(1991). 125. Kaufmann. R.; Kirsch, D.; 355-385 (1994). 126. Kallfmann, R.; Spengler, B 7, 902-910 (1993). 127. Brown, R. S., and Lennon 128.c61by,s.M..Kin8 T B 868 (1994). 54 116. Siuzdak, G., Proc. Natl. Acad. Sci. USA, 91, 11290-11297 (1994). 117. Busch, K. L., J. Mass Spectrom., 30, 233-240 (1995). 118. Zaluzec, E. J .; Gage, D. A.; and Watson, J. T., Protein Expression and Purification, 6, 109-123 (1995). 119. Cohen, S. L., and Chait, B. T., Anal. Chem, 68, 31-37 (1996). 120. Preston, L. M., Murray, K. K., and Russell, D. H., Biol. Mass Spectrom., 22, 544- 550(1993) ' 121. Vorm, 0., and Roepstorfi, P., and Mann, M., Anal. Chem, 66, 3281-3287 (1994). 122. Vorm, 0., and Mann, M., J. Am. Soc. Mass Spectrom., 5, 955-958 (1994). 123. Beavis, R. C., and Chait, B. T., Proc. Natl. Acad. Sci. USA, 87, 6873-6877 (1990). 124. Hillenkamp, F., Karas, M., Beavis, R. C., and Chait, B. T., Anal. Chem, 63, 1193- 1198A (1991). 125. Kaufmann, R.; Kirsch, D.; and Spengler, B., Int. J. Mass Spectrom. Ion Proc. 131, 355-385 (1994). . 126. Kaufmann, R.; Spengler, B.; and Lutzenkirchen., Rapid Commun. Mass Spectrom., 7, 902-910 (1993). 127. Brown, R. S., and Lennon, J. J., Anal. Chem, 67, 1998-2003 (1995). 128. Colby, S. M., King, T. B., and Reilly, J. P., Rapid Commun. Mass Spectrom., 8, 865- 868(1994) 129. Vestal, M. L., Juhasz, P., and Martin, S. A., Rapid Commun. Mass Spectrom., 9, 1044-1050(1995) 130. Whittal, R. M., and Li, L., Anal. Chem, 67, 1950-1954 (1995). 15111111581..Rosk9~ M- T" 3” A.,Anal. Chem, 68. 941'946 1 132Carr. S. A., in Mass Spectre Carr. S. A., Eds; Humana PICS 133.Billeci,1. M.. and Stults. J. '1 114. Goldberg, D. E., Slater. A. Henderson G. B.. J. Exp. Mea 135. Specht B.. Oudenampsen-K and Spener, 1.. J. Biotechnol. 136.Robertson J. G.. Adams. ( V'illafranca. J. J., Biocherrristr 137. Chang, 1. Y., Schindler. P. (1995). 138.1180, P-C., Leykam, J .. Anc 219, 9-20 (1994). 139‘ Burlingame, A. L., Curr. Q 140.Wilcox, M. D., Schey, K Biophys. Res. Commun. 212 14 1. Zhao, Y., and Chait, B. T. 142.Chait,B.T. Wang R B , , ,, e2 143. Patterson, S. D 144. ' Erdjument,~Bromage H ' , and Aeber Protein Sci., 3, 2435-2446 ( 55 131. Juhasz, P., Roskey, M. T., Smirnov, I. P., Haff, L. A., Vestal, M. L., and Martin, S. A., Anal. Chem, 68, 941-946 ( 1996). 132. Carr, S. A., in Mass Spectrometry in the biological sciences, Burlingame, A. L., Carr, S. A., Eds; Humana Press: Totowa, NJ, (1996) 133. Billeci, T. M., and Stults, J. T., Anal. Chem, 65, 1709-1716 (1993). 134. Goldberg, D. E., Slater, A. F. G., Beavis, B. T., Chait, B. T., Cerami, A., and Henderson, G. B., J. Exp. Med, 173, 961 (1991). 135. Specht, B., Oudenampsen-Kruger, E., Ingendoh, A., Hillenkamp, F., Leaius, A. G., and Spener, F., J. Biotechnol., 33, 259-269 (1994). 136. Robertson, J. G., Adams, G. W., Medzihradszky, K. F., Burlingame, A. L., and Villafranca, J. J ., Biochemistry, 33, 11563-11575 (1994). 137. Chang, J. Y., Schindler, P., and Chatrenet, B., J Biol. Chem, 270, 11992-11997 (1995). 138. Liao, P-C., Leykam, J ., Andrew, P. C., Gage, D. A., and Allison, J ., Anal. Biochem., 219, 9-20 (1994). 139. Burlingame, A. L., Curr. Opin. Biotechnol., 7, 4-10 (1996). 140. Wilcox, M. D., Schey, K. L., Busman, M., and Hil'debrandt, J. 1)., Biochem. Biophys. Res. Commun., 212, 367-374 (1995). 141. Zhao, Y., and Chait, B. T., Anal. Chem, 66, 3723-3726 (1994). 142. Chait, B. T., Wang, R., Beavis, R. C., and Kent, S. B., Science, 262, 89-92 (1993). 143. Patterson, S. D., and Aebersold, R., Electrophoresis, 16, 1791-1814 (1995). 144. Erdjument,-Bromage, H., Lui, M., Sabatini, D. M., Snyder, D. H., and Tempst, P., Protein Sci., 3, 2435-2446 ( 1994). 145.Clauser, K R., Hall. S. C.. Sr Epstein. L. B., and Burlingan (1995). 146Youngquist. R. S.. Fuentes Commun. Mass Spectrom.. 8. ' l47.1\’ordhoff, E.. Kirpeltar 1.. (1996). 148.11mbach. P. A.. Cram P. F. 102(1995). 149.1itzgera1d. M. C .. Parr. C (1993). 150. Wu, K 1.. Shaler. T. A.. anc 151.Nordhofi°. E.. Kirpekar. F. Kristiansen, K, Roepstorff. 1 152.Nordhoff, 13., Karas, M., Lem) A., Muth, J ., Meier. (1995). 153.141.6614, F., Nordhoff, E. Karas, M., and Hillenkamp 56 145. Clauser, K. R., Hall, S. C., Smith, D. M., Webb, J. W., Andrews, L. E., Tran, H. M., Epstein, L. B., and Burlingame, A. L., Proc. Natl. Acad Sci. USA. 92, 5072-5076 (1995). 146.Youngquist, R. S., Fuentes, G. R., and Lacey, M. P., and Keough, T., Rapid Commun. Mass Spectrom., 8, 77-81 (1994). 147. Nordhoff, E., Kirpekar, F ., and Roepstorff, P., Mass Spectrom. Rev., 15, 67-138 (1996). 148. Limbach, P. A., Crain, P. F., and McCloskey, J. A., Curr. Opin. Biotechnol., 6, 96- 102 (1995). 149. Fitzgerald, M. C., Parr, G. R., and Smith, L. M., Anal. Chem, 65, 3204-3211 (1993) 150. Wu, K. J ., Shaler, T. A., and Becker, C. H., Anal. Chem, 66, 1637-1645 (1994). 151.Nordhoff, E., Kirpekar, F., Karas, M., Cramer, R., Hahner, S., Hillenkamp, F., Kristiansen, K., Roepstorff, P., Lezius, A., Nucleic Acids Res., 22, 2460-2465 (1994). 152. Nordhoff, E., Karas, M., Cramer, R., Hahner, S., Hillenkamp, F., Kirpekar, F ., Lezius, A., Muth, J., Meier, C., and Engles, J. W., J Mass Spectrom., 30, 99-112 (1995) 153. Kirpekar, F ., Nordhoff, E., Kristiansen, K., Roepstorff, P., Lezius, A., Hahner, S., Karas, M., and Hillenkamp, F., Nucleic Acids Res., 22, 3866-3870 (1994). ASTRATEGY TO LOCA' BY SPECIFIC CHEMICA' ASSISTED LASER DESOl MA! 1 Introduction Cysteine sulfhydryl gror using its free sulfhydryl (SH) 8 cysteine proteases, as the chelati reshufiling enzymes. Pinpointi: becomes an important strategy ft 1418e majority of new protein se cloned cDNA (or, in the case flagments), differentiating betwc (disulfide bonds) is still cumbers conventional techniques, cannot Classical approaches for modification of free sulfhydryls d - . envatlzed protein chemically r b . 9 LC fractionation of the p1 11. 2 ). Recently, peptide mappi CHAPTER 2 A STRATEGY TO LOCATE CYSTEINE RESIDUES'IN PROTEINS BY SPECIFIC CHEMICAL CLEAVAGE FOLLOWED BY MATRIX- ASSISTED LASER DESORPTION/IONIZATION TIME-OF—F LIGHT MASS SPECTROMETRY I. Introduction Cysteine sulfliydryl group (thiol) contributes to protein biological functions by using its free sulfhydryl (-SH) group in the active site for enzyme. catalysis such as in cysteine proteases, as the, chelating site for metal ions, and as the active site of disulfide reshufiling enzymes. Pinpointing the number and location of cysteine residues thus becomes an important strategy for determining protein structure. While it is true that the large majority of new protein sequences are now deduced from nucleotide sequences of cloned cDNA (or, in the case of prokaryotic proteins, from cloned genomic DNA fragments), differentiating between free cysteine residues (sulfhydryl groups) and cystines (disulfide bonds) is still cumbersome and tedious, because the cDNA technique, like other conventional techniques, cannot identify posttranslational modifications. Classical approaches for localizing protein free sulfhydryl groups usually involve modification of free sulfhydryls before and after reduction of a protein, degradation of the derivatized protein chemically or enzymatically into smaller peptide fragments, followed by HPLC fractionation of the peptides and Edman degradation of the derivatized peptides (1, 2). Recently, peptide mapping by mass spectrometry of proteolytic digests before and 57 afterHPLC fractionation also has the sulfhydryl derivatization und disulfide bond exchange is not ar becleaved at both sides of cys cysteine. Multiple-step enzymat case which is tedious and requil dcrivatizing reagents have to be and to avoid the retention time ( a1113411112111 sequencer. Althoug 1110111310813th of sulfhydryl ; feasibility remains to be tested. 11. Chemical Cleavage at CY“ Cleavage at cysteine re: cyanolysis of disulfide bonds However, because of the severa and the elimination of thiocyan not acceptable to protein chemis Jacobson et al (7) s SPeciflcally cyanylates cysteine Side of the cyanylated cystein WHO-terminal peptide and a : (Figur 62.1). f a protein conta: 58 after HPLC fractionation also has been reported (3-5). However, this procedure requires the sulflaydryl derivatization under a carefully controlled conditions so that sulfliydryl- disulfide bond exchange is not an issue. Secondly, this approach requires that a protein be cleaved at both sides of cysteine residues to give peptides that contain only one cysteine. Multiple-step enzymatic or chemical digestions usually are performed in this case which is tedious and requires a sample size on the order of nanomoles. Thirdly, derivatizing reagents have to be chosen to facilitate HPLC separation with UV detection and to avoid the retention time overlap of derivatized cysteine and other amino acids on an Edman sequencer. Although other approaches such as methods based on affinity chromatography of sulfhydryl groups have also been proposed (see chapter 1), their feasibility remains to be tested. 11. Chemical Cleavage at Cysteine Residues Cleavage at cysteine residues of peptide chains under alkaline conditions after cyanolysis of disulfide bonds was first observed by Catsimpoolas and Wood (6). However, because of the several side reactions, such as the reversibility of the cyanolysis and the elimination of thiocyanylate, the yields were low and the cleavage reaction was not acceptable to protein chemists. Jacobson et al (7) showed that 2-nitro-5-thiocyanobenzoic acid (NTCB) specifically cyanylates cysteine thiols. Subsequent cleavage occurs on the N-terminal side of the cyanylated cysteinyl residue under mildly alkaline conditions to form an amino-terminal peptide and a series of 2-iminothiazolidine-4-carboxylyl (ITC) peptides (Figure 2.1). If a protein contains 11 cysteine residues, the cleavage reaction results in the 11,.\"—' ”C — (A1 C59“: (3118-91 NE 11 11,18- ~ 4 (B) Cleavag pH 9, 37°C //0 H3N+ ‘ ’ ‘T C\ 11H H3N+~ 4130-1411 -C - Figure2l. Reaction betwee , (A) Cyanylation ar. cleavage reaction under alkal 59 sH 0 CH 0 11 1 2 11 //0 H3N+—- -—C —NH—CH —C —NH-—- ——c\ _ o- HOOC . \ (A) Cyanylation NTCB pH8-9; rt, 15' ’ 02N@-SCN V NEC -S O CH O 11 I 2 11 //0 H3N+ — - ——C —NH—-CH —C —NH —- —C \ _ \ t 0 0H0 (B) Cleavage _ pH 9, 37°C, >16hr 0H V O HN S ' // O H3N+——-—C\ + Y I II //0 OH HN C -NH - - —- \ . O' ('sz /O H3N+—_ - --CO—NH -—C —_CO --NH-- - —C< (side reaction: B-elimination) 0‘ Figure 2.1. Reaction between cysteine residue and 2-nitro-5-thiocyanobenzoic acid (NTCB), (A) Cyanylation and (B) Cleavage reaction. B-Elimination competes with cleavage reaction under alkaline conditions. fi formation of n+1 peptide fragme and location of cysteine residues reaction can come to completion proteins tested, Degani and Pat alkaline conditions, competes wi peptides showed marked variant depending on the structural prop ems; of NTCB and a low tom the side reaction of displacemen thiol groups (8). Other reagents for cyan) (9) and by Bnocklehurst et al (1 111GB, but their utility needs to Jacobson et al (7) pro mechanism implicates that hyt influence of rising pH on the rat ofOH' is shown in Figure 2.2. ' .out the nucleophilic attack (th reaction probably proceeds only This generates a much more I cl’CIization and cleavage, withor 60 formation of n+1 peptide fragments, mass analysis of the fragments indicates the number and location of cysteine residues. Although the original paper claimed that the cleavage reaction can come to completion with little side reactions for most of the peptides and proteins tested, Degani and Patchornik (8) found B-elimination, occurring also under alkaline conditions, competes with the cleavage as an adverse reaction. The cyanylated peptides showed marked variations in their tendency to undergo B-elimination reaction depending on the structural properties of the respective peptides. Experimentally, a large excess of NTCB and a low total concentration of thiol groups must be applied to avoid the side reaction of displacement of CN- from S-cyanocysteine residues by the unreacted thiol groups (8). Other reagents for cyanylation of SH groups have been prepared by Wakselman (9) and by Brocklehurst et al (10, 11). These reagents may offer some advantages over NTCB, but their utility needs to be further examined. Jacobson et al (7) proposed a mechanism for the cleavage reaction. This mechanism implicates that hydroxide ion catalyzes the cleavage, as shown by the influence of rising pH on the rate of cleavage reaction. A possible explanation of the role of OH‘ is shown in Figure 2.2. The amide nitrogen is apparently too weak a base to carry .out the nucleophilic attack (the pKa of a protonated amide is -4.0 or less), and the reaction probably proceeds only after attack of OH‘ on the carbonyl carbon of the amide. This generates a much more basic nitrogen, which then can participate in concerted cyclization and cleavage, without formation of an acyliminothiazolidine intermediate. C l'IJ‘Iq+ - ' - H ‘ \ H3N*—- — x} + ._ . .— H,N c\ Figure 2.2. A me 61 NEC -S O CH Q q I 2 II // H3N+ —- — —NH—CH —C —NH—- —c\ 0' OH- v HEB N\\ /S\ O G O ,, / /NH——‘—CO—NH— —C\ H3N+— “(I3 \A 0' OH HN s H3N+ — - ———C/;) + Y (ll) //0 \OH HN C —NH—- ——C\ 0' Figure 2.2. A mechanism for base catalyzed cleavage reaction. fi One featme of the reactir related to the easy displacement nitrothiophenolate. The cyanyla are unreactive to NTCB. The sulflrydryls can thus be achieve reaction without prior reduction While potentially a very used for sequence determination by the iminothiamlidineearboxy to remove the ITC group from cleavage by NTCB has been to ntunber of side reactions that cc the reversibility of the cyanolysi Because cysteines are 1 usually produces large fragmen‘ polyacrylamide gel electrophore using this approach are often co Recently, Papayannopr ,\ Spectrometry to sequence the l isolated from Sarcophaga bul. could be used to sequence Pet 62 One feature of the reaction is that the reactivity of NTCB toward a thiol group is related to the easy displacement by the surfur nucleophile of the good leaving group p- nitrothiophenolate. The cyanylation is selective to free sulfhydryl groups, disulfide bonds are unreactive to NTCB. The selective cyanylation and subsequent cleavage of free sulfhydryls can thus be achieved in the presence of disulfide bonds by carrying out the reaction without prior reduction of the protein (7). While potentially a very useful method, this cleavage reaction has seldom been used for sequence determination because all butthe N-terminal peptide becomes blocked by the irninothiazolidinecarboxylyl (ITC) group. Thus far, there is no convenient method to remove the ITC group from the cleavage products (12). Another reason that Cys cleavage by NTCB has been used infrequently is that the reaction conditions cause a number of side reactions that compete with the desired cleavage reaction. These include the reversibility of the cyanolysis reaction and the B-elimination of thiocyanylate (8). Because cysteines are relatively scarce in proteins, cleavage at these residues usually produces large fragments which can be mass mapped by sodium dodecyl sulfate- polyacrylamide gel electrophoresis (SDS-PAGE) (13-32). However, peptide assignments using this approach are often complicated by its poor mass accuracy ( error > 5%). Recently, Papayannopoulos and Biemann (34) used CID tandem mass spectrometry to sequence the NTCB cleavage reaction products of a protease inhibitor isolated from Sarcophaga bullata. Their work demonstrated that mass spectrometry could be used to sequence peptides from the NTCB cleavage reaction in. spite of the fi blocked N-terminus. However. and hence cysteine-rich proteins, We developed a simple a location of both cystine and fro chemical cleavage reaction desc peptides by MALDI-TOF MS (1 ordenatrn'ed protein is first allo chain at free sulfliydryl sites. disulfide bonds by tris(2-carbc reagent for this purpose since, 1: thiol groups that may react wit thermltiug peptide fragments deduce the number and locatic the same cleavage reaction aft on the number and location of of the protein. The selectiy sensitivity, speed, and mass a approach. We also demonstrr to remove excess reagents an spectra. Experimental condi sulflrydryl cyanylation, and Zetabind membranes. 63 blocked N-terminus. However, tandem MS is practically limited to low-mass peptides, and hence cysteine-rich proteins, because of the effective mass limit of CID (< 3000 Da). We developed a simple and sensitive methodology to recognize the number and location of both cystine and free sulfhydryls in peptides and proteins using the specific chemical cleavage reaction described above, followed by mass mapping of the resulting peptides by MALDI-TOF MS (35). In these analyses, as shown in Figure 2.3, a peptide or denatured protein is first allowed to react with NTCB to selectively cleave the peptide chain at free sulfhydryl sites. The reaction mixture is then subjected to reduction of disulfide bonds. by tris(2-carboxyethyl)phosphine hydrochloride (TCEP) (36), a useful reagent for this purpose since, unlike other common reducing reagents, it does not contain thiol groups that may react with NTCB under the reaction conditions. Mass mapping of the resulting peptide fragments by MALDI-MS provides information from which one can deduce the number and location of the free sulfhydryls. Parallel experiments involving the same cleavage reaction after reduction of disulfide bonds provide mass spectral data on the number and location of total cysteines, which can be used to confirm the sequence of the protein. The selectivity of the NT CB cleavage reaction combined with the sensitivity, speed, and mass accuracy of MALDI-TOF MS are attractive features of this approach. We also demonstrated the reactions performed on a Zetabind membrane used to remove excess reagents and other salts and buffers in order to improve the MALDI spectra. Experimental conditions are described for protein disulfide bond reduction, sulflrydryl cyanylation, and cleavage reactions performed either in solution or on Zetabind membranes. / HS2t \ coo Cyanylation & Cleavage ”K? )H,‘ 39 ZOK—i / 40 \jii/COO- Reduction by TCEP 19 NH + 10 3 2 KSH —/ 3. 0 ¥_SH _f_/ 30 MALDI Analysis CDAP/FCEP approach to locate free sulfhydrst) Figure 2.3. Chemical modi r\ HS 20 C00‘ Cyanylation & Cleavage 19 NH; K180 _/ Reduction by TCEP V 19 10 NH; 20 \SH/ 39 30 MALDI Analysis V CDAP/TCEP approach to locate free‘ sulfhydryl(s) Reduction by TCEP Cyanylation & Cleavage ‘l NH3+ 20 V29 30 39 V MALDI Analysis v TCEP/CDAP approach to identify total cysteines Figure 2.3. Chemical modification to identify free sulflrydryls and total cysteines. ll]. Experimental Section MALDI-TOF MS MALDI mass spectra w Vestec Corp, Houston, TX) t nitrogen laser (337 nm, 3-ns pul 26W. Data were acquired in ‘ recorder with Z-ns resolution. or internal calibration using star (1 pmol of each standard) obt experiments were performed to C0,, leaukee, WI) as the nu (vlv) solution of acetonitrile/ar peptide or protein samples ant allowed to air dry before beir immobilized on an inert Zetal mounted membrane containing matrix/protein solution was a acetonitrile/O. 1% TFA solutio introduction into the mass spec III. Experimental Section MALDI-TOF MS MALDI mass spectra were obtained on a Vestec LaserTec Research (VT 2000, Vestec Corp, Houston, TX) time—of-flight (TOF) mass spectrometer equipped with a nitrogen laser (337 nm, 3-ns pulse). The accelerating voltage in the ion source was set to 26 kV. Data were acquired in the linear mode of operation, using a SOD-MHz transient recorder with 2-ns resolution. Time-to-mass conversion was achieved by either external or internal calibration using standards of bradykinin (m/z 1061.2) and insulin (m/z 5734.5) (1 pmol of each standard) obtained from Sigma Chemical Co. (St. Louis, MO). All experiments were performed using or-cyano-4-hydroxycinnamic acid (Aldrich Chemical Co., Milwaukee, WI) as the matrix. Saturated matrix solutions were prepared in a 50% (v/v) solution of acetonitrile/aqueous 0.1% TFA and mixed in equal volumes with the peptide or protein samples and applied to a stainless-steel probe tip. The mixture was allowed to air dry before being introduced to the mass spectrometer. For a protein immobilized on an inert Zetabind membrane, matrix solution was applied to a probe- mounted membrane containing the adsorbed protein or reaction products (37). The matrix/protein solution was allowed to dry under the saturated vapor of the 50% acetonitrile/O. 1% TF A solution and cocrystallized on the membrane surface prior to introduction into the mass spectrometer. Chemicals Tris(2-carboxyethyl)phos Rockford, IL. Guanidine hy Biochemicals (Indianapolis, MHRQEAVDCLKKFNARRKI peptides and proteins, 2—nitro-5- pruchased from Sigma and use NTCB was prepared in eitl grnnidine-HCl solution and th aqueous solution was prepared base, and stored under N2 in -1 suitable concentration before us Analysis of Free sulfhydryls l Peptide or protein samp M tris-HCl buffer (pH 8.0) Cyanylation of the free sulth: excess of NTCB (over free sul: then raised to 9.0 by adding protein was accomplished afie bonds was performed at 37°C Chemicals Tris(2-carboxyethy1)phosphine hydrochloride. (TCEP) was purchased from Pierce, Rockford, IL. Guanidine hydrochloride was a product of Boehringer-Mannheim Biochemicals (Indianapolis, IN). Peptides DRVYIHPCHLLYYS and MHRQEAVDCLKKFNARRKLKGA were purchased from Bachem, California. Other peptides and proteins, 2-nitro-5 -thiocyanobenzoic acid (NTCB), and Trizma buffer, were purchased from Sigma and used without further purification. The 0.01 M solution of NTCB was prepared in either 0.1 M tris-HCl buffer or the buffer made in 6 M guanidine-HCI solution and the pH of the solution was adjusted to pH 8.0. TCEP aqueous solution was prepared as 0.1 M stock solution and adjusted to pH 8.0 with tris base, and stored under N2 in -20°C freezer. The stock solution was further diluted to a suitable concentration before use. Analysis of Free sulfhydryls for Proteins in Solution (NTCB/TCEP procedure) Peptide or protein samples (1 nmol to 100 pmol) were solubilized in 2~5 pl of 0.1 M tris-HCl buffer (pH 8.0) containing 6 N guanidine-HCI as denaturing agent. Cyanylation of the free sulfliydryl groups was achieved by adding a 5-10 fold molar excess of NTCB (over free sulfhydryls) and reacting at 37°C for 30 minutes. The pH was then raised to 9.0 by adding 3.0 M tris base. Cleavage of the cyanylated peptide or protein was accomplished after incubation at 37°C for 16 hours. Reduction of disulfide bonds was performed at 37°C for 30 minutes by providing a 5-10 fold molar excess of TCEP reducing reagent (over volmne of a 50% (v/v) acetonitn' Confirmation of Total Cystein The above protein sampl TCEP (over disulfide bonds) at protein. Then a 5-10 fold mol modify the total cysteines (nati with3.0Mtrisbaseand the m' dihrted 50—fold with a 50% (v/v MS. Microscale Analysis of Proteir Figure 2.4 illustrates a treatment and processing of membrane (OAS-um pore siz Products Inc, Denver, CO) war A sample containing 10 tris-HCl buffered (pH 8.0) 6 molar excess of NTCB soluti The membrane-attached prob water that served to prevent 67 TCEP reducing reagent (over disulfide bonds). Samples were diluted with a 50-fold volume of a 50% (v/v) acetonitrile/O. 1% TF A solution prior to analysis by MALDI-MS. Confirmation of Total Cysteines (T CEP/NT CB procedure) The above protein samples were allowed to react with a 5-10 fold molar excess of TCEP (over disulfide bonds) at 37°C for 30 minutes to reduce the disulfide bonds in the protein. Then a 5-10 fold molar excess of NTCB (over total cysteines) was applied to modify the total cysteines (native and nascent sulfliydryls). The pH was adjusted to 9.0 with 3.0 M tris base and the mixture was incubated at 37°C for 16 hours. The sample was diluted 50-fold with a 50% (v/v) acetonitrile/O. 1% TFA solution for analysis by MALDI- MS. Microscale Analysis of Proteins Immobilized on an Inert Membrane Figure 2.4 illustrates a schematic flowchart for various protocols of chemical treatment and processing of the sample prior to analysis by MALDI. A Zetabind membrane (OAS-um pore size and SO-um thickness, purchased fiom Life Science Products Inc., Denver, CO) was fixed to a probe tip as described previously (3 7). A sample containing 10 to 15 pmol of protein (0.5111 volume, dissolved in 0.05 M tris-HCl buffered (pH 8.0) 6 N guanidine-HCI) was applied to the membrane. A SO-fold molar excess of NTCB solution was then added to the protein on the membrane surface. The membrane-attached probe tip was put in a closed vessel containing a few drops of water that served to prevent drying of the reaction mixture droplet on the membrane surface. The vessel was incuba h'l'CB. The reaction mixture immersion in 0.01% NH.Ol-l f other salts. Afier air drying, 1.0 ul HCI solution was applied to th for 16 hours to promote cleav SO-fold molar excess (~l ul, applied to the membrane to 37"C for another 30 minutes. io0.1% TFA aqueous solution A 1.5-pl aliquot of 8 acid in a 50% acetonitrile/O. l ‘M to resolubilize the peptides. period) in a vessel saturated solution so that the peptide: effectively resolubilized fror subsequently cocrystallized ‘ analyzed directly by MALDI. Parallel experiments membrane to confirm the to 68 surface. The vessel was incubated at 37°C for 30 minutes to promote cyanylation with NTCB. The reaction mixture was allowed to dry on the membrane and then washed by immersion in 0.01% NH4OH for 30 seconds to remove the excess NTCB reagents and other salts. After air drying, 1.0 1.11 of 0.05 M sodium borate-buffered (pH 9.0) 6 N guanidine- HCl solution was applied to the membrane. The probe tip was incubated again at 37°C for 16 hours to promote cleavage in a closed vessel containing a few drops of water. A SO-fold molar excess (~1 pl, in this case) of TCEP (over. disulfide bonds) was then applied to the membrane to reduce the disulfide bonds. The probe tip was incubated at 37°C for another 30 minutes. After air drying, the membrane was washed by immersion in 0.1% TFA aqueous solution for 30 seconds to remove the reagents and buffer salts. A 1.5-pl aliquot of a saturated matrix solution of or-cyano-4-hydroxycinnamic acid in a 50% acetonitrile/O.1% TFA aqueous solution was applied to the dried membrane to resolubilize the peptides. The solution was allowed to dry slowly (over one hour period) in a vessel saturated by the vapor of the 50% acetonitrile/0.1% TFA aqueous solution so that the peptides and proteins immobilized on the membrane could be effectively resolubilized from the membrane surface by the matrix solution and subsequently cocrystallized with the matrix. The resulting peptide mixtures were analyzed directly by MALDI-MS (37). . Parallel experiments with the TCEP/NTCB procedure also were performed on the membrane to confirm the total number of cysteines in the protein. 593,36 ha mUHZ Alllrlllll Alllllllll 9923.0 6 m m: 521% dogged“? 3v Ame... 0G NH - E08 \ :5 Pa uunwomue woumohu 530k“ 69 .oSfiQEoE tog am no Hogan—088m 9:8.on .«o $3128 2885? How 080:8 3859850 4mm 059m Ill'lllll fleece an a: 520 A3) has... > 58%. 6.? azma amok actuated Gov mung... 065m t 0er / owmaéflo A0 amen 0mm e $9 $2: Zommu sow / Ellie] :22 Mesa nouaNEnEOme 6W $ng 5553 he a: :83 Amv A \ £8.on cassava Emmet» @320 flfimbooo gwagopoa gem mu 3on Comm :3 mOHZ coverage 2v 330% £805 lV. Results and Discussion A. Cyanylation and Cleavage Preliminary experimen by MALDI-MS were conduct amino acid sequences and cy expected masses of their expec Typical MALDI mass cleavage products, obtained b under the described experim cyanylation, a cyano group re the original peptide (Figure 2. the observed value exactly m at 37°C for 16 hours, the cyan: 2.53), two of them are cleave respectively. The peak at m/z with the cleavage reaction. ' the mass shift from the cyany Likewise, Figure 2.6 : remaining three peptides. T peptide mixtures. Most of t] such as the cyanylated/melt 70 IV. Results and Discussion A. Cyanylation and Cleavage of Peptides by NTCB Preliminary experiments involving NTCB cleavage of analytes prior to analysis by MALDI-MS were conducted with sulfhydryl-containing peptides having different amino acid sequences and cysteine locations. Table 2.1 lists the peptides and the expected masses of their expected products after cyanylation and cleavage. Typical MALDI mass spectra of the cyanylated peptide (MW 2427.9) and its cleavage products, obtained by reacting 100 pmol of peptide with 500 pmol of NTCB under the described experimental conditions, are shown in Figure 2.5. After the cyanylation, a cyano group replaces hydrogen, resulting in a mass increase of 25 Da from the original peptide (Figure 2.5A). The expected cyanylated peptide has a m/z of 2453.9; the observed value exactly matches the expected data. After incubation in pH 9.0 buffer at 37°C for 16 hours, the cyanylated peptide is replaced by three new components (Figure 2.5B), two of them are cleavage products with the expected m/z of 1099.5 and 1373.5, respectively. The peak at m/z 23194.3 corresponds to the B-elimination product competing with the cleavage reaction. The B-elimination product can be easily identified because the mass shift from the cyanylated peptide is -59 Da. Likewise, Figure 2.6 shows the MALDI spectra of the cleavage products from the remaining three peptides. These spectra reveal general features of MALDI spectra of Peptide mixtures. Most of the major peaks can be assigned to expected specific species, such as the cyanylated/uncleaved peptide, the B-elimination products, and the specific cocoa“: 95.2 a 71 mdmg . memo Tbomm Yemom v.83“ <©M§<§q0§mmA§m memo mooa wee: v.32 v.83 mtwqdmogmm meow fl IVS @an M @603 mgmm fl Zdemmgmxwoh a is car age ..fiww .fim . . . . 