.nnJ-n-nr 1.‘ nu... .0, .7 “nu"...«yw .v ~' '1. "H.015 llllllllilHHIIHIIHIHIIHIHJMHlllllllHHlllllllllllll 31293 01701 9187 This is to certify that the dissertation entitled Capture anal 1 dreutf camp.4 of tie Intermeol ates 0f Refdolmq anal ReduCtfl/e unfoldaaj of Proteins 63 Cyomytati 0M Metkodalpjfi and MMS3 SPECtYOMefY‘fi presented by Yliaj Ya‘tj has been accepted towards fulfillment of the requirements for Doctor of F" (”filigree 1n ($91169ng Date Tune I6 ; 1198 LIBRARY Michigan State Unlveralty PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 1 JAN 13 2033 CAPTURE AND IDENTIFICATION OF THE INTERNIEDIATES OF REFOLDING AND REDUCTIVE UNFOLDING OF PROTEINS BY CYANYLATION METHODOLOGY AND MAss SPECTROMETRY By Ying Yang A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1998 ABSTRACT CAPTURE AND IDENTIFICATION OF THE INTERMEDIATES OF REFOLDING AND REDUCTIVE UNFOLDING OF PROTEINS BY CYANYLATION METHODOLOGY AND MASS SPECTROMETRY By Ying Yang Cyanylation chemistry of cysteine residues combined with mass spectrometry provides promising methodology for providing structural information of cysteine residues in the proteins. The sulfliydryl groups of proteins can be specifically cyanylated by a cyanylation reagent, the N-terminal side of cyanylated cysteine residues can be cleaved in alkaline condition, and the masses of peptide fragments (one N-terminal side peptide and a series of iminothiazolidinc-blocked peptides) can be analyzed by mass spectrometry. Matching the masses of the cleaved peptides can be used to recognize the structural information of cysteine residues in the proteins. Two types of mass spectrometry have been used to analyze proteins. One is matrix- assisted laser desorption ionization time of flight mass spectrometry (MALDI-TOF-MS); the another is electrospray ionization mass spectrometry (ESI-MS). We found that MALDl-MS showed frequent, and sometimes complete, suppression of a signal from iminothiazolidine-blocked peptides; because such a phenomenon can compromise our analytical strategy, we have investigated the use of ESI for these analyses. The results indicate that the responses of our blocked peptides to E81 and MALDI are frequently complementary. Conventional methodology for determining the pairing in the disulfide bonding structure of proteins relies on the use of appropriate proteases to cleave the peptide backbone in between the cysteine residues. The possibility of a proteolytic cleavage site becomes less likely when proteins contain closed or adjacent cysteins. We have investigated the applicability of a novel methodology involving partial reduction of disulfide bonds, cyanylation of sulfliydryl groups, chemical cleavage of partially reduced/cyanylated protein, and mass mapping of cleavage products to the structural analysis of proteins containing adjacent cysteines in two model protein, long R3 insulin- like growth factor-I (LR3IGF-I) and insulin-like growth factor-I (IGF-I). These proteins contain three disulfide bonds, two of which involve adjacent cysteines. The disulfide structures of the proteins were unambiguously determined by the methodology. A new methodology based on cyanylation chemistry of cysteine residue and mass spectrometry has been used to capture and identify folding intermediates of long insulin- like growth factor I (LR3IGF-I) and insulin-like growth factor I (IGF-I). The refolding was quenched at different time points by adjusting the pH to 2.0-3.0 with a 1N HCl solution of 1-cyano-4—dirnethylamino—pyridinium (CDAP) which trapped intermediates via cyanylation of free sulfhydryl groups. The disulfide structm'e of the intermediates was determined by partial reduction/cyanylation/chemical cleavage/mass mapping. The folding intermediates of LR3IGF-I and IGF-I were successfully trapped and characterized. The time-dependant distribution and disulfide structure of the folding intermediates of LR3IGF-I and IGF-I were compared based on their structural features. ACKNOWLEDGMENTS I wish to thank my mentor, Dr. J. Throck Watson, for his assistance and advice. I especially would like to thank my daughter, J. Jinjing Yang, my husband, J ie Yang, for their support and understanding during this project. iv TABLE OF CONTENTS LIST OF TABLES ................................................................................. vii LIST OF FIGURES ................................................................................ viii CHAPTER 1. INTRODUCTION .................................................................................... l I. Introduction .................................................................................... 1 II. Oxidative Folding and Trapping of Intermediates ........................................ 3 A. Disulfide Bonds 1n Protein Folding .................................................... 5 B. Trapping of the Folding Intermediates ................................................ 12 III. Recognizing the Location of Cysteines and Cystines ................................... 18 A. Classical Approach ...................................................................... 18 B. Affinity Chromatography ............................................................... 22 C. Cleavage at Cysteine residue .......................................................... 23 IV. Assignment of Disulfide Bonds ............................................................ 24 A. Conventional Fragmentation Strategy ................................................ 25 B. Fragmentation of Proteins .............................................................. 25 C. Purification of Cystinyl Peptides ...................................................... 27 D. Identification of Cysteine-containing Peptides ...................................... 27 B. 8-8 Assignment of Adjacent Cysteines .............................................. 30 V. Matrix-Assisted Laser Desorption/Ionization Mass Spectrometry (MALDI-MS) and Electrospray Mass Spectrometry (ESI-MS) ........................................ 32 A. Matrix-Assisted Laser Desorption/Ionization Mass Spectrometry (MALDI-MS) ........................................................................... 33 B. Electrospray Ionization Mass Spectrometry (ESI-MS) ........................... 39 C. The Comparison of MALDI and E81 ................................................ 45 VI. References .................................................................................... 47 CHAPTER 2 A COMPARISON OF MALDI AND ESI FOR ANALYSIS OF IMINO'I'HIAZOLIDINE-BLOCKED PEPTIDES IN DISULFIDE MAPPING ........................................................................................... 58 I. Introduction .................................................................................. 58 II. Analytical Methodology ..................................................................... 61 A. Cyanylation and Chemical Cleavage Reaction ...................................... 61 B. Localization of Cysteines and Cystines by ESI-MS ................................ 63 III. Experimental Section ........................................................................ 64 IV. Results and Discussion ..................................................................... 68 A. Recognizing the Location of Cysteine and Cystines of B-Lactoglobulin A. . ...68 B. Recognizing the Location of Cysteines and Cystines in Ovalbumin ............. 76 C. Evaluation of CDAP and NTCB ...................................................... 85 V. Conclusions ................................................................................... 88 VI. References .................................................................................... 90 Chapter 3 DISULFIDE MASS MAPPING IN PROTEINS CONTAINING ADJACENT CYSTEINES WITH CYANYLATION/CLEAVAGE METHODOLOGY ................................................................................. 93 I. Introduction ................................................................................... 93 II. Methodology for Assignment of Disulfide Linkages ................................... 95 III. Experimental Section ........................................................................ 97 IV. Results and Discussion .................................................................... 100 A. Disulfide Mapping of IGF-I .......................................................... 100 B. Disulfide Mapping of LR3IGF-I ..................................................... 112 V. Conclusions ................................................................................. 121 VI. Future Work ................................................................................. 121 VII. References ................................................................................. 125 Chapter 4 TRAPPING AND IDENTIFICATION OF INTERMEDIATES DURING THE REFOLDING OF LR’IGF-I AND IGF-I ...................................................... 128 I. Introduction ....................... 128 H. The strategy for Trapping Folding Intermediates ...................................... 133 III. The experimental Section ................................................................. 135 IV. Results and Discussion .................................................................... 140 A. Refolding and Reductive Unfolding of LR3IGF-I ................................. 140 B. Refolding and Reductive Unfolding of lGF-I ...................................... 157 V. Isolation and Refolding of the Folding Intermediates of LR3IGF-I ................. 176 VI. Comparison of the Folding Process of LR3IGF-I and IGF-1 ......................... 178 VH. Conclusions ................................................................................ 178 VIII. References ................................................................................. 186 vi Table 1.1 Table 2.1 Table 2.2 Table 2.3 Table 3.1 Table 3.2 Table 4.1 Table 4.2 Table 4.3 Table 4.4 LIST OF TABLES Comparison of ESI-MS and MALDI-MS .......................................... 46 Expected and observed masses of cleavage product of B-lactoglobulin A ............................................................................................ 69 Expected and observed masses of cleavage product of ovalbumin ............. 77 Comparison of CDAP and NTCB ................................................... 89 Calculated and observed m/z values for possible fragments resulting fiom the cleavage reaction of IGF-I chains at sites of designated cysteine pairs ...... 109 Calculated and observed m/z values for possible fragments resulting from the cleavage reaction of LR3IGF-I chains at sites of designated cysteine pairs .......................................................................................... 115 The expected and observed masses of cleavage products of one-disulfide intermediate of LR3IGF-I ............................................................ 148 The expected and observed masses of cleavage products of the purified/cyanylated two-disulfide intermediate of LR’IGF-I .................. 151 The expected and observed masses of cleavage products of singly reduced/cyanylated isomers of the three-disulfide intermediate of LR’IGF-I .......................................................................................... 156 The expected and observed masses of the fragments related to the disulfide assignment of the intermediates of IGF-I ......................................... 