00553.0. 28 888 Lmifi 20388 m0 H Z 05 Saw 829.qu wgewcooebafifism 8m $533.“ mo 83? N? 333030 .mm 035. _______._._fi._.— 5 0 1000 Figure 2.5. MALDI mass : reaction with a 5-fold mola products. ++ indicates pea for the correlation of calcul 72 2453.9 A 4A A A AAAA. A_ A .A A AA A —‘ AA —A - ‘ ‘ A A L I m '— - u ' 1100.2 1373.9 Wimgr J hag 500 who 15 o 2(360 2500 3000 m/z ' Figure 2.5. MALDI mass spectra of peptide RYVVLPRPVCFEKGMNYTVR after reaction with a 5-fold molar excess of NTCB, (A) cyanylated peptide and (B) cleavage products. -H- indicates peaks corresponding to doubly charged species. See Table 2.1 for the correlatiOn of calculated and observed m/z values. 73 1507.1 A t t 15491) 160715 1746.2 B 0‘. O\ 2.: 3 ++ I a AM J.,] A 1659.7 C . 2568.0 627.3 9865 r i 2 ++ M:_WJ:1A - +Lt.' A A 4A..) L .._ 1000 1500 2000 2500 3000 m/z Figure 2.6. MALDI spectra of cleavage products of peptides (A) TCVEWLRRYLKN, (B) DRVYIHPCHLLYYS, and (C) MHRQEAVDCLKKFNARRKLKGA, after reaction with 5-fold molar excess of NTCB. See Table 2-1 for the identities of the peaks. cleavage produce Wit“ 3 ms: used. Mass mapping by MAL the peaks. However, the rt35130I greatly, depending on the natu cleavage products. m/z 986.5 amount however, the respon different. Another observatior elimination of cyanylated pept of B-elirnination products (it: cyanylated cysteine residue ( amino acid residues. such as Products, perhaps though 3' underwent little B-eliminatior DRVYIHPQHLLYYS. Exp changing reaction conditions NTCB/analyte were not succ have the same pH dependenc elimination products can als. reactions affect the yield of information on the sulflrydry Shut tures 0n the cleavage rea 74 cleavage products, with a mass accuracy of better than 0.05% if internal standards are used. Mass mapping by MALDI, therefore, unambiguously recognizes the identities of the peaks. However, the responses of peptides on MALDI, i. e., the peak intensities, vary greatly, depending on the nature of the peptides. For example, in Figure 2.6C, the two cleavage products, m/z 986.5 and m/z 1659.7 respectively, should be of equal molar amount, however, the responses of the two fragments on MALDI are considerably different. Another observation is that, in addition to the expected cleavage products, [3- elimination of cyanylated peptides always competes as a main side reaction. The yields of B-elimination products depend on the structures of amino acids adjacent to the cyanylated cysteine residue (19). Our experiments indicate that certain neighboring amino acid residues, such as Pro, His, or Phe, may increase the yield of B-elimination products, perhaps though steric hindrance. Therefore, while TQVEWLRRYLKN underwent little B-elimination, the [El-elimination product predominated for the peptide DRVYIHPQHLLYYS. Experiments carried out to minimize this side reaction by changing reaction conditions such as pH, temperature, reaction time, and molar ratio of NTCB/analyte were not successful, because both cleavage and B-elimination reactions have the sarrie pH dependency. Fortunately, both cyanylated/uncleaved peptides and [3- elimination products can also be detected by MALDI-MS. Therefore, although these reactions affect the yield of the NTCB cleavage reaction, they provide complementary information on the sulfhydryl location and cleavage pattern. The effects of amino acid structures on the cleavage reaction will further be discussed in detail in Chapter 3. Price (14) reported tha' during the cyanylation of sulfh) between protein sulfhydryl gror also observed this side reactior experimental conditions. No mixed disulfides has been obse B. The Effects of pH and Tel Although the cleavage systematic study and optimize reaction. The experimental co protocol proposed by J acobsor 9.5 buffer for 16-80 hours alt to promote the cleavage (38). by conducting the cleavage re and 9.5 buffer solutions, 1 cValidation and cleavage car using tIiS-HCl or sodium bor; be cOmpleted in a 5-fold m minutes at 37°C, the completr Promote both cleavage and B 75 Price (14) reported that another concurrent reaction may occur less frequently during the cyanylation of sulfliydryl groups and lead to the formation of a mixed disulfide between protein sulfhydryl group and NTCB reagent. Denslow and Nguyen (3 8) recently also observed this side reaction. But such a side reaction seems to be minor under our experimental conditions. No trace of the products corresponding to the formation of mixed disulfides has been observed on the analysis by MALDI-MS. B. The Effects of pH and Temperature on the Kinetics of the Cleavage Reaction Although the cleavage reaction by NTCB was proposed two decades ago (7), little systematic study and optimization have been carried out regarding to the kinetics of the reaction. The experimental conditions used today are exactly the same as that used in the protocol proposed by Jacobson et al (7). Typically, the cleavage is performed in pH 8.0 ~ 9.5 buffer for 16-80 hours although higher temperature was also occasionally employed to promote the cleavage (3 8). We have studied the effects of pH on the cleavage reaction by conducting the cleavage reaction on the above model peptides in pH 7.5, 8.0, 8.5, 9.0, and 9.5 buffer solutions, respectively. Time-course studies indicated that both cyanylation and cleavage can be carried out under mildly alkaline conditions (pH 8-9) using tris-HCl or sodium borate as buffers. While cyanylation of sulfliydryls can easily be completed in a 5-fold molar excess of NTCB solution even at pH 7.5 within 30 minutes at 37°C, the complete cleavage of the cyanylated peptide chains takes as long as 16‘ hours even at pH 9.0. Experimental results showed that higher pH can greatly Promote both cleavage and B-elimination, while the relative yields of the two competitive reactions do not vary Significa resulted in incomplete cleavag resulting from long hours Of in' It should be pointed out that thi from peptide to peptide, depen tocysteines. Efficient cleavag The cleavage reactior temperature (e. g. 50°C) has were usually conducted in me reaction system by dialysis or analysis on microscale (pico although higher temperature . after long hours of incubatior of solvents. In addition, or re“milling reagents (e. g., it mperatures have to be caret CI . o Identification of Free Sr Proteins may contain f 011118. The NTCB/TCEP an to ' determine the free sulth) 76 reactions do not vary significantly over a wide pH range. The lower pH (< 7.5) often resulted in incomplete cleavage (cyanylated/uncleaved product) and other side reactions resulting from long hours of incubation, as will be discussed in greater detail in chapter 3. It should be pointed out that the kinetics of both cleavage and B-elimination vary greatly from peptide to peptide, depending considerably on the structure of amino acids adjacent to cysteines. Efficient cleavage requires careful monitoring of the reaction. The cleavage reaction was previously performed at 37°C, although higher temperature (e. g., 50°C) has also been used. However, previous cleavage experiments were usually conducted in macroscale (mg) after removing the excess reagents from the reaction system by dialysis or other purification procedures (7, 13-20). Our results from analysis on microscale (picomole of samples in microliters of solution) indicated that although higher temperature can accelerate the cleavage, it is not recommended because after long hours of incubation, the reaction vial can easily get dry due to the evaporation of solvents. In addition, our cleavage reaction is performed without removal of the remaining reagents (e. g., NTCB), the side reactions caused by NTCB at higher temperatures have to be carefully considered. C. Identification of Free Sulfliydryls and Total Cysteines in Proteins Proteins may contain free sulfluydryls, disulfide bonds, or a combination of both forms. The NTCB/TCEP and TCEP/NTCB procedures described in Figure 2.3 were used to detennine the free sulfl’rydryls and total cysteines, respectively. Figure 2.7 lists the 1-—-'-T—“"'T SH Sl Sp' 8 1 l 25 m: 22 1 SH 66 i 1 k l 10 s [3 macaw SH SHQ Figure 77 AcNH 18 39 44 47 77 97 l | I I T SH SH SH SH SH Spinach Ferredoxin (MW. 10483) S S s .19 l 25 56 I 95 I 212 22 | I 63 I 153 200 SH S S Papain (MW. 23426) S ‘ H 66 I i i 160 162 I 106 119 121 1 S S B-Lactoglobulin A (MW. 18368) 11 3O 68 i i 344 367 382 385 | | 73 120 I l SH SH (3 03 SH SH Ovalbumin (MW. 42699) 1 “i4 141 SH oc-Hemoglobin A (MW. 15053) 93 112 1 4 6 la la. B-Hemoglobin A (MW. 15954) Figure 2.7. Structural features of model proteins. Table 2.2. Mass assignrr 4—q NTCl # Fragment % ———— Protein Ferredoxin‘ I - I 7 18-38 3943 44.-46 47-76 77-97 18-43 44-76 Papain 1-24 25-212 Ovalbumin 1_ 10 11-29 30-366 367-381 382-385 I-29 I-29 367-385 367-335 B'LactOg. l- 12 lObulin A 0 121-162 Omen‘(Iglobint‘ ‘Hmoglobrn 78 Table 2.2. Mass assignment of peptide fragments after different treatments NTCB/TCEP Treatment TCEP/NTCB Treatment Pr oteln Fragment [M+H]+CI - [M+I-1]+obi Fragment [M+H]+calc [M+I‘I]+ob§_ Ferredoxina 1-17 1839.1 1333.8 18-38 2352.2 2347.7 3943 517.5 ndb 44-46 320.3 320.0 47-76 3337.6 3337.6 77-97 2332.5 2332.0 18-43 2792.00 2792.8 44-76 3580.8c 3581.5 Papain 1-24 2621.5 2622.1 1-21 2372.7 2371.9 225-212 20848 nd 2224 290.3 288.4 25-55 3547.0 3546.7 56-62 880.9 880.9 63-94 3794.2 3792.4 95-152 6309.2 6308.9 153-199 5039.7 5038.9 200-212 1503.7 1503.5 Ovalbumin 1-10 1011.1 nd 1-10 1011.1 nd 11-29 2378.7 2378.5 11-29 2378.7 2377.4 30-366 37684.7 nd 30-72 4762.5 4761.4 367-381 1715.1 1715.2 73-119 5427.1 5426.1 382-385 429.5 429.4 120-366 27581.1 nd 1-29 3312.8c 3312-3 367-381 1715.1 1714.3 129 337134 3371-5 382-385 429.5 nd 367-385 20676.: 2067-7 129 3312.8° 3311.2 367-385 2126 68 2126-6 1-29 3371.8d 3370.3 . 367-335 2067.6c 2068.1 B-Lactog— 1420 4905.7 4901.1 1-65 7247.4 7238.2 lobulin A 121-162 13506.7 13510 66-105 4651.6 4653.5 106-118 1461.6 1462.1 119-120 274.3 nd 121-159 4551.3 4548.1 160-162 396.5 nd or-Hemoglobina 1-103 1 1073-6 1 1065 B-Hemoglobin 104-141 4097-3 4093-0 192 9919.0 9915.0 93-1 1 1 2207.6 2207.0 112-146 3828.4 3831.6 _ a Only NTCB treatment was applied to spinach ferredoxin and hemoglobin samples since the proteins only contain free sulfliydryl groups. b Not detected by MALDI-MS. ° B-elimination products. d Cyanylated/uncleaved peptides. structural features of the prote observed masses of the fragmer I. Spinach Ferredoxin Spinach ferredoxin cont biological systems. some of thi uith Feand/or M0 (41). The applied to the protein solution complexes. The MALDI mass of ferredoxin reacted with 2.5 the MALDI probe tip). sh! conesponding to 39-43. The good agreement, with a relatii the TOP instrumnt, the peak with the peak at m/Z 2332.5 Ill/Z 2347.7 deviated from thi 1 he peak at m/z 1820.1 is at 183 ‘ 9-1) dunng the cleavage 1 m/ 22792.8 and 3581.5 cm 44-76' ‘ 1n Wthh B-elimination 79 structural features of the proteins studied by the methodologies. The calculated and observed masses of the fragments from these proteins are listed in Table 2.2. 1. Spinach Ferredoxin Spinach ferredoxin contains 5 free sulfhydryls and no disulfide bonds (3 9, 40). In biological systems, some of the free sulfl1ydryl groups in ferredoxin can form complexes with Fe and/or Mo (41). Therefore, prior to treatment with NTCB, 1 mM EDTA was applied to the protein solution to liberate the free sulfhydryls from possible S-Fe or S-Mo complexes. The MALDI mass spectrum in Figure 2.8, recorded fiom a 100-pmol sample of ferredoxin reacted with 2.5 nmol of NTCB (~1.5 pmol of the analyte was applied to the MALDI probe tip), shows all the protonated peptide fragments except that correSponding to 39-43. The calculated and observed [M + H]+ values (Table 2.2) are in good agreement, with a relative mass deviation of < 0.1%. But, due to poor resolution of the TOF instrument, the peak at m/Z 2352.5 (corresponding to fragment 18-3 8) overlaps with the peak at m/z 2332.5 (corresponding to fragment 77-97). The observed peak at m/Z 2347.7 deviated from the calculated value by 48 Da, indicating much higher error. The peak at m/Z 1820.1 is attributable to the dehydration product of fragment 1-17 (m/z 1839.1) during the cleavage reaction, as observed by Papov and Biemann (42). Peaks at m/z 2792.8 and 3581.5 correspond to the incomplete cleavage of fragments 18-43 and 44-76 in which 1.3-elimination occurs at Cys39 and Cys47, respectively. Intensity \ 320.0 Relative: \ \1 019 Q ///18ZOJ rooo 2r Figure 2.8, MALDI mass S 5001111101 of NTCB Asterisks ( i the identiti a . Apprt )lndicates peak 38 of marked pea 80 2332.0 2347.7 1 \320.0 Relative Intensity ,//1820J \1838 8 E *2792.8 3337.6 ; *3581.5 J “ - LL 1000 2000 3000 4000 5000 6000 m/z Figure 2.8. MALDI mass spectrum of 100 pmol of spinach ferredoxin after reaction with 500 pmol of NTCB. Approximately 1.5 pmol of analyte was applied to the probe tip. Asterisks (*) indicates peaks corresponding to B-elimination products. See Table 2.2 for the identities of marked peaks and the correlation of calculated and.observed m/z values. 2. Papain Papain, a member Of a study because it has one 0““ 15). The NTCB-(TCEP 1310' sulfhydryl (Figure 2.9A). 86‘ 0525 should give two PCP‘i‘ bond (CysZZ-Cys66). The rec and produces peptide fiagmen m’z 2621.5 and 20848.0. res was observed corresponding conesponding t0 peptide ch probably due to suppression 1 Another interesting 01 reaction mixture from papair frtrgment 1-24 at m/z 2622 iminothiazolidinyl residue a1 are Still linked by an inter-c1 at C1825. Thus, observatic diSUIfide bond might have be 011 B-lactoglobulin A (MW 1 at Cylel, B-lactoglobulin iI - ° ' ner charm disulfide bond (1 2. Papain Papain, a member of a family of over 40 thiol proteases (43), was chosen for this study because it has one cysteine (Cys25) at the active site and three disulfide bonds (44, 45). The NTCB/TCEP procedure was utilized to recognize the locus of the free sulfhydryl (Figure 2.9A). Because papain is a single chain polypeptide (44), cleavage at Cys25 should give two peptide chains which remain linked by an inter-chain disulfide bond (Cys22-Cys66). The reduction of disulfide bonds by TCEP cleaves the two chains and produces peptide fragments 1-24 and 25-212 having calculated values for [M + H]+ at m/z 2621.5 and 20848.0, respectively. Experimentally, an intense peak at m/z 2622.1 was observed corresponding to the peptide chain 1-24. The expected high-mass peak- corresponding to peptide chain 25-212 was not seen in the MALDI mass spectrum, probably due to suppression by the low-mass peptide fragment present in the sample. Another interesting observation is that the MALDI analysis of the NTCB cleavage reaction mixture from papain prior to disulfide bond reduction also showed the peptide fragment 1-24 at m/z 2622.1. In principle, only an intact protein derivatized by an iminothiazolidinyl residue at Cys25 should be detected because the two peptide chains are still linked by an inter-chain disulfide bond (Cy522-Cys66) after the NTCB cleavage at Cys25. Thus, observation of the peak at m/z 2622.1 indicates that the inter-chain disulfide bond might have been cleaved during the MALDI process. Similar experiments on B-lactoglobulin A (MW 18368) (46) gave a similar result. After cleavage with NTCB at Cys121, B-lactoglobulin A should produce two peptide chains that are linked by an inter-chain disulfide bond (Cys66-Cysl60). However, the MALDI mass spectrum of B- l-24 l E. 0) ad W 25212 5000 10000 15000 20000 25000 m/z ' B 121 .9 a 8 .8 _°>’ 63-94 :3. 32 3 9‘. M H N w). 0\ 1000 2000 3000 4000 5000 6000 7000 8000 m/z Figure 2.9. MALDI mass spectra of papain after (A) NTCB/TCEP treatment and (B) TCEP/NTCB treatment. See Table 2.2 for the identities of other marked peaks and ,the correlation of calculated and observed m/z values. lactoglobulin A after reaction resulting from the cleavage 0f the recent report by P3116150" undergo "prompt fragmentatio 111). Total cysteine assessm procedure. Most of the expect seven cysteine residues were peaks unable to be assigned tc inthe commercial sample. T: their observed masses. In th Elimination were found, whic POlypeptide chain at all the c 3. Ovalbumin Ovalbumin (48‘50) studied by the proposed PTO N TCB/TCEP procedure (Fi gave two more peaks (at m/ an dcleavage of the disulfid a . greement Wrth the previou Furrh ermore, two clusters 83 lactoglobulin A after reaction with NTCB showed peaks corresponding to fragments resulting from the cleavage of the inter-chain disulfide bond. Our observations support the recent report by Patterson et al that the inter-chain disulfide bond in peptides can undergo "prompt fragmentation" or “in-source fragmentation” in the MALDI ion source (47). Total cysteine assessment of papain was performed according to the TCEP/NTCB procedure. Most of the expected fragments corresponding to the cleavage at each of the seven cysteine residues were found in the MALDI mass spectrum (Figure 2.9B). The peaks unable to be assigned to any of the fragments were most likely due to an impurity in the commercial sample. Table 2.2 also lists the calculated masses of the fragments and their observed masses. In this case, only a few fragments of incomplete cleavage or B- elimination were found, which indicates that the NTCB reagent can effectively cleave the polypeptide chain at all the cysteine residues in papain. 3. Ovalbumin Ovalbumin (4850), containing four free cysteines and one disulfide bond, was studied by the pr0posed procedures. Compared to the mass spectrum obtained from the NTCB/TCEP procedure (Figure 2.10A), that obtained with the TCEP/NTCB treatment gave two more peaks (at m/z 4761.4 and 5426.1) that were due to reduction, cyanylation, and cleavage of the disulfide bond Cys73-Cy5120 (Figure 2.108). This conclusion is in agreement with the previous assignment of the disulfide bond linkage in ovalbumin (50). Furthermore, two clusters of peaks were observed in the MALDI mass spectra which ’_____.____——————— 367-381 \ 367-385* /’367'385* _85 367-381 N 3675485" 1000 2000 Figure 2.10. MALDI mas TUSP/NTCB treatment. ' Peptrdes and/or their B-eli and the correlation 1 84 367-38l\ * .3 A (i \O m * O\ In 01 oo .1. <13 ... [\ g / as \ * I 9(- 83 33 UM flea-.4344 4 -4 -.._ T.,...” 367-381V * ., 2’8 B m if. ‘9 cost 32 * (\l \/ {it 1000' 2000' 3000' 4000 5000 6000 m/z Figure 2.10. MALDI mass spectra of ovalbumin after (A) NTCB/TCEP and (B) TCEP/NTCB treatment. * indicates peaks corresponding to cyanylated/uncleaved peptides and/or their B-elimination products. See Table 2.2 for the identities of marked Peaks and the correlation of calculated and observed m/z values. correspond to the incompltite peak at m/z 3371.5 is due 10 Cysll (expected Ml 33718) at Cysll (expected m/z 3312. peptide and its Belimination ' and 2067.7 are likely due uncleaved Cys382 and its ovalbumin contains two ph observed [M+H]' ion of the phosphate group is still att indicating that phosphoryla exIltrirnental/ reaction condi ”I1 be used to detect the 1 segment, because it is prime Shift. This example demon: be used to confirm a protein 4. B-Lactoglobulin A B-Lactoglobulin A (46). After NTCB/TCEP t1 12 1 are expected, with a 1 E I Xpemnentally, the fragmt 85 correspond to the incomplete cleavage and/or B-elimination of sulfliydryl groups. The peak at m/z 3371.5 is due to the peptide chain 1-29 containing a cyanylated/uncleaved Cysll (expected m/z 3371.8), while m/z 3311.1 corresponds to the B-elimination product at Cysll (expected m/z 3312.8). The mass difference between the cyanylated/uncleaved peptide and its B-elimination product is 59 Da. For the same reason, peaks at m/z 2126.6 and 2067.7 are likely due to the peptide chain 367-3 85, containing a cyanylated] uncleaved Cys3 82 and its [3-elirnination product, respectively. It is known that ovalbumin contains two phosphorylated residues at serines 68 and 344 (50). The observed [M+H]+ ion of the fragment 30-72 (expected at m/z 4761.5 Da) shows that the phosphate group is still attached to Ser68 after cyanylation and cleavage reactions, indicating that phosphorylated proteins should be amenable to analysis under these experimental/ reaction conditions. On the other hand, the mass mapping of the fragments can be used to detect the possible posttranslational modification in a specific protein segment, because it is principally possible to recognize the modification from the mass shift. This example demonstrated that total cysteine determination by this approach can be used to confirm a protein’s structure. 4. B-Lactoglobulin A B-Lactoglobulin A contains five cysteines, of which Cys12l is a free cysteine (46). After NTCB/TCEP treatment, two fragments corresponding to the cleavage at Cys 121 are expected, with a m/z of 13506.7 (1-120) and 4905.7 (121-162), respectively. Experimentally, the fragment 1-120 presents as a small peak (Figure 2-11A), probably due to the suppreSSion from c two expeaed fragments. 1'12 the other two fragments 121‘] cyanylation and cleavage at C 160 is insufficiently stable if thiolate ions that subsequen Given the fact that B-lactoglr observed nonspecific cleav: exchange and subsequent c mixture against 20% acetic molecular weight salts and 0 detected by MALDI (Fig. 2- The peak at m/z 18390 is 111 containing fragment 1-120 dOUbly Charged ion. The IV Shows fragments 1-65, 66-1 a[111160-162, which fall it it agments are essentially MALDI responses of indi‘ S. . lrnrlar MALDI spectra of the disulfide bonds and sul: 86 due to the suppression from other components (peptides and salts). In additional to the two expected fragments, 1-120 and 121-162, MALDI can also identify trace amounts of the other two fragments 121-159 (m/z 4551.2) and 1-65 (m/z 7247.0), attributable to both cyanylation and cleavage at Cys66 and Cysl60, implying that the disulfide bond pair 66- 160 is insufficiently stable in the presence of NTCB "reagent and is hydrolyzed to form thiolate ions that subsequently participate in the cyanylation and cleavage reactions. Given the fact that B-lactoglobulin A is particularly liable to SH/S-S exchange (51), the observed nonspecific cleavage at Cys66 and Cys160 could also result from such exchange and subsequent cyanylation and cleavage. After dialysis of the cleavage mixture against 20% acetic acid for 12 hours (MW cut-off: 3000 Da) to remove low molecular weight salts and other contaminants, the fragment 1-120 can be unambiguously detected by MALDI (Fig. 2-11B), confirming that Cys121 is a cyanylation/cleavage site. The peak at m/z 18390 is likely due to the reoxidization during the dialysis of sulfliydryl- containing fragment 1-120 and 121-162, while the peak at m/z 9191 is obviously its doubly charged ion. The MALDI spectrum after TCEP/NTCB treatment (Figure 2-11C) shows fragments 1-65, 66-105, 106-118, 121-159, and two smallest fragments, 119-120 and 160-162, which fall in the region reserved for matrix peaks. Therefore, all the fragments are essentially detected, despite of the greatly variable differences in the MALDI responses of individual fragments. Denslow and Nguyen (3 8) have reported similar MALDI spectra of cleavage products of B-lactoglobulin obtained by reduction of the disulfide bonds and subsequent cyanylation and cleavage of the sulfhydryl groups. 2000 4000 6000 Figure 2.11. MALDI n NTCB/Dialysis/TCEP, correlation of calculatec 87 121-162 A a ;i .._1 xi 1-65 ++ 1-120 1-162 2.. ._- ‘TJA‘“““T 1-162 L120 121-162 1-65 -++ 121'15\9 66-105 / C 2000 4000 6000 8000 10000 12000 14000 16000 18000 20000 m/z Figure 2.11. MALDI mass spectra of b-lactoglobulin after (A) NTCB/TCEP, (B) NTCB/Dialysis/TCEP, and (C) TCEP/NTCB treatment. See Table 2.2 for the correlation of calculated and observed m/z values. 4. Hemoglobin Human hemoglobin A one free cysteine at 104, whi 112. Since hemoglobin A car commercially available. The mixture of 01- and B-hemoglc the expected fragments can ‘ observation is that each of tl chain, is accompanied by a p The identity of these peaks Optimization of the deterrnin 0104-141 1393—1 11 -146 Relative Intensity 81 1 2 fl: 4. Hemoglobin Human hemoglobin A contains 0t- and [3- chains (52). oe-Hemoglobin A contains one fi'ee cysteine at 104, while B-hemoglobin A contains two free cysteines at 93 and 112. Since hemoglobin A can be easily oxidized in the air. The native hemoglobin is not commercially available. The hemoglobin we acquired was a service sample containing a mixture of or- and B-hemoglobin A variants. After NTCB cyanylation and cleavage, all the expected fiagments can be unambiguously identified (Figure 2.12). An interesting observation is that each of the fragments, 1-92 and 93-111 of B chain, and 1-103 of A chain, is accompanied by a peak which shows a 102 Da mass increase fiom that fragment. ' The identity of these peaks are unclear. Because of the limited sample available, the optimization of the determination was not pursued further. insulin oelO4-14l .E .9 .4 .0 £1 :3 E a) «t, 0', '81 a .2; T) 94 ' 1-92 2000 4000 ‘ 6000 8000 10000 12000 14000 16000 m/z Figure 2.12. MALDI mass spectrum of oe- and B-hemoglobin mixtures after reaction with NTCB. See table 2.2 for the correlation of calculated and observed m/z values. E. Choice of Reducing Rea: Several reducing rea reduction of protein disu dithiothreitol (DTT) (54)~ ar However, thiol-containing re be avoided because excess interferes with analyses tarboxyethyl)ph0sphine (TI disulfide bonds (36, 55, 56) used in the presence of an 0: isstable not only in acidic reducing activity than or reaction solution is therefon Sufficient for complete red larger amounts of TCEP d MALDI—MS. F. o 0 Stability of Disulfide ll Stoichiometric expe for complete cyanylation ( Performed under the same E. Choice of Reducing Reagent Several reducing reagents have been deve10ped for the efficient and selective reduction of protein disulfide bonds. Among these, B-mercaptoethanol (53), dithiothreitol (DTT) (54), and sodium borohydride (NaBH4) (55) are traditionally used. However, thiol-containing reducing reagents such as B-mercaptoethanol and DTT must be avoided because excess reagents will react with NTCB. Sodium borohydride interferes with analyses by MALDI-MS (55). Tributylphosphine and tris(2- carboxyethyl)phosphine (TCEP) have proven to be selective and reactive towards disulfide bonds (36, 55, 56). However, tributylphosphine is water-insoluble and must be used in the presence of an organic solvent. TCEP is a water-soluble reducing reagent and is stable not only in acidic solution, but in basic solution as well. TCEP shows higher ‘ reducing activity than DTT over a wide pH range (57). The pH adjustment of the reaction solution is therefore unnecessary. Usually, a 5-fold molar excess of TCEP is sufficient for complete reduction of disulfide bonds in denaturing conditions although larger amounts of TCEP do not appear to have any adverse effect on the analysis by MALDI-MS. F. Stability of Disulfide Bonds during the NTCB Reaction Stoichiometric experiments indicated that the NTCB/thiol ratio of 5:1 is sufficient for complete cyanylation of thiol groups. Because both cyanylation and cleavage were Performed under the same condition without removal of the NTCB reagent, the excess of NTCB reagent can always along period of incubation, The mechanism of th acting as a nucleophile to a cyanylated product, which 0 amino peptide bond (7). I groups. Selective cyanylati presence of cystines beca specificity is not without li containing peptides indica incubation time result in cle the stability study of disul: excess of NTCB reagent u disulfide bond. The same 11 excess of NTCB reagent i: hours. MALDI spectra, as generally stable in the pre: incubation at 37°C, cor cyanylation/cleavage of the NTCB exceeds SO-fold und correspond to the cleavage disulfide bonds undergo l 90 NTCB reagent can always have a chance to attack the unreacted sulfhydryl groups during a long period of incubation, which ensures the cyanylation goes completion. The mechanism of the NTCB reaction involves the dissociated thiolate anion (-S') acting as a nucleophile to attack the cyanate moiety under alkaline conditions to form a cyanylated product, which can then undergo cleavage catalyzed by hydroxide ions at the amino peptide bond (7). In principle, the NTCB reagent is specific for free sulfhydryl groups. Selective cyanylation/cleavage at free sulfhydryl groups can be performed in the presence of cystines because the disulfide bonds do not react with NTCB. However, this specificity is not without limits under the reaction conditions. Experiments on disulfide- containing peptides indicate that a large excesses of NTCB reagent and excessive incubation time result in cleavage at disulfide bond residues. This has been confirmed by the stability study of disulfide-containing peptide(s) in the presence of different molar excess of NTCB reagent under the same buffer conditions. Somatostatin contains one disulfide bond. The same molar amount of somatostatin was mixed with 2, 5, 10, 50-fold excess of NTCB reagent in pH 9.0 buffer, respectively, and incubated at 37°C for 16 hours. MALDI spectra, as shown in Figure 2.13, show that while the disulfide bond is generally stable in the presence of a 10-fold molar excess of NTCB after 16 hours