165 vii Figure 1.1 Figure 1.2 Figure 1.3 Figure 2.1 Figure 2.2 Figure 2.3 Figure 2.4 Figure 2.5 Figure 2.6 Figure 2.7 Figure 2.8 Figure 2.9 LIST OF FIGURES Folding pathway of BPTI .......................................................... 13 MALDI-TOF-MS instrument ...................................................... 37 ESI-MS instrument and droplet production in ESI interface .................. 42 Mechanisms of cyanylation and cleavage reaction ............................. 62 Structure of B-lactoglobulin A .................................................... 68 MALDI spectra of cleavage products of B-lactoglobulin A following reduction/cyanylation/clcavage (panel A) or cyanylation/cleavage/reduction (panel B) ............................................................................... 71 HPLC chromatogram of cleavage products following procedure TCEP/CDAP ......................................................................... 72 E81 spectra of cleavage products from procedure TCEP/CDAP of B- lactoglobulin A ...................................................................... 73 HPLC chromatogram of cleavage products following procedure CDAP/TCEP ......................................................................... 74 E81 spectra of cleavage products from procedure CDAP/TCEP of B- lactoglobulin A ...................................................................... 75 Structure of ovalbumin ............................................................. 76 MALDI spectra of cleavage products of ovalbumin following reduction/cyanylation/cleavage (panel A) or cyanylation/cleavage/reduction (panel B) .............................................................................. 79 viii Figure 2.10 Figure 2.11 Figure 2.11 Figure 2.12 Figure 2.13 Figure 2.14 Figure 2.15 Figure 3.1 Figure 3.2 Figure 3.3 HPLC chromatogram of mixtures of cleavage product following procedure TCEP/NTCB of ovalbumin ........................................................ 80 E81 spectra of cleavage products following procedure TCEP/NT CB of ovalbumin ............................................................................ 81 Continue .............................................................................. 82 HPLC chromatogram of mixture of cleavage products following the procedure NTCB/TCEP of ovalbumin ............................................ 83 E81 spectra of cleavage products following the procedure NTCB/TCEP of ovalbumin ............................................................................ 84 HPLC chromatogram of mixture of cleavage product from procedure CDAP/TCEP of ovalbumin ........................................................ 86 E81 spectra of cleavage product from procedure CDAP/TCEP of ovalbumin .......................................................................................... 87 Amino acid sequence and disulfide structure of IGF-I ....................... 101 Overview of the methodology used to assign the disulfide linkages in IGF-I ......................................................................................... 104 HPLC separation of denatured IGF-I and its partially reduced/cyanylated isomers. Separation was carried out on a Vydac C18 column at a flow rate of 1.0ml/min 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 land 2 represent singly reduced/cyanylated species. Peak 3 represent doubly reduced/cyanylated species, as determined by MALDI-TOF analysis .............................................................................. 106 Figure 3.4 Figure 3.5 Figure 3.6 Figure 3.7 Figllre 3.8 The MALDI mass spectra of peptide mixtures resulting from the cleavage of the two singly reduced/cyanylated IGF-I isomers, corresponding to the HPLC peaks 2 and 3 in Figure 3.2, respectively. The symbols “itz” and * represent the iminothiazolidine derivatives and protonated B-elimination products, respectively ............................................................ 110 Amino acid sequence and disulfide structure of LR3IGF-I. The bold letters at N-terminal sites indicate the 13 amino acid extension. The bold R represents the replacement of E3 of IGF-I ..................................... 113 HPLC separation of denatured LR’IGF-I and its partially reduced/cyanylated isomers. Separation was carried out on a Vydac C18 column at a flow rate of 1 ml/min 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 ......................................................... 1 14 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 3.4, respectively. The symbols “itz” and * represent iminothiazolidine derivatives and B-elimination products, respectively. The peaks with question marks are discussed in the text. . ..1 l7 LC/ESI-MS results of peptide mixture from the cleavage of a singly reduced/cyanylated LR3IGF-I isomer, corresponding to HPLC peak 1 in Figure 3.6. Figure 3.8 A is the RTIC of the cleavage products. Figure 3.8 B is the ESI mass spectra of the peptides corresponding to the RTIC peaks in Figure 3.8 A. The symbol “itz” represents iminothiazolidine derivative...119 Figure 3.9 LC/ESI-MS results of peptide mixture from the cleavage of a singly reduced/cyanylated LR3IGF-I isomer, corresponding to HPLC peak 2 in Figure 3.6. Figure 3.9 A is RTIC of the cleavage products. Figure 3.9 B is the E81 mass spectra of the peptides corresponding to the RTIC peaks in Figure 3.9 A ........................................................................ 120 Figure 3.10 The overview of proposed non-chromatographic approach .................. 123 Figure 4.1 Amino acid sequences and disulfide structures of LR3IGF-I and IGF-1. The numbers in the parentheses indicate the location of cysteine residues of IGF-I. The 13 hold letters at the N-terminus indicate the 13-amino acid extension in LR3IGF-I. The bold R represents the replacement of E3 of IGF-I ................................................................................ 131 Figure 4.2 The overview of the processes of refolding and reductive unfolding of proteins .............................................. 134 Figure 4.3A HPLC separation of CDAP-trapped intermediates during the time course of refolding of LR3IGF-l ............................................................ 141 Figure 4.33 HPLC separation of CDAP-trapped folding products after folding reached to the equilibrium, increased the concentration of GSSG to 100 mM and incubated for another 30 minutes ................................................ 143 Figure 4.4 HPLC separation of CDAP-trapped intermediates in the time course of reductive unfolding of LR3IGF-I ................................................ 144 Figure 4.5 MALDI spectrum of peptide mixture resulting from the cleavage of the xi Figure 4.6 Figure 4.7 Figure 4.8 Figure 4.9 purified and cyanylated intermediate corresponding to peak 3 in Figure 3.4. The symbol “itz” represents iminothiazolidine derivatives .................. 147 HPLC separation of intact intermediate (peak 1 here) corresponding to HPLC peak 2 in Figure 4.3 and one of its expected partially reduced/cyanylated isomers (peak 2). The peak with a question mark did not correspond to any specific cleavage fragments to the disulfide structure of the intermediates as analyzed by MALDI ...................................... 150 MALDI spectra of peptide mixtures resulting from the cleavage of the purified/cyanylated intermediate corresponding to HPLC peak 2 in Figure 4.3 and peak 1 in Figure 4.6 (A) and one of its partially reduced/cyanylated isomers (peak 2 in Figure 4.6) (B). The symbols “itz” and * represent iminothiazolidine derivatives and B-elimination species, respectively. Itz60-(65)-83 represents that the residue in parenthesis (residue 65) is cyanylated but remains uncleaved. Itz60—(65*)-83 represents that the B- elirnination occurs at residue 65 ................................................. 152 HPLC separation of residual intact intermediate corresponding to HPLC peakl in Figure 4.3 and its partially reduced/cyanylated isomers .............................................................................. 154 MALDI spectra of peptide mixtures resulting from cleavage of partially reduced/cyanylated isomers corresponding to HPLC peaks 2 and 3 in Figure 4.8 of the three-disulfide intermediate (peak 1 in Figure 4.3). The symbol “itz” indicate iminothiazolidine derivatives. The question marked peaks are discussed in the text ................................................................ 155 xii Figure 4.10 Figure 4.11 The disulfide structure of folding intermediates of LR’IGP-I ............... 158 HPLC separation of CDAP-trapped intermediates during the time course of refolding of IGF-I ................................................................. 160 Figure 4.12A HPLC separation of CDAP-trapped intermediates in the time course of reductive unfolding of IGF-I ..................................................... 162 Figure 4.123 HPLC separation of CDAP-trapped intermediates in the unfolding after the Figure 4.13 Figure 4.14 Figure 4.15 Figure 4.16 equilibrium ........................................................................ 164 HPLC separation of residual intact three—disulfide intermediate corresponding to HPLC peak 1 in Figure 4.11 and its partially reduced/cyanylated isomers. Peaks 2 and 3 each represent two-disulfide species. The question marked peaks are discussed in the text ............... 167 MALDI spectra of peptide mixtures resulting from the cleavage of the two singly reduced/cyanylated isomers, corresponding to HPLC peaks 2 and 3 in Figure 4.13, respectively. The symbol “itz” represents iminothiazolidine derivatives ........................................................................... 168 HPLC chromatogram of partial reduction/cyanylation mixture showing residual intact intermediate (two-disulfide species) corresponding to HPLC peak 3 in Figure 4.11 and one of the expected partially reduced/cyanylated species (peak 2). The question marked peak is discussed in the text ...... 169 MALDI spectra of peptide mixtures resulting from the cleavage of the purified and cyanylated intermediate corresponding to peak 1 in Figure 4.14 (also to peak 3 in Figure 4.11) (A) and the partially reduced/cyanylated isomer corresponding to HPLC peak 2 in Figure 4.14 (B). The symbols “itz” xiii Figure 4.17 Figure 4.18 Figure 4.19 Figure 4.20 Figure 4.21 Figure 4.22 and * represent iminothiazolidine derivatives and B-elimination peak, respectively. The peak with a question mark cannot be assigned to any specific cleavage fragment of the intermediate .............................. 170 MALDI spectra of peptide mixtures resulting from the cleavage of purified and cyanylated two-disulfide intermediates corresponding to peaks 4 and 5 in Figure 4.11, respectively. The symbol “itz” represents iminothiazolidine derivatives ......................................................................... 172 The MALDI spectrum of cleavage products of the purified/cyanylated one- disulfide intermediate corresponding to HPLC peak 6 in Figure 4.11. The symbol “itz” represents iminothiazolidine derivatives ....................... 174 The disulfide structure of refolding intermediates of IGF-I ................. 175 HPLC separation of CDAP-trapped intermediates after 30-minutes refolding starting from the purified 2-disulfide intermediate (11) or 3- disulfide intermediate (III) of LR3IGF-I in the same refolding buffer as used in refolding of reduced/unfolded LR3IGF-I .................................... 177 The predicted folding pathway of LR3IGF-I ................... A ............... 180 The predicted folding pathway of IGF-I ....................................... 183 xiv CHAPTER 1 ' INTRODUCTION 1. Introduction Recent advances in mass spectrometry have created new technological capabilities that are applicable to the study of peptides and proteins involved in biological processes. The discovery and development of ionization techniques for large bio-oligomers at significantly higher ionizing efficiencies permits the detection and measurement of very large intact or fiagmented biopolymeric substances (1). These tools are electrospray ionization (ESI) (2) and matrix-assisted laser desorption/ionization (MALDI) (3). They have become among the most powerful methods yet available for the macromolecular characterization of living systems ranging from the measurement of the molecular weights and purity of large polar biopolymeric substances to delineation of the detailed sequence and structure of components of the complex mixtures obtained fi'om their selective enzymatic or chemical degradation. These ionization techniques have been revolutionary in spurring research in protein biochemistry, glycobiology, and biotechnology, and most recently in DNA sequence analysis. Over the next few years these technologies are destined to become as commonplace in the conduct of biomedical r “catch as HPLC did in the early 19808. The synthesis of proteins in organisms starts by transcription of the DNA sequences to. messenger RNA (mRNA). In eukaryotic organisms, mRNA is processed to remove inn-0m and is then translated on the ribosomal complex to synthesize the protein. To be b . lologically active, all proteins must adopt specific folded three-dimensional structures after they depart from the ribosome (4, 5). The genetic information for the protein specifies only the primary structure, the linear sequence of amino acids in the polypeptide backbone. Many purified proteins can spontaneously refold in vitro after being completely unfolded, so the three-dimensional structure must be determined by the primary structure (6). How this occurs has come to be known as ‘the protein folding problem’. It had been primarily of academic interest, but with the advent of the recombinant DNA revolution and its potential for producing therapeutic proteins, often in an insoluble, unfolded, inactive, and useless form, has made it also of great practical importance. For many years, scientists have endeavored to understand the process by which protein folding occurs, that is, the folding pathway. In order to understand the folding pathway of a protein, it is necessary to trap and characterize the folding intermediates (7-9). A particularly difficult aspect in the study of the protein folding is the fact that intermediates may be short-lived and therefore hard to isolate and analyze structurally and fimctionally. Disulfide-containing proteins provide an opportunity to isolate and characterize trapped intermediates by quenching the disulfide pairing chemically during the time course of folding (4,10-12). Furthermore, the study of folding process of disulfide-containing protein is also important because disulfide formation often confounds the high-level expression and renaturation of recombinant proteins (5). The sufliydryl groups of cysteine residues are among the most reactive side chains in proteins. In 1973, Jacobson et al. (13) introduced specific cyanylation of sulfhydryl groups of cysteine residues followed by chemical cleavage at cyanylated cysteine residues. This specific reaction with the combination of MALDI-MS has been recently used to characterize the cysteine status of proteins (14, 15). The methodology has the advantages of being simple, fast, and sensitive. In this dissertation, several aspects related to the analysis of cysteine status in a protein, as well as in folding intermediates, based on the combination of cyanylation methodology and mass spectrometry will be discussed. In chapter 2, two ionization technologies, MALDI and ESL will be compared for their capabilities of analyses of iminothiazolidine-blocked peptides resulting from the chemical cleavage reaction at cyanylated cysteine residues. In chapter 3, assignment of disulfide linkages of proteins containing adjacent cysteines will be discussed. In chapter 4, the trapping method and disulfide structures of intermediates in folding and reductive unfolding processes including the folding pathway of Long R3 insulin—like growth factor-I will be investigated. The purpose of this chapter is to introduce basic aspects of MALDI and ESI-MS, and describe current methodologies for trapping folding intermediates, locating cysteines and cystines in proteins, as well as recognizing disulfide linkages. II. Oxidative Folding and Trapping of Intermediates The general rules governing protein folding will be established only by characterizing for many proteins the conformations and energetics of structures at particular points along their folding pathways. Description of the properties of a native state is relatively simple because it is a single, long-lived conformation. Intermediates and unfolded states are much more difficult to characterize because they may be short- lived or exist as a distribution of possible conformers. Two useful methods are currently available to tackle this problem. A newly developed technique of pulsed-label NMR (16- 19) permits trapping and identification of amide groups that are engaged in the structured elements. This technique can, in principle, be applied to all types of proteins. However, intermediates trapped by this method are not amenable to chromatographic purification. Therefore, the intermediates cannot be further characterized structurally and functionally. Disulfide bonds introduce nonlocal topological restraints in the folding of proteins, and their formation provides valuable signals in tracing the folding pathway of cysteine- containing proteins (20, 21) based on two important phenomena. First, it is known that the formation of disulfides is coordinated with the folding of protein (22). That is, during folding, disulfidcs reshuffle (break and reform) readily in order to adapt to more favorable conformations. Second, disulfide-folding intermediates, trapped either reversibly or irreversibly, can be separated and isolated by chromatographic systems. This allows examination of the number, diversity, structure, and properties of intermediate species. Data obtained from these analyses are indispensable for the construction of protein folding pathways (23). In this method, the disulfide-bonded intermediates which form during the folding process are chemically trapped in a time course manner. The trapped intermediates are isolated and structurally characterized, the kinetics of the interconversion is determined, and the different components are coordinated with a folding pathway. Therefore, the main concerns in a study of disulfide- containing protein folding are trapping of folding intermediates by quenching the disulfide formation as well as the determination of disulfide structure of intermediates. A. Disulfide Bonds in Protein Folding Disulfide bonds are comparatively labile and cleave readily to thiols by reduction, and thiols are readily oxidized to disulfides. Cystine can be readily formed by oxidation of cysteine which is reversibly obtained by reduction (Equation 1.1) CO H oxidation 2 PIS/Y 2 ‘ 5 HOZCY\S—S/Ycoz N reduction ”2 NH2 NH2 cysteine cystine Equation 1.1. Cysteine and cystine l. Disulfide Formation Disulfides do not form spontaneously between thiols, even when in close proximity, unless there is an appropriate oxidant. While molecular oxygen (in the presence of trace amounts of heavy metals) may adequately serve this role for some proteins, the most common source of oxidizing equivalents is a low molecular weight disulfide. Protein thiol oxidation to the disulfide results from a thiol/disulfide exchange process (24) that transfers oxidizing equivalents from the low molecular weight disulfide to the protein (Equation 1.2). — — r 'i‘lg Ililg ‘flg S' 4' f—S ——-—> "IS- Re I «— is Rn Re Rn Re L _ _a,°? ? Equation 1.2. The mechanism of thiol/disulfide exchange M Thiol/disulfide exchange occurs via direct attack of a nucleophilic thiolate anion on one of the sulfurs of the disulfide bond (25, 26). The transition state is rather symmetrical with approximately equal bond formation to the nucleophilic sulfur (KS) and the leaving sulfur (RigS') and with a relatively small amount of negative charge on the central sulfur (RS) (27). The rate constant for the reaction increases as the basicity of the attacking thiolate (R..S') nucleophile increases (pKa increases) and as the basicity of the leaving thiolate (Rth') decreases (pKa decreases). The pKa of the typical cysteine sulfllydryl group is about 8.6; however, this may vary considerably from protein to protein due to the effects of the local environment. At pH of 8.6, the typical intermolecular step of thiol/disulfide exchange between small molecules or unhindered protein thiols and disulfides will occur with a second-order rate constant of about 20 M'ls'l (11). Since thiol/disulfide exchange occurs via the attack of a nucleophilic thiolate anion, the rate will increase with increasing pH until the attacking thiol is predominantly in the thiolate form. Upon adding a disulfide reagent to a reduced protein, two sequential thiol-disulfide exchange reactions are required to form one protein disulfide bond. The first is the simple chemical reaction between the disulfide reagent and one of the cysteine thiol groups, to generate the mixed disulfide (Equation 1.3). '3” -SSR K1 + % RSSR I + RSH -SH K-1 -311 Equation 1.3. The formation of mixed disulfide bond The formation reaction of a mixed disulfide is bimolecular in either direction, so the rates should be dependent upon the concentrations of the disulfide and thiol forms of the reagent. The second step is the one in which a second cysteine thiol group reacts with the mixed disulfide to form the protein disulfide bond (Equation 1.4). '33R Kintra ‘ \S + RSH -SH Equation 1.4. The formation of disulfide bond This step is intramolecular thiol/disulfide exchange in the forward direction, and its rate depends primarily upon the tendencies of the thiols of different cysteine residues to come into proximity of the mixed disulfide, which is primarily controlled by the protein conformation. Therefore, this step provides the most useful information about protein conformation and folding, and King, can be calculated from the observed rates of disulfide formation and reduction with different reagents (11). Once a disulfide bond has formed in a protein with more than two cysteine residues, this disulfide bond can undergo intramolecular rearrangements via processes analogous to the intramolecular step in disulfide formation. Again, the rate of a rearrangement reaction depends on the extent to which the conformation of the polypeptide favors an attack of the thiolate on the existing disulfide bond (Equation 1.5). Equation 1.5. Intramolecular rearrangement of a disulfide bond The observed rate of disulfide formation depends on several factors, including the degree to which the protein thiols are ionized, the rate