of incubation at 37°C, complete splitting of the disulfide bond and subsequent Cyanylation/cleavage of the nascent cysteine residues take place when the molar excess of NTCB exceeds 50-fold under the same conditions. The masses of the resulting fragments correspond to the cleavage at one or both cysteine sites. One possible explanation is that disulfide bonds undergo hydrolysis under mildly alkaline conditions (58, 59). The inta 4km degradation pr W., fi—fi—‘fl" 1200 1300 1400 Figure 2.13. Stability stt of NTCB reagent. After at 37°C for 16 hours, 801 91 intact somatostatin \ 2 fold WW 5 fold 10 fold /l degradation products \ ‘ 50 fold N 1200 1300 1400 1500 1600 1700 1800 1900 2000 m/z Figure 2.13. Stability study of somatostatin in the presence of different molar excesses of NTCB reagent. After exposed to 50-fold excess of NTCB reagent in pH 9.0 buffer at 37°C for 16 hours, somatostatin decomposes completely. resulting thiolate anions atta cleave the peptide chain NTCBII‘CEP procedure is NTCB must be applied to e purified by chromatography subsequent cleavage reacti total cysteines (T CEP/NTC prior to the NTCB reaction. Another commonly bond exchange that usually 9.0, 161101118 incubation) di not affect the assignment srtlfltydryls become blocke hand, the described methot bond pairing patterns. G. Analysis by MALDI-ll Because of the wid obtained fi'om NTCB cleaV are common. Previous 1 employed sodium dodecl Unfortunately, its accuracl 92 resulting thiolate anions attack the NTCB reagent to form cyanylated products that further cleave the peptide chain at the original disulfide bond sites. Therefore, if the NTCB/TCEP procedure is used for sulfhydryl—rich proteins where a high molar excess of NTCB must be applied to ensure complete cyanylation, the cyanylated proteins are better purified by chromatography or dialysis to remove the excess NTCB reagent prior to the subsequent cleavage reaction. This limitation is not a problem for the determination of total cysteines (TCEP/NTCB procedure) since all the disulfide bonds must be reduced prior to the NTCB reaction. Another commonly encountered problem in cysteine characterization is disulfide bond exchange that usually occurs at pH > 8.0. Under our experimental conditions ( pH 9.0, 16 hours incubation) disulfide bond exchange cannot be avoided. However, it should not affect the assignment of 'free sulflrydryl groups and/or total cysteines since free sulfhydryls become blocked immediately after cyanylation with NTCB. On the other hand, the described method cannot directly be employed for the assignment of disulfide bond pairing patterns. G. Analysis by MALDI-MS Because of the widely variable occurrence of cysteine in proteins, the fragments obtained from NTCB cleavage vary greatly in size. Fragments with a mass over 5000 Da are common. Previous reports on the mass mapping of NTCB cleavage products employed sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE). Unfortunately, its accuracy of molecular weight determination is usually limited to :t 5- 10%. In many cases, pep masses of two or more pep Incomplete cleavage of prot and the SDS-PAGE techni products corresponding t cyanylated/uncleaved pepti MALDI-TOF MS determine molecular weig 0.01% and thereby provid comparison of calculated assignment is better than i for calibration. This accura MALDI spectra art though peaks for a sod Occasionally, matrix addur that the MALDI signal intt because the reaction produ peptides may not give a discrimination effects are r of salts, or due to differe peptides are more suppress 93 10%. In many cases, peptide assignments are inconclusive or are ambiguous if the masses of two or more peptides obtained from the NTCB digestion are similar (19, 21). Incomplete cleavage of protein chains is often observed with the NTCB reaction (60-62), and the SDS-PAGE technique has difficulty distinguishing among incomplete cleavage products corresponding to the peptide with an unreacted sulfhydryl group, a cyanylated/uncleaved peptide, or a B-elimination product. MALDI-TOF MS has been a valuable complement to SDS-PAGE. It can determine molecular weights as high as 300,000 Da with the mass accuracy of i01- 0.01% and thereby provide an unambiguous answer to questions mentioned above. A comparison of calculated values and our observed data indicates that the error of mass assignment is better than i 0.05% for most of the peptides if internal standards are used for calibration. This accuracy is sufficient for the mass mapping of cleavage products. MALDI Spectra are dominated by peaks for singly protonated species [M+H]+, though peaks for a sodium adduct [M+Na]+ often appear at a lower intensity. Occasionally, matrix adduct peaks are observed. A limitation of our present method is that the MALDI signal intensities vary considerably from peptide to peptide in a mixture because the reaction products are directly analyzed without purification by HPLC. Some peptides may not give a detectable signal during analysis of a mixture. These discrimination effects are either due to MALDI contaminants such as high concentrations of salts, or due to differential responses of the peptides, or both. Usually high-mass Peptides are more suppressed and yield weaker signals. H. Reactions of Proteins 1 One of the great i techniques, such as FAB a However, the high concei interferes with a successful form a thick cake or a resi Samples containing 00an of the sample to minimize analyte often can be obser where only trace quantities Recently, various tr supports for direct analysis and proteins on an inert tr soluble MALDI contamina while the analyte is retaine matrix solution. Moreove proteins are possible on the The in situ disulfi‘ reaction for protein samph Zetabind membrane and a of which have been emplc 64). Experiments with ti 94 H. Reactions of Proteins Immobilized on A Zetabind Membrane One of the great advantages of MALDI-MS over other mass spectrometric techniques, such as FAB and E81 mass spectrometry, is its tolerance of contaminants. However, the high concentration of denaturing agents used in our procedure still interferes with a successful analysis by MALDI-MS. The protein-matrix rrrixtures often form a thick cake or a residue that does not completely cocrystallize on the probe tip. Samples containing contaminants frequently give no or very weak signals. Upon dilution of the sample to minimize the effect of the contaminants, a MALDI signal from the analyte often can be observed. Diluting the sample is, however, self-defeating in cases where only trace quantities of the analyte are available. Recently, various transfer membranes (3 7, 63, 64) have been studied as sample supports for direct analysis of proteins by MALDI-MS. The immobilization of peptides and proteins on an inert transfer membrane can be used to facilitate removal of water- soluble MALDI contaminants, such as high concentrations of denaturing agents and salts, while the analyte is retained when the membrane is immersed in water prior to adding the matrix solution. Moreover, in situ chemical and enzymatic modifications of adsorbed proteins are possible on the membrane. The in situ disulfide reduction by TCEP, cyanylation by NTCB, and cleavage reaction for protein samples as small as 10 pmol have been performed on a nylon-based Zetabind membrane and a porous polyethylene membrane (purchased from Fisher), both of which have been employed for direct MALDI-MS analysis of high-mass proteins (37, 64). Experiments with the NTCB reaction on the two membranes showed comparable results. However, the pol operationally difficult to ha The MALDI mass reactions with 10 pmol of } membrane, are shown in P NTCB reagents was appl cyanylation of sulfhydryl: membrane by washing, W] the high concentration of subsequent analysis by I disulfide bond sites, in the the excess NTCB reagent performed in solution and spectra for papain. The c complete for ovalbumin observed in the MALDI rr Cysll. However, the res poorer for the analytes imr adequately controlled. One problem with maintaining an aqueous s< This problem can be mini containing a few drops c 95 results. However, the polyethylene membrane is a porous thin film on which it was operationally difficult to handle the chemical reactions. The MALDI mass spectra of the reaction products, obtained respectively from reactions with 10 pmol of papain and 15 pmol of ovalbumin immobilized on a Zetabind membrane, are shown in Figures 2.14 and 2.15. A SO—fold molar excess of TCEP and NTCB reagents was applied to ensure complete reduction of disulfide bonds and cyanylation of sulflrydryls. Because all reagents can readily be removed from the membrane by washing, while the protein and/or peptide reaction products are retained, the high concentration of TCEP and NTCB do not present any large problem during subsequent analysis by MALDI-MS. Moreover, the possible cleavage reaction at ' disulfide bond sites, in the presence of a high molar excess of NTCB, was minimized as the excess NTCB reagent was washed out after the cyanylation reaction. The reactions performed in solution and on the surface of the Zetabind membrane gave ahnost the same spectra for papain. The cleavage reaction performed on the membrane was even more complete for ovalbumin relative to that performed in the solution; no peaks were observed in the MALDI mass spectrum that corresponded to the incomplete cleavages at Cysl 1. However, the resolution and accuracy of mass assignment can be substantially poorer for the analytes immobilized on the membrane if the power level of the laser is not adequately controlled. One problem with performing microscale chemical reactions on a membrane is in maintaining an aqueous solution of the reactants on the membrane for an extended period. This problem can be minimized by performing reactions in a closed vessel (see Fig. 2-3) containing a few drops of water to saturate the air space with water vapor so that the 1-24 Relative ' 5000 Figure 2.14. MALDI n treatment on a Zetabind 96 1-24 .42 U) r: 8 .3 0 .> :3. d) 94 ""1 -:..: J4-‘ramt‘vfiew‘ that: - :- :.-:-M> 5000 10000 15000 ’ 20000 25 000 m/z Figure 2.14. MALDI mass Spectrum of 10 pmol of papain after in situ NTCB/TCEP treatment on a Zetabind membrane. - 367-381 Relative Intensity Figure 2.15. MALDI m treatment on a Zetabind 97 367-381 \ 9: .12 7‘ {D m G l\ 8 5 m g 1‘. :2: e» 3 d) 00 m m 1006 2060 3060 4060 5060 6000 m/z Figure 2.15. MALDI mass spectrum of 15 pmol of ovalbumin after in situ TCEP/NTCB treatment on a Zetabind membrane. droplet of aqueous reactior reach completion. The resolubilizatir subsequent cocrystallizati successful analysis by MA an aqueous solution of tl membrane and the analy saturated with the vapor ( drying/cocrystallizing prc improves the resolution in After some of the Nguyen (38) reported a s after cyanylation on PVl MALDI-MS. In their pr ribonuclease, bovine B-la membrane. After rinsing by placing the membrane 98 droplet of aqueous reaction mixture will be sustained long enough to allow the reaction to reach completion. The resolubilization of reaction products from the membrane surface and subsequent cocrystallization of the analyte/matrix mixture are the key to eventual successful analysis by MALDI-MS. To assure the complete resolubilization of samples, an aqueous solution of the matrix in 50% acetonitrile/0.1% TFA was applied to the membrane and the analyte-matrix mixture was allowed to dry- slowly in a container saturated with the vapor of the 50% acetonitrile/O.1% TFA aqueous solvent. This long drying/cocrystallizing process also helps produce fine analyte-matrix cocrystals and improves the resolution in the MALDI mass spectrum. After some of the results presented in this chapter were published, Denslow and Nguyen (3 8) reported a similar approach to cleave blotted proteins at cysteine residues after cyanylation on PVDF membranes. The resulting fragments were analyzed by MALDI-MS. In their protocol, the cyanylation of 50 pmol each of bovine pancreatic ribonuclease, bovine B-lactoglobulin, and bovine insulin was carried out on, a PVDF - membrane. After rinsing the membrane to remove buffers, cleavage was then carried out by placing the membrane in alkaline buffer and the cleavage products were extracted into MALDI solvent by sonication. Although most of the expected fragments were detected, the efficiency of cleavage is somewhat less than that performed in free solution. Several partial cleavage fragments were found, some of them correspond to B—elimination or mixed disulfide bonds (38). These results support our conclusion that the cleavage reaction can be achieved 0 that the MALDI analysis 0 V. Conclusions We have demonstr treatment of a protein an simple and sensitive me sulfliydryl residues in the follouing TCEP and NTC posttranslational modifica reaction products. We als remove excess reagents a NTCB reagent is unique protocol provides the adv: high sensitivity. 99 reaction can be achieved on membranes. Moreover, our experiments further demonstrate that the MALDI analysis of cleavage products can directly be performed on membranes. V. Conclusions ‘ We have demonstrated here that the combination of NTCB and TCEP chemical treatment of a protein and subsequent mass-mapping by MALDI-TOF MS provides a simple and sensitive methodology for determining the number and location of free sulfhydryl residues in the presence of disulfide bonds. On the other hand, mass-mapping following TCEP and NTCB procedures can be used to confirm the primary sequence and posttranslational modification of a protein and to identify the cleavage products and side reaction products. We also demonstrated that an inert Zetabind membrane can be used to remove excess reagents and other salts to improve the MALDI spectra. The use of the NTCB reagent is unique in that it specifically targets the sites being analyzed. This protocol provides the advantages of fast analysis, easy operation, high mass accuracy, and high sensitivity. Vl. References l. Darbre, A. Practical } 149. 2. canoe. Q-aand W 3, Hirayama, K., and AI‘ (Matsuo, T., CaPFiOHa Ltd, (1994). p. 299-3j 4. Robertson, J. G.. Ad Villafranca J. J Bioc 5. Glocker,M. 0., Arbor Acad. Sci. USA., 91, 5 . Catsirnpoolas,N., and . Jacobson, G. R., Sch: 248,6583-6591 (1973 - Degani, Y., and Patch . Wakselman, M., Gui Chem. Commun., 1, 2 10. Brocklehurst, K., Ma. A., Mushiri, M. S., an 11. Brocklehurst, K., Ma] 12' SChaffer, M. H. and (1976). 100 VI. References 1. 10. Darbre, A. Practical Protein Chemistry: A Handbook, John wiley & Sons, (1985) p. 149. Chin, C. C. Q., and Wold, F., Anal. Biochem., 214, 128-134 (1993). Hirayama, K., and Akashi, S., in Biological Mass Spectrometry: Present and Future (Matsuo, T., Caprioli, R. M., Gross, M. L., and Seyama, Y, Eds), John Wiley & Sons Ltd., (1994). p. 299-312. Robertson, J. G., Adams, G. W., Medzihradszky, K. F ., Burlingame, A. L., and Villafranca, J. J ., Biochemistry, 33, 11563-11575 (1994). Glocker, M. 0., Arbogast, B., Milley, R., Cowgill, C., and Deinzer, M. L., Proc. Natl. Acad. Sci. USA., 91, 5868-5872 (1994). Catsimpoolas, N., and Wood, J. L., J. Biol. Chem, 241, 1790-1796 (1966). Jacobson, G. R., Schaffer, M. H., Stark, G. R., and Vanaman, T. C., J. Biol. Chem. 248, 6583-6591 (1973). Degani, Y., and Patchornik, A., Biochemistry, 13, 1-11 (1974). Wakselman, M., Guibe-Jampel, E., Raoult, A., and Busse, W. D., J. Chem. Soc. Chem. Commun., 1, 21(1976). Brocklehurst, K., Malthouse, J. P. G., Baines, B. S., Blenkinsop, R. D., Churcher, J. A., Mushiri, M. S., and Ormerod, F ., Biochem. Soc. Transactions, 6, 261 (1978). 11. Brocklehurst, K., Malthouse, J. P. 6., Biochem, J., 175, 761 (1978). 12. Schaffer, M. H., and Stark, G. R., Biochem. Biophys. Res. Comm, 71(4), 1040-1047 (1976). 13. Casey, R., and Lang. 4 14. Price N. C., Biochem. 15. Schaffer, M. H., and (1976). 16111, H. S., and Gracy. 17. Thomas, M. L., Janat USA, 79, 1054-1058 ( 18. Roberts, D. D., and G 19. Matsudaira, P., lakes, 82, 6788-6792 (1985) 20. Mahboub, S., Richard (1986). 21. Bahler, M., Benfenati 108,1841-1849(198‘. 22. Minard, P., Desmad Biochem., 185, 419-4 23' NefSky, 3., and Brets 24.13shdat, Y., Chapot, l\ 25' Katayama. E, J. Bioc 26. Sutherland, C., and V 27. Tang, 0, Yuksel, K. 283(1), 12.19 (1990). 101 13. Casey, R., and Lang, A., Biochem. J., 145, 251-261 (1975). 14. Price N. C., Biochem. J., 159, 177-180 (1976). 15. Schaffer, M. H., and Stark, G. R. Biochem. Biophys. Res. Comm. 71, 1040-1047 (1976) i 16. Lu, H. S., and Gracy, R. W., Arch. Biochem. Biophys., 212(2), 347-359 (1981). 17. Thomas, M. L., Janatova, J., Gray, W. R., and Tack, B. F ., Proc. Natl. Acad. Sci. USA, 79, 1054-1058 (1982). 18. Roberts, D. D., and Goldstein, I. J ., J. Biol. Chem, 259, 909-914 (1984). 19. Matsudaira, P., Jakes, R., Cameron, L., and Atherton, B, Proc. Natl. Acad. Sci. USA. 82, 6788-6792 (1985). 20. Mahboub, S., Richard, C., Delacourte, A., and Han, K., Anal. Biochem., 154, 171-182 (1986) 21. Bahler, M., Benfenati, F., Valtorta, F., Czemik, A. J ., and Greengard, P., J. Cell Biol, 108, 1841-1849 (1989). 22. Minard, P., Desmadril, M., Ballery, N., Perahia, D., and Mouawad, L., Eur. J. Biochem., 185, 419-423 (1989). 23. Nefsky, B., and Bretscher, A., Proc. Natl. Acad. Sci. USA, 86, 3549-3553 (1989). 24. Eshdat, Y., Chapot, M. P., and Strosberg, A. D., FEBS Letters, 1-2, 166-170 (1989). 25. Katayama, E., J. Biochem., 106, 988-993 (1989). 26. Sutherland, C., and Walsh, M. P., J. Biol. Chem, 264, 578-583(1989). 27. Tang, C., Yuksel, K. U., Jacobson, T. M., and Gracy, R. W., Arch. Biochem. Phys, 283(1), 12-19 (1990). 28. May, J. M., Buchs, A. 29. Wines, B. D., and Ear (1991) 30. 80111, S., and Friedman 31.Altarnirano, M. M., P (1992). 32. Goold, R., and Baines 33. Stark, G. R. in Metht (1977), Academic Pre 34. Papayannopoulos, 1.,A 35. Wu, J., Gage, D. A.,; 36.er3, J. A., Butler, (1991). 37.2e11uzec, E. J., Gag. Spectrom., 5, 230-23‘ 38. Denslow, N. D., and 11241. 39. Matsubara, H” and S 40. Matsubara, H., and S 102 28. May, J. M., Buchs, A. and Carter-Su, C., Biochem., 29, 10393-10398 (1990). 29. Wines, B. D., and Easterbrook-Smith, S. B., Molecular Immunology, 28(8), 855-863 (1991) 30. Som, S., and Friedman, S., J. Biol. Chem, 266, 2937-2945 (1991). 31. Altamirano, M. M., Plumbridge, J. A., and Calcagno, M. L., Biochem., 31, 1153-1158 (1992) 32. Goold, R., and Baines, A. J ., Eur. .J. Biochem., 224, 229-240 (1994). 33. Stark, G. R. in Methods Enzymol. (Jakoby, W. B., and Wilchek, M., Eds), vol. 47, (1977), Academic Press, San Diego, p. 129. 34. Papayannopoulos, LA, and Biemann, K., Protein Sci., 1, 278-288 (1992). 35. Wu, J ., Gage, D. A., and Watson, J. T., Anal. Biochem., 235, l61-174(1996). 36. Burns, J. A., Butler, J. c., and Whitesides, G. M., J. Org. Chem, 56, 2648-2650 (1991) 37. Zaluzec, E. J ., Gage, D. A., Allison, J ., and Watson, J. T., J. Am. Soc. Mass Spectrom., 5, 230-237 (1994). 38. Denslow, N. D., and Nguyen, H. P., in Techniques in Protein Chemistry V11, (1996), p.241. 39. Matsubara, H., and Sasaki, R.M., Proc. Natl. Acad. Sci. USA., 57, 439-445 (1967). 40. Matsubara, H., and Sasaki, R.M., J. Biol. Chem. 243, 1732-1757 (1968). 41. Chakrabarti, P., Biochem., 28, 6081-6085(1989). 42. Papov, V. V., and Biemann, K. Proceedings of the 42nd ASMS Conference on Mass Spectrometry and Allied Topics, Chicago, IL, May29-June3, (1994), pp.688. 43.Thibault, P., PleasanC< Mass Spectrom. Ion PI 44. Amon, R., in Methods Press, San Diego: "01' 45. Cohen, L. W., Coghlat 46. Mckenzie, H. A. R3 (1972). 41. Patterson, S., and Kati 48. Tompson, E.0.P., and 49.1akatera, K., and Wat 50. Nisbet, A.D., Saundry J. Biochem., 115, 335 51. Phelan, P., and Malthu 52. Voet, D., and Voet, J. 53. Kawamura, M., and 1 54'K09183berg, K., in M Academic Press, San 55.Za1uzec, E. J., Gage, (1994). 56- FiSCher, W. H., Rivit 225-228 (1993). 103 43. Thibault, P., Pleasance, S., Laycock, M.V., MacKay, R.M., and Boyd, R.K., Int. J. Mass Spectrom. Ion Proc., 111, 317-353 (1991). 44. Arnon, R., in Methods Enzymol. (Perlmann, G. E., and Lorand, L., Eds), Academic Press, San Diego, vol. 19, (1970), pp. 226-244. 45. Cohen, L. W., Coghlan, V. W., and Dihel, L. C., Gene, 48, 219-227(1986). 46. Mckenzie, H. A., Ralston, G. B., and Shaw, D. c., Biochem., 11(24), 4539-4547 (1972). 47. Patterson, S., and Katta, V., Anal. Chem, 66, 3727-3732(1994). 48. Tompson, E.O.P., and Fisher, W.K., Aust. J. Biol. Sci., 31, 433-442 (1978). 49. Takatera, K., and Watanabe, T., Anal. Chem, 65, 3644-3646(1993). ' 50. Nisbet, A.D., Saundry, R.H., Moir, A.J.G., Fothergill, LA, and Fothergill, J .E., Eur. J. Biochem., 115, 335-345 (1981). 51. Phelan, P., and Malthouse, J. P. G., Biochem. J., 302, 511-516 (1994). 52. Voet, D., and Voet, J. G., Biochemistry, John Wilet & Sons, (1990), p. 218-219. 53. Kawamura, M., and Nagano, K., Biochim. Biophys. Acta, 774, 188 ( 1984). 54. Konigsberg, K., in Methods in Enzymol., (Hirs, C. H. W., and Tirnasheff, S. N., Eds), Academic Press, San Diego,vol. 25, pp. 185. 55. Zaluzec, E. J., Gage, D. A., Watson, J. T., J. Am. Soc. Mass Spectrom. 5, 359-366 (1994) 56. Fischer, W. H., Rivier, J .E., and Craig, A.G., Rapid Commun. in Mass Spectrom. 7, 225-228 (1993). 57. Han, J. C., and Han, G. Y., Anal. Biochem., 220, 5-10 (1994). 58. Allen, G., in Lahore (Bordon, R. 11., and Kn 59.Torchinsky, Y. M., in Oxford, (1981). 60. May, J. M., Buchs. A.. 61.Tang, C. Y., Yuksel, Biophys., 283, 1.2-19 (l 62. Goold R, and Baines, 63.Vestling, MM, and Fe 64. Blacldedge, J .A., and A 104 58. Allen, G., in Laboratory Techniques in Biochemistry and Molecular Biology, (Burdon, R. H., and Knippenberg, P. H., Eds), Elsevier, Vol.9, (1989), pp. 318. 59. Torchinsky, Y. M., in Sulfur in Protein; (Metzler, D., Translation Ed.), Pergamon, Oxford, (1981). 60. May, J. M., Buchs, A., and Carter-Su, C., Biochemistry, 29, 10393-10398 (1990). 61. Tang, C. Y., Yuksel, U., Jacobson, T. M., and Gracy, R. W., Arch. Biochem. Biophys., 283, 12-19 (1990). 62. Goold, R., and Baines, A. J ., Eur. J. Biochem., 224, 229-240 (1994). 63. Vestling, M.M., and Fenselau, 0., Anal. Chem, 66, 471—477 (1994). 64. Blackledge, J .A., and Alexander, A.J., Anal. Chem, 67, 843-848 (1995). FURTHER STUD CYSTEINE RESIDU MS: 01". I. Introduction It has been known 1 the N-terminus of cysteine amino-tenninal peptide ant (1,2). The cleavage reactit 37°C for a prolonged pf Polypeptide chain at the cy bonds are turreactive to thi frequency of the occurrenc insize and can usually be electrophoresis (SDS-PAG CYanylated cysteine can a thiocyanate and dehydroal mildly alkaline conditions, a moderate degree of cleav near ly to completion for a P . atchonnk (9) found that d CHAPTER 3 FURTHER STUDY ON THE LOCALIZATION OF PROTEIN CYSTEINE RESIDUES BY CHEMICAL CLEAVAGE AND MALDI- MS: OPTIMIZATION AND APPLICATION I. Introduction It has been known for some time that peptide chains can selectively be cleaved at the N-terminus of cysteine residues after cyanylation of sulflaydryl groups to form an amino-terminal peptide and a series of 2-iminothiazolidine-4-carboxylyl (ITC) peptides (1, 2). The cleavage reaction can be carried out in mildly alkaline conditions (pH8-10) at 37°C for a prolonged period of incubation (12-80h). Selective cleavage of the polypeptide chain at the cyanylated cysteine residues may be achieved because disulfide bonds are unreactive to the cyanylation and cleavage. Because of the widely variable frequency of the occurrence of cysteine in proteins, the resulting fragments vary greatly in size and can usually be mass-mapped by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) (3-7) or more recently by mass spectrometry (2, 8). Cyanylated cysteine can also undergo a base-catalyzed B-elimination reaction to form thiocyanate and dehydroalanine. Since both cleavage and [J-elimination occur under mildly alkaline conditions, the two reactions are competitive, sometimes resulting in only a moderate degree of cleavage of a peptide bond. Although the cleavage seemed to go nearly to completion for a variety of proteins in Jacobson’s experiments, Degani and Patchornik (9) found that dehydroalanyl residues were formed in variable amounts from 105 the [l-elirnination of cyany pelimination over cleav: peptides. Higher pH was Nakagawa er a1 (10) rCC€ hormone, hPTH [1-84]. if biologically active or-amic' to produce with an 80% j recombinant techniques. been extensively used wit In chapter 2, we d inproteins using this spe TOF MS. This methodo SimPle, fast, and sensitive Proteins. One of the inhe reaction takes too long to exPosure to alkaline 3011 results. In this chapter, v function 0f PH, solvents. amino acids adjacent to show that a higher pH . ne . vertheless, 1n contras' 106 the B-elimination of cyanylated cysteine residues at pH 8-10. Increasing pH favored the B-elimination over cleavage, depending on the respective structural properties of peptides. Higher pH was therefore avoided in order to suppress the B-elimination. But Nakagawa et al (10) recently reported the cleavage of cyanylated human parathyroid hormone, hPTH [1-84], in a high concentration of ammonia in an attempt to produce a biologically active oc-amidated peptide. By the specific cleavage reaction they were able to produce with an 80% yield the or-amidated peptide which could not be produced by recombinant techniques. In all, the protocol (1, 9) developed by Jacobson et al (1) has been extensively used with little modification over the past two decades. In chapter 2, we described a new methodology to localize free sulfhydryl groups in proteins using this specific cleavage reaction followed by mass-mapping by MALDI- . TOF MS. This methodology is advantageous over conventional approaches in that it is simple, fast, and sensitive, and is especially useful for locating sulfhydryl groups in large proteins. One of the inherent problems with this procedure, however, is that the cleavage reaction takes too long to accomplish. Various side reactions may rise after long hours of exposure to alkaline solution, which brings about some uncertainty into the analytical results. In this chapter, we present our systematic studies on the cleavage reaction as a function of pH, solvents, and amino acid structures for polypeptides containing different amino acids adjacent to N-terminus of cyanylated cysteines. Our experimental results show that a higher pH can greatly promote both cleavage and B-elimination reactions, nevertheless, in contrast to the previous reports, for most of the peptides studied, the extent of B-elimination d Ammonia, a stronger nucl of peptide chains and min most of the peptides stur itithin an hour at room ter effects of peptide structur The optimized conditions in a variety of proteins wi H. Experimental Sectio Mass Spectrometry MALDI mass spe Spectrometer (PerSeptivf VSL-337ND nitrogen la the ion source was set t1 Time-to-mass conversio Standards of bradykinin skeletal myoglobin (m/z All experiments were Chemical Co, Milwauk In a 50% (v/v) solution 107 extent of B-elimination does not increase significantly even at a pH as high as 12. Ammonia, a stronger nucleophile than hydroxyl anions, greatly accelerates the cleavage of peptide chains and minimizes side reactions related to the prolonged incubation. For most of the peptides studied, the cleavage in l M NH4OH solution can be complete within an hour at room temperature. Based on the results from over a dozen peptides, the effects of peptide structure on the rate and yield of the cleavage reaction were evaluated. The optimized conditions have been applied to the recognition of free sulfhydryl groups in a variety of proteins with both known and unknown cysteine structures. II. Experimental Section Mass Spectrometry MALDI mass spectra were obtained on a Voyager Elite time-of-flight (TOF) mass spectrometer (PerSeptive Biosystems Inc., Framingham, MA) equipped with a model VSL-337ND nitrogen laser (Laser Science, Newton, MA). The accelerating voltage in the ion source was set to 20 kV. Data were acquired in the linear mode of operation. Time-to-mass conversion was achieved by external and/or internal calibration using standards of bradykinin (m/z 1061.2), bovine pancreatic insulin (m/z 5734.5), and horse skeletal myoglobin (m/z 16,952) obtained from Sigma Chemical Co. (St. Louis, MO). All experiments were performed using oc-cyano-4-hydroxycinnamic acid (Aldrich Chemical Co., Milwaukee, WI) as the matrix. Saturated matrix solutions were prepared in a 50% (v/v) solution of acetonitrile/aqueous 0.1% TFA, and mixed in equal volumes with peptide or protein 1 mixture was allowed to air Chemicals Guanidine hydroc (Indianapolis, IN). l-C peptides TCVEWLRRYI ovalbumin, and rabbit 111 used without further puri DRVYIHPCHLLYYS, .4 American Peptide Com; California Inc. The 0.10 HCl was freshly prepare prepared in 0.1 M citrate Cyanylation of SH Gro To the peptide 2 cyano-4-dilnethylamino- sulfhydryl content. Cya 1 00m temperature for 1l reversed-phase HPLC l manually, the masses of 108 with peptide or protein samples, and applied to a stainless-steel sample plate. The mixture was allowed to air dry before being introduced into the mass spectrometer. Chemicals Guanidine hydrochloride was a product of Boehringer-Mannheim Biochemicals (Indianapolis, IN). 1-Cyano-4-dimethylamino-pyridinium tetrafluoroborate (CDAP), peptides TCVEWLRRYLKN and RYVVLPRPVCFEKGMNYTVR, spinach ferredoxin, ovalbumin, and rabbit muscle creatine phosphokinase were purchased from Sigma and used without further purification. Acetonitrile and TFA were of HPLC grade. Peptides DRVYIHPCHLLYYS, ALLETYCATPAKSE, and SLRRSSCFGGR were products of American Peptide Company. The rest of the peptides were purchased from Bachem California Inc. The 0.10 M CDAP solution in pH 3.0, 0.1 M citrate buffer-4M guanidine- HCl was freshly prepared prior to use. The 1 mM protein and peptide solutions were prepared in 0.1 M citrate buffer, pH 3.0, containing 4 M guanidine-HCI. Cyanylation of SH Groups To the peptide and protein solutions was added a 10-fold molar excess of 1- cyano-4-dimethylamino-pyridinium tetrafluoroborate (CDAP) solution over the sulfllydryl content. Cyanylation of sulfhydryl groups was accomplished by incubation at room temperature for 10-15 min. The modified proteins or peptides were purified by reversed-phase HPLC under gradient elution. The HPLC fractions were collected manually, the masses of which were determined by MALDI-MS. Those corresponding to the cyanylated proteins at further use. Cleavage of the Cyanl’la The Cyanylated p6 4M Mme-HG solut: equal volume of 4. M gut 0.25 M tris-HCl buffer, t The pH 8.0 and pH 9.0 \ the cleavage reaction, w' ambient temperature for were taken and diluted tc MS. The rest of the so HPLC fractions were co of the cleavage and B-eli under the peaks identifle The cleavage c PhOSPhOkinase at a cone tris-HCI buffer, pH 9.0, NHJOH solution-4 M g Aliquots of 1 pl of th CH3CN/0.1%TFA for ' 109 the cyanylated proteins and peptides were dried in a speed vac and kept in a freezer for fiirther use. Cleavage of the Cyanylated Peptides and Proteins The cyanylated peptide samples were reconstructed to a concentration of 1 mM in 4 M guanidine-HCI solution. A portion of the above solutions (~5 pl) was mixed with an equal volume ’of 4 M guanidine-HCI solution in pH 8.0, 0.25 M tris-HCl buffer, pH 9.0, 0.25 M tris-HCI buffer, 0.02 M NaOH (pH 12), and 2 M NH4OH solution, respectively. The pH 8.0 and pH 9.0 vials were incubated in 37°C water bath for 18 hours to promote the cleavage reaction, while the vials with NaOH and NH4OH solutions were placed at ambient temperature for 1 hour. After the reaction, l-ul aliquots of the above solutions were taken and diluted to 100 ill with 1:1 CH3CN/0. 1%TF A for analysis by MALDI-TOF MS. The rest of the solutions was acidified and injected into an HPLC column. The HPLC fractions were collected and identified by MALDI-TOF MS. The relative yields of the cleavage and B-elimination products were calculated by integrating the HPLC areas under the peaks identified by MALDI. The cleavage of cyanylated proteins, ferredoxin, ovalbumin, and creatine phosphokinase at a concentration of 0.5 mM, was carried out by incubation both in 0.1 M tris-HCl buffer, pH 9.0, containing 4 M guanidine—HCI at 37°C for 18 hours and in 1 M NH4OH solution-4 M guanidine-HCI at ambient temperature for one hour, respectively. Aliquots of 1 pl of the above solutions were taken and diluted to 100 pl with 1:1 CH3CN/0.1%TFA for analysis by MALDI-TOF MS. The rest of the solution was acidified and then injectet identified by MALDI-TO HPLC Separation The purification cleavage products were using Waters model 6001 11m. The column used th size, 300-A pore, 4.6x2. solution and CH3CN co: from 5-50% B in 50 r conditions for individua major HPLC fractions . determined by MALDI-l HI. Results and Discus A‘ Cyanylation of SH The cJ'anylation tluocyanobenzoic acid ( was believed to be spe fo ' tmatlon of mixed disr 110 acidified and then injected to the HPLC column. The HPLC fractions were collected and identified by MALDI-TOF MS. ' HPLC Separation The purification of the cyanylated proteins and peptides and the separation of cleavage products were achieved by reversed-phase HPLC with linear gradient elution using Waters model 6000 pumps controlled by a PC computer. UV detection was at 215 nm. The column used throughout the study was a Vydac C18 (#218TP54, 10 um particle size, 300-A pore, 4.6x250 mm). Mobile phases A and B contain 0.1% TFA aqueous solution and CH3CN containing 0.1% TFA, respectively. Typically, a gradient elution from 5-50% B in 50 minutes was applied in preliminary experiments. The HPLC conditions for individual proteins and peptides were then modified and optimized. The major HPLC fractions were collected manually, and the masses of the fractions were determined by MALDI-MS. III. Results and Discussion A. Cyanylation of SH Groups by CDAP The cyanylation of sulfliydryl groups was traditionally accomplished by 2-nitro—5- thiocyanobenzoic acid (NTCB) under mildly alkaline conditions (pH 8-10). The reagent was believed to be specific to sulfliydryl groups, although side reactions, such as the formation of mixed disulfide bonds between the NTCB and protein SH groups, were also reported (9. 11, 12)‘ Sine under alkaline conditions be difficult even though t 1-cyano-4-dimethylamin‘ cyanylation of SH group reagent has not been user The CDAP is ad\ groups ill the presence c minimize sulfhydryl/(iii peptides are stable und subsequent cleavage rear showed that complete cy molar excess of the rea temperature. Our study the published results. I excess of CDAP (~50-f POUIS) could result in t1 Product can be separ. implicating the notable M988 analysis of the 311 corresponding Cyanylatt b y CDAP needs to be reac ' tron. After extensir 111 reported (9, 11, 12). Since both the cyanylation and the consequent cleavage are achieved under alkaline conditions, the separate study of cyanylation and cleavage reactions might be difficult even though the reaction conditions are carefully controlled. Another reagent, 1-cyano-4-dimethylamino-pyridinium tetrafluoroborate (CDAP), was proposed for the cyanylation of SH groups under slightly acidic conditions (pH 3-7) (14). However, this reagent has not been used as extensively. as the NTCB. The CDAP is advantageous over the NTCB for specific cyanylation of protein SH groups in the presence of disulfide bonds, because the acidic conditions can effectively minimize sulfhydryl/disulfide exchange. Furthermore, the cyanylated proteins or peptides are stable under acidic conditions, the kinetics of the cyanylation and the subsequent cleavage reaction can be studied independently. Previous papers (10, 14, 15) showed that complete cyanylation by the CDAP can be carried out using three to five fold molar excess of the reagent over free SH groups under very mild conditions at room temperature. Our study with sulfhydryl-containing peptides listed in Table 3.1 confirmed the published results. Further experiments with the model peptides showed that a large excess of CDAP (~50-fold over peptide SH groups) and excessive incubation time (> 2 hours) could result in the formation of an unidentified side reaction product. The side product can be separated from the corresponding cyanylated peptide on HPLC, implicating the notable differences between the hydrophobicity of the two products. Mass analysis of the side product by MALDI shows a mass increase of 43 Da from the corresponding cyanylated peptide. Therefore, like cyanylation by NTCB, the cyanylation by CDAP needs to be performed under controlled conditions to minimize the side reaction. After extensive examination with different peptides, we optimized the reaction Table 3.1. Calculated cleavage reaction amino acid sequenc PHCKRM RGPCRAFI SRNRCNDQ SLRRSSCFGGR MSRPACPNDKYE YK'ITICGKGLSATV EKPLQNFTLCFR ALLETYCATPALSE TCVEWLRRYLKN DRVYIHPCHLLYYS KRNPGSNKRFPSNCGI RYVVLPRPVCFEKGM MHRQBAVDCLKKFNJ 112 Table 3.1. Calculated m/z values of fragments for sulfliydryl-containing peptides after the cleavage reaction IM+HJ + amino acid sequence M.W. , , , fragment I fragment II B-ellmlnatlon PHCKRM 771.0 253.3 562.7 738.0 RGPCRAFI 919.3 329.3 636.0 887.3 SRNRCNDQ 991.2 532.6 503.6 958.2 SLRRSSCFGGR 1225.7 705.9 564.8 1192.7 MSRPACPNDKYE 1410.0 561.7 892.3 1377.0 YK’I‘TICGKGLSATV 1441.9 625 .8 861.1 1407.9 EKPLQNFTLCFR 1493.9 1088.4 450.5 1460.9 ALLETYCATPALSE 1496.6 709.7 831.9 1463 .6 TCVEWLRRYLKN 1580.9 120.1 1506.8 . 1547.9 DRVYIHPCHLLYYS 1779.4 900.1 924.3 1746.4 KRNPGSNKRFPSNCGRD 1947.3 1517.8 476.5 1914.3 RYVVLPRPVCFEKGMNYTVR 2427.9 1099.5 1373.5 2394.9 MHRQEAVDCLKKFNARRKLKGA 2600.4 986.1 1659.3 2567.4 conditions: the cyanylati CDAP (over sulfhydryl Under these conditions reaction was observed. B. Optimization of Cl Although the sel of cyanylated cysteine r today are almost the (1). Degani and Patcho reactions and the effects elimination are both bas while B-elimination seer conditions (pH 8-9) W incubation time differs . hours, depending on the also affect the rate of Cl cleavage, no appreciab conclusion made by Dr dipeptides and tripeptidt model peptides represe 113 conditions: the cyanylation was carried out by treatment with a 10-fold molar excess of CDAP (over sulfhydryl groups) in pH 3.0 buffer for 15 min at ambient temperature. Under these conditions the cyanylation reaction went to completion and very little side reaction was observed. B. Optimization of Cleavage Reactions Although the selective chemical cleavage of peptide chains at the N-peptide bonds of cyanylated cysteine residues was proposed long ago, the experimental conditions used today are almost the same as those described in the protocol developed by Jacobson et al (1). Degani and Patchornik (9) studied the pH [dependence of cleavage and B-elimination reactions and the effects of amino acid structures on both reactions. The cleavage and [3- elimination are both base catalyzed and therefore competitive under alkaline conditions, while B-elimination seems to be favored over cleavage at higher pH. Mildly alkaline conditions (pH 8-9) were suggested for the cleavage reaction in early study. The incubation time differs significantly under the above conditions, ranging from 16 to 80 hours, depending on the structures of proteins under study. Structural features of proteins also affect the rate of cleavage reaction (16). Proteins have to be denatured prior to the cleavage, no appreciable cleavage was observed for native proteins. However, the conclusion made by Degani and Patchornik was based on the study on a few simple dipeptides and tripeptides with special structural factors. The behavior of cysteines in the model peptides represents to a less extent general structural features of cysteine residues in peptides and proteins. better understanding and We have system filnction of pH and reac amino acids adjacent to were first cyanylated an cyanylated peptides, afi 12.0, and 1 M NILOH The cleavage mixtures fractions identified by side reactions were eval Results of the k cleavage can be accom within one hour at amt 37°C in pH 8-9 buffers. the greatest concerns e: favor B-elimination ow However, our study on that it is not the case. . differ significantly fo: conditions. Some of 114 in peptides and proteins. Further examination of the cleavage reaction is needed for the better understanding and optimization of the cleavage reaction. We have systematically studied the rate and yield of the cleavage reaction as a function of pH and reaction medium for a variety of polypeptides containing different amino acids adjacent to the cyanylated cysteines. The peptides under study (Table 3.1) were first cyanylated and the completion of cyanylation was monitored by MALDI. The cyanylated peptides, after purified by HPLC, were subjected to cleavage at pH 8.0, 9.0, 12.0, and 1 M NH4OH (pH~11) solutions containing 4 M guanidine-HCI, respectively. The cleavage mixtures were separated by HPLC, the fractions collected, the masses of the fractions identified by MALDI. The relative yields of the cleavage products and other side reactions were evaluated by measuring HPLC peak areas. Results of the kinetics study showed that, for most of the peptides studied, the cleavage can be accomplished in both ammonia solution and 0.01 M NaOH solution within one hour at ambient temperature, whereas it requires 18 hours of incubation at 37°C in pH 8-9 buffers. Although elevated pH can accelerate cleavage reactions, one of the greatest concerns expressed in previous papers was that a higher pH would largely favor B-elimination over cleavage, causing a lower yield of the cleavage reaction. However, our study on the relative yields under various pH and reaction media indicated that it is not the case. As shown in Table 3.2, the relative yields of B-elimination do not differ significantly for most of the peptides studied under the four experimental conditions. Some of the peptides show even higher cleavage yields in 1 M NH4OH or Table 3.2. Relative PHCKRMd RGPCRAFI SLRRSSCFGGR MSRPACPNDKYE DRVYIHPCHLLYYS KRNPGSNKRFPSNC< RYVVLPRPVCFEKGI MHRQEAVDCLKKFI‘ ‘ in each case,B-elirnin " side reactions other t1 ° peptide undergoes ap‘ d incomplete conversio 115 Table 3.2. Relative yields (%) of B-elimination under different cleavage conditions pH 8.0, pH 9.0, pH 12 , 1M NH 40H, amino acid sequence 37°C,18h 370C,18h rt, 1h rt, 1h PHCKRM“ 9.9a 15.9 25.1 14.8 RGPCRAFI 31.6b 87.5 82.5 70.1 SLRRSSCFGGR 2.8 2.9 3.9 4.6 MSRPACPNDKYE 6.1 5.8 1 1.3 3.9 YK'IVI‘ICGKGLSATV 16.8 18.2 22.2 20.3 EKPLQNFTLCFR 15.2c 29.9 49.8 20.9 ALLETYCATPALSE ~100 ~100 ~100 ~100 TCVEWLRRYLKN 20.9 13.8 13.1 18.3 DRVYIHPCHLLYYS 56.2 90.7 98.5 93.8 KRNPGSNKRFPSNCGRD 5.0 5.0 5.0 5.0 RYVVLPRPVCFEKGMNYTVR 34.3 41.2 52.4 51.5 MHRQEAVDCLKKFNARRKLKGA 12.8 1 1.1 18.4 13.2 3‘ in each case,B-elimination is a main side reaction b side reactions other than B-elirnination and dimerization are observed ° peptide undergoes appreciable dimerization d incomplete conversion 11,1 11.14;. Figure 3.1. (B) Peptide 116 SH l O CH 0 ll 2 n //0 H3N+— - —C -NH—CH —C —NH— - —C \ - O (A) Cyanylation CD AP M N-<——\N+—CN pH3-7; rt, 15' ’ 62 9 _ BF4 V NEC —S O \CH O n l 2 I //0 H3N+ — —C —NH-CH —C —NH — —c \ - O :NH3 (B) Cleavage 1 M NH4OH, ~1h, rt. v O HN S // O H3N+ *' —C\ + Y 11 //0 NH2 HN c—NH—- —fc\ 0' Figure 3.1. (A) Cyanylation of sulfliydryl group by CDAP and (B) peptide bond cleavage catalyzed by ammonia NaOH solutions. The?" reaction under stronger a] The mechanism c 2) indicates that the hydi rising pH on the rate of nucleophilic attack of ( mechanism, a stronger condition that the size nucleophilic group to t requirement. Ammonia the hydroxyl anion, the . carbonyl carbon would as shown in Figure 3.1. hydroxyl anion catalyze Structure replaces carbo: Although comp; and 1 M NH40H solutit 1 M NH4OH solution p 0'01 M OH‘ is available agent will greatly accel 1s a better nucleophile t anion in terms of the c 117 NaOH solutions. These results demonstrate the promise to carry out the cleavage reaction under stronger alkaline conditions without sacrificing the cleavage yields. The mechanism of cleavage reaction proposed by Jacobson et al (1) (see chapter 2) indicates that the hydroxide ion catalyzes the cleavage, as shown by the influence of rising pH on the rate of cleavage reaction. The reaction probably proceeds only after nucleOphilic attack of OH‘ on the carbonyl carbon of the amide. According to the mechanism, a stronger nucleophile should promote the cleavage reaction, given the condition that the size of the nucleophilic group is small so that the attack of the nucleophilic group to the carbonyl atom is sterically feasible. Ammonia meets the requirement. Ammonia has a stronger nucleophilic property than hydroxyl anion. Like the hydroxyl anion, the size of ammonia is small. Nucleophilic attack of ammonia to the carbonyl carbon would facilitate the cleavage giving an or-amidated N—terminal peptide, as shown in Figure 3.1. Therefore, the N-terminal fragments obtained by ammonia and hydroxyl anion catalyzed cleavage have one mass difference in mass spectrometry (amide structure replaces carboxyl structure). Although comparable results were obtained for peptide cleavage in 0.01 M NaOH and 1 M NH4OH solutions, the latter is preferred as it provides several advantages. First, 1 M NH4OH solution provides much more nucleophilic species (:NH3 ~1 M), while only 0.01 M OH’ is available in pH 12 NaOH solution. The high concentration of nucleophilic agent will greatly accelerate the concerted cleavage reaction. Second, although ammonia is a better nucleophile than hydroxyl anion, it has a smaller proton affinity than hydroxyl anion in terms of the capability of catalyzing B-elimination reaction. Therefore, the B- elimination is anticipated from our experimental 1 NaOH solution are high solution containing 4 M chain is more stable ar Finally, the excess v01: cleavage. The representatix i111 M NH40H solutior gave comparable result elimination reactions a1 typical MALDI spectra NH40H solution. Ther amain competitive sidc observed. One such mi flee SH group again, rePorted by others (9, 1 Which probably arises grOUps. However, th neEligible. 118 elimination is anticipated to be minor in 1 M NH4OH. This tendency has been supported from our experimental results (Table 3.2) that the yields of B-elimination in 0.01 M NaOH solution are higher than that in 1 M NH4OH solution. Third, the 1 M NH4OH solution containing 4 M guanidine-HCI has a moderate pH (~11), at which the peptide chain is more stable and alkali-induced damage of peptide chains can be prevented. Finally, the excess volatile ammonia can easily be removed from the mixture after cleavage. The representative chromatograms of peptides, after cleavage in pH 9.0 buffer and in 1 M NH4OH solution, are presented in Figure 3.2 to Figure 3.5. The two conditions gave comparable results in terms of relative yields of competitive cleavage and B- elimination reactions and other possible side reactions. Figure 3.6 to Figure 3.9 show typical MALDI spectra of these peptides, after cleavage in pH 9.0 buffer and in 1 M NH4OH solution. The results from both HPLC and MALDI confirm that B-elimination is a main competitive side reaction with cleavage reaction. Other side reactions were also observed. One such minor reaction was the conversion of cyanylated peptides to form a free SH group again, indicating the reversibility of the cyanylation reaction as also reported by others (9, 10). Another minor side reaction was the dimerization of peptides, which probably arises from nucleophilic attack of nascent thioate anions on thiocyano groups. However, these two side reactions were always minor and in most cases negligible. Figure 3.2. HPLCt in (A) pH 9.0, 37°C 1,11, [1, CM, and D uncleaved species, 119 11 A I L L B d 1 ,. __ 11 B 1 grimy” 5 1'0 1'5 2? 25 min Figure 3.2. HPLC chromatograms of cleavage products of SLRRSSCFGGR in (A) pH 9.0, 37°C, 18h, and (B) 1 M NH4OH, rt, 1h. The peaks marked with I, II, B, C/U, and D are fragment 1, fragment II, B-elimination product, cyanylated/ uncleaved species, and peptide dimer, respectively. 120 U 10 15 20 25 30 Figure 3.3. HPLC chromatograms of cleavage products of DRVYIHPCHLLYYS in (A) pH 9.0, 37°C, 18h, and (B) 1 M NH4OH, It, 1h, respectively. A II M 1 w i3 B I H CH] WW j i j I 35 Figure 3-4- HPLt 1n(A111H 9.0, 37 121 II (in; CM Figure 3.4. HPLC chromatOgrams of cleavage products of EKPLQNFTLCFR I 20 min 25 ' 30 in (A) pH 9.0, 37°C, 18h, and (B) 1 M NH4OH, rt, 1h, respectively. “We 35 HPLt MHRQEAVDCL 40H, rt, 1h, r- 122 II C/U 1 II C/U WA I I I I 1 I I 6 8 10 12 14 16 18 20 Figure 3.5., HPLC chromatograms of cleavage products of peptide MHRQEAVDCLKKFNARRKLKGA in (A) pH 9.0, 37°C, 18h, and (B) 1 M NH4OH, rt, 1h, respectively.- b 500 1000 Figure 3.6. MALDI M) pH 9.0, 37°C, 1,11, D, CM, and B I'C uncleaved peptide, ar N ++III 111 +“\/ 1000 Figure 3.7. cleavage in (A) pH Symbol 4+ represer 123 11 A B C/U 2.111221 h_/. 11 B [3 9U “Jinn 41 m 500 , 1000 1500 2600 2500 3000 m/z Figure 3.6. MALDI spectra of the cyanylated peptide SLRRSSCFGGR, after cleavage in (A) pH 9.0, 37°C, 18h and (B) 1 M NH4OH solution, rt, 1h, respectively. Symbols 1, H, D, C/U, and [3 represent fragment 1, fiagment II, peptide dimer, cyanylated/ uncleaved peptide, and its B-elimination product, respectively. [3 A 11 ++ HIL/ when» ' B B I 11 _ “in mnl L 500 1000 1500 m2 2000 2500 3000 Figure 3.7. MALDI spectra of the cyanylated peptide DRVYIHPCHLLYYS,'after cleavage in (A) pH 9.0, 37°C, 18h and (B) 1 M NH4OH solution, rt, 1h, respectively. SYmbol +1- represents a doubly charged species. V, .L__ l g 500 1001 Figure 3.8. MALDI sp in (A) pH 9.0, 37°C, 1: IP represents the intac \ ++I ++ M 1000 Figure 3.9, after cleavage in (A) 124 11 At B IP 2’ .. .._. r .1 IL 11 o B 115: I ethnrn ,m_ r- W I , 560 _ 1060 1500 . 2600 2500 3000 m/z Figure 3.8. MALDI spectra of the cyanylated peptide EKPLQNFTLCFR, after cleavage in (A) pH 9.0, 37°C, 18h and (B) 1 M NH4OH solution, rt, 1h, respectively. Symbol IP represents the intact peptide. II +.+_-L. _ --JiLr M 500 10b0 (goo 2660 2560 3000 m/z Figure 3.9. MALDI spectra of the cyanylated MHRQEAVDCLKKFNARRKLKGA, after cleavage in (A) pH 9.0, 37°C, 18h and (B) 1 M NH4OH solution, rt, 1h, respectively. C. Effects of Amino 1" Degani and Patc rates of B-elimination dinitrophenylcysteine Negatively charged earl modified cysteine resid' take place, retard the 1 esterified amino acid a rate, presumably becau: the functional groups encountered if the cyar C-terminal side. In mt the cyanylated cystein Slmctural diversity on acids. This conclusic adjacent to the cyany undergo B-elimination r’v‘Sidues may have the The structural We“ Significantly t denatured peptides ex with rigid or bulky s 125 C. Effects of Amino Acid Structures on Cleavage and B-Elimination Reactions Degani and Patchornik (9) showed that the effects of amino acid structures on the rates of B-elimination closely resemble those found in the B-elimination of S-2,4- dinitrophenylcysteine compounds whose B-elimination is also base-catalyzed. Negatively charged carboxylate groups located in the vicinity of the C-oc hydrogen of the modified cysteine residue, where the attack of the catalyzing hydroxyl ion is supposed to take place, retard the rate of the elimination. On the other hand, the presence of an esterified amino acid adjacent to the cysteine residue greatly enhances the elimination rate, presumably because of the electron-withdrawing effect of the ester group. However, the frmctional groups mentioned in Degani and Patchornik’s (9) examples are rarely encountered if the cyanylated cysteine residue in peptides does not locate in either N— or C-terminal side. In most cases, the group located in the vicinity of the C-oc hydrogen of the cyanylated cysteine is an amide back bone of peptide chain, which imposes little structural diversity on the C-oc hydrogen, regardless of structural features of the amino acids. This conclusion was supported by the observation that the same amino acid adjacent to the cyanylated cysteine presents considerable variation in the tendency to undergo B~elimination, whereas different amino acids adjacent to cyanylated cysteine residues may have the same relative yield (Table 3.2). The structural features of the amino acid on the N-terminal side of cyanylated cysteine significantly affects both rate and yield of the cleavage reaction. For most of the denatured peptides examined, the cleavage products are predominant. But amino acids with rigid or bulky side chains, such as Pro and Tyr, are more resistant to cleavage, giving B-elimina‘ion as Son and Friedman,S r cySteine residues after C the reaction, in which tl atom to provide the req SCN group must be ori nitrogen atom of the c steric hindrance 0f the residue, resulting in ‘ neighboring amino ac reaction, which was de itith the identical amii other hand, is less dept cysteine. However, tendency and the rela variation in pH has Products, whereas di tendency to form cle dominates the relative It should be n. 22 amino acid residue role in the cleavage e 126 giving B-elimination as a main product (Table 3.2). This observation is consistent with Som and Friedman’s report (17) that the -Pro-Cys- was protected from cleavage at cysteine residues after cyanylation. This is understandable in terms of the mechanism of the reaction, in which the hydroxyl anion (or ammonia) must attack the scissile carbonyl atom to provide the required anionic tetrahedral intermediate, and the carbon atom of the SCN group must be oriented appropriately for nucleophilic attack by the activated amide nitrogen atom of the cyanylated cysteine. Both steps are significantly retarded by the steric hindrance of the side chain of the amino acid adjacent to the cyanylated cysteine residue, resulting in the decrease of the cleavage reaction rate. In additional to neighboring amino acids, the peptide’s solution structure also affects the cleavage reaction, which was demonstrated by the slightly different yields obtained from peptides with the identical amino acid on the N-tel’minal side of cysteine. B-Elimination, on the other hand, is less dependent on the structure of the amino acid adjacent to the cyanylated cysteine. However, both cleavage and B-elimination reactions show the same pH tendency and the relative rate can be partially compensated. This explains why the variation in pH has little effect on the relative ratio of cleavage and B-elirnination products, whereas different neighboring amino acids give a large variation in the tendency to form cleavage products. It is the structure of the peptide, not pH, that dominates the relative rate and yield of cleavage and B-elimination reactions. It should be noted that the data presented here are deduced from peptides with 6- 22 amino acid residues. For large proteins, solution structures may play a more important role in the cleavage, even if the protein is denatured. Many exceptions may be observed. 1), Application of Opt: The Optimized peptide chains at cya methodology was exam in proteins with diff phosphokinase were ev Figure 3.10 sho NH,0H solution and i1 “CXperimental Section’ 47, and 77, respective Experimentally, we or (m/z 320.3 and 517.5 cleavage products und 0f incubation at 37°C 1 an overlapped peptide not observed if the cl 1820.1 in pH 9.0 bufi 1.17 (tn/z 1839.1) du dehydration did not 01 One of the gt: a nd subsequent clear D. Application of Optimized Methodology The optimized conditions described above greatly facilitated the cleavage of peptide chains at cyanylated cysteine residues. The feasibility of the improved methodology was examined by experiments carried out to localize free sulflrydryl groups in proteins with different cysteine status. Ferredoxin, ovalbumin, and creatine phosphokinase were evaluated under different cleavage conditions to compare the results. Figure 3.10 shows the MALDI spectra of spinach ferredoxin after cleavage in 1. M NH4OH solution and in pH 9.0 buffer under the experimental conditions described in the “experimental Section”. Ferredoxin contains five free sulfhydryls at positions 18, 39, 44, 47, and 77, respectively (18). Six fragments are anticipated after complete cleavage. Experimentally, we could detect by MALDI four out of six, two low-mass fragments (m/z 320.3 and 517.5, respectively) were missing. The mass spectra obtained from cleavage products under both conditions show a similar pattern. However, after 18 hours of incubation at 37°C in pH 9.0 buffer, a peak at m/z 3584.0 was observed, attributable to an overlapped peptide 44-76 with B-elimination at Cys47. This overlapped peptide was not observed if the cleavage was performed in 1 M NH,OH solution. The peak at m/z 1820.1 in pH 9.0 buffer is probably attributable to the dehydration products of fragment 1-17 (m/z 1839.1) during prolonged exposure to the alkaline conditions (2), while this dehydration did not occur in l M NH4OH solution. One of the great advantages of the proposed methodology is that the cyanylation and subsequent cleavage are specific to free cysteine residues, selective cleavage of 128 18 39 44 47 77 1-'-QC|3P---SCR---SCIJSSCA---TC'A---97 SH SH SH SH SH Fragment Expected m/z 1-17 1839.1 18-38 2352.2 39—43 517.5 44-46 320.3 47-76 3337.6 77-97 2332.5 2332” 18-43 2792.0 44-76 3580.8 2347.7 3337.6 / A g :2 53 E? 4.5, quantitative cyanylation of SH groups thus may not be achieved in cases where cyanylation is slow due to structural factors of proteins. Another unexplored cyanylation reagent, thiocyanopyridine (TCP) (23), may provide advantages over the CDAP and NTCB. The TCP, like NTCB, can cyanylate sulfhydryl groups though the exchange of was- (2,2'-dipj Figure 3.14. c] thiocyanopyrid 137 @—S—S—NO_ ———>Rss—-@ + HS—@ (2,2'-dipyridyldisulfide) (thiol-form) H/ (thione-form) l RSH + NCS—@ ————> RSCN + HS—@ (thiocyanopyridine) (thiol-form) Figure 3.14. Chemical modification of SH groups by 2 ,2 d—ipyridyldisulfide and thiocyanOpyridine, respectively. the cyano group- BU‘ reactive to SH group expectation is deduo dipyridyldisulfide (2-1 with thiol groups in a resonance structure of compensated for by protonation of its ring shares the same struct (Figure 3.14). The potentially powerful this reagent is retarde In all, the use reaction conditions : Siutpler and more pla 138 the cyano group. But we expect the TCP, due to its structural features, will be more reactive to SH groups and can be used in lower pH than the NTCB reaction. This expectation is deduced from the comparison of structural similarity between 2,2’- dipyridyldisulfide (2-PDS) and TCP (Figure 3.16). The former can quantitatively react with thiol groups in a wide pH range (pH 1-9), as the releasing group is stabilized by resonance structure of the thione. The low nucleophilicity of thiol groups at acidic pH is compensated for by an increased electrophilicity of the 2-PDS as a result of the protonation of its ring nitrogen (pKa ~3). The releasing group of TCP after cyanylation shares the same structure as the product of 2-DPS, both having the stable thione structure (Figure 3.14). The structural similarity between 2-PDS and TCP makes TCP a potentially powerful reagent for cyanylation. Unfortunately, the further exploration of this reagent is retarded at this moment, because the TCP is not commercially available. In all, the use of a new cyanylation reagent, together with the optimized cleavage reaction conditions and new membrane techniques, is promising to provide an even simpler and more plausible application of our cyanylation methodology. VI. References 1. Jacobson, G. R., E 248, 6583-6591 ( lS 2. Wu, J., Gage, D. A U) . Nefsky, B., and BI 4. Thomas, M. L., J USA, 79, 1054-10 S. Roberts, D. D.,an 6. Matsudaira, F., 13 82, 6788-6792(19 7. Mahboub, S., Ric 182(1986). 