of reaction of thiolate with reagent, the stability of the mixed-disulfide intermediate, and the rate of formation of the protein disulfide from the mixed disulfide (28). A redox buffer of a mixture of a low molecular weight disulfide and its corresponding thiol usually is used for the formation of disulfide bonds. During oxidative folding, the disulfide component of the redox buffer serves to provide oxidizing equivalents for protein disulfide formation. The thiol component of the redox buffer serves to reduce native or non-native disulfide bonds that may otherwise lock the protein in misoxidized forms or to catalyze thiol/disulfide rearrangements. Since the reaction of disulfide formation is reversible, the capacity to form a disulfide bond will depend on the equilibrium and the rate constants for the individual steps and the concentration and oxidation potential of the redox buffer (24). There are two kinds of redox buffer that can be used for disulfide formation, linear disulfide and cyclic disulfide. Two of them are most commonly used, glutathione and dithiothreitol. Glutathione is a tripeptide linear disulfide reagent. It exists in two forms: reduced (GSH) and oxidized (disulfide, GSSG). Glutathione universally occurs in the tissues of animals, plants, and microorganisms at comparatively high concentration (from 0.4 to 12 mM, predominantly in the form GSH) (29). GSI-I/GSSG system attributes an important role in the formation of disulfide bonds in proteins during their synthesis in vivo (30). Therefore, it is a most pertinent (in viva) electron donor and acceptor to be used for the formation of disulfide bonds in vitro. The formation of disulfide with GSSG has the advantage of being rapid at alkaline pH, simple, extremely specific, and easy to control (31). 4 Oxidized dithiothreitol (DTT), a cyclic disulfide, has been employed in studies of the mechanism of protein folding. Disulfide formation using DTT is much more discriminating than with a linear reagent. The rrrixed disulfide is energetically unfavorable and rapidly dissociates intramolecularly, so it should not accumulate significantly (32). Only very favorable protein disulfides are formed with this reagent, at a rate that should be proportional to the rate of the intramolecular step of the reaction (33). 2. Oxidative-Folding Problem In the late 19603 and early 19703, a number of investigators began to determine the mechanisms by which unfolded proteins regain their native conformations in vitro. These studies have largely been motivated by the recognition that the number of possible conformations accessible to a polypeptide chain is so great that finding the native conformation by a completely random search would be improbably slow. This suggested that there might be specific pathways of protein folding and that characterizing partially folded intermediates associated with these pathways might provide a key to understanding the relationship between the amino acid sequence and the three- dimensional structure of a protein. In 1962, Haber and Anfinsen (34) first used the oxidative folding of ribonuclease and other small, disulfide-containing proteins to show that the primary sequence contained enough information to specify the correct three-dimensional structure of a protein and that the driving force for protein folding was the thermodynamic stability of the native structure. Two cysteines must be within covalent bonding distance to form a disulfide bond; therefore, identifying the intermediates that are actually populated during the folding reaction gives clues about the structural features and interactions that guide the unfolded protein to its native structure (35). At least for the disulfide-containing proteins that spontaneously refold, the overall driving force is the thermodynamic stability of the native structure-native disulfide pairings give the most stable protein (7). The interactions that specify which cysteines will be paired and in what temporal order could be similar to those in the fully folded protein (native interactions), or they may represent interactions that are not found in the native protein (non-native interactions). Compared to the necessary rate of protein synthesis and secretion in viva, uncatalysed in vitra oxidative folding is often a very slow and inefficient , process. Consequently, in viva, oxidative folding is catalysed (36, 37) and possibly coupled to protein translation/translocation (38). The rules that govern the oxidative folding of a protein in vitra may or may not be the same as those that guide in viva folding. The folding of even simple disulfide-containing proteins illustrates the potential complexity of oxidative folding. The mechanisms that direct the formation of a three- dirnensional native structure from an unfolded protein must now accommodate the 10 chemistry of disulfide formation. For a protein with six cysteine residues that form three disulfide bonds, there are 15 ways of pairing the cysteines to generate molecules with three disulfide bonds. The complexity rapidly increases when more disulfides are involved. After more than 30 years since the first study of the oxidative folding of ribonuclease, the knowledge of the mechanism of protein folding is still incomplete. There are several central questions concerning the oxidative folding of proteins. (a) Do disulfide bond formation determine what structures will form or do other interactions in the protein dictate which disulfide bonds will form? (b) Are all of the specific interactions that guide the folding process present in the native structure or do non-native interactions play a role? (c) Does oxidative folding proceed through one pathway or do multiple, alternative pathways exist? ((1) Why do some disulfide-containing proteins fold with high efficiency into their native structures while others do not? (e) Do catalysts of oxidative folding alter the mechanism of folding or simply increase the rate and extent of native structure formation? (1) Is the mechanism of oxidative folding in vivo the same as that in vitra? 3. Folding Pathway of Bovine Pancreatic Trypsin Inhibitor (BPTI) The most detailed and informative folding pathway elucidated thus far is that of bovine pancreatic trypsin inhibitor (BPTI), a small single-domain 58-residue amino acid protein stabilized by three disulfide bonds. The folding pathway of BPTI (21) is 11 illustrated in Figure 1.1. The reduced BPTI is a random coil, and it can efficiently refold by sequentially passing through one-disulfide intermediates, two-disulfide intermediates, and finally to the native protein. The folding experiments of BPTI (4, 11, 39, 40) revealed two striking features of the BPTI folding mechanism. First, the process is distinctly non- random: only a small fraction of the possible disulfide-bonded intermediates are detected during folding or unfolding. Second, disulfide formation is not a simple sequential process: the kinetically preferred mechanism includes intramolecular rearrangements involving intermediates with non-native disulfide bonds (i.e, those not present in the native protein). This finding raised the possibility that the amino acid sequence might specify the structures of folding intermediates which have different conformations from that of the final folded protein. This possibility would add new complications to the already formidable problem of predicting the three-dimensional structure of a protein. B. Trapping of the Folding Intermediates To identify disulfide-bonded folding intermediates, rapid quenching of the folding reaction is necessary to trap the intermediates in chemically stable forms that can be physically separated and analyzed. The criteria for a good trapping agent include that it block the thiolates of an intermediate rapidly, cOmpletely, and without modifying the protein at sites other than thiols. Several methods have been used to trap the intermediates. 12 Others H H . H 1-1 (30-51. 538) <-<> <=> - N C (54101ng (3e51, 5-55) (30-51. 5-55. 14-38) / (30-518. 514) N<§> SH Reduced C SH (30.51) (30-51. 14-38) Figure 1.1. Folding pathway of BPTI 1. Acid Trapping The folding of proteins are conducted under alkaline condition, usually at around pH 8.5. The pKa of the typical cysteine sulthydryl group is about 8.6. Therefore, the thiolate of the intermediates can be rapidly protonated by adding acid in the folding solution to a pH less than 3 and the further thiol/disulfide exchange can be stopped (Equation 1.6). Equation 1.6. Trapping folding intermediates by acidification Quenching by acidification is rapid and occurs at the diffusion control rate (109 M'ls“) (41). Subsequent reversed-phase HPLC at pH 2 is well suited for the separation of acid-quenched intermediates without introducing significant rearrangements during the separation. A practical advantage of acid quenching is its reversibility. As a result, it is possible to purify an acid-quenched intermediate and subsequently to allow further rearrangement or folding to occur so that more detailed information of the folding pathway can be obtained (4, 42, 43). The folding intermediates of several proteins have been trapped by this method (4, 44-47). While trapping by acidification is rapid, it does not completely stop thiol/disulfide exchange; it just slows it down by reducing the concentration of the reactive thiolate anion. Assuming that the pKa of the thiol is 8.5, an intramolecular thiol/disulfide exchange occurring with a half-life of lms at pH 8.5 would 14 occur with a half-life of approximately 1600s at pH 2 (41). Therefore, samples must be analyzed promptly after the acidification and the intermediates trapped by acid must be chemically modified before the disulfide bond structure can be defined. In this respect, it is a cumbersome procedure. 2. Trapping by Alkylation Folding intermediates can also be trapped by addition of an alkylation reagent, such as iodoacetate, iodoacetomide, and 4-vinyl pyridine (48-51). Iodoacetate is the most commonly used. The thiol groups of the intermediate will be alkylated by the alkylating reagent and further thiol/disulfide exchange will be stopped (Equation 1.7). -S-CHz-COOf + ICHZCOO' ——> -s Equation 1.7. Trapping folding intermediates by alkylation At a concentration of 0.2 M and pH 8.7, iodoacetate reacts with sterically accessible sulfhydryl groups with a half-life of 0.3 s (23). This is much faster than many of the intermolecular thiol/disulfide exchange reactions during oxidative folding; at 1mM GSSG, the half-life for reaction of GSSG with an exposed SH at pH 8.7 should be about 30 8. However, rearrangement of intermediates during trapping with iodoacetate has been observed for both BPTI (4) and ribonuclease (52). Quenching by 0.2 M iodoacetate is comparable to or, in some cases, slower than the rate of some of the intramolecular 15 rearrangements that interconvert intermediates of the same oxidation state (4, 53). When quenching by alkylation is slow compared to rearrangement, and the alkylating reagent reacts preferentially with one of the intermediates, alkylation will lead to overestimation of the amount of the intermediate that reacts fastest with the alkylation reagent. This could be a more severe problem for partially structured intermediates where steric hindrance would retard the rate of alkylation of some thiols (54). Although a high concentration of iodoacetate can be applied to mininize such a side reaction, modification of other functional groups by a high concentration of iodoacetate may provoke other problems (55). 