8. Papayannopoulos 9- Degani, Y., andl 10. Nakagawa, S., T; Nishimura, 0., F‘ 11.DeI1slow, N, D, Academic Press, 12. Price, N. C., BiOt 13.Liao, 1H,, and r 14. Wakselrnan, M., 15. Wu, J., and Wat: VI. References 1. 10. 11. 12. 13. 14. 15. Jacobson, G. R., Schaffer, M. H., Stark, G. R., and Vanaman, T. C., J. Biol. Chem. 248, 6583-6591(1973). Wu, J., Gage, D. A., and Watson, J. T., Anal. Biochem., 235, 161-174(1996). . Nefsky, B., and Bretscher, A., Proc. Natl. Acad. Sci. USA, 86, 3549-3553(1989). Thomas, M. L., Janatova, J ., Gray, W. R., and Tack, B. F ., Proc. Natl. Acad. Sci. USA, 79, 1054-1058(1982). Roberts, D. D., and Goldstein, I. J ., J. Biol. Chem, 259, 909-914(1984). Matsudaira, P., Jakes, R., Cameron, L., and Atherton, B, Proc. Natl. Acad. Sci. USA. 82, 6788-6792(1985). Mahboub, S., Richard, C., Delacourte, A., and Han, K., Anal. Biochem., 154, 171- 182(1986). Papayannopoulos, LA, and Biemann, K., Protein Sci., 1, 278-288(1992). Degani, Y., and Patchornik, A., Biochemistry, 13, 1-11(1974). Nakagawa, S., Tamakashi, Y., Hamana, T., Kawase, M., Taketomi, S., Ishibashi, Y., Nishimura, 0., Fukuda, T., J Am. Chem. Soc, 116, 5513-5514(1994). Denslow, N. D., and Nguyen, H. P., (in Techniques in Protein Chemistry VII, Academic Press, Inc., (1996), p.241-248. Price, N. C., Biochem. J., 159, 177-180(1976). Liao, T-H., and Wadano, A., J Biol. Chem, 254(19), 9602-9607(1979). Wakselman, M., and Guibe-Jampel, 13., J. C. 3. Chem. C0mm.,21-22(1976). Wu, J ., and Watson, J. T., Protein Sci., 6, 391-398(1997) 16. Stark, G, R., in Me 17. Som, S., and Fried: 18. Matsubara, H., and 19. Nisbet, A.D., Saur J. Biochem., 115, I 20. Putney, S., Herli Belagaje, R., Bier 259(23), 14317-14 21.Nakagawa, S., Ta 0., Fukuda, T., Bi 22. Preston, L. M., h 550(1993). 23. Brocklehurst, K., As Mushiri, M, S 140 16. Stark, G, R., in Methods Enzymol., 47, 129-132(1977). 17. Som, S., and Friedman, S., J. Biol. Chem, 266(5), 2937-2945( 1991). 18. Matsubara, H., and Sasaki, R.M., J. Biol. Chem, 243, 1732-1757(1968). 19. Nisbet, A.D., Saundry, R.H., Moir, A.J.G., Fothergill, L.A., and Fothergill, J.E., Eur. J. Biochem., 115, 335-345(1981). 20. Putney, S., Herlihy, W., Royal, N., Pang, H., Aposhian, H. V., Pickering, L., Belagaje, R., Biemann, K., Page, D., Kuby, S., and Schimmel, P., J. Biol. Chem, 259(23), 14317-14320(1984) 21. Nakagawa, S., Tamakashi, Y., Ishibashi, Y., Kawase, M., Taketomi, S., Nishimura, 0., Fukuda, T., Biochem. Biophys. Res. Commun., 200, 1735-1741(1994). 22. Preston, L. M., Murray, K. K., and Russell, D. H., Biol. Mass Spectrom., 22, 544- 550(1993). 23. Brocklehurst, K., Malthouse, J. P. G., Baines, B. S., Blenkinsop, R. D., Churcher, J. A., Mushiri, M. S., and Ormerod, F., Biochem. Soc. Trans, Vol. 6, 261-263(1978). I. Introduction However, of the two, chain. Cysteine is c the codons UGC an dehydrogenation of proteins. Inside the glutathione. Next to observation suggests d disulfide bonds in a generally lost complett Many biologit disulfide bonds. Beca it is necessary to knov any unique contributi With the advent of r CHAPTER 4 A NOVEL METHODOLOGY FOR ASSIGNMENT OF DISULFIDE BOND PAIRINGS IN PROTEINS I. Introduction Proteins are synthesized in cells by a stepwise process in which amino acids are added, one by one, from the N-temrini of the chains. Most of the proteins are found to contain SH and/or S-S groups, which are associated with cysteine and cystine residues. However, of the two, only cysteine is incorporated directly into the initial polypeptide chain. Cysteine is carried to the ribosome by its corresponding tRNA, which recognizes the codons UGC and UGU. Protein S-S groups arise by subsequent pairing and dehydrogenation of SH groups. Disulfide bonds occur frequently in extracellular proteins. Inside the cell the sulflrydryl group is maintained in a reduced state by glutathione. Next to tryptophan, cyst(e)ine is the most conserved amino acid. This observation suggests that cyst(e)ine is very important for structure and fimction. If all of disulfide bonds in a protein are reduced, both the tertiary structure and function are generally lost completely. Many biologically and therapeutically important proteins (peptides) contain disulfide bonds. Because the disulfide bond is an important element of protein structure, it is necessary to know the locations of disulfide bonds in order to understand more fully any unique contributions they may make to protein (peptide) structure and function. With the advent of recombinant DNA techniques for mutagenesis and expression of 141 cloned genes, it has b1 disulfide bonds into p that the products of proteins, or characteri an important aspect in methods for quanti determination of the cleavage of protein c such as cyanogen br bond. The resulting sequences or molecu mass spectrometry ( protein leads to the n Although thi: it is limited in man: their amino acid cor requirement may no protein is very sma cleaved between ev more than one disu residue by an enzyr 142 cloned genes, it has become possible to make altered structures, such as insertion of new disulfide bonds into proteins to increase stability. In these cases, one must either verify that the products 0f genetic manipulations are identical to those found in the native proteins, or characterize their differences. The assignment of disulfide bonds is therefore an important aspect in the structural characterization of proteins. Although there are good methods for quantifying the; number of disulfide bonds in proteins, the unambiguous determination of the location or pairing of disulfide bonds continues to challenge protein chemists. - Current methodology (1, 2) for assignment of disulfide bonds in proteins involves cleavage of protein chains between half-cystinyl residues with specific cleavage reagents, such as cyanogen bromide and trypsin, to obtain peptides that contain only one disulfide bond. The resulting mixture of peptides is separated, and the amino acid compositions, sequences or molecular masses of the peptides are determined by Edman degradation or mass spectrometry or both. Assignment of these peptides to specific segments of the protein leads to the recognition of disulfide crosslinkages. Although this approach is well established and has been used with much success, it is limited in many respects. First, disulfide-containing peptides can be identified by their amino acid composition or sequence only if they are purified to homogeneity. This requirement may not be achieved in the case of large proteins, or when the quantity of protein is very small. Second, the above methodology requires the protein chain be cleaved between every half-cystinyl residues so that the resulting peptides contain no more than one disulfide pair. If the protein is not cleaved between every half-cystinyl residue by an enzyme or chemical reagent, the peptide obtained ‘by the cleavage must be cleaved again using enzymatic or chemic proteins are often res difficult or irnpossibl spaced or adjacent cy are frequently used to disulfide scrambling approach may not be The problem of dis proteins, irrespective methodologies such been proposed to av obtained by these nor Gray (5, 6) h: highly bridged sma experiments, peptide the partially reduced positions of alkylate related to the disulfi obviously tedious, i protein using this ap; cleaved again using another enzyme or chemical reagent. However, even multiple-step enzymatic or chemical digestions rarely yield a full set of diagnostic fragments because proteins are often resistant to some specific cleavage reagents. It will be even more difficult or impossible to find an enzyme which is likely to cleave between two closely spaced or adjacent cysteines (1). It is likewise important to note that the conditions that are frequently used to cleave a protein chain are also the conditions that frequently lead to disulfide scrambling (3). As a result, the location of disulfide bonds determined by this approach may not be the same as the location of disulfide bonds in the native protein. The problem of disulfide exchange plagues all attempts to locate disulfide bonds in proteins, irrespective of the general method currently used. Although supplementary methodologies such as non-specific fragmentation by partial acid hydrolysis (4) have been preposed to avoid disulfide bond scrambling, it is difficult to deal with the data obtained by these non-specific techniques. Gray (5, 6) has described an approach for analyzing disulfide linkage patterns in highly bridged small peptides with close or adjacent cysteine residues. In his experiments, peptides were partially reduced under controlled conditions, the isomers of the partially reduced protein separated by HPLC, the nascent free thiols alkylated, and the positions of alkylated cysteines recognized from the results of sequence analysis and related to the disulfide bond pair that had been reduced and cyanylated. However, it is obviously tedious, if not impractical, to sequence an alkylated high-mass peptide or protein using this approach. [1. Our Novel Strat In chapter 2, free cysteine groups chemical reaction be selectively cyanylate cysteinyl residue can peptide and a series mapped to the seque react with NTCB groups. In chapte conditions, and stud‘ elimination. The 0 makes this methodo. interpretation of the c In this chapr assignment of disul: cleavage of partially resulting peptides by for a simple protein r reduced by tris(2-c: produce a mixture . Conditions can be 0; II. Our Novel Strategy for Disulfide Assignment In chapter 2, we developed a simple methodology for recognizing the location of free cysteine groups in peptides and proteins (7). This approach employs a Specific chemical reaction between sulflrydryls and 2-nitro-5-thiocyanobenzoic acid (NTCB) to selectively cyanylate cysteine thiols (8). The N—terminal peptide bond of the modified cysteinyl residue can then be cleaved under alkaline conditions to form an amino-terminal peptide and a series of 2-iminothiazolidine-4-carboxylyl peptides which can be mass mapped to the sequence of the original molecule by MALDI-MS. _ Disulfide bonds do not react with NTCB and therefore do not interfere with the determination of sulfhydryl groups. In chapter 3, we further improved and optimized the cleavage reaction conditions, and studied the effects of pH and amino acid structure on cleavage and [i- elimination. The optimized cleavage conditions greatly facilitated the cleavage and makes this methodology much more attractive in terms of speed of the analysis and interpretation of the data. In this chapter, we report a novel strategy (outlined in Figure 4.1) for the assignment of disulfide bond pairings in proteins using the above specific chemical cleavage of partially reduced and cyanylated protein isomers with mass mapping of the resulting peptides by MALDI-TOF MS (9). In this methodology, as shown in Figure 4.2 for a simple protein containing only two disulfide bonds, the denatured protein is partially reduced by tris(2-carboxyethyl)phosphine (TCEP) in a buffer solution at pH 3.0 to produce a mixture of residual intact protein and isomers of partially reduced species. Conditions can be optimized so that the predominant products are isoforms in which only Figure 4.1. I of disulfide 1: 145 Denatured Protein partial reduction TCEP, pH 3 Partially Reduced Protein Isomers cyanylation CDAP, pH 3 Cyanylated Protein Isomers HPLC separation Fractionation of Identified by , ———> Protern Isomers MALDI-MS cleavage 1 M NH4OI-I V reduction TCEP V A Mixture of Peptide Fragments MALDI-MS Figure 4.1. Descriptive overview of our proposed methodology for assignment of disulfide bond pairings in proteins. nee + 1 Protein 10 1.... S.“ + 1 Protein 1° H‘ClO Ire: rrcu ITC1( (( Figure 4.2. Chemicr protein of interest is each of which corre: protein containing 11 [TC stands for iminr 146 50 partial reduction; TCEP @ pH 3 SH SH Intact + 1 I 20 r 40 50 + 1 SI 20 T 40 5 0 Completely Protein 10 g __39__§ 10 SH 3 0 SH Reduced Species cyanylation; CDAP @ pH 3 S N s___ Intact + 1 lIoC 2'0 SICN 4'0 50 + 1 I 2:) 3480 4‘0 50 + Completely Reduced Protein S 30 S 10 SCN . SCN /Cyanylated Specres iHPLC fractionation . . . I er identrficatron . , b MAI _ lcleavage; 1N NH4OH y DI MS 1 9 1 L.“ 19 1TC10 %Q 29 S_Sr S—S ITCZO 30 39 ITC30 4'0 50 ITC4O 50 reduction; TCEP SH 1 9 1 19 §H 10 SH ITCIO 2." 29 rrC4o—-——so SH MS analysis (Cys10-Cys3 0) (CysZO-Cys40) Figure 4.2. Chemical overview of our methodology. Partial reduction means that the protein of interest is reduced under controlled conditions into a mixture of isoforms, each of which corresponds to reduction of only one of its disulfide bonds; for a protein containing 11 cystines, n isomers of the singly-reduced protein will result. ITC stands for iminothiazolidinyl carboxyl residue at the amino terminus. asingle disulfide b cyanylated by l-cyano the same buffer cond then separated by re MALDI-MS to dete +52 Da correspond corresponds to doubl shill from the mass cleavage in aqueous disulfide bonds, are mass-mapped by to the location of cleavage. A prim obtained from our ap protein isomers are I needed to define a confirm an assignm related to disulfidr performed in an aci adjacent or close llrerefore, two prr methodology. TI number of modr 147 a single disulfide bond has been reduced. Nascent sulfliydryls are immediately cyanylated by 1-cyano-4-dimethylamino-pyridinium tetrafluoroborate (CDAP) (10) under the same buffer conditions. The partially reduced and cyanylated protein isomers are then separated by reversed-phase HPLC, followed by analysis of HPLC fractions by MALDI-MS to determine which isomers are singly reduced/cyanylated. Those shifted by +52 Da correspond to a singly reduced/cyanylated species; a shift of +104 Da corresponds to doubly reduced/cyanylated species, etc. Those isomers with a 52-Da mass shift from the mass of the intact protein are dried and subjected to specific chemical cleavage in aqueous ammonia. The cleaved peptides, which may be linked by residual disulfide bonds, are then completely reduced to give a mixture of peptides that can be mass-mapped by MALDI-MS. The masses of the resulting peptide fragments are related to the location of the paired cysteines that had undergone reduction, cyanylation, and cleavage. A primary advantage of this approach is its underlying simplicity. The data obtained from our approach are straightforward and easily interpreted because only a few protein isomers are produced, and each is relevant. Typically, only n-l intermediates are needed to define an n-bridge system. Redundant information is always obtained to confirm an assignment. Secondly, we describe a practical way to circumvent problems related to disulfide bond scrambling because both reduction and cyanylation are performed in an acidic medium. Finally, our approach can be used for the assignment of adjacent or closely spaced cysteines for which conventional methodology fails. Therefore, two problems associated with current methodologies are solved using our methodology. The feasibility of the new approach is demonstrated by analyzing a number of model proteins with various disulfide bond linkages. The reported experimental conditio cyanylation, HPLC se III. Experimental S Mass Spectrometry MALDI mass the ion source was mode of operation. calibration using 5 5734.5), and horse 5 (St. Louis, MO). [- acid (Aldrich Chem were prepared in a ‘ equal volumes with plate. The mixtur spectrometer. 148 experimental conditions are optimized for partial reduction of the protein, sulflrydryl cyanylation, HPLC separation, and cleavage. III. Experimental Section Mass Spectrometry MALDI mass spectra were obtained on a Voyager Elite time-of-flight (TOF) mass Spectrometer (PerSeptive Biosystems Inc., Framingham, MA) equipped with a model VSL-337ND nitrogen laser (Laser Science, Newton, MA). The accelerating voltage in the ion source was set to 25 kV. Data were acquired in the. positive or negative linear mode of operation. Time-to-mass conversion was achieved by external and/or internal calibration using standards of bradykinin (m/z 1061.2), bovine pancreatic insulin (m/z 5734.5), and horse skeletal myoglobin (m/z 16,952) obtained from Sigma Chemical Co. (St. Louis, MO). All experiments were performed using oc-cyano-4-hydroxycinnarnic acid (Aldrich Chemical Co., Milwaukee, WI) as the matrix. Saturated matrix solutions were prepared in a 50% (v/v) solution of acetonitrile/aqueous 0.1% TF A, and mixed in equal volumes with peptide or protein samples, and applied to a stainless-steel sample plate. The mixture was allowed to air dry before being introduced into the mass Spectrometer. Chemicals Tris(2—carbox Chemical Co. (Rock Mannheim Biocherni A, bovine milk or-Iac inhibitor (BPTI), ci tetrafluoroborate (C also purchased fro Acetoninile and TF pH 3.0 was prepare with little deteriorati freshly prepared prio Partial Reduction 1 Ten-nmol p: citrate buffer (pH disulfide bonds war in the protein (eg. mol of ribonuclea temperature for l1 149 Chemicals Tris(2-carboxyethyl)phosphine (TCEP) hydrochloride was purchased from Pierce Chemical Co. (Rockford, IL). Guanidine hydrochloride was a product of Boehringer- Mannheim Biochemicals (Indianapolis, IN). Bovine pancreatic ribonuclease A type HI- A, bovine milk oc-lactalbumin, soybean trypsin inhibitor (STI), bovine pancreatic trypsin inhibitor (BPTI), citric acid, sodium citrate, and 1-cyano-4—dimethylamino—pyridinium tetrafluoroborate (CDAP) were purchased from Sigma and used without further purification. Recombinant LR3IGF-I (11271), recombinant human epidermal growth factor (hEGF, containing a Met at the N-terminal as the initiator of biosynthesis) were also purchased from Sigma and purified by reversed-phase HPLC before use. Acetonitrile and TFA were of HPLC grade. The TCEP solution in 0.1 M citrate buffer at pH 3.0 was prepared as 0.10 M stock solution and stored under N2 at -20°C for weeks with little deterioration. The 0.10 M CDAP solution in 0.1 M citrate buffer at pH 3.0 was freshly prepared prior to use. Partial Reduction of Proteins Ten-nmol protein samples were solubilized and denatured in 10 ul of 0.1 M citrate buffer (pH 3.0) containing 1 6M guanidine-HCI. Partial reduction of protein disulfide bonds was carried out by adding an equivalent of TCEP for the cystine content in the protein (e.g., 40 nmol of TCEP was reacted with 10 nmol of ribonuclease A as 1 mol of ribonuclease A contains 4 moles of cystine), followed by incubation at room temperature for 10-15 minutes. Depending on the protein understudy, the amount of reducing agent may reduction products. computer. UV de (#218TP54, 10pm IOum particle size, proteins were sligh fractions were coll: determined by MA sample size of 10—r columns; the use 1 sizes. 150 reducing agent may be adjusted to ensure the singly reduced isoforms are predominant reduction products. Cyanylation of Nascent Sulfhydryls To the partially reduced protein mixture was added a 20-fold molar excess of CDAP solution .over the total cysteine content. Cyanylation of the nascent sulfhydryl groups was accomplished by incubation at room temperature for another 10-1 5 minutes. HPLC Separation of Partially Reduced and Cyanylated Protein Isomers Partially reduced and cyanylated species were separated by reversed-phase HPLC with linear gradient elution using Waters model 6000 pumps controlled by a PC computer. UV detection was at 215 nm. The columns were either a Vydac C18 (#218TP54, 10pm particle size, 300-A pore, 4.6x250 mm) or a Vydac C4 (#214TP54, lOum particle size, 300-A pore, 4.6x250 mm). The HPLC conditions for individual proteins were slightly different (see chromatograms shown below). The major HPLC fractions were collected manually and the masses of the collected protein isomers were determined by MALDI-MS. Appropriate fractions were then dried for further use. The sample size of IO-nanomoles was used for convenient detection from conventional HPLC columns; the use of microbore columns should allow the use of much smaller sample sizes. Cleavage of Singly To the dried aqueous solution to of the peptide chain removed in a vacu Complete Reductio Truncated p reduced by reacting minimize the possib acetonitrile/0.1% T IV. Results and Di A. Partial Reducti There are tr reduction and parti bonds in a protein denaturing conditio selectively split by three disulfide bon (ll). Sequential 1 151 Cleavage of Singly Reduced and Cyanylated Protein Isomers To the dried HPLC fraction was added 2ul of 6 M guanidine-HCI in 1M NH4OH aqueous solution to dissolve the protein residue and then Sul of 1 M NH4OH. Cleavage of the peptide chain was performed at room temperature for 1 hour. Excess ammonia was removed in a vacuum system. Complete Reduction of Remaining Disulfide Bonds Truncated peptides, still linked by residual disulfide bonds, were completely reduced by reacting with Zul of 0.1 M TCEP solution at 37°C for 30 minutes at pH 3-5 to minimize the possibility of reoxidation. Samples were diluted with lOOul of a 50% (v/v) acetonitrile/O. 1% TFA solution prior to analysis by MALDI-MS. IV. Results and Discussion A. Partial Reduction of Proteins There are two terminologies frequently used in disulfide reduction: sequential reduction and partial reduction. The former refers to the situation in which disulfide bonds in a protein are opened one by one by using different reducing agents and/or denaturing conditions. For example, only the most labile disulfide bond in papain can be selectively split by the addition of 2-mercaptoethanol in the presence of 8M urea, but all three disulfide bonds can be reduced in 6M guanidine-HCI by the same reducing agent (11). Sequential reduction is practically limited in use because it is difficult, if not impossible, to choose protein can be cleave latter term, partial re small portion of eac desired products are bridges. As illustra controlled so that ea resulting in a mixtur is understandable same protein molec of each disulfide b which are negligi Consequently, the 1 minor singly reduce can easily be ach stoichiometry of rec Although e' reduced to a similar of various disulfidr different, dependir conformational cm (12-15). Our m1 152 impossible, to choose appr0priate conditions so that only one of the disulfide bonds in a protein can be cleaved at a time without the reduction of the other disulfide bonds. The latter term, partial reduction, refers to limited, controlled conditions under which only a small portion of each of the disulfide bonds, 6. g., 10% of each, is reduced. That is, the desired products are those in which reduction has Opened some, but not all disulfide bridges. As illustrated in Figure 4.3 for ribonuclease A, reduction conditions can be controlled so that each of the four disulfide bonds is reduced to a minor extent (S. 10%), resulting in a mixture of four singly reduced isomers as a main reduced form. This result is understandable as the probabilities of reducing two and three disulfide bonds in the same protein molecule (under the above conditions where we observed a 10% reduction of each disulfide bond) are 10%x10% = 1% and 10%x10%x10% = 0.1%, respectively, which are negligible in comparison to obtaining the singly reduced isoforms. Consequently, the mixture of partially reduced protein contains major intact proteins, minor singly reduced isoform, trace doubly reduced isoform, and so on. Practically, this can easily be achieved by controlling incubation time, reaction temperature, and stoichiometry of reducing agents. Although every disulfide bond has the same redox potential and should be reduced to a similar extent from the standpoint of thermodynamics, the reduction kinetics of various disulfide bonds in solutions of native or partially denatured protein are quite different, depending on accessibility of the disulfide bonds to reducing agents, the conformational energy of the disulfide bonds, and the redox potential of the environment (12-15). Our methodology requires that individual disulfide bonds be broken at 26 26 SH (m 58 SH (11 Figure . 153 | 1—1 j 124 26 40 58 65 72 84 95 110 Ribonuclease A partial TCEP, pH 3, reduction 15', rt \l | J_—" I 124 26 40 58 65 72 84 95 1 10 (major) ["1 84 4O 1—_l 95 | I W SH SH SH (minor) (minor) 58 r—j 110 I 65 72 1 I I I T SH SH SH SH (minor) (minor) + doubly reduced isoforms (trace) + triply reduced isoforms (very trace) OOOOOOOOOOOOOOOOOOOOO Figure 4.3. Conceptional scheme to illustrate "Partial Reduction". comparable rates. T study were denatur accessibility of reduc Water-solubl bonds (5-7, 9). Red bond scrambling. F controlled which m The mixture reduction states su conducted with ribo and LR’IGF-I, sho cystine content in ii 15min at room tem 20—fold excess of T From the limited 11 control is not critic more readily be co content may not b protein of unkno 154 comparable rates. To minimize structural diversity of disulfide bonds, proteins under study were denatured by dissolution in 6 M guanidine so that differences in the accessibility of reducing agent to each disulfide bond was minimized. Water-soluble TCEP has proved to be an excellent reducing agent for disulfide bonds (5-7, 9). Reduction by TCEP can be carried out at pH 3.0 to suppress disulfide bond scrambling. Furthermore, at pH 3.0, the reduction of disulfide bonds is kinetically controlled which makes partial reduction possible (5). The mixture of partially reduced proteins contained intermediates with different reduction states such as singly reduced and doubly reduced isoforms. Experiments conducted with ribonuclease A, insulin, soybean trypsin inhibitor, oc-lactalbumin, hEGF, and LR3IGF-I, showed that by using approximately an equivalent of TCEP for the total cystine content in the proteins, about 5~10% of each disulfide bond was reduced within 15 min at room temperature. However, for bovine pancreatic trypsin inhibitor, at least a 20-fold excess of TCEP was required to achieve 10% reduction of each disulfide bond. From the limited number of proteins we have examined it appears that stoichiometry control is not critical, although BPTI was an exception. The extent of the reduction can more readily be controlled by the reaction time and temperature. However, as the cystine content may not be known a priori, the extent of reduction should be monitored for a protein of unknown cystine content. B. Cyanylation of N Jacobson et specifically cyanylat cyanoAdirnethylam‘ same purpose (10). chemistry. We emp free and total cystei selectivity of the conditions which or has been reported conditions (9, 10, 1 in peptide and pro cyanylation by CD15 cg, pH 3-5 at ror unnecessary to rem CDAP. However, over peptide sulthy could result in me control the stoichi hand, CDAP reac reaction system w cyanylation is pos unknown protein. 155 B. Cyanylation of Nascent Sulfhydryls Jacobson et al. (8) first showed that 2-nitro-5-thiocyanobenzoic acid (NTCB) specifically cyanylates cysteine thiols in mildly alkaline solutions. Another reagent, 1- cyano—4-dimethylamino-pyridiniurn tetrafluoroborate (CDAP) was later proposed for the same purpose (l0). However, only NTCB has been extensively applied in protein chemistry. We employed the NTCB reaction to characterize the number and location of free and total cysteine groups in peptides and proteins (7). In spite of the reactivity and selectivity of the NTCB reagent, the cyanylation must be performed under alkaline conditions which may permit some disulfide bond scrambling. CDAP, on the other hand, has been reported to be a selective and reactive cyanylation reagent under acidic conditions (9, 10, 16, 17). The utility of the CDAP for cyanylation of sulthydryl groups in peptide and proteins has been extensively investigated in chapter 3. Complete cyanylation by CDAP was possible by using a 5-fold molar excess under mild conditions, e.g., pH 3-5 at room temperature, which is compatible with TCEP. It is therefore unnecessary to remove excess TCEP, change buffer, or readjust the pH prior to using CDAP. However, as shown in chapter 3, a large excess of CDAP (~50-fold molar excess over peptide sulfhydryls) and excessive incubation time (>2 hours at room temperature) could result in modification of other amino acid side chains. Therefore, it is necessary to control the stoichiometry of CDAP in order to minimize side reactions. On the other hand, CDAP reacts instantly with TCEP even at pH 3.0, the excess of TCEP in the reaction system will be killed upon addition of CDAP, and no firrther reduction during Cyanylation is possible. This is definitely an advantage for the control of reduction in an unknown protein. However, it should be noted that a larger amount of CDAP has to be applied in case mor structure, such as B? C. HPLC Separafio The mixture intact protein and a reversed-phase HPL bond(s), i. e., partial 4.4 is a typical ribonuclease A. Be are four other peaks greater than that r reduced/cyanylated the same, suggestir disulfide bonds in other proteins will The separ hydrophobicity of structure, exposes protein’s hydroph reduced disulfide thus, should elute support this spec 156 applied in case more TCEP is required for partial reduction of proteins with a tight structure, such as BPTI. C. HPLC Separation of Partially Reduced and Cyanylated Protein Isomers The mixture obtained from the above reactions contains a majority of residual intact protein and a minority of partially reduced/cyanylated isoforms. The capacity of reversed-phase HPLC to separate these protein isomers with various residual disulfide bond(s), i. e., partially reduced isoforms, is demonstrated in following examples. Figure 4.4 is a typical chromatogram of the partially reduced/cyanylated isoforms of ribonuclease A. Besides the main peak that corresponds to residual ribonuclease A, there are four other peaks (marked 1-4), each corresponding to a species having a mass ~52 Da greater than that of the original protein. These peaks are attributable to singly reduced/cyanylated protein isomers. The intensities of the four peaks are approximately the same, suggesting that under our reaction conditions, the reducing rates of the four disulfide bonds in denatured ribonuclease A are comparable. The HPLC separation of other proteins will be discussed in individual sections below. The separation mechanism in HPLC is based on the difference in the hydrophobicity of analytes (5, 18). Opening a given disulfide bond disrupts protein structure, exposes the protein’s interior hydrophobic amino acids, and increases the protein’s hydrophobicity to different extents. Therefore, isomers with one or more reduced disulfide bond(s) should tend to interact more strongly with HPLC columns and, thus, should elute at later retention time than the original protein. The experimental data support this speculation. Almost all partially reduced isomers show longer retention 20 30 Figure 4.4. HPL‘ reduced/cyanylat phase HPLC on : gradient 20-40% in CH3CN. Pealr MALDI-TOF an 157 Intact 1 2 3 4 20 3o 40 50 30 i0 8?) Figure 4.4. HPLC separation of denatured ribonuclease A and its partially reduced/cyanylated isomers. Ten-nmol of the protein were separated by reversed- phase HPLC on a Vydac C18 column at a flow rate of 1.5 ml/min with a linear gradient 20-40% B in 90 minutes, where A = 0.1% TFA in water and B =~0.1% TFA in CH3CN. Peaks 1-4 represent singly reduced/cyanylated species, as determined by MALDI-TOF analysis. times than their intac from the HPLC data, dramatically change the unusually broad most likely due to di suggested by Gray ( and cyanylation, e column in the water this case for faster propanol within a re It should b isomers, it is urine below for ribonucle two respective prot unambiguous assigr described methodol D. Cleavage of PE One of the bond of cyanylatr usually employed hours at 37°C) in 2 158 times than their intact original proteins. The differences in hydrophobicity, as reflected from the HPLC data, indirectly suggests that reduction of individual disulfide bonds can dramatically change the three-dimensional structure of a protein. Another observation is the unusually broad HPLC peak for some proteins (e.g., or-lactalbumin, STI), which is most likely due to different conformations that show slightly different retention times, as suggested by Gray (5). Some hydrophobic proteins (e.g., STI), eSpecially after reduction and cyanylation, exhibited a long retention time or even irreversible retention on the column in the water-acetonitrile mobile phase. l-PrOpanol was substituted for CH3CN in this case for faster elution. Most of the proteins can be eluted by 40% aqueous l- propanol within a reasonable time (19). It should be pointed out that although HPLC is able to separate the protein isomers, it is unnecessary for all components to be baseline separated. As described below for ribonuclease A, even with poorly resolved chromatographic peaks representing two respective protein isomers, interpretation of MALDI data is still possible for the unambiguous assignment of disulfide bonds. This provides a valuable advantage of the described methodology for analyzing complicated mixtures. D. Cleavage of Peptide Chains One of the key features of our methodology is cleavage at the N—terminal peptide bond of cyanylated cysteinyl residues. Previously described methods (7, 8, 20-23) usually employed mildly alkaline conditions (pH~9) and long hours of incubation (>16 hours at 37°C) in attempts to control the B-elimination reaction, a side reaction that results in a lower yield of elimination over the cleavage kinetics as amino acids adjacent show that higher pH for most of the significantly even at within an hour at to 1M NH40H (pH~1 our experiments, 1 from the reaction 3 using higher pH i reversed. Modifica conditions describe After cleav: linked by the remai the peptide mixtu subsequent analysi E. Interpretation Unlike the products, MALD .159 in a lower yield of the cleavage reaction. An elevated pH was reported to favor B- elimination over the cleavage reaction (20, 21). We have systematically studied the cleavage kinetics as a function of pH for cyanylated polypeptides containing different amino acids adjacent to the N-terminus of cyanylated cysteines. The results in chapter 3 show that higher pH can greatly accelerate the cleavage and B-elimination reactions, but for most of the peptides studied, the extent of B-elimination does not increase significantly even at a pH as high as 12, a condition at which cleavage can be complete within an hour at room temperature. Cleavage can be accomplished in 0.01 M NaOH or l M NH4OH (pH~1 1) solutions containing denaturing agent; both give similar results. In our experiments, 1 M NH4OH is preferred over 0.01 M NaOH because it can be removed from the reaction system after completion of the cleavage reaction. One side reaction in using higher pH is the neutralization of -COOH groups; however, this can be easily reversed. Modification or cleavage of other side chains was rarely observed under the conditions described above. After cleavage at cyanylated cysteinyl residues, the truncated peptide chains, still linked by the remaining disulfide bonds, can be easily reduced by excess TCEP. Finally, the peptide mixture is diluted to minimize the adverse effect of guanidine on the subsequent analysis by MALDI. E. Interpretation of MALDI Data Unlike the MALDI spectra of a complicated mixture of enzymatic digestion products, MALDI data derived from analysis of the cleavage reaction mixture of a cyanylated protein ar cyanylated cysteinyl reduced/cyanylated corresponding to the related to the positio to deduce the disul cleavage, provides otherwise would ha assignment. Ribonucleas four disulfide bon Cys72. Figure 4.5 ,___E 26 Table 4.1 lists the peptide chains at cyanylated. Figur cleavage of isome: peaks 1-4, respect 160 cyanylated protein are easy to interpret. Cleavage of the peptide chain takes place only at cyanylated cysteinyl sites, which in principle yields three fragments for each singly reduced/cyanylated protein isomer (and sometimes two overlapped fragments corresponding to the B-elimination at one cysteine site). The mass of each fragment is related to the position of the two cyanylated cysteinyl residues which in turn can be used to deduce the disulfide bond. linkage. B-Elimination, an alternative to peptide chain cleavage, provides mass spectral data corresponding to overlapped peptides (that otherwise would have cleaved) and serves as a confirmation for the disulfide bond pairing assignment. Ribonuclease A (Mr = 13,683) (24) contains 124 amino acids that are linked by four disulfide bonds: Cy326-Cys84, Cys40-Cys95, CysS8-CysllO, and Cys65- Cys72. Figure 4.5 shows the disulfide structure of ribonuclease A. 1 J m l 1,4 26 40 58 65 72 84 95 1 10 Figure 4.5. Disulfide structure of ribonuclease A. Table 4.1 lists the calculated m/z values for possible fragments due to cleavage of the peptide chains at different sites depending on which disulfide bond was reduced and cyanylated. Figure 4.6A-D are four MALDI spectra of peptide mixtures resulting from cleavage of isomers of singly reduced/cyanylated ribonuclease A corresponding to HPLC peaks 1-4, respectively. Table 4.1. CysS 8-1 Cys4C 161 Table 4.1. Calculated and observed rrr/z values for possible fragments resulting from the cleavage reaction of ribonuclease A chains at sites of designated cysteine pairs Reduction Fragment Calculated m/z Observed m/z of Disulfide 125 2702.8 2705.3 26-83 6547.3 6548.5 Cys26-Cys84 84-124 4526.0 4527.4 1-83 9176.2 9176.7 26-124 10995.3 10998.6 1-64 7083.9 7083.8 65-71 789.8 nd ' Cys6S-Cys72 72-124 5906.5 5907.7 1.71 7795.7 7790.0 65-124 . 6618.3 6617.9 1-57 6353.1 6351.1 58-109 5767.4 5766.8 Cys58-Cysl 10 110-124 1659.8 1659.8 1-109 12042.4 12036.7 58-124 7349.1 nd 1-39 4413.9 4414.4 4094 6063.6 6061.2 Cys40-Cys95 95-124 3302.7 3303.7 194 10399.5 10430.5 40-124 9288.3 9293.4 110-124 95' ++ L..— Figure 4.6. Th of the four sing HPLC peaks 1 doubly-charge peaks l~124 in occurred at on information. 5 162 1—25 84-124 A ++ ,L 26'l83 1183* 26-124* 1424 W qt. JN-___ 1-64 72124 B * ++ g] * .,', : ++ ‘ will A 1-124 Wk - ..x 4 wvf .._ +fi—w-JM 110-124 C ++ 58-109 ;57 ‘7“ \I i 1409* In...“ A A. 95124 /1-39 D 4094 * fl. ++ j A 3‘. WM M A 5000 10000 15000 m/z Figure 4.6. The MALDI mass spectra of peptide mixtures resulting from the cleavage of the four singly reduced/cyanylated ribonuclease A isomers, corresponding to the HPLC peaks 1-4 in Figure 4.4, respectively. The symbols ++ and * represent the doubly-charged species and protonated B-elimination products, respectively. The peaks 1-124 in ( A) and (B) represent intact proteins in which B-elimination only occurred at one cysteinyl residue. These products do not provide any specific information. See also Table 4.1 for calculated and observed m/z values. The mass spectrum HPLC peak 1 in Ft fragments 1-25, 26 4526.0) with a rela peptide chain cleav m/z 9176.7 corres cleavage at Cy584 10998.6 is another at Cy584. Overall, deduced. With simi CysSS-CysllO, al The MALDI spr represented by Hi 7083.8, correspor fragment 65-71 i: these two fragme elimination prod confirmation of t 65~124, while thi The HPI carefully collectt 163 The mass spectrum in Figure 4.6A corresponds to the cleavage products represented by HPLC peak 1 in Figure 4.4. Three peaks at m/z 2705.3, 6548.5, and 4527.4 are due to fragments 1-25, 26-83, and 84-124, respectively (expected m/z 2706.8, 6547.3, and 4526.0) with a relative mass deviation of <0.05%. From these data, one can deduce that peptide chain cleavages occur at Cy826 and Cys84. Additionally, the MALDI peak at m/z 9176.7 corresponds to an overlapped peptide, 1-83, resulting from peptide chain cleavage at Cys84, but B-elimination at Cy826. Likewise, the MALDI peak at m/z 10998.6 is another overlapped peptide, 26-124, with cleavage at Cy326, but B-elimination at Cys84. Overall, a disulfide bond linkage between Cy826—Cys84 can be unambiguously deduced. With similar strategy, two other disulfide bond linkages, Cys40-Cys95 and Cys58-Cys110, also can be recognized from Figure 4.6C and Figure 4.6D, respectively. The MALDI spectrum in Figure 4.6B, corresponding to the cleavage products represented by HPLC peak 2 (Figure 4.4), shows two main peaks at m/z 5907.7 and 7083.8, corresponding to fragment 72-124 and 1-64, respectively. Another expected fragment 65-71 is missing. However, it is still possible to deduce Cys65-Cys72 from these two fragments because no other combination gives such masses. Two minor [3- elimination products, m/z 6617.9 and 7790.0, are particularly informative for confirmation of the assignment in this case. The peak at m/z 6617.9 represents residues 65-124, while that at m/z 7795.7 covers residues 1-71. The HPLC peaks 1 and 2 in Figure 4.4 were not resolved completely. Even carefully collected HPLC fractions from one component still contained a small amount of the other componen the cleavage produ corresponding to assignment of the r possible because 0 In the earl optimized, the com collected as a broa fraction is given in spectrum A and B pairs simultaneous pairs in a mixture. and doubly reduce bond pairs from 1 gives more compl that the assignmer more complicate: responses from t1 isoforrn might be correct assignme' 164 the other component. This is reflected in the MALDI spectra (Figures. 4.6A and B) of the cleavage products of these two fractions, each of which contains small fragments corresponding to the cleavage products of the other fraction. The unambiguous assignment of the respective disulfide bond pairs in the presence of another isomer is still possible because only a few fragments are produced. In the earlier stage of this research, the HPLC conditions were not fully optimized, the compounds represented by peaks 1 and 2 in Figure 4.4 coeluted and were collected as a broad peak. The MALDI spectrum of the cleavage products of that broad fraction is given in Figure 4.7. It is apparent that this spectrum is a perfect overlap of the spectrum A and B in Figure 4.6. From this spectrum, one can assign two disulfide bond pairs simultaneously, demonstrating the capability of the methodology to assign disulfide pairs in a mixture. On the other hand, if the HPLC peak contains a mixture of both singly and doubly reduced/cyanylated isoforms, Caution should be used in assigning disulfide bond pairs from the fragments. Because a doubly reduced/cyanylated protein isomer gives more complicated fragments than a singly reduced/cyanylated protein, it is expected that the assignment from such a mixture will be challenging. The situation could be even more complicated, because, due to the discrimination of MALDI responses, the MALDI responses from the cleavage products of even trace amount of doubly reduced/cyanylated isoforrn might be more intense than those of the singly reduced/cyanylated isoform. The correct assignment may be retarded in such a case. 1-25 Figure 4.7. Th of a mixture 01 charged specie 165 1-25 72-124 84-124 1-64 71* g 26-83 1' .4 1-83* 26—r24* 1'1%4 5600 10000 15000 m/z Figure 4.7. The MALDI mass spectrum of peptide mixtures resulting from the cleavage of a mixture of HPLC peaks 1 and 2. The symbols ++ and * represent the doubly- charged species and protonated B-elimination products, respectively. We have obs chain usually is less 5767.4, and m/z 60 for the middle frag reduces the yield of oi-Lactalb linked as 4 disulfid and isomers of ‘ isomers have 5' different disulfide peaks 1-4 in Figur corresponding M calculated m/z va lactalbumin chain Although mass increase 01 protein, MALDI 517.6) which we not detected, p0: it impossible to 166 We have observed that the middle fragment from cleavage of a cyanylated peptide chain usually is less abundant in the MALDI spectrum (e.g., m/z 6547.3, m/z 789.8, m/z 5767.4, and m/z 6063.6 in Table 4.1). A possible explanation for this observation is that for the middle fragment, B-elimination is able to occur at either side, which significantly reduces the yield of the expected cleavage product. or-Lactalbumin (123 amino acids, Mr = 14,175) also contains 8 cysteines that are linked as 4 disulfide bonds (25). Figure 4.8 shows the HPLC separation of oc-lactalbumin and isomers of its partially reduced/cyanylated species. Peaks for four singly reduced/cyanylated or-lactalbumin isomers (~52-Da mass shift from original molecule, marked 1-4) are observed in addition to that for residual or-lactalburnin. Again, the four isomers have similar abundances which suggest comparable reduction rates for the different disulfide bonds in denatured oc~lactalbumin. The four HPLC fractions (HPLC peaks 1-4 in Figure 4.8) were subjected to cleavage under the described conditions. The corresponding MALDI mass spectra are shown in Figure 4.9A-D. Table 4.2 lists the I calculated m/z values for possible fragments resulting from the cleavage reaction of or- lactalbumin chains at sites corresponding to different cysteine pairs. Although the compound represented by HPLC peak 1 in Figure 4.8 showed a mass increase of ~52 Da, corresponding to an isomer of a singly reduced/cyanylated protein, MALDI-MS (Figure 4.9A) only detected one expected cleavage product (m/z 517 .6) which was due to fragment 120-123. Two other fragments, 15 and 6-119, were not detected, possibly because of signal suppression. The insufficient information makes it impossible to deduce a disulfide bond linkage from the MALDI data. Figure 4.8. in Cyanylated 130 of 1.5 [111/min 111 Water and I reduced/cyan: 167 Intact 10 2'0 30 40 50 60 70 Figure 4.8. HPLC separation of denatured oc-lactalbumin and its partially reduced! cyanylated isomers. Separation was carried out on a Vydac C4 column at a flow rate of 1.5 ml/min with a linear gradient 40-50% B in 90 minutes, where A = 0.1% TFA in water and B = 0.1% TF A in CH3CN. Peaks 1-4 represent singly reduced/cyanylated species, as determined by MALDI-TOF analysis. Table 4.2. ( resulting frr designated « Reductio of Disulf Cys6-Cy: Cys28-C Cys6l-( Cys73 168 Table 4.2. Calculated and observed m/z values for possible fragments resulting from the cleavage reaction ofa—lactalbumin chains at sites of designated cysteine pairs Reduction Fragment Calculated m/z Observed m/z of Disulfide 1-5 618.7 nd 6-119 13135.7 nd cys6-c.y3120 120-123 517.6 517.6 1-118 13676 nd 6-123 13575 nd 1-27 3125.6 3122.5 28-110 9525.6 9517.1 CysZS-Cysl 11 111-123 1620.8 1620.1 1-110 12573 12560 28—123 11068 nd 1-60 6918.9 6914.4 61-76 1800.8 1798.9 Cys6l-Cys77 77-123 5552.5 5557.5 1-76 8641.5 8648.6 61-123 7275.3 7280.6 1-72 8258.1 8264.5 73-90 2098.3 2099.0 Cys73-Cys91 91-123 3915.6 3912.8 1-90 10278 10303 73-123 5936.0 nd 120-123 / L...__._.. 111-123 1-27 r E --76 (:1 111—123 E? 61—76 73-90 Kim. \ Figure 4.9. The the four Singly Peaks L4 in Fi Charged Specie: for the calculat 169 120-123 A / ..._—A— 4— #— 4 4 Aw——~_l 111-r23 1-27 B \O 3 77-123 1-60 I \o l l 1-7628-110 _ 1-110 77-123 ' R q, C E ‘T S 3 1-60 \/ \61-123 1-76 :r*- ".._“...r- A .._. “‘M 91-123 D 73-90 E, i “I 1 1-72 1-90 ' A: A___. 5600 10000 15000 m/z Figure 4.9. The MALDI mass spectra of peptide mixtures resulting from the cleavage of the four singly reduced/cyanylated oc-lactalbumin isomers, corresponding to the HPLC peaks 1-4 in Figure 4.8., respectively. The symbols ++ and * represent the doubly— charged species and protonated b-elimination products, respectively. See also Table 4.2 for the calculated and observed m/z values. The HPLC p peak broadening. H all the fragments re singly reduced/cyar 1620.1 and 3122.5 peak at rn/z 9517. usually shows 10v elimination. Furti product l-llO. Tl Peaks at m/z 179 respectively, resul 76 (m/z 8640.3) 1 1“formation is co deduced firm the The Cy36 aIiSmg from anal 3m Figme 4.8). the fragments 6 Products, 61.12; is linked to Cy: fraction 2_ Fig Products in HP 170 The HPLC peaks 2 and 3 in Figure 4.8 were not baseline separated due to severe peak broadening. However, the MALDI-MS data interpretation permitted recognition of all the fragments resulting from the cleavage of the mixture of the two isomers of the singly reduced/cyanylated protein. As seen in Figure 498, two intense peaks at m/z 1620.1 and 3122.5 can be assigned to fragments 111-123 and 1-27, respectively. A third peak at m/z 9517.1 matches the middle fragment 28110 which, as mentioned above, usually shows lower abundance due to its opportunity for double participation in B- elirnination. Furthermore, the peak at m/z 12560 is an indication of the B-elimination product 1-110. Thus, a disulfide bond, Cys28-Cysl 11, can be deduced. The other three peaks at m/z 1798.9, 5552.1, and 6909.9 match fragments 61-76, 77-123, and 1-60, respectively, resulting from cleavage at Cys6l and Cys77. The B—elimination product 1- 76 (m/z 8640.3) further confirms that Cys77 is a cleavage site. The combination of this information is consistent with the assignment. Thus, two disulfide bond pairs can be deduced from the mass spectrum of an impure HPLC fraction. The Cys61-Cys77 pair is further confirmed by the mass spectrum in Fig. 4-9C, arising from analysis by MALDI-MS of the cleavage products in HPLC fraction 3 ( peak 3 in Figure 4.8). The three main peaks at m/z 1798.9, 5557.5, and 6914.4 correspond to the fragments 61-76, 77—123, and 1-60, respectively. In addition, two B-elimination products, 61-123 at m/z 7280.6 and 1-76 at m/z 8648.6, provide confirmation that Cys61 is linked to Cys77. The peak at m/z 1620.0 is likely due to some “carry-over” from fraction 2. Figure 4.9D was obtained from analysis by MALDI-MS of the cleavage products in HPLC fraction 4 (Figure 4.8). The three MALDI peaks at m/z 2099.0, 3912.8, and 8264.5 elimination product Cys73 is linked to C The other e 811 (Mr. 19977) cc and Cysl36-Cysl4 MALDI analysis < indicating the samr had a very poor i between STI and ( memw but several impur 8amilles were sul 4-10, the elevated difficult. Table Preliminary MA: given in Figure 4.11A) correSpc indicating Cyslj they are not real detected for frag rs 1n3uffici em 1 171 3912.8, and 8264.5 represent fragments 73-90, 91-123, and 1-72, respectively. The B- elimination product (m/z 10300.2) is due to fragment 1-90. These data indicate that Cys73 is linked to Cys91. The other example, soybean trypsin inhibitor (STI), is also demonstrated here. STI (Mr. 19977) contains 180 amino acids linked by two disulfide bonds, Cys39-Cys86 and Cysl36-Cysl45. ( another variant of STI contains 181 amino acids with Mr. 20095. MALDI analysis of commercial STI gave a m/z 19988 Da as a predominant peak, indicating the sample contains 180 amino acids). Being a more hydrOphobic protein, STI had a very poor HPLC behavior in CH3CN/HZO mobile phase. The strong interaction between STI and C18 or C4 column makes the HPLC separation impractical. The HPLC analysis of STI was improved greatly by using n-prOpanol as a mobile phase modifier, but several impurities were still poorly resolved from the STI peak. The purified STI samples were subjected to partial reduction and cyanylation. As illustrated in Figure 4.10, the elevated base line makes the separation of intact STI and its reduced isoforms difficult. Table 4.3 lists the m/z values of the expected and observed fragments. Preliminary MALDI analysis of the cleavage products from HPLC peaks 1 and 2 are given in Figure 4.11A and B. The MALDI peaks at m/z 4036.3 and 5017.2 (Figure 4.11A) correspond to the fragment 145-180 and an overlapped peptide 136-180, indicating Cys136 and Cys145 have been cleaved. Other peaks are also observed, but they are not readily related to the cleavage at any cysteine sites. Since no signals were detected for fragments 1-136 and 136-144, the information obtained by MALDI analysis is insufficient to draw any conclusion. Figure 4.11B was obtained by the MALDI Figure 4.10. r Partially reduc COlumn at a fl Where A = 0_j TFA. Peaks MALDI-T01 172 Intact 20 25 30 35 40 Figure 4.10. HPLC separation of denatured soybean trypsin inhibitor and its partially reduced! cyanylated isomers. Separation was carried out on a Vydac C4 column at a flow rate of 1.2 mllmin with a linear gradient 15-50% B in 30 minutes, where A = 0.1% TFA in water and B = 80% l-propanol/20% H20 containing 0.1% TFA. Peaks 1 and 2 represent singly reduced/cyanylated species, as determined by MALDI-TOF analysis. Table 4.3. resulting 1 at sites of Reducti 0f Disu Cys39-C Cysl3€ Cys39 Cys13 173 Table 4.3. Calculated and observed m/z values for possible fragments resulting from the cleavage reaction of soybean trypsin inhibitor chains at sites of designated cysteine pairs Reduction Fragment Calculated m/z Observed m/z ofDisulfide 1-38 4058.4 4059.1 Cys39-Cys86 39-35 5258.1 5259.9 86-180 10747 10752 1-135 14971 nd 136-144 1057.1 nd Cysl36-Cysl45 145-180 4035.5 4036.3 13(1-144) 15951 nd 8(136-180) 5015.6 5017.2 1-38 4058.4 4057.1 39-85 5258.1 5257.8 95????sz 86-135 5740.4 5738.8 ys ys [3(136-144) 1057.1 nd 8(145—180) 4035.5 4037.9 145—180 136-l 1-38 4000 Figure 4.11. T1 of the two sing correSp0nding 1 96319 are duet 174 145-180 A 136-180 1-38 B 39-85 All I / 86-180 4000 6000 8000 - 10000 121100 14b00 m/z Figure 4.11. The MALDI mass spectra of peptide mixtures resulting from the cleavage of the two singly reduced/cyanylated soybean trypsin inhibitor (STI) isomers, corresponding to the HPLC peaks 1 and 2 in Figure 4.10, respectively. The unidentified peaks are due to impurities in the sample. ' 3000 3500 Figure 4.12. "I cleavage of un isomers. Befo remove CXCgSS 175 1-38 145-180 \1 . O 10 22 or “"3 L 3000 3500 4000 4500 5000 5500 6000 6500 m/z Figure 4.12. The MALDI mass spectrum of peptide mixtures resulting from the cleavage of unseparated, partially reduced/cyanylated soybean trypsin inhibitor (STI) isomers. Before the cleavage, the mixture was dialysed against 1% HAO solution to remove excess reagents. See table 4-3 for possible fragments and their m/z values. analysis of the clea' 5259.9, 10752 are unidentified peaks data, a Cys39-86 10 In the pres remove excess rea: poor chromatogra] separation, instead two disulfide bc reduced/cyanylate _Sp€Ctrum of the t fragments cones; at both of the disr attributable to ; reduced/Cyatnylat at each of the fc Would be difficul F‘ Application As We re bond linkage 13} Under alkaline c Closely Spaced 176 analysis of the cleavage products from HPLC peak 2. The fragments at m/z at 4059.1, 5259.9, 10752 are due to the fragments 1-38, 39-85 and 86-180, respectively. Other unidentified peaks are apparently due to impurities in the HPLC fraction. From these data, a Cys39-86 linkage is deduced. In the present procedure, the cyanylated proteins are separated by HPLC to remove excess reagents and to separate the resulting isomers. For simple proteins with poor chromatographic performances, such as STI, it could be better to avoid HPLC separation, instead, use dialysis to get rid of excess reagents. Since STI only contains two disulfide bonds, the cleavage products from the mixture of the partially reduced/cyanylated isoforms were expected to be reasonably simple. The MALDI _ spectrum of the cleavage products from the unseparated isomers (Figure 4.12) shows fragments corresponding to the cleavage at each of the two disulfide bond sites and also at both of the disulfide bonds (see Table 4.3). For example, the fragment at m/z 5738.8 is attributable to fiagment 86~135. Apparently, the mixture contains the doubly reduced/cyanylated STI, which gives rise to the fragments corresponding to the cleavage at each of the four cysteine residues. The unambiguous assignment of disulfide pairs would be difficult in such cases. F. Application to Proteins Containing Closely Spaced or Adjacent Cysteines As we reviewed in chapter 1, two problems remain in the recognition of disulfide bond linkage by conventional methods; one is the disulfide bond scrambling occurring under alkaline conditions and another is the failure to cleave protein chains between two closely spaced or adjacent cysteines. Our methodology can circumvent problems with disulfide bond excl conditions. Equall containing closely s the proteolytic clea of cystine and cyan The feasibi containing closely human recombinar (LR3IGF-l), and ii Recombin; resembled its nat initiator 0f proteir the disulfide bon. assignment of dis by an 15811 residr containing Only c thsnsncgp; Figue 4.13. epidermal gr 177 disulfide bond exchange, because the chemical reactions were performed under acidic conditions. Equally important is that our methodology is also applicable to proteins containing closely spaced or adjacent cysteines, because our approach does not depend on the proteolytic cleavage between cysteine residues, but depends on the partial reduction of cystine and cyanylation of the corresponding nascent cysteine residues. The feasibility of our methodology to recognize disulfide linkage in proteins containing closely spaced and adjacent cysteine residues is demonstrated here using human recombinant epidermal growth factor (hEGF), LONG R3 insulin growth factor-I (LR3 IGF-I), and insulin as examples. Recombinant hEGF used in this research was purchased from Sigma. It resembled its native form (53 amino acids), but with a Met at the N-terminus as an initiator of protein biosynthesis (26, 27). Figure 4.13 shows the amino acid sequence and the disulfide bond linkage in recombinant hEGF. Conventional methodologies for the assignment of disulfide pairs in hEGF is tedious because Cys31 and Cys33 are separated by an Asn residue, which is usually resistant to proteolytic digestion to give peptides containing only one disulfide bond. * J . 1.. 1 MN SD SEéPLSI-IDGYCILHDGVCMYIEALDKYA WGYIGER QYRDLKWWELR - 1 6' 1 4 2 0 3 l 3 42 5 3 Figure 4.13. Primary structure and disulfide bond linkage of recombinant human epidermal growth factor (hEGF) (MW. 6347.2) Table 4.4. resulting 11 designated Reductior of Disulfi Cys6-Cys Cysl4-Cy. Cys33-C} Cys6-Cy: Cysl4-Cj Cys6~Cy Cys33-C Cysl4- Cys33- Cys6.r Cysl4 Cys33 178 Table 4.4. Calculated and observed m/z values for possible fragments resulting from the cleavage of recombinant hEGF chains at sites of designated cysteine pairs Reduction of Disulfide Fragment Calculated m/z Observed m/z -1-5 682.7 nd 6-19 1541.6 1540.1 CYS6‘CYS20 2053 4217.9 4218.3 6-53 5676.5(5735.5) 5734.4 113 1555.6 1554.6 Cys 14-Cys3l 1430 1970.3 1969.2 31-53 2916.3 2917.1 132 3694.1 3692.9 3341 1020.1 1020.2 Cys33-Cys42 4253 1721.0 1722.0 .141 4641.2 4643.1 33-53 2664.1(2723.1) 2666.3(27259) -1-5 682.7 11d 6-13 915.9 915.9 Cys6-Cy520 14-19 667.7 nd (”51443531 20-30 1344.6 1345.0 31-53 2916.3 2916.0 14-30 1936.3 1934.7 46 682.7 nd 6-19 1541.6 1541.1 Cys6-Cys20 2032 1560.8 11d Cy833-Cys42 3341 1020.1 1020.0 4253 1721.0 1721.7 3353 2664. 