3. Other Trapping Methods Acid and alkylation trapping described above were attacked by Rothwarf and Scheraga (56, 57), who insisted that the criteria for a good blocking agent are that it block quickly, completely, and without modifying the protein at sites other than thiols. They proposed a new alkylating reagent, 2-aminoethyl methanetlriosulfonate [(NH2)C2H58S02CH3] (AEMTS) (Equation 1.8). AEMT S is a thiosulfonate reagent (58) which reacts with free thiols in the solution rapidly; even thiols that are essentially buried should still be accessible by local or global fluctuations of the protein within the 2 min blocking time (59). 16 + -S-S-CH2-CH2-NH3 f.’ . + CH3-fi-S-CH2-CH2-NH3 —> O Equation 1.8. Trapping folding intermediates by AEMTS There are several advantages of trapping by AEMTS (57). It can reversibly block the thiol groups in the intermediates so that it permits the refolding process to be restarted fi'om isolated intermediates, an invaluable tool in the determination of refolding pathways. In addition, blocking with a reagent such as AEMTS provides a basis for separation of intermediates in a predicable manner due to the addition of a positive charge to a thiol group by AEMTS. It is especially important for very complicated refolding processes containing large numbers of chemically distinct species. Recently, Thannhauser et a1. combined AEMTS trapping and amino acid analysis to quantitatively determine the number of cysteines and cystines in the folding intermediates of hirudin (60). While trapping by AEMTS is rapid and specific, disulfide scrambling is still questionable since the blocking of a thiol group by AEMTS is a thiol/disulfide exchange reaction. Recently, a new trapping reagent was suggested to study the folding of macrophage colony stimulating factor (M-CSF) (61). The protein-folding intermediates were trapped by selective modification of bis-cysteine suflrydryls with phenylarsonous acid derivatives (PAA) (Equation 1.9). PAA-derivatives form cyclic dithioesters in the presence of two Snitably spaced sulflrydryl groups of the intermediates. Although the trapping method 17 was found to have the advantage of a broad derivatizing pH range, it depends too much on the structural accessibility of bis-cysteine SH groups, therefore, it is limited in practical use. Equation 1.9. Trapping folding intermediates by PAA III. Recognizing the Location of Cysteines and Cystines Practically, all the methods used to recognize the location of cysteine and cystine residues in proteins can be used for their location in the intermediates. Localization of cysteines and cystines in proteins is the oldest and best-studied field in protein chemistry. A large number of reagents and procedures have been developed and used to address this problem and there is no indication that the search for better methods is over yet. A. Classical Approach In the classical approach, one derivatizes free sulfhydryl groups with UV absorbing, radioactive or fluorescent groups under the condition to prevent sulflrydryl-disulfide exchange, and then cleaves the protein by chemical or enzymatic method. The resulting peptide fiagments are separated by and collected from chromatography. The collected fractions showing specific UV, radioactive or fluorescent detection are recognized, 18 mapped to sequence by the Edman technique and /or mass spectrometry, and related to specific segments of protein. A key to the success of the classical approach is to choose an appropriate derivatization reagent for sulflrydryl groups. The reagent should be soluble in the reaction medium and label sulfliydryl groups selectively, rapidly, and irreversibly under mild conditions (ideally under weak acidic conditions to avoid sulfllydryl/disulfide 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. Iodoacetate and pyridylethylation are the most extensively used alkylating derivatization methods for sequencing of peptides (62-68). Recently, Lee et a1. (67) identified the position of the sulflrydryl group of human glucocerebrosidase by modifying the sulflrydryl group with ["C] iodoacetic acid without reduction of disulfide bonds, digesting the protein with both protease-V8 at pH 4.0 followed by oc-chymotrypsin at pH 7.5, and finally sequencing the peptide based on detected radioactivity. Krieglstein et a1. (68) determined the location of sulfliydryl groups and disulfide bridges of botulinum neurotoxin type A. They converted five Cys residues to S-pyridylethyl cysteine residues; afier mercaptolysis converted all nine Cys residues to the S-pyridylethylated form. After confirming the predicted number of Cys residues by amino acid analysis, the positions of the five Cys residues carrying sulflrydryl groups and the four involved in disulfide l9 bridges were determined by comparing the elution patterns in HPLC of the cyanogen bromide mixtures of the exclusively alkylated and the mercaptolyzed-alkylated neurotoxin. The chromatographically isolated components were identified by N-terminal amino acid sequence analysis. While iodoacetate has the disadvantages of coelution with PTH-Gln and PTH-Glu on HPLC afier Edman degradation of modified protein and makes identification problematic, pyridylethylation results the problem of N-tenninal alkylation of protein because of the high reactivity of 4-vinylpyridine (69, 70). The other alkylating reagents have been developed for determination of cysteine residues (69-73). The alkylating reagent, 3-Bromopropylamine (72, 73), offers the advantage of easy identification of its derivative of cysteine residue by protein sequencing and amino acid analysis over other alkylating reagents in the modification and subsequent identification of cysteine residues by protein sequencing. J ue & Hale have developed the conditions for concomitant on- sequencer reduction and alkylation of cysteines in proteins with tri-n-butylphosphine and 3-bromopropylamine before sequencing the proteins (71). While alkylation by iodoacetates continues to be useful, there has been far greater interest in the use of this chemistry as a mechanism for introducing a larger molecule which can serve as structural probe. Examples include the fluorescent reagents, 5- (iodoacetamido)fluorescein (IAF) (74) and 5-[2~((iodoacetyl)amino)-ethyl]naphthalene- l-sulfonic acid (1,5-IAEDANS) (75-77). Bishop et at. has used IAF to selectively label two cysteine residues on the surface of CaI-ATPase of sacroplasmic reticulum (SR), digest with trypsin, purify the tryptic peptides by size-exclusion and reverse-phase HPLC, and identify the cysteine residues by 20 amino acid sequencing (78). Recently, a combination of 1,5-IAEDANS labeling of cysteine residues, enzymatic digestion, and MALDI-PSD has been developed to assign the location of free cysteine residue of testis angiotensin-converting enzyme (tACE) (79). The fiee cysteine residue was located by labeling with 1,5-IAEDAN S, isolating the fluorescent peptide from enzymatic digests by HPLC, and analyzing its sequence by MALDI-PSD. Labeling with IAEDANS or IFA suffers fiom less selectivity and reactivity toward sulflrydryls, as well as the fact that the reagents themselves are fluorescent before reaction with thiols. A highly reactive and specific reagent for protein thiols, 4- (aminosulfonyl)-7-fluoro-2,1.3-benzoxadiazole (ABD-F) has been developed (80). The selectivity of ABD-F is superior to that of other sulfhydryl-reactive fluorophores. In addition, unlike most other reagents previously utilized, ABD-F is nonfluorescent before reaction with protein thiols. The ABD-Cys adduct absorbs maximally at 385 nm and has a fluorescence emission maximum at 520 nm, allowing simple and sensitive detection in the presence of an excess of unlabeled protein. ABD-F has been utilized for analysis of cysteine residues in the proteins (81-83). The sulflrydryl and disulfide bond locations in the proteins were established using tributylphosphine (Bu3P) as the reducing reagent and ABD-F as the sulfllydryl reagent (83). The reaction was first run in the absence of reducing reagent, and all the peaks, both fluorescent and nonfluorescent ones were collected from HPLC of the tryptic peptides. An aliquot of each of the nonfluorescent peaks was then treated simultaneously with B113P and ABD-F since the two reagents do not react with each other. Sequencing the fluorescent peptides established the positions of the cysteine residues and disulfide bonds. 21 The ability to reduce disulfides and block the resulting thiols simultaneously in a single reaction, the stability of ABD-Cys to acid hydrolysis, and the recovery, albeit in low yield, of PTH-(ABD-)Cys on sequencing, along with the characteristic fluorescence of ABD—thiol derivatives, make it a favorable methodology. B. Afl'mity Chromatography The sulflrydryl group is the only functional group in a protein for which conditions are available for the formation of a stable covalent bond which can be split under mild conditions (84). A thiol-containing proteins can be attached on an affinity column. After the immobolized protein is cleaved chemically or enzymatically on the affinity column, the nonsulfhydryl-containing peptides do not bind to the column and can be washed away. The attached sulflrydryl-containing fragments are then removed from the column and subjected to further structural characterization by Edman degradation and/or mass spectrometry. The affinity interactions that can be used for the sulfhydryl-containing proteins include (85): (i) binding to affinity media that contain heavy metals, (ii) binding to affinity media with reactive disulfides that undergo facile thiol-disulfide exchange, and (iii) binding to affinity media that contain chelated zinc. While these methods were originally used to purify sulfhydryl-containing proteins, this approach was recently used to locate cysteine residues in proteins by mass mapping of bound sulflrydryl-containing peptides (86). This approach is powerful for locating sulflrydryl groups in high-mass proteins because the HPLC separation of the digests is unnecessary. But conditions must be established and optimized under which all the sulfllydryls can be attached to the column. 