1(2723. 1) 2664.9(2725.1) 113 1555.6 1555.2 14-30 1970.3 1970.1 Cysl 4-Cy831 3 1-32 260.2 . . nd Cys33,cys42 3341 1020.1 1020.6 42-53 1721.0 1722.5 3353 2664.1(2723. 1) 2666.9(2726.4) -1-5 682.7 nd 6-13 915.9 915.8 1419 667.7 nd Cys6-Cys20 20.30 1344.6 1344.3 Cys 14-Cy331 31-32 260.2 nd Cys33-Cys42 33.41 1020.1 1020.1 4253 1721.0 1722.0 14-30 1936.3 1935.8 33-53 2664. 1(2723. 1) 2665.9(2727.3) 15 Figure 4.14. Cyanylated i: rate of 1.0 111 15-50 min, . reuresent sir anal1’Sis. 179 Intact I 2 L7 MW ' 15 20 25 '30 35 40 45 min Figure 4.14. HPLC separation of denatured hEGF and its partially reduced/ cyanylated isomers. Separation was carried out on a Vydac C18 column at a flow rate of 1.0 mllmin with a linear gradient 15-35% B in 15 min, and 35-55% B from 15-50 min, where A = 0.1% TFA in water and B = 0.1% TFA in CH3CN. Peaks 1-3 represent singly reduced/cyanylated species, as determined by MALDI-TOF analysis. After partial obtained by HPLC the first peak is int: doubly reduced/cy Table 4.4 lists the 1 peptide chains at reduced and cya lactalbumin, the r1 experimental cor characterized. Fi cleavage of singlj respectively. Fig isoforms and the The mas represented by p ‘0 fragments 6 cYtlnylated/uncl. is not detectabl. chain cleavage Cleavage produ .41, and 42_53 4.15B). Additi 180 After partial reduction and cyanylation of hEGF, eight well-resolved peaks were obtained by HPLC (Figure 4.14). Mass analysis of the peaks by MALDI indicated that the first peak is intact protein, while peaks 1-3, 4-6, and 7 are singly reduced/cyanylated, doubly reduced/cyanylated, and completely reduced/cyanylated isoforms, respectively. Table 4.4 lists the expected and observed m/z values for fragments due to cleavage of the peptide chains at different cysteine sites depending on which disulfide bond(s) were reduced and cyanylated. Unlike the chromatograms of ribonuclease A and or- lactalbumin, the reduction of two disulfide bonds in hEGF was also observed under the experimental conditions and, therefore, the corresponding fractions were also characterized. Figure 4.15A-C are MALDI spectra of peptide mixtures resulting from cleavage of singly reduced/cyanylated hEGF isomers corresponding to HPLC peaks 1-3, respectively. Figure 4.16A-D are the MALDI mass spectra of doubly reduced/cyanylated isoforms and the completely reduced/cyanylated isoform, respectively. The mass spectrum in Figure 4.15A corresponds to the cleavage products represented by HPLC peak 1 in Figure 4.14. The peaks at m/z 1540.1 and 4218.3 are due to fragments 6-19 and 20-53, respectively. The peak at m/z 5734.4 is due to a cyanylated/uncleaved fragment 6-5 3 with an expected m/z of 5738.5. The fragment -1-5 is not detectable by MALDI in any case. From these data, one can deduce that peptide chain cleavage occurs at Cys6 and CysZO. Therefore, Cys6 is linked to CysZO. The cleavage products of HPLC peak 2 shows m/z values attributable to fragments -1-32, 33- ,41, and 42-53, respectively, suggesting Cys33 and Cys42 are cleavage sites (Figure 4.15B). Additionally, two overlapped peptides, 33—53 and -1-41 confirm the conclusion. 6-19 42 33-41 ++ 1’2 4:41 1 LE. 1000 Figure 4.15. “the three si 1‘3 in Fig. 4. SWiles, B-eli also Table 4,; 181 6-53* peak 1 -1-53 6-19 ++ 20-53 ++ 42-53 peakZ * *- 33-41 (13% ('1 m,” .132 ++ -1-41 31-53 peak 3 2m 4.3 8 v-< <- 3 :++|| l l ll (0 L _ r - _ r a. A. A... . 1000 2000 3000 4000 5000 6000 7000 m/z Figure 4.15. The MALDI mass spectra of peptide mixtures resulting from the cleavage of the three singly reduced/cyanylated hEGF isomers, corresponding to the HPLC peaks 1-3 in Fig. 4.14, respectively. The symbols ++ , *, and ** represent the doubly-charged species, B-elimination products, and cyanylated/uncleaved products, respectively. See also Table 4.4 for the calculated and observed m/z values. Overall, a disulfide The cleavage prodt 4.15C). In additir disulfide bond link detected, correspor 5) confirm that thr sites. However,l bring about confu The inter reduced/cyanylatt reduced/cyanylat Strategy iS help correSponding to sites”, that is, th (Figure 4.16A) fragments 6-19, CYS42, reSpecti Observed. Obv Strategy, the M the disulfide ll MALDI spectr that CleaVage C g'> fragment .1 182 Overall, a disulfide bond linkage between Cys33-Cys42 can be unambiguously deduced. The cleavage products from HPLC peak 3 present a little more complicated case (Figure 4.15C). In addition to the desired fragment -1-13, 14-30, and 31-53 implicating the disulfide bond linkage between Cysl4-Cys31, two fragments, 31-41 and 42-53, were also detected, corresponding to the undesired cleavage at Cys42. Further experiments (chapter 5) confirm that the cleavage at Cys42 is minor in comparison with the cleavage at other sites. However, because of the high responses on MALDI, these two fragments could bring about confusion in data interpretation. The interpretation of MALDI data from the cleavage products of doubly reduced/cyanylated isoforms is never as straightforward as that from singly reduced/cyanylated isomers, because more fragments are observed. A slightly different strategy is helpful for the assignment. Instead of looking at the MALDI peaks corresponding to the cleavage sites, the absence of such peaks represents the “uncleaved sites”, that is, those still linked by a disulfide bond. For example, the MALDI spectrum (Figure 4.16A) of the cleavage products from HPLC peak 4 shows peaks due to fragments 6-19, 33-41, and 42-53, indicating the cleavage at Cys6, CysZO, Cys33, and Cys42, respectively. No fragments due to the cleavage at Cysl4 and Cys31 can be observed. Obviously, these two cysteines are linked as a disulfide pair. With the same strategy, the MALDI spectra of the cleavage products from HPLC peaks 5 and 6 reveal the disulfide linkage Cys33—Cys42 and Cys6-Cy320, respectively. The analysis of the MALDI spectrum of the cleavage products from the completely cyanylated hEGF shows that cleavage occurs at each of the cysteine positions, although some of the fragments (e. g‘a fragment -1-5) might be missing on MALDI. 42-5 33-41 owl —r k t. ‘3.“ 1 .go 3 éa 5- o T 33-41 ++ 20-30 1000 Figure 4.16. ’ 0f the threec cyanylated h The SMbols and cyanylat and Observe. lR’Z 4253 peak 4 31-53 42-53 peak 5 "" o «5g 319: 1430* J-J a .9135: ML - . _ 42-53 3341 peak 6 41+ ‘1'13 14-30 ("in 42-53 3341 peak 7 .... at: +v—4 r M ‘5 M m Ar)" 20.30 M~ __ 1000 2000 3000 4000 5000 6000 7000 m/z Figure 4.16. The MALDI mass spectra of peptide mixtures resulting from the cleavage of the three doubly reduced/cyanylated hEGF isomers and a completely reduced/ A cyanylated hEGF,corresponding to the HPLC peaks 4-7 in Fig. 4-14, respectively. The symbols ++ , *, and ** represent doubly-charged species, B-elimination products, and cyanylated/uncleaved products, respectively. See also Table 4.4 for the calculated and observed m/z values. Recombina replacing glutamat LR3IGF-I shows hi contains adjacent scheme, assigned « below, has never recombinant s0ur< as for lGF-l isola have a major infl bonded isomers c Figure 4 reduced/cyanyla With the tYpica First, the major the Singly redur lemme Smalle: most of the pi Second, Only t‘ 2in Figure 4.1 184 Recombinant LR3IGF-I is a variant of human IGF-I that contains arginine replacing glutamate-3 as well as an amino terminal extension of 13 amino acids (28). LR3IGF-I shows higher biological activities than its analog, IGF-I. Like IGF-I, LR3IGF-I contains adjacent cysteine residues at 60 and 61 positions. The disulfide bonding scheme, assigned on the basis of homology to the insulin (or IGF-I) sequence and shown below, has never been verified experimentally. In preparing authentic LR3IGF-I from recombinant sources, it is important to confirm that the disulfide bond linkage is the same as for IGF-I isolated from natural sources, since a mismatching of disulfide bonds could have a major influence on any biological activity, as has been observed with disulfide- bonded isomers of insulin (29). 1 19 31 6061 65 74 l j r 89“ Figure 4.17 shows the HPLC separation of LR3IGF-I and isomers of its partially reduced/cyanylated species. The chromatogram shows two exceptions in comparison with the typical chromatographic pattern of the partially reduced/cyanylated isoforms. First, the major peak representing the intact protein shows longer retention .than one of the singly reduced isoforms. This result indicates that the hydrophobicity of the protein become smaller than that of the intact protein after opening one disulfide bond, whereas most of the proteins show larger hydrophobicity after splitting the disulfide bond(s). Second, only two singly reduced/cyanylated protein isomers were observed (peaks 1 and 2 in Figure 4.17), even though the protein contains three disulfide bonds. Stronger b 5 10 Figure 4.17 cyanylated rate of 1.0 1 TFA in wa Tedllced/cy 185 Intact 5 1h 1T5 20 35 30 3'5 40 45 min Figure 4.17. HPLC separation of denatured LR’IGF-I and its partially reduced/ cyanylated isomers. Separation was carried out on a Vydac C1 8'colu1rm at a flow rate of 1.0 mein with a linear gradient 30-50% B in 45 minutes, where A = 0.1% TFA in water and B = 0.1% TFA in CH3CN. Peaks 1 and 2 represent singly reduced/cyanylated species, as determined by MALDI-TOF analysis. reducing condition prolong the reduct isomers, but the tl isomer from the retention time as MALDI-MS 0f tl disulfide bond (C under the reductir understandable f stability of IGFJ in LR3IGF-I) is might eXplain th Table 4,; Cleavage reactir The two HPLC under the descr Figure 4-18 A 2 fragments 65-3 the cleavage re m/z 2752.9 an _ Where B‘elimi; Cleavage sites, 186 reducing conditions (increase the ratio of reducing agent, apply higher temperature, and prolong the reduction time) result in the formation of doubly reduced/cyanylated protein isomers, but the third disulfide bond still refuses to reduce. At first we assume that the isomer fiom the reduction and cyanylation of the third disulfide bond had the same retention time as its intact protein and, therefore, coelutes with the intact one. But MALDI-MS of the cleavage products indicates that it is not the case. Apparently one disulfide bond (Cys31-Cys74, as concluded from the next paragraph) is inert to reduction under the reduction conditions, even if the protein has been denatured. This conclusion is understandable from the thermodynamic point of view. A recent study on the disulfide stability of IGF-I showed that the disulfide bond 18-61 (corresponding to disulfide 31-74 in LR3IGF-I) is superstable and preferentially formed in the folding process (30), which might explain that this disulfide pair is structurally stable enough to resist the reduction. Table 4.5 lists the calculated m/z values for possible fragments resulting from the cleavage reaction of LR3IGF-I chains at sites corresponding to different cysteine pairs. The two HPLC fractions (HPLC peaks 1 and 2 in Figure 4.17) were subjected to cleavage under the described conditions. The corresponding MALDI mass spectra are shown in Figure 4.18A and B. The peaks at m/z 2188.7 and 6372.9 in Figure 4.18A are due to fragments 65—83 and 1-59, respectively. The fragment 60-64, a middle fragment during the cleavage reactions, was missing from the MALDI spectrum. However, the peaks at m/z 2752.9 and 6936.5, representing two overlapped peptide fragments 60-83 and 1-64, where B-elimination occurs at Cys65 and Cys60, respectively, confirm that 60 and 65 are cleavage sites. From these data, one can deduce that Cys60 is linked to Cys65. The Table 4.5 resulting designao Reduc 0f Dis Cys60 Cysl' Cysi Table 4.5. Calculated and observed m/z values for possible fiagments resulting from the cleavage reaction of LR3IGF-I chains at sites of designated cysteine pairs Rafa“ Fragment Calculated m/z Observed m/z of Dlsulfide 1-59 6371.2 6372.9 60-64 637.7 nd Cys60-Cys65 65-83 2188.6 2188.7 0(1-64) 6931.9 6936.5 [3(60-83) 2749.3 2752.9 1-18 1978.4 1976.9 19-60 4538.0 4538.6 Cys 1 9-Cys61 61-83 2681.1 2682.5 [3(1-60) 6439.4 6441.9 13(19-83) 7142.1 7139.1 1-30 3223.8 31-73 4964.5 Cys3 1-Cys74 74-83 10092 no reduction [30-73) 8111.3 501-33) 5897.7 1000 2 Figure 4.18. Of the W0 Si peaks 1 and B-eliminatic also Table 4 188 65-83 A 1-59 L A All 60-83* 1-64* 1-83 61-83 ’ B 1-18 19-60 \7 922/ *0 “? Bibi 521A A 1000 2000 3000 4000 5000 6600 7600 8000 9000 10000 m/z Figure 4.18. The MALDI mass spectra of peptide mixtures resulting from the cleavage of the two singly reduced/cyanylated LR3IGF-I isomers, corresponding to the HPLC peaks 1 and 2 in Figure 4-17, respectively. The symbol * represent the protonated B-elimination products. The peaks with question marks are discussed in the text. See also Table 4.5 for the calculated and observed m/z values. cleavage products 60, and 61-83, rt 4.18B). Addition linked to Cys61. can be easily assi digestion. As i; assume adjacent disulfide structuj lGF-l and 161?. Shovvn). 11 is als to study the refo it shoulc Confusing in ha 103 Da from th the peak with fragment 19-6( Caused by the exchange b ctr unexplainable, Cl’S-Gly. Stru. than disulfide other hand, 11 189 cleavage products from HPLC peak 2 show m/z values attributable to fragments 1-18, 19- 60, and 61-83, respectively, suggesting Cysl9 and Cys61 are cleavage sites (Figure 4.18B). Additionally, an overlapped peptide, 1-60, confirm the conclusion that Cysl9 is linked to Cys61. Overall, the disulfide bond structure containing two adjacent cysteines can be easily assigned without multiple enzymatic digestions and/or multiple step Edman digestion. As insulin growth factor contains a number of protein analogs which all assume adjacent cysteine structures, the data presented here are of great importance to the disulfide structure study of such a protease. As evidence, preliminary experiments on IGF-I and IGF-II, both containing adjacent cysteines, show similar results (data not shown). It is also expected that the approach described here will be sufficiently powerful to study the refolding intermediates of the IGF analogs (3 0). It should be pointed out that the question-marked peaks in Figure 4.18B are very confusing in nature. The peak with a single question mark showed a mass increase of 103 Da from the fragment 1-18, suggesting a cysteine attached to the fragment, whereas the peak with double question marks showed a mass decrease of 103 Da from the fragment 19-60, suggesting a cysteine removed from the fragment. The latter might be caused by the cleavage at Cy559, rather than Cys60, assuming that the disulfide bond exchange between two adjacent cysteine residues occurs. The former one is unexplainable, because such an exchange can never occur at Cysl9, which has the -Lys- Cys-Gly- structure. Obviously structural heterogeneity of recombinant proteins, other than disulfide exchange, would be critical to wield occurrence of such fragments. On the other hand, unambiguous assignment of the disulfide structure is evident even if such undesired fragme native disulfide b Insulin cc intrachain (Cys 1 residues (Cys } cyanylated insulf Ach B ch The chr disulfide bond Preferred. 0t partially reduc reduced/cyany disulfide bond A prol gives a good mode, Whilei exl3ecte,d and for fragment: 190 undesired fragments are present, because the fragments from specific cleavage at the native disulfide bond are sufficient to make a positive conclusion. Insulin contains two interchain (Cys A7-Cys B7 and Cys A20-Cys B19) and one intrachain (Cys A6-Cys A1 1) disulfide bridges, and includes a pair of adjacent cysteine residues (Cys A6, Cys A7). The HPLC chromatogram of the partially reduced] cyanylated insulin isoforms is shown in Figure 4.19. l 7 l . 20 A chain GIVEQgCASVCSLYQLBNYCN ‘ I 11 / B chain FVNQHLCllGSHLVEALYLVgGERGFFYTPKA The chromatogram shown by Gray (5) indicates that the reduction rates of the disulfide bonds in native insulin vary widely and that one route of reduction was preferred. Our chromatogram (Figure 4.19) of denatured insulin and isomers of its partially reduced/cyanylated form shows a peak for each of the three possible singly reduced/cyanylated species. Our data indicate that under denaturing conditions, the three disulfide bonds undergo reduction at comparable rates. A problem with the analysis of insulin fragments by MALIDI is that its A-chain gives a good signal in the negative ion mode, but a very weak signal in the positive ion mode, while the B-chain gives a good response in both modes (31). Table 4.6 shows the expected and observed m/z values (deprotonated molecules detected in negative mode) for fragments resulting from the cleavage reaction of insulin chains at sites corresponding 15 Figure 4.19 cl’itnylated rate of 1.0 TFA in wa reduced/c} marked A 191 Intact - 1.1 tum 15 2O 25 30 35 40 45 50 Figure 4.19. HPLC separation of denatured insulin and its partially reduced/ cyanylated isomers. Separation was carried out on a Vydac C18 column at a flow rate of 1.0 ml/min with a linear gradient 20-50% B in 40 minutes, where A = 0.1% TFA in water and B = 0.1% TFA in CH3CN. Peaks 13 represent singly reduced/cyanylated species, as determined by MALDI-TOF analysis. The peaks marked A and B are insulin A-chain and B-chain, respectively. Table 4.6 fragment of design Reduc 0f Dis CysA6-l CysA7- CysA2' 192 Table 4.6. Calculated and observed m/z values [M-H]' for possible fragments resulting from the cleavage reaction of insulirchains at sites of designated cysteine pairs Reduction . Fragment Calculated m/z Observed m/z of Disulfide Al-S 543.6 nd A6-10 505.6 nd Al 1-21 1373.8 1374.2 CysA6-CysA1 1 A6-21 1803.4 1804.8 A1-21 2340.0 2339.6 B1-30 3399.9 3400.3 Al-6 646.7 nd A7-21 1734.9 1735.7 C A7-C B7 ys ys B1-6 755.9 nd B7-3O 2685.1 2686.3 A1-l9 2123.7 2122.6 A20-21 261.3 nd CysAZO-CysB 19 B 1-18 2043 .4 2041.5 B19—3O 1401.5 1402.3 B1-30 3400.0 (3366.0) 3400.1 (3366.1) All-21 Figure 4.20. the cleavage HPLC peak B‘eliminatic Table 4.6 fC 193 131.30 A _ A1-21 A11 21 A6-21* \ kw A7-21 B 13730 A1-19 .. r. o a :3 ._'.. C Q) m / 11319-30 l | B1-18 1000 w 1500 2000 2500 3000 3560 4000 m/z Figure 4.20. The negative ion MALDI mass spectra of peptide mixtures resulting from the cleavage of the three singly reduced/cyanylated insulin isomers, corresponding to the HPLC peaks 1-3 in Figure 4.19, respectively. The symbols * and ** represent the B~elimination products and cyanylated/uncleaved peptides, reSpectively. See also Table 4.6 for the calculated and observed m/z values. to different cystei mass spectra. Fig in Figure 4.19) sl resulting in the fi Al-S and A610 plagued by the cleavage. Howe of the adjacent 4.19 show Simil corresponding 1 1735.7 is attrib‘ 0f the insulin . SPecific fragme described here proteins COHtai Al-19, B 1-18 linked 10 CySE V' Picomole: So far, by the detect methOdologie diSUlflde bOn‘ 194 to different cysteine pairs. Figures 4.20A-C are the corresponding negative ion MALDI mass spectra. Figure 4.20A (corresponding to the cleavage products of the HPLC peak 1 in Figure 4.19) shows that chemical cleavage occurs at Cys6 and Cysll of the A chain, resulting in the fragment A1 1-21 and A6—21. Two low-mass fragments corresponding to Al-S and A6-10 are not readily detected by MALDI. Furthermore, the mass spectrum is plagued by the unidentified impurity peaks, probably resulting from the nonspecific cleavage. However, it is still convincing to conclude that cleavage indeed occurred at one of the adjacent cysteine sites, CysA6. The cleavage products from HPLC 2 in Figure 4.19 show similar results (Figure 4.20B). In addition to an expected peak at m/z 2686.3, corresponding to fragment B7-30 (calculated m/z 2685.1), another intense peak at m/z 1735.7 is attributed to fragment A7-21 (calculated m/z 1734.9), resulting from cleavage of the insulin A—chain at Cys A7 which is adjacently linked to Cys A6. These two specific fragments suggest that Cys A7 be linked to Cys B7. Therefore, the methodology described here shows the potential for the assignment of disulfide bond pairings in proteins containing close or adjacent cysteine residues. Figure 4.20C gives fragments Al-19, B 1-18, B19-30, plus an overlapped peptide B1-30. It is obvious that CysA20 is linked to CysBl9. V. Picomole Scale Reactions So far, the sample size used in our protocol is ~10 nmol. The sensitivity is limited by the detection limit of our HPLC systems. Although it is better than conventional methodologies, our methodology described here should have the potential to assign disulfide bonds at the picomole level of proteins. One improvement in the methodology —¥— 30 Figure 4.: Species b: Safneast 195 awn 41mm“. ____, ,J Lm_m 30 40 5'0 60 70 minute Figure 4.21. separation of ribonuclease A and its partially reduced/cyanylated species by microbore C18 HPLC column (1x100mm). HPLC conditions are the same as those in Figure 4.4, except the flow rate is 0.05 ml/min. is to use a microbr microscale is a cl Twice as much th The CDAP conce used in protocol ‘ the lower conce reduced/cyanylat mm C18 microb was obtained 0: preliminary resn samples. V1. Character The an: facets of simpl the case of mul only one of 11 cYsteines from masses of the reduced isofo: itself to 00m] ehmination p 196 is to use a microbore column HPLC to separate the protein isoforms. To run chemistry at microscale is a challenge to our protocol. Appropriate modifications have to be made. Twice as much the TCEP reducing agent had to be used to promote the partial reduction. The CDAP concentration for the subsequent cyanylation is five times as much as that used in protocol to prevent the autohydrolysis of CDAP reagent, which is accelerated in the lower concentration medium. Figure 4.21 shows the separation of the partially reduced/cyanylated isoforms of 500 picomoles of ribonuclease A obtained on a 1x100 mm C18 microbore column (thanks to the generosity of Dr. Gage). A better separation was obtained on the microbore column, while the retention time is shorter. These preliminary results indicate that our chemistry can be used at microscale level for protein samples. VI. Characteristics of the Methodology The analytical advantage of our proposed methodology accrues from several facets of simplicity. First, the optimized procedure for partial reduction assures that in the case of multiple disulfides there will be significant quantities of all isoforms in which only one of the cystines has been reduced. Second, cleavage of a cyanylated pair of cysteines from a newly reduced cystine yields only three fragments. Third, the sum of masses of the three cleavage fragments is equal to the mass of the cyanylated singly reduced isoform plus the mass of two molecules of water, a feature that readily lends itself to computerized assessment of the experimentally determined data. Fourth, [3- elimination products are usually observed which confirms the assignment of cleavage data by peaks thi sites. This is an consistency of d likelihood that a? is quite low, evr resolvable by TV reduced and cyz results reported is the risk of dis Because our e scrambling are evidence that : indeed observe anticipated fro Signal MALDI is stil does make as: those Shown i electrOSpray problem, Fu Which may b batch analySl 197 data by peaks that represent sequences of residues that overlap one of the two cleavage sites. This is another feature that lends itself to computerized assessment of the internal consistency of data interpretation. Fifth, for relatively large proteins (> 10 kDa) the likelihood that alternative disulfide pairings would lead to the same mass spectral pattern is quite low, even if some mass Spectral peaks for two alternative pairings may not be resolvable by MALDI. Sixth, complete HPLC separation of the isomers of the singly reduced and cyanylated species is unnecessary for identification as demonstrated in our results reported here. Seventh, one of the severe problems in conventional methodology is the risk of disulfide bond scrambling which usually occurs in an alkaline medium (1-3). Because our experiments are performed under acidic conditions, opportunities for scrambling are kinetically suppressed. Complete analysis of the MALDI data gives no evidence that scrambling occurred during our procedure. Some unexpected peaks are indeed observed in our spectra, but their masses do not appear to be related to fragments anticipated from scrambled disulfide bonds. Signal suppression for some components of a complex mixture in the analysis by MALDI is still a problem (7). The suppression of signal from certain peptide fragments does make assignments risky when they are represented by minor peaks such as some of those shown in Figure 6. The complementary nature of analytical results by MALDI and electrospray (ESI) (32) would seem to provide one promising way to deal with this problem. Furthermore, ESI provides ready access for analysis of mixtures by LC-MS, which may be an effective alternative to problems of analyte suppression encountered in batch analysis. VII. Conclusion We have chemical modifi subsequent mas alternative meth known primary minimizing disr Furthermore, t1 cysteines for w sensitive. We: ionization to th biological sour 198 VII. Conclusions We have demonstrated that the combination of partial reduction of a protein, chemical modification of the sulfliydryl groups, cleavage of the peptide chain, and subsequent mass-mapping by MALDI-TOF MS provides a simple and effective alternative methodology for the assignment of disulfide bond pairings in proteins of known primary structure. Our novel approach offers an important advantage of minimizing disulfide bond scrambling, a concern in most conventional methodologies. Furthermore, this approach may be applied to proteins containing close or adjacent cysteines for which conventional approaches fail. The procedure is simple, fast, and sensitive. We are currently applying this methodology in combination with electrospray ionization to the characterization of disulfide bond pairings in proteins from a variety of biological sources. VIII. Reference 1. Hirayama, K (Matsuo T, < p299-312. 2. Smith, D. L. 3. Ryle,A.P., 4. Zhou,Z.,an 5. Gray, W. R. 6. Gray, W. R. 7. Wu ,J., Gag 8. Jacobson,( 248,6583-1 9- W181. and 10- Wakselma 11.8hapira,E 12. Creighton. 13.Kuwajimg 29, 8240-: 14. Reeve, J, 15-T€l14.5) would catalyze the hydrolysis of the CDAP reagent. The final pH of the reaction was monitored by a pH meter. C. Separation of Folding Intermediates by HPLC Folding intermediates of hEGF trapped at 30min, 3h, 24h, and 48h, respectively, were separated by HPLC (Figure 5.4). In order to interpret these chromatograms, structural information of the fractionated intermediates was analyzed by MALDI-TOF MS. A total of 18 peaks were identified at different time points. Peaks 1, 3, and 4 represent 3-disulfide species. Peaks 2, 5, 6, 7, 10, and 12 represent 2-disulfide bond species. Peaks 13, 14, 16, and 17 are l-disulfide bond species. Peak 18 is the reduced/unfolded hEGF. [Peaks 8, 9, 11, and 15 contain both 2- and 1-disulfide species. The results revealed that the intermediates consist of eight l-disulfide isomers and nine 2- 30min 3h 24h 48h Flgure : differer 218 % 18 /Il-A /R 30mm l‘A l-B I'C l-D ‘ is ‘Z 16 l J 56 7 8 910“ 12 1415 17 P 3h r 4 24h . . 3 I Ml ‘— J\~_ w 4 __ N \ 48h Ill-B M zofoo ' 36.00 1 401.09 minutes Figure 5.4. HPLC separation of CDAP-trapped folding intermediates of hEGF at different time courses. disulfide isomer are some overlaj It was a] of each disulfie folding are ess disulfide specie of the species after 3 hours 0 the 2-disulfrde almost disappe species were 6 are similar to Chang’s pape: 3- or l-disulf This is not sr than it should nOn-native 3 MALDI-MS cthOmatOgrau As a (iodoacetate, Profiles Shc intermediate 219 disulfide isomers. Most l-disulfide species eluted after 2-disulfide species, while there are some overlaps or reversed order. ' It was apparent that along the folding process, equilibrium existed among isomers of each disulfide species. The HPLC profiles at 10- (data not shown) and 30-min of folding are essentially the same although the relative ratio of the most populated 2- disulfide Species and the reduced/unfolded species is different. At 30 minutes, about 40% of the species presented as the reduced/unfolded isomer, whereas it almost disappeared after 3 hours of folding. The 3-disulfide species began to appear as a main peak, while the 2—disulfide species dominated. After 24 hours of folding, the 2-disulfide species almost disappeared and the native hEGF dominated. After 48 hours, only the 3-disulfide species were detected, representing complete folding. The HPLC profiles presented here are similar to some of Chang’s results (29), however, there are some differences. In Chang’s paper, thirteen 2—disulfide isomers were identified, some of them overlaped with ,. 3- or 1-disulfide species. We were able to detect nine discernible 2-disulfide isomers. This is not surprising as the sensitivity of our detection system was substantially lower than it should be. Unlike Chang’s chromatogram in which the two peaks representing the non-native 3-disulfide isomers contain a considerable amount of 2-disulfide species, MALDI-MS analysis showed that the corresponding peaks (3 and 4) in our chromatograms represented pure isomers. As a comparison, the HPLC profiles obtained by different trapping techniques (iodoacetate, acid, or CDAP) at different time points are shown in Figure 5.5. The three profiles show almost the same pattern in terms of the appearance of disulfide intermediates. It should be noted however that, in Chang’s experiments, the HPLC Flt- . wommmuwrmaflflmv vommmbrgoxx oumwoomowo 220 mom: a essences maze Baggage Ea .vommwb.Eom 6335-880382 mo 8an Una: we nemfimmaov .m.m oSmE . . m Z a Z Z d . = . _ J52}? 133.431 ti \ . seem : = we a _ momma? commebiado eoammbéo< 380882 profiles of acid- counterparts. l isomers trapper from the under. or less. Howe disulfide inter disulfide inten a1 (29) assum means that all the hydrophol farfetched. cYanylation d SCN groups: acid and CD [A intermediates intermediates populated 2. peak could r characterizat our results imerthedlate measuremer 221 profiles of acid—trapped intermediates did not fully resemble those of iodoacetate-trapped counterparts. It is understandable in terms of the difference in the hydrophobicity of isomers trapped by different manners. The hydrophobicity of alkylated protein differs from the underivatized isomers. Therefore, the relative retention time could change more or less. However, it is beyond understanding that the majority of iodoacetate-trapped 2- disulfide intermediates were eluted within three fractions, whereas acid-trapped 2- disulfide intermediate(s) were accumulated within one fraction (marked as 11). Chang et al (29) assumed all the three fractions coeluted in acid-trapped HPLC profile, which means that all the three isomers had the same hydrophobicity before modification, while the hydrophobicities were substantially different after alkylation. This eXplanation is farfetched. Fortunately, the hydrophobicity of hEGF isomers before and after cyanylation does not change greatly (the polarity of SH and SCN is similar while SH and SCN groups are both small species), direct comparison of HPLC profiles obtained from acid and CDAP trappings is reasonable. It is noteworthy that the HPLC chromatogram of intermediates trapped by the CDAP technique is similar to that representing the intermediates trapped with acid. In both cases, only one peak corresponding to the most populated 2-disulfide intermediate(s) was obtained. To exclude the possibility that the peak could contain more than one 2-disulfide species, as suggested by Chang (29), the characterization of the disulfide structure of the species was carried out (see below) and our results indicated unambiguously that the peak contains ONLY one 2-disulfide intermediate. This result confirms that acid trapping provides a more accurate measurement of folding process. The two major isomers observed by Chang in the chromatogram caused by alkj It was of each disu intermediates fractions for behaved simf previous rep characterized equilibrated CXperiments, (Figure 5.4, convert the ; alkaline me. Tyr, and a r thiols to cat D. Disulfir Tran one is “Pei Alternative a COlOI' 01- 222 chromatogram representing iodoacetate trapping could be from thiol/disulfide exchange caused by alkylation under alkaline condition. It was apparent that along the folding process, equilibrium existed among isomers of each disulfide species. For instance, the concentration of 1- and 2-disulfide intermediates ascended and then descended as folding progressed, but the relative ratio of fractions for each kind of isomers remain constant. Scrambled 3-disulfide species behaved similarly during the folding. These data supported the conclusion drawn by previous reports (29) that, like other proteins, the folding pathway of the hEGF was characterized by a sequential flow of unfolded EGF (R) through three groups of equilibrated intermediates, namely, 1-, 2-, and 3-disulfide (scrambled) isomers. In our experiments, about 40% of the folding intermediates remained as scrambled species (Figure 5.4, 48-h sample). The exposure to air for another 48 hours cannot automatically convert the scrambled species to the native form (furthermore, long hours of exposure to alkaline medium results in minor oxidation of some amino acids such as Met, His and Tyr, and a mass shift from the intact protein would be observed), due to the lack of free thiols to catalyze the disulfide reshuffling. D. Disulfide Mapping of Well Populated Intermediates Traditionally, this can be achieved by a number of strategies. The most common one is “peptide mapping” by isolating and analyzing every enzyme-fragmented peptide. Alternatively, it can be done by selective labeling of disulfide bonds after reduction with a color or fluorescent thiol-specific reagents (17, 31). Both methods need microgram amounts of inte and numerous a As we 6 of cyanylated sulfhydryl gror structure deter defrning disull chemistry beh intermediates I successfully cl 3-Disulfrde Ir Afterz Mass analysis peak showed confirmecl th; were two nor collected ant Strategy desc 33 Straightfo isomers We“ CDAP Was t Table DOSSlblC CYa 223 amounts of intermediates, multiple enzymatic digestions, HPLC separation of peptides, and numerous attempts of sequence analysis. As we demonstrated in chapters 2-4, the mass mapping of the cleavage products of cyanylated proteins provides specific information on the number and location of sulflaydryl groups and disulfide bond pairsf This strategy is also applied to the disulfide structure determination of the folding intermediates of hEGF. The procedures for defining disulfide structure in 1-, 2- and 3-disulfide species differ slightly, but the chemistry behind the methodology is exactly same. A total of seven well populated intermediates (marked with III-A, III-B, II-A, I-A, I-B, LC, and I-D in Figure 5.4) were successfully characterized by our strategy. 3-Disulfide Intermediates After 48-h folding (Figure 5.4, 48-h), three well populated species were separated. Mass analysis by MADLI indicated all three species are 3-disulfide isomers. The first peak showed the identical retention time as the native hEGF. . Disulfide mapping confirmed that the isomer is the native hEGF (data not shown). The latter two peaks were two non-native isomers with scrambled disulfide structures. The two peaks were collected and subjected to the partial reduction/cyanylation/cleavage/mass mapping, a strategy described in chapter 4. The control of partial reduction is important but is never as straightforward as the protocols used in chapter 4. The molar amounts of protein isomers were estimated from the HPLC peak area and the concentration of the TCEP and CDAP was thus calculated. Table 5.1 shows the expected nr/z of fragments resulting from cleavage of 15 possible cyanylated disulfide isomers. Figure 5.6 shows the HPLC separation of partially Table 5.1. 3 of 15 singly Reductie of Disuli 6-14 6-21 14 1. 224 Table 5.1. Expected m/z values for possible fragments resulting fiom the cleavage of 15 singly reduced/cyanylated hEGF isomers Reduction Reduction F t m/ F t m/ of Disulfide ragmen Z 0 f D1S 111 fi d6 ragmen Z 1.5 551.5 1-13 1423.4 6-14 6-13 915.9 14-42 1441 3160.6 1453 4837.6 4253 1721.0 1.5 551.5 1-19 2047.1 6-20 619 1539.6 2031 2030 1344.6 20-53 4214.0 3163 2913.3 1-5 551.5 149 2047.1 6-31 6-30 2840.2 2033 2032 1560.8 31-53 2913.3 33-53 2697.2 15 5515 1-19 2047.1 633 6-32 3056.4 2042 2041 2536.9 3353 2697.2 4253 . 1721.0 1.5 551.5 130 3347.7 642 6-41 4031.5 31-33 31-32 261.3 4253 1721.0 33-53 2697.2 143 1423.4 130 3347.7 1420 14-19 667.7 3142 3141 1236.3 20-53 2913.3 42-53 1721.0 1—13 1423.4 1-32 3563.9 1431 1430 1968.3 3342 33-41 1020.1 31-53 2913.3 4253 1721.0 1.13 1423.4 14-33 1432 2184.5 3353 2697.2 10.00 Figure 5.6 225 Ill-A W) 10.00 ’— 20.‘oo 7 30.30 T who T 5000 minutes Figure 5.6. HPLC separation of III—A and its partially reduced/cyanylated species. ++ ++ +1 33-4l 1000 Figure 5. 0f the th] COHespO. and e re] 226 42-53 A 6-41' ++ l [A l 141* 42-53 33-53 B 1 1-13 ++ 14—32 H \l 14-53* I WW. .5... 42-53 31.-53 C 33-41 * m V: S; 1.19 8 W . f_ .4 f4; W A 1000 2000 3000 4000 5000 6000 m/z Figure 5.7. The MALDI mass spectra of peptide mixtures resulting from the cleavage of the three singly reduced/cyanylated species of non-native hEGF III-A, A-C corresponding to the HPLC peaks 1-3 in Figure 5-6, respectively. The symbols ++ and * represent the doubly-charged species and B-elirnination products, respectively. reduced/cyany] 1-3 were singlj reduced EGF distribution of disulfide bond completely rec products corre spectrum (Fig fragments 6-t fragment 1-5 Cys6 is conne From and 33-53 we S-TB). An 0 Therefore, C haglnent 42. degree of c Cyanylation, intensity 0ft the asstghme Although fr: MC peak 227 reduced/cyanylated protein isomer III-A. Mass analysis of the peaks showed that peaks 1-3 were singly reduced/cyanylated species, the peak eluted at 42.5 min was a completely reduced EGF which should be typically present as a small peak. This atypical distribution of the partially reduced species suggests that the opening of the first two disulfide bonds facilitate the reduction of the last one, resulting in the accumulation of completely reduced species. Figure 5.7 shows the MALDI mass mapping of the cleavage products corresponding to HPLC peaks 1-3 in Figure 5.6, respectively. The MALDI spectrum (Figure 5.7A) of the cleavage products of HPLC peak 1 in Figure 5.6 shows fragments 6-41, 42-53, and an overlapped B-elimination product 1-41. The small fragment 1-5 was missing. It is clear that cleavage occurs at Cys6 and Cys42. Therefore, Cys6 is connected to Cys42 as a disulfide bond. From the cleavage products of HPLC peak 2 in Figure 5.7, fragments 1-13, 14-32, and 33-53 were detected, indicating peptide chain cleavage at Cysl4 and Cys33 (Figure 5.7B). An overlapped B-elimination product, 14-53, corroborates with the conclusion. Therefore, Cysl4 is linked to Cys42. However, this assignment was questioned by the fragment 42-53, indicating that Cys42 also underwent cleavage. Apparently a minor degree of cleavage must have occured at Cys42, either by the disulfide split and cyanylation, or by some disulfide exchange during cleavage process, or both. The peak intensity of this fragment is so strong that further experiments were performed to confirm the assigmnent. Although fragment 42-53 is indicated in the MALDI spectra of cleavage products from HPLC peaks 1-3, the absolute concentration of the fragment is quite different in the ro.oo\ Figu] redue 2 in] 228 fragment 42-53 A 42-53 . L. 10.00 ' 20Too T 30:00 ' 4600 ’ 50T00 minutes Figure 5.8. Comparison of HPLC profiles of cleavage products of two singly reduced/cyanylated III-A isomers. (A) and (B) are represented by peaks 1 and 2 in Figure 5-6, respectively. three cases. Tl the same HPLt that the fractio. fragment repre while the fragr so small corr fragment inder fact, the peak mixture of cle 1% contamin. the data inter] The n the fragment pair (Cys20- Again, the M The 1 Salne strateg; Species, Pe: anEllysis by Wmmfi elimination - 229 three cases. The cleavage products from HPLC peaks 1 and 2 were re-fractionated under the same HPLC conditions (Figure 5.8). The MALDI analysis of the fractions indicated that the fraction eluting at 17.5 min corresponded to fragment 42—53. It was clear that this fragment represented one of the major cleavage products in HPLC peak 1 (Figure 5.7A), while the fragment 42-53 in the mixture of the cleavage products from HPLC peak 2 was so small compared to other fragment(s) that it could be neglected. However, this fragment indeed plagued every MALDI spectrum due to its high response. As a matter of fact, the peak corresponding to fragment 42-53 was still intense on MALDI even if the mixture of cleavage products from HPLC peak 1 was diluted by 100-fold. That means 1% contamination by fragment 42—53 could generate a false information and jeopardize the data interpretation. The mixture of the cleavage products from HPLC peak 3 showed the presence of the fragments 1-19, 20-30, 31-53 (Figure 5.7C), indicating the linkage of last disulfide pair (CysZO-Cys31) which is already deducible after the first two pairs were defined. Again, the MALDI spectrum is also plagued by a trace of the undesired fragment 42-53. The disulfide structure in hEGF isomer III-B was subjected to analysis by the same strategy. Figure 5 .9 shows the HPLC separation of the partially reduced/cyanylated Species. Peaks 1-3 are three singly reduced/cyanylated isomers, as confirmed by mass analysis by MALDI. The MALDI analysis of cleavage products from these peaks are Shown in Figure 5.10A-C. Figure 5-10A shows the fragments 6-13, 14-53, and a [3- elimination product 1-13. It is clear that Cys6 is hooked to Cysl4. A small peak roho‘T Figure 230 III- B 1 2 10.00 ‘ 20.00 ’ 30T00 T 40.00 T 5000 minutes Figure 5.9. HPLC separation of III-B and its partially reduced/cyanylated species. 6-13 1’1‘: 33-41 31-4' --A ++ h 1000 — Figure 5. 3 sing1y 1 to the HI 6-13 A 14-53 *42-53 52 'I" ++ .. l .__._.22 - 2- 42-53 B 33-41 33-53* -. L'i’- 1311* 31-41 42’53 31_53 C 1-19 0 «- C? m E?» ”9 4 ++ I g; 1-30 1000 2000 3000 4000 5000 6000 m/z Figure 5.10. The MALDI mass Spectra of peptide mixtures from the cleavage of the 3 singly reduced/cyanylated species of non-native hEGF III-B, A-C corresponding to the HPLC peaks 1-3 in Figure 5 .9, respectively. corresponding 5.108) of the fragments 1-2 elimination p: conclusively : therefore, car pair, Cy320-( fragments in and a B-elimi the assignme the MALDI- 31-53 can 1 fractionation HPLC/RSI e 2‘DtSlllfrde In or he disulfid. 55% and > HPLC isola 11 Shows fl the 0figinal 232 corresponding to the fragment 42-53 can be ignored. The MALDI spectrum (Figure 5.10B) of the cleavage products corresponding to HPLC peak 2 in Figure 5.9 shows fragments 1-32, 33, 41, 42-53, and two overlapped peptides corresponding to the B- elimination products 1-41 and 33-53, respectively. Disulfide bond linkage 33-42 can be conclusively assigned. In this spectrum, the fragment 42-53 presents as a huge peak and, therefore, cannot be ignored. From the disulfide structure of the first two pairs, the third pair, CySZO-Cys31, can be deduced, which is confirmed by the analysis by MALDI of the fragments in Figure 5.10C. In addition to the expected fragments, l-19, 20-30, 31-53, and a B-elimination product 1-30, two undesired fragments, 31-41 and 42—53, can confuse the assignment, as these fragments indicate that Cys42 is also a cleavage site. Frankly, the MALDI is limited in this case for an unambiguous assignment, although the fragment 31-53 can be used as evidence that Cys42 is a minor cleavage site. The HPLC fractionation of the cleavage products combined with the MALDI identification or HPLC/ESI experiments should be informative to clarify the uncertainty. 2-Disulfide Intermediates In our folding experiment, only one 2-S-S intermediate (II-A) was collected and the disulfide bond structure characterized. This most abundant intermediate counts for 55% and >70% of the total population at 30-min and 3-h folding, respectively. The HPLC isolated, cyanylated II-A was subjected to partial reduction/cyanylation. Figure 5- 11 shows the chromatogram of the partially reduced/cyanylated products. In addition to the original 2-disulfide species, two l-disulfide and one O-disulfide species arise from the as? Figure 5.1 233 ll-A R . 1 2 10.00 ' 20.00 ' 30:00 T 45.00 ' 50.00 I minutes ‘ Figure 5.11. HPLC separation of II-A and its partially reduced/cyanylated species. partial reducti disulfide oper disulfides ope It shoe more than on of the partial] pure intermee one species, complicated ' Althc products of ' easier to 100: in Il-A Will ; Preducts rel: Spectra. Th, Peak HA 3] Cleavage oc feSidUes. T whatsoever The 2 in Figure indicatingt‘ 234 partial reduction and cyanylation. That is, the HPLC peak 1 (Figure 5.11) has one disulfide opened, peaks 2 and 3 have two disulfides opened, and peak 4 has all three disulfides opened, resulting in a familiar partial reduction pattern. It should be pointed out that the HPLC fraction II-A in Figure 5.4 could contain more than one unresolved species, as suggested by Chang (29). From the HPLC pattern of the partially reduced protein isomers, it is clear that the original speak II-A represents a pure intermediate, other species, if any, are very minor. If the fraction contains more than one species, the partially reduced products of those species Should be much more complicated than the present pattern. Obviously it is not the case. Although the disulfide structure of the II-A can be directly mapped from cleavage products of the species represented by HPLC peaks 2 and 3 in Figure 5.11, it is much easier to locate first the cyanylated cysteine residues. The localization of the SH groups in II-A will greatly help the subsequent identification of the MALDI data of the cleavage products related to the reduced disulfide bonds, as the latter usually give more complex spectra. The MALDI spectrum (Figure 5.12A) of the cleavage products corresponding to peak II-A shows fragments 6-19, 20—53, and B-elimination product 6-53, indicating that cleavage occurred at cysteine residues 6 and 20, which represent two unoxidized SH residues. The peak 1-53 represents an intact II-A that did not undergo any cleavage whatsoever under our eXperimental conditions. The MALDI Spectrum (Figure 5.12B) of the cleavage products from HPLC peak 2 in Figure 5.11 shows fragments 6-19, 33-41, 42-53 and a B-elimination product 33-53, indicating that another disulfide pair, 33—42, must have been reduced, cyanylated and 235 A 6-19 1-53* y2-53 6-53* ++++ 20-53 3341 42-53 B 0.9 '7 ' 33-53* ++ so ML. ..._ idem/n .llLMA. .. .3” 31-53 6-13 C 42-53 1000 2000 3000 4000 5000 6000 m/z Figure 5.12. The MALDI mass spectra of peptide mixtures from the cleavage of (A) the cyanylated hEGF II-A, (B) and (C) the partially reduced/cyanylated species of II-A, corresponding to the HPLC peaks 1 and 2 in Figure 5.11, respectively. 236 cleaved. Therefore, Cys33 is connected to Cys42. Likewise, the MALDI spectrum (Figure 5.12C) of the cleavage products from HPLC peak 3 in Figure 5.11 shows fragments 6-13, 20-30, 14-30, and 31-53, indicating cleavage at cysteine residues 6, 14, 20, and 31, respectively. There is a minor degree of undesired cleavage at Cys42, resulting in the formation of fragments 31-41 and 42-5 3. In summary, intermediate II-A, the most populated species during the early stage of folding, contains a native structure, Cysl4-Cys31 and Cys33-Cys42. l-Disulfide Intermediates The recognition of disulfide structures in 1-disulfide intermediates ' is straightforward. The strategy is based on the mass mapping of cleavage products of the cyanylated 1-disulfide species. Instead of looking at the MALDI peaks corresponding to the cleavage Sites, the absence of such peaks represents the “uncleaved sites”, that is, those still linked by a disulfide bond. For a protein containing 6 cysteine residues, there are fifteen possibly l-disulfide intermediates. The possible fragments corresponding to cleavage at cyanylated cysteine residues of the 15 isomers are listed in Table 5.2. Since some of the fragments may not be detected by MALDI, the mass mapping of the products from the completely reduced/cyanylated hEGF was used as a “figureprint” spectrum of the cleavage products. It is immediately clear that the fragment 1-5 cannot be detected by MALDI, while fragment 42-53 is very sensitive to MALDI detection. This profile helps the interpretation of MALDI data from other intermediates, especially if ambiguity arises. The MALDI spectra of the cleavage products of the four well populated 1- disulfide intermediates are shown in Figure 5.13. Figure 5.13A is the MALDI spectrum 237 Table 5.2. Fragments of cleavage products corresponding to 15 cyanylated l-disulfide intermediates of hEGF Disulfide Disulfide Z Linkage Fragment m/ Linkage Fragment m/z 149 2047.1 1-5 551.5 2030 1344.6 6-19 1539.6 6-14 31-32 262.3 1442 2030 1344.6 3341 1020.1 3132 262.3 4253 1721.0 3363 2697.2 143 1423.4 1.5 551.5 1430 1968.3 6-13 915.9 6-20 31-32 262.3 2031 14-32 2184.5 3341 1020.1 3341 1020.1 4253 1721.0 4253 1721.0 1-13 1423.4 1.5 551.5 1449 667.7 6-13 915.9 6-31 2032 1560.8 2033 14-30 - 1344.6 3341 1020.1 3141 1236.3 42-53 1721.0 4253 1721.0 143 1423.4 14 551.5 14.19 667.7 6-13 915.9 6-33 2030 1344.6 2042 1430 1968.3 3141 1236.3 31-32 262.3 4252 1721.0 3363 2697.3 143 1423.4 16 551.5 1419 667.7 613 915.9 642 2030 1344.6 3133 14-19 667.7 3132 262.3 2041 2536.9 33-53 2697.2 4253 1721.0 1.5 551.5 1.5 551.5 630 2840.2 6-13 915.9 1420 31-32 262.3 3142 14-19 667.7 3341 1020.1 2032 1560.8 4253 1721.0 3353 2697.2 1-5 551.5 16 551.5 6-19 1539.6 613 915.9 1431 2032 1560.8 3342 14-19 667.7 3341 1020.1 2030 1344.6 4253 1721.0 3163 2913.3 16 551.5 619 1539.6 1433 2030 1344.6 3141 1236.3 4253 1721.0 238 42-53 3341 LA ++ 1'13 14-30 \ I l 3353* 3353 LR _‘ 4253 ms ER ++ 14-30 31-53 42-53 LC 3341 is a. «a a 14 - “2.3:..JL- - _+__A- * 3341 42-53 I-D ++ ‘I ll 20-30 “'19 33-53* 1000 2000 3000 4000 5000 6000 m/z Figure 5.13. Disulfide mapping of l-disulfide folding intermediates of hEGF. 239 of cleavage products corresponding to HPLC I-A in-Figure 5.4. Fragments 1-13, 14-30, 33-41 and 42-53 were detected, implying chemical cleavage at a cyanylated cysteine at position 14, 31, 33, and 42, respectively. Obviously, no cleavage occurs at Cys6 and Cys20. Otherwise, fragments 1-5, 6-13, 14-19, and 20-30 would arise, giving m/z at 551.5, 915.9, 667.7, and 1344.6, reSpectively. Therefore, Cys6-Cys20 linkage is deduced. It is notable that this intermediate has a native disulfide structure, which was not detected in Chang’s experiments (29). The second l-disulfide intermediate (I-B) Showed MALDI peaks corresponding to fragments 6-13, 14-30, 33-53, and 42-53, respectively. Cleavage occurs at Cys6, 14, 31, 33, and 42, but not Cy820. Extensive experiments indicated that the cleavage at Cys42 was very minor in comparison with the cleavage at other sites. Since fragment 33-53 is very intense, a reasonable conclusion can still be drawn that Cys42 does not present as a free sulfhydryl. Therefore, Cys20 is connected to Cys42. The MALDI spectrum for I-C (Figure 5.13C) is Simple to interpret. Fragments 6-30, 33-41, and 42-53 show that Cys6, 31, 33, 42 are cleavage sites. Therefore, the disulfide pair 14-20 can be deduced. By the same strategy, the MALDI spectrum (Figure 5.13D) of I-D shows peaks for the fragments 1-19, 20-30, 33-42, and 42-53, suggesting the Cys6-Cysl4 linkage. The disulfide structures of the last two intermediates are in agreement with the structures proposed by Chang et al (29). By our strategy, we were able to recognize the disulfide structure in LA and LB, neither of which was assigned by Chang et al. Among the four well populated l-disulfide intermediates, only one has a native disulfide structure. 240 IV. Aspects for Further Improvement One of the most important features of our methodology is the simplicity. Only a few fragments were produced, each of them is relevant tolthe assignment of (disulfide structures. Some undesired fragments may be observed, but they rarely affect the assignment of disulfide bonds because our methodology only relies on the recognition of specific cleavage at cysteine residues. However, if the side products correspond to the cleavage at undesired cysteine site(s), the data interpretation would no longer be straightforward. This Situation is more complicated by the fact that the MALDI response is not proportional to concentration or amount of analytes. Even a trace amount of an impurity may give a high MALDI response, as seen by the fragment 42-53 in the case of hEGF. Fortunately, HPLC can provide quantitative information on the relative concentration of analytes, if ambiguity arises. The most straightforward way is to perform HPLC/ESI analysis of cleavage products so that information on both quantities and masses of the analytes can be obtained. Our CDAP trapping provides advantages over both iodoacetate trapping and acidic quenching. Effective trapping can be achieved in most cases as demonstrated by the MALDI analysis of the trapped intermediates. However, the CDAP concentration and the pH must be controlled because of the possible side reactions (see chapter 3) and the instability of CDAP at higher pH (hydrolysis). As the available SH groups vary in the folding course and the concentrations of various intermediates differ greatly, the optimal CDAP concentration is difficult to control. It should be a good investment to study more stable and specific novel reagents that can cyanylate SH groups in acidic solution. ThiocyanOpyridine (TCP) (32) (see chapter 3) may act as such a reagent. Another 241 . ”am: we 824608.485 mango.“ Banana :95 :33 mo $525.6. 03:35 .36 8,5me 242 reagent, 4-thiocyanatoaniline (TCNA), was introduced in Biemann’s group several years ago (33). This reagent was claimed to be a usefiil alternative to NTCB. But the capability of this reagent to modify SH groups in acidic solution remains to be examined. V. Conclusions To summarize, seven species of well populated intermediates in the refolding of hEGF have been isolated and characterized, which included four l-disulfide, One 2- disulfide, and two 3—disulfide scrambled species (Figure 5.14), of which only two have native disulfide structure. Among the seven intermediates, I-A and LB are newly identified; the disulfide structures of the other five intermediates are the same as these published previously (29). These results demonstrate the feasibility of our methodology. The significance of these intermediates to the folding pathway of hEGF needs to be further specified. Our trapping technique, based on the cyanylation of free thiol groups under acidic conditions, showed similar results to those of acid trapping in terms of the distribution of intermediates during the folding process. This technique circumvents problems associated with the traditional iodoacetate trapping and greatly facilitates the subsequent characterization of disulfide structures of isolated folding intermediates. Our strategy for the determination of disulfide structures by chemical cleavage and mass mapping of the fragments is much simpler, faster, and conclusive, although the undesired cleavage at Cys42 indeed occurred in some cases. This methodology opened a new door to the application of our chemistry. The preliminary experimental results presented in chapters 4 and 5 also show the potential to T————_————”“_* -- 243 characterize disulfide structures of folding intermediates of even more complicated proteins, such as those containing adjacent cysteine residues (e. g., IGFs). 244 VI. References 1. Ellis, R. J., Curr. Opin. Struc. Biol., 4, 117-122(1994). 2. Frydman, R. B., Nirnmesgern, E., Ohtsuka, K., and Hartl, F. U., Nature, Land. 370, 111-117(1994). 3. Kim, P. S., and Baldwin, R. L., Annu. Rev. Biochem., 59, 631(1990). 4. Matthews, C. R., Annu. Rev. Biochem., 62, 653(1993). 5. Ptitsyn, O. B., Curr. Opin. Struc. Biol.,.5, 74( 1995). 6. Li, Y.-J., Rothwarf, D. M., and Scheraga, H. A., Nat. Struct. Biol., 2, 489(1995). 7. Jennings, P. A., and Wright, P. E., Science, 262, 892-896( 1993). 8. Balbach, J ., Forge, V., van Nuland, N. A. J., Winder, S. L., Hore, P. J ., and Dobson, C. M., Nat. Struct. Biol., 2, 865-870(1995). 9. Xu, X., Rothwarf, D. M., and Scheraga, H. A., Biochemistry, 35, 6406-6417(1996). 10. Frech, C., and Schmid, F. X., J. Mol. Biol., 251, 135-149(1995). 11. Creighton, T. E., and Ewbank, J. J., Biochemistry, 33, 1534-1538(1994). 12. Chang, J .-Y., Biochemistry, 35, 11702-11709(1996). 13. Glocker, M. O., Arbogast, B., Milley, R., Cowgill, C., and Deinzer, M. L., Proc. Natl. Acad. Sci., USA, 91, 5868-5872(1994). 14. Hober, S., Forsberg, G., Palm, G., Hartmanis, M., and Nilsson, B., Biochemistry, 31, 1749-1756(1992). 15. Creighton, T. E., and Goldenberg, D. P., J. Mol. Biol., 179, 497( 1984). 16. Darby, N. J., Norin, P. E., Talbo, G., and Creighton, T. E., J. Mol. Biol., 249, 463(1995) 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. Chang, J .Y., Biochemistry, 35, 11702-11709(1996). 245 Weissman, J. S., and Kim, P. S., Science, 253, 1386(1991). Weissman, J. S., and Kim, P. S., Proc. Natl. Acad. Sci. USA, 89, 9900(1992). Rothwarf, D. M., and Scheraga, H. A., Biochemistry, 32, 2671(1993). Rothwarf, D. M., and Scheraga, H. A., Biochemistry, 32, 2680(1993). Rothwarf, D. M., and Scheraga, H. A., Biochemistry, 32, 2690(1993). Rothwarf, D. M., and Scheraga, H. A., Biochemistry, 32, 2698( 1993). Creighton, T. 13., Science, 256, 111-112(1992). Weissman, J. S., and Kim, P. S., Cell, 71, 841-851(1992). Rothwarf, D. M., and Scheraga, H. A., J. Am. Chem. Soc., 113, 6293(1991). Goto, Y., and Hamaguchi, J. Mol. Biol., 146, 321(1981). Happersberger, P., Cowgill, C., Deinzer, M. L., and Glocker, M. 0., Proceedings of the 44th ASMS Conference on Mass Spectrometry and Allied Topics, Portland, Oregon, May, 1996, p1050. Chang, J.-Y., Schindler, P., Ramseier, U., and Lai, P.—H., J. Biol. Chem, 270(16), 9207-9216(1995). Torella, C., Ruoppolo, M., Marino, G., and Pucci, F., FEBS Letters, 352, 301- 306(1994). Chang, J .-Y., J. Biol. Chem, 268, 4043-4049(1993). Brocklehurst, K., Malthouse, J. P. G., Baines, B. S., Blenkinsop, 1R. D., Churcher, J. A., Mushiri, M. S., and Ormerod, F., Biochem. Soc. Trans, Vol. 6, 261-263(1978). Zaia, J ., Papayannopoulos, I. A., and Biemann, K., Proceedings of the 39th ASMS Conference on Mass Spectrometry and Allied Topics, Nashville, TN, May 1991. '0 HICHIGRN STAT ruiyijlryiyhijl“ ht...