22 C. Cleavage at Cysteine Residue The cleavage of the peptide bond at cysteine residue after cyanylation was first introduced by Catsimpoolas and Wood (87). In this reaction, the conversion of an SH group into an SCN group was achieved in two stages by successive treatment of a protein with Elhnan’s reagent and then with cyanide. Jacobson and Stark (88, 89) showed that 2- nitro-S-thiocyanobenzoic acid (NT CB) specifically cyanylates cysteine sulfllydrls, which subsequently cleave at 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 (itz) peptides. Walkselman and Guibe—Jampel proposed another cyanylating reagent, 1-cyano-4-dimethylamino-pyridinimn (90). This reagent has the advantages of water solubility and reactivity at acidic condition. The cyanylation/chemical cleavage reaction of cysteine residue can be utilized to recognize structural information for the protein. If a protein contains 11 cysteine residues, the cleavage reaction results in the formation of n+1 peptide fi'agments, mass alignment of which indicates the number and location of cysteine residues. While SDS-PAGE can be used to mass map the peptide fiagments (91 -93), it suffers the poor mass accuracy (error > 5%) (89, 94). Daniel et al. used electrospray mass spectrometry to determine the masses of the NTCB cleavage products of E. cali dihydroorotase to identify the location of the active cysteine residues in the protein (95). Recently, Wu, et al. (96) combined the cyanylation reaction, specific chemical cleavage at cysteine residues, and MALDI-MS to quantitatively determine the number and location of cysteine and cystine residues in the proteins. This method has the advantages of being precise, simple and highly sensitive. However, MALDI-MS has a 23 signal suppression phenomenon which may result in incomplete assignments; the cyanylation reaction with NTCB is performed at alkaline condition, at which sulflrydryl- disulfide exchange may proceed for some of the proteins. IV. Assignment of Disulfide Bonds Disulfide bonds between cysteine residues, to generate cystine, are a common post- translational modification of proteins, principally those that operate in the extracellular milieu. A full description of the covalent structure demands that the connectivity of the bridged cysteines be analyzed. Although the amino acid sequences of proteins can be deduced from the corresponding cDNA sequences, the state and connectivity of the cysteine residues after the translation cannot be predicted accurately. Characterization of disulfide linkages in proteins has become an essential part of the analysis of recombinant DNA products, because molecules having incorrect disulfide linkages may have much lower activity than that of the desired product. Therefore, it is important to verify that recombinant materials have the correct disulfide linkages. Although disulfides are chemically rather simple, two factors can make the analysis difficult. First, the number of possible isomers grows rapidly as the number of bridges increases——1, 3, 15, 105, 945, etc. Second, base and reducing agents can catalyse exchange among partners, obscuring the original connectivity (97). Native proteins under physiological conditions are deceptively stable, but onCc they are denatured or nicked by proteases, disulfide exchange occurs when even a trace of thiol is present. 24 A. Conventional Fragmentation Strategy The general strategy for locating disulfide bonds in proteins by conventional methods involves several steps: (1). Digest proteins by chemical reagents and/or enzymes between the half-cystinyl residues avoiding disulfide exchange. (2). Separate fiagments also under conditions that stabilize the disulfide bonds. (3). Detect disulfide-containing peptides. (4). Identify subfi'agments, and hence individual cysteine residues that are connected. The conventional fragmentation approach has been well established for recognizing disulfide bond structure of proteins (98-100). The characteristic aspects of the strategy will be discussed below. B. Fragmentation of Proteins A The aim of fi'agmentation of a protein for disulfide mapping is to produce by enzymatic or chemical digestion a definitive set of peptide fragments only containing one disulfide bond. If a protein is not cleaved between every half-cystinyl residue by an enzyme or a chemical reagent, the peptides obtained by the cleavage must be cleaved again using another enzyme or chemical reagent. It is easy to identify the peptide generated by highly specific cleavages of proteins (101-103), because the N- or C- terrninal residues of the peptides are known. The general risk of disulfide exchange occurs during digestion; even though the protein, enzymes, and reagents may be initially the of thiols, the latter can arise by base catalyzed degradation of disulfides. Under any reasonable conditions very little thiol will be produced, but only catalytic amounts are 25 needed to cause extensive disulfide rearrangement during long incubations. Therefore, the choice of a proper enzyme should satisfy three requirements: the specificity of the enzyme; the capability of cleaving between every half-cystinyl residue; and the conditions to minimize disulfide scrambling. Cyanogen bromide, trypsin, and Staphylococcus aureus V8 protease are available to cleave proteins specifically. Specific chemical cleavage at the C-terminal side of Met by cyanogen bromide (CNBr) has been widely applied for initial cleavage of proteins due to its weak acidic reaction condition and volatile reagent and by-products. Since 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 (101). Trypsin has also been used fi'equently because it cleaves on the C- terrninal side of Arg or Lys residues with high specificity. Unfortunately, trypsin has maximum activity at pH 8.3, and is not active in acid. As a result, disulfide scrambling may occur during enzymolysis. Despite this shortcoming, trypsin is still the most widely used enzyme for the digestion of proteins because of its high specificity. Greatly increased use is now being made of the V 8 protease from S. aureus. This enzyme specifically cleaves on the C-terminal side of Glu or Asp residues in the pH range of 4.0 to 7.8. Unfortunately, if proteins cannot be cleaved by specific cleavage reagents, they must be cleaved by non-specific reagents. Pepsin, thermolysin and partial acid hydrolysis are uSeful for this purpose. The major advantage of using pepsin and partial acid hydrolysis to generate peptides is that disulfide bonds in proteins are fairly stable under mild acidic conditions (104, 105). However, the protein digests using such nonspecific 26 cleavages are complex mixtures, which make the separation and identification much more difficult. C. Purification of Cystinyl Peptides The purification of disulfide-bonded peptides constitutes a special problem. The difficulties in this area relate to the instability of disulfide bonds. Under the condition of mild acidity, i.e, pH 2.0-6.5, the disulfide bond is relatively stable. This requirement restricts the choice of isolation techniques. 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. However, the purification of digests from high-mass proteins is still a challenging task even though microbore or capillary HPLC columns are use because so many peptide fragments rrright be produced after hydrolysis. Gel filtration and ion-exchange chromatographic technologies are applicable for purification of digestion mixtures. The techniques are based on the chemical modification of cysteinyl residues to cysteic acids after the reduction of disulfide bonds in cystine peptides. Non-cystine peptides may be separated with great case from the more highly charged cysteic acid peptides generated from cystine peptides. D. Identification of Cystine-Containing Peptides The conventional approach for the identification of cystine-containing peptides after the purification is very tedious and time consuming. The peptides, which may contain 27 intra- or intermolecular disulfide bonds, are reduced, alkylated, isolated by HPLC again, and identified by their sequence with Edman degradation (106, 107). The sofi ionization methods in mass spectrometry have proven useful for locating disulfide bonds in proteins. One of the methods is by fast atom bombardment (F AB-MS) (108-112). Disulfide—containing peptides suitable for analysis by FAB-MS can be produced by treating proteins with enzyme. An aliquot of the digest can be analyzed by FAB-MS to determine the molecular weights of the peptides which are related to specific segments of the native protein; then following reduction or oxidization of the disulfide bonds with chemical reagents, such as the reducing reagent dithiothreitol or the oxidizing reagent performic acid. The sample is analyzed again by FAB-MS. The disulfide assignment can be completed by comparing the two FAB-MS results. An example is shown in the following scheme: 11111“ 378 Ala Gly Cys Lys Ala Gly Cys Lys '3 Reduction in \ > 3 SH l Thr Phe Thr Ser Cyi Thr Phe Thr Ser Cys 1111-1" 933 MHI 558 This method soon became an accepted method and was successfully applied to model proteins such as insulin (109), hen egg-white lysozyme and bovine ribonuclease A (111), and duck egg-white lysozyme (110). However, it needs specific enzymes to cleave the protein to form peptides only containing one disulfide bond, and the fragments should 28 not be large since FAB-MS is not suitable for high mass analysis, and it cannot be used for close or adjacent cysteine analysis. The another MS method is tandem mass spectrometry (MS/MS) employing high- energy collision-induced dissociation (CID) or low-energy CID, which produces peptides by enzyme digestion. The peptides containing a disulfide bond are selected as the precursor ion which is sequenced by MS/MS, and after reduction the peptides are sequenced again by MS / MS; the disulfide linkage can be assigned by comparing the two results (113-117). Electrospray ionization (E81) (2) and matrix-assisted laser desorption ionization (MALDI) (3) have revolutionized the role of MS for the analysis of proteins and peptides. Peptide-mapping by E81 (118, 119) or MALDI-MS (120, 121) are being used increasingly in the characterization and identification of disulfide bonds. The strategy used here is exactly the same as in 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. Recently, MALDI-MS with postsource decay was used to generate fragment ions from peptide fragments containing heteropeptides linked together by two disulfide bonds (122). Postsource decay analysis of these peptide fragments generates a series of singly charged fragment ions that, in addition to the peptide sequence ions, provide useful information for assigning disulfide arrangement in highly bridged disulfide-linked 29 peptides. The assignment was made possible by fragmentation at peptide bonds between two Cys residues in a peptide that constitutes the highly bridged fragment, while retaining the disulfide linkage to the other peptide. Fragmentation using other types of instruments, such as quadrupole ion-trap mass spectrometry with CID dissociation, usually did not generate such fragment ions. The method was used to facilitate the assignment of disulfide structure in a tumor necrosis factor binding protein, which contains 162 amino acids and 13 disulfide bonds. The conventional methodology requires the cleavage between half-cystinyl peptides by enzyme(s). If proper enzyme(s) cannot he find for this purpose, partial hydrolysis by acid can be used (105). Partial acid hydrolysis is performed under controlled conditions in which only limited peptide fiagments are formed. Partial acid hydrolysis is attractive as disulfide bonds are particularly stable in dilute or weak acids. However, it generates extremely complex digests. Unless one is dealing with a relatively small molecule, the problem is almost intractable without complicated chromatographic separation and mass spectrometry technique. E. S-S Assignment of Adjacent Cysteines The major problem of the conventional methodology is that it is unlikely to find a cleavage reagent that can cleave between the two adjacent or close spaced cysteine residues and produce peptide fragments that only contain one disulfide bond. In this case, alternative methodologies should be used. Different approaches have been reported to solve the problem, which include the combination of organic synthesis of the possible isomers of a disulfide-containing peptide 30 and high-performance tandem mass spectrometry (123) and multiple steps of Edeman degradation (124-129). The selective reduction of disulfide bonds can be applied to the problem (130-133). The selective reduction is based on the observation that disulfide bonds in a protein can often be reduced sequentially to thiols. The disulfide bonds of the protein are reduced stepwise by controlling the reduction conditions (temperature, concentration of reducing reagent, time, and denaturing conditions). The thiols so formed are labeled with a reagent, such as alkylating and fluorescent reagents. The protein is then reduced completely, and subsequently cleaved enzymatically. The location of the disulfide bond is determined by isolating and identifying the labeled peptides. The selective reduction only can be applied for the proteins containing disulfide bonds that have different stability. An alternative method for analyzing highly bridged proteins is based on the partial reduction of disulfide bonds by his-(2-carboxyethyl)-phosphine (TCEP) at pH 3, purification of the reduced protein by HPLC, alkylation of free thiols, analysis by sequencing, providing explicit assignment of disulfide bonds that had been introduced (134,135). This method has the advantages of being suitable for analyzing adjacent cysteines and tightly clustered cysteines, and preventing the sulfhydryl-disulfide exchange when performing partial reduction. The disadvantage is the limit of the analysis of the alkylated protein by sequencing. It is time consuming and hard to analyze large proteins. Recently an approach using a combination of the partial reduction by TCEP and cyanylation of free thiols by 1-cyano-4-dimethylamino-pyridinium (CDAP) 31 tetrafluoroborate at acidic condition was reported (136). Following the cleavage of the peptide bond on the N-terminal side of cyanylated thiols, the mass mapping of peptide fragments was performed by MALDI-MS. This method has the advantages of minimizing disulfide exchange, the possibility to analyze the proteins containing close or adjacent cysteines, and being fast, simple, and sensitive. V. Matrix-Assisted Laser Desorption/Ionization Mass Spectrometry (MALDI-MS) and Electrospray Mass Spectrometry (ESI-MS) In the last twenty years, important achievements in organic mass spectrometry have been obtained by the development of a number of new ionization techniques for the analysis of large, polar and thermally labile compounds. The general problem to be solved was how to generate intact gas-phase ions of the intact molecule fi'om compounds that are usually. solids and degrade or decompose when being exposed to thermal heating. Progress was achieved with the discovery of the desorption ionization techniques; plasma desorption mass spectrometry (PDMS) (137) and fast atom bombardment mass spectrometry (FAB MS) (138). The introduction in 1988 of two new techniques, matrix- assisted laser desorption/ionization (MALDI) (3) and electrospray ionization (ESI) (139), 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, 1. e., the desorption and ionization of large and labile biomolecules, and shown considerable promise in characterizing biomolecules by accurate determination of 32 molecular mass up to a few hundred thousand daltons. In combination with biochemical techniques, numerous applications have been achieved (1). A. Matrix-Assisted Laser Desorption/Ionization Mass Spectrometry (MALDI-MS) The first attempts to make use of laser ionization for the analysis of organic compounds date back to the 1970s. Especially, the desorption of ions containing the intact molecule of complex compounds, like stachyose and digitonin (140), may be regarded as a milestone. However, significant dissociation occurs in larger molecules because of the higher energy needed to desorb the large molecules, and the energy goes directly into the analyte. The major advance in laser desorption was the introduction of matrices (MALDI) by Hillenkamp and Karas in the late 19803 (141). The key to the MALDI process is to embed the analyte macromolecule in a high molar excess of a suitable matrix of molecules having a strong absorption at the laser wavelength. This induces an efficient transfer of the laser-pulse energy to the analyte and results in a soft desorption process, i. e., little or no fragmentation. The matrix also provides photoexcited acid or base sites for ionization of sample molecules during ion/molecule collisions (142). Although the mechanism of desorption and ion formation is complicated and unresolved, the formation of ions to nearly 106 Da and the high sensitivity of MALDI (subfemtomole) has led to the explosive commercialization of instruments in not only mass spectrometry laboratories, but also in biochemical facilities and laboratories where bench scientists are able to use the instrumentation. 33 1. Matrix and Sample Preparation The matrix is the key to the successfirl analysis of high-mass biomolecules. While several hundred of organic compounds have been investigated (143, 144), only a few of these are widely applicable for peptide and protein analysis: a-cyano-4-hydroxycinnamic acid (4-HCCA), 1,5-dihydroxyoxybenzoic acid (DHB), and 3,5-dimethoxy-4- hydroxycinnamic acid (sinapinic acid) (145). It was considered important that the laser wavelength match the desorption maximum of the matrix compound. All the matrices above have strong UV absorption in the 320-350 nm range and can be used with a much cheaper nitrogen laser (337 nm) or frequency-doubled Nd-YAG laser (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 (146). A number of different methods have been developed for growing analyte-matrix crystals. The original method described by Karas and Hillenkamp has been named the ‘dried-droplet’ method (141). It entails drying a droplet of a solution containing the matrix (1-10 mmol) and the protein (1-10 umol). A simple variant on the “dried droplet” method was proposed by Vorrn, et al. (147). They first placed a thin layer of the matrix compound onto a metal surface and then put a droplet of the analyte containing solution on top of the layer of matrix compound. As the droplet dries it dissolves some of the matrix which crystallizes onto the substrate matrix layer when the droplet dries 34 completely. The crystallized matrix layer is very thin, allowing for more reproducible mass measurements and producing ions for < 10 laser shots on a spot. It is possible to grow large, analyte-matrix crystals under near equilibrium condiitons, rather than a rapidly drying droplet (148). Supersaturated matrix solutions containing analyte will form crystals that can be used directly in an ion source. Supersaturation can be achieved either by heating and cooling or by slow evaporation. A recently developed technique of producing crystals for use in MALDI ion sources involves growing a thick fihn of matrix crystal (149). The substrate for the crystals (usually a piece of metal) is first covered with the matrix material, by rapid drying of a solution of the matrix dissolved in an organic solvent. The small matrix crystals that cover the surface are then crushed and smeared over the substrate surface, resulting in a well-adhered layer of stressed, crystalline matrix. A saturated solution containing the matrix and analyte is then placed onto the surface and allowed to dry slightly. The stressed crystalline material on the substrate acts to seed crystal formation at many sites on the surface, resulting in the rapid growth of a rather uniform polycrystalline film of analyte-matrix crystals over the surface. The liquid drOp can then be removed by blotting. The resulting film is very strongly adhered to the substrate and it can be washed thoroughly to remove any contaminants that were present in the analyte solution. The mechanism of “matrix assistance” for the transfer of analyte molecules to the gas phase and their ionization remains incompletely understood; the choice of matrix and method application is still empirical. Sample preparation by different matrices and solvents (150), matrix additives (151), and evaporation rate (147, 152) affects the 35 resolution and sensitivity of MALDI. Optimal results require parallel analyses under different conditions. 2. Instrumentation To date, MALDI of large molecules has been obtained mainly in combination with the field-free time-of-flight (TOF) mass analyzer. A schematic drawing of MALDI-TOP instrument is shown in Figure 1.2. The solid deposit of sample-matrix is irradiated by a pulsed laser, the most common one is the Nd-YAG 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 analyte molecule results in the desorption and ionization of the co-crystallized 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 25KV) 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 of all ions is zero) by the electric field toward a long (1-2 m) field-free time-of-flight (TOF) analyzer. The time-of-flight mass analyzer is simple, cheap, and well-suited to the pulsed nature of laser desorption in MALDI (153). It has virtually no upper mass range and is therefore compatible with MALDI, which can produce very high m/z ions. Another 36 EOEEEE wfituOHrarEE .N... 0.59". J. 3509 .28qu .//.._o_ _ / _ A I— BE 29.“. 7 #4 , _ ._ %% nae % _ _ _ 4./ “noun ‘ x a. Emm I..— _ \~\ \\ x 528mm _ amour... \\ E239... _ U \\ m \ Emmmmwwl Hut .9325 Emam 37 advantage of 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 massses and is described by the relationship: tof=L/v=L(m/22 eV)”2, where L is the length of flight tube, v is the ion velocity, m is the mass of the ion, and V is the acceleration potential. Thus, low-mass 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 spectrum. 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)”2 and can be used to convert the x-axis of the spectrum (ion arrival time, tot) into a m/z ratio axis (a conventional mass spectrlun). The TOF analyzer is highly efficient because all ions of different m/z ratio arising fi'om 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. This limits the capacity to detect certain protein modifications and protein sequence variations, especially at high mass. In spite of this, it is possible to achieve a mass accuracy of 0.1-0.01% for proteins 38 with molecular masses between 1 and 40 KDa, and with somewhat poorer accuracy for proteins above 40 KDa (153). Two techniques were introduced to solve the low-resolution problem of linear TOF mass analyzer, reflectron-TOF (reTOF) (154) and delayed ion extraction (DE) (155). The reflectron-TOF consists of two linear field-fiee region and an ion mirror, which compensates for different, nonzero initial kinetic energies of ions of the same m/z. Delayed ion extraction allows the ions to disperse in the source region (acceleration region just above the surface) owing to their initial velocity, while the density of neutrals is decreased by pumping them away. The number of ion-molecular collisions before extraction of the ions into the field-flee drift region is therefore reduced, which decreases the width of the translational energy distribution. Simultaneously, ions having higher initial kinetic energy move further away from the extraction field in the source region and thus are given less “kick”; those of lower kinetic energy experience a higher extraction field. This approach has proven so successful that limitations in mass resolution are now focused on detection systems (the time resolution). B. Electrospray Ionization Mass Spectrometry (ESI-MS) Electrospray ionization-mass spectrometry (ESI-MS) has its origins in research that long preceded the current flurry of activity. The study of the electrospray phenomena extends back perhaps over two and one-half centuries to the work of Bose (156), and certainly to that of Zeleny early in this century (157). The seminal research into the use of electrospray as an ionization method for macromolecules was due to Malcom Dole and co-workers (158, 159), who performed extensive studies into the electrospray process and 39 defined many of the experimental parameters. Experimental evidence was presented by Dole for ionization of zein (Mr. ~50,000) (160) and lysozyme (161). However, interpretation of their results was problematic because a mass spectrometer was not available, and only ion retardation and ion mobility measurements were obtained. In 1984, ESI combined with mass spectrometry was first reported, essentially simultaneously, by both Yarnashita and Fenn (162) and Aleksandrov et al. (163, 164). In 1988, Fenn and co-workers (165) first reported ESI-MS spectra of intact multiply protonated molecules of proteins up to Mr 40,000. Since its introduction in 1988 as a tool for the ionization of large biomolecules, the practice of ESI-MS in medical, biotechnological, and pharmaceutical arenas has grown impressively. One obvious reason for this is the speed with which commercial instrumentation became available, and particularly the ease of retrofiting existing quadrupole mass spectrometers for ESI-MS. As evidenced by the growing number of ESI-MS related publications, routine analysis by ESI-MS is becoming increasely common in basic research, quality control, and bioanalytical laboratories. l. The Electrospray Process Electrospray ionization requires continuous flow of solution (generally from a small diameter capillary tube) in the presence of a high electric field. Liquid is expelled fi'om the tube as charged droplets which have a charge-to-volume ratio that can approach the physical maximum allowed for droplet stability, i. e., the Rayleigh limit (166), which by definition is reached when droplet surface charge density exceeds the liquid’s surface tension (167). Droplet instability is manifested in a “coulomb explosion” or “droplet 40 fission” event. It is generally believed solvent is lost through evaporation until Coulomb explosions produce smaller droplets. The small droplets evaporate rapidly and the protonated molecules are released into the gas phase. It is unclear whether ions escape from droplets (i. e., field ionization) or solvent evaporates to leave ions (1. e., droplet evaporation), and the two mechanisms will likely be experimentally undistinguishable. In contrast to all other mass spectrometric ionization methods, the change from the liquid into gas phase is very gentle and probably occurs with at least one solvation shell still surrounding the analyte molecule. Fragmentation during this ‘desorption event’ has hardly been observed, and it is now thought that even aspects of the tertiary structure of a protein survive into the gas phase, at least under certain circumstances (168). 2. Instrumentation The generic representation of an ESI source shown in Figure 1.3 incorporates several features that are found on many ESI sources. The electrospray capillary typically consists of a small i.d. (<100 um) metal or glass capillary biased at :tl-S KV (positive potential for cations, negative potential for anions) relative to the desolvating capillary. Recent studies suggest that significant improvements in sensitivity can be realized with very small diameter (<20 um) electrospray capillaries and the use of very low flow rates (169). A solution containing the analyte of interest is forced into the electrospray capillary at a flow rate of 0.01-10 til/min (depending on analyte concentration and solution conditions) and is nebulized by the resulting electrospray plume. A common solvent for the analysis of peptide cations is 50/50 MeOH/Hzo (or acetonitrile/H20) with 1-2% HOAc (or formic acid) and a protein/peptide concentration of 10-5 — 10'7 M. 41 Mass Analyzer Electrospray Plume Skimmer . Cones Electrospray , -" Capillary ‘ —// ”I ‘ D—esolvating\\ Capillary +High Voltage ion Optics __,. ——>[M+nH]n+ _ solvent droplet formation of evaporation fission at desolvated Ions Rayleigh by further field limit desorption and/or ion evaporation Figure 1.3. ESI-MS instrument and droplet production in ESI interface 42 Electrospray devices also are used to couple separation techniques, including capillary electrophoresis (170) and high performance liquid chromatography (171) to mass spectrometry. Direct coupling of the separation technique and the ionization technique is accomplished in a number of ways including the addition of a conductive sheath liquid or the application of an electrically conductive coating to the terminus of the separation capillary (171). It has been recognized that the electrospray ionization process is driven by droplet desolvation and that desolvation at the molecular level is essential to obtain high quality spectra. Thus, a desolvation capillary often is added to expedite desolvation of the highly charged droplets generated in the electrospray plume. This capillary typically has an internal diameter of 400 pm to 1 mm, and a length of 10-25 cm. Glass inlet capillaries are employed widely in conjunction with a countercurrent flow of heated, dry nitrogen to enhance the rate of droplet desolvation and prevent large amounts of material from entering the spectrometer. Both ends of the glass capillary are gold coated and electrically biased to define a finite potential difference between the capillary ends and the neighboring source elements. Next, a skimmer cone, or more commonly, a series of skimmer cones with a 300nm to 1.5mm diameter aperture is used to transmit to the mass spectrometer as many ions fiom the sample as possible while decreasing the pressure in each subsequent vacuum stage of the chamber. An electrical bias on the skimmer cone(s) serves to focus the [ion beam and defines the average kinetic energy of the ion beam. Ion transmission from the source to the mass analyzer is enhanced greatly by the addition of a few simple ion lenses that guide and focus the ion beam into the mass analyzer. 43 In conjunction with electrospray ionization a quadrupole mass spectrometer is commonly used as a result of relatively low cost and ease of interfacing with many commercial systems. A quadrupole mass analyzer consists of four equidistant rod electrodes arranged as two pairs to which time-varying ac and dc potentials are applied. Only ions having a mass-to-charge ratio within a very narrow m/z range are transmitted by the quadrupole. Ions having a m/z outside this range move in unstable orbits and collide with the electrodes. Ions striking the detector transfer charge that registers as a current that is proportional to ion abundance and is amplified subsequently. Correlating ion abundance values with the values of applied dc bias and amplitude of ac voltage gives a mass spectrum. The relatively high sensitivity and rapid scan rates obtainable with commercial quadrupole mass spectrometers resulted in their widespread use. 3. Interpretation of ESI-MS Spectra As a majority of ionization techniques today produce single charged ions (and, thus, mass spectra in which the measured m/z is equivalent to the mass of analyte), the multiple-charge states common to ESI-MS spectra might seem a bit confusing. However, their interpretation often is straightforward once the appropriate procedure is applied. As first demonstrated by Mann, Meng, and Fenn (172), a simple algorithm can be applied to single component spectra based on the assumption that each peak differs from its neighboring peak by only charge. The value of the charge state of a selected lower m/z peak (210,.) is determined according to the equation: Ztow = [(m/z)hi]/ {[(m/z)hi-(m/z)tow]/n} , where n is the number of peaks between the selected lower m/z peak [(m/z)tow] and a selected higher m/z peak [(m/z)m] being considered. From the known charge state of a single peak calculated by the equation, the charge states of all the remaining peaks are directly calculated by simple addition or subtraction. From measured m/z and the calculated charge 2, the ionic mass, 111,, is Obtained easily, m1 = m/z x z. The mass of the corresponding neutral analyte, Mr: Mr = [(m/z) x z] i nMa, Ma is the mass of charged adduct. As each charge state yields an independent mass measurement, mass accuracy often is improved considerably by averaging over the entire charge envelope: Mr = (l/N) am, where Mr is the calculated molecular weight from a single charge state and N is the number of charge states being averaged. C. The Comparison of MALDI and ESI While MALDI and ESI are widely used techniques in analysis of large biomolecules, they have different and complementary natures (Table 1.1), which has made it increasingly important that researchers have access to both types of instrumentation. Independently, E81 and MALDI-MS can help answer many questions; yet together they present a formidable research tool with new levels of sensitivity, accuracy, and mass range. Their utility for mass measurement meets the needs of chemists and biologists alike, facilitating routine characterization in small molecule synthesis, protein synthesis, and compounds obtained directly from biological matrices. The use of ESI and MALDI mass spectrometry extends beyond simple characterization. Noncovalent interactions, protein and peptide sequencing, DNA sequencing, protein folding, in vitra drug analysis, and drug discovery are among the areas to which ESI and MALDI-MS have been applied. 45 Table 1.1. Comparison of ESI-MS and MALDI-MS ESI-MS MALDI-MS Mass limit, Da ~200,000 >3.000.000 (practical) (70,000) (1 50,000) HPLC/MS capable Tolerant of HIM Advantages Multiple charging concentrations of salts Capable of observing Highest mass capability noncovalentcomplexes Tolerant of mixtures directly from water Femtomole sensitivity Femtomole-to-picomole Being developed as a 100' sensitivity for sequence analysis Disadvantages Multiple charging can be Typically low resolution confusing with mixtures (4500) Typically need < "M salt linear TOF without DE concentrations for 90°d Not amenable to LC/MS signals Not tolerant of mixtures Suitable Peptides Peptides compounds Proteins Proteins Carbohydrates Glycoproteins Nucleotides Carbohydrates Nucleotides Oligonucleotides Phosphoproteins Small chargeable molecules Oligonucleotides Phosphoproteins Small chargeable molecules Heterogeneous samples 46 VI. References l. Burlingame, A. L., Boyd, R. K., and Gaskell, S. J ., Anal. Chem, 68, 599R (1996). 2. Fenn, J. B., Mann, M., Meng, C. K., Wong, S. F., and Whitehouse, C. M., Science, 246, 64 (1989). 3. Karas, M., and Hillenkamp, F., Anal. Chem, 60, 2299 (1988). 4. Weissman, J. S., Kim, P. S., Science, 253, 1386 (1991). 5. Pain, R. H., Trends Biochem Sci, 12, 309 (1987). 6. Scheraga, H. A., Konishi, Y., 001, T., Adv. Biophys, 18, 21 ( 1984). 7. Kim, P. S., and Baldwin, R. L., Annu. Rev. Biochem, 59, 631 (1990). 8. Matthews, C. R., Annu. Rev. Biochem, 62, 653 (1993). 9. Ptitsyn, O. B., Curr. Opin. Struct. Biol, 5, 74 (1995). 10. Creighton, T. B., J. Mol. Biol., 87, 563 (1974). 11. Creighton, T. B., Goldenberg, D. P., J. Mol. Biol, 179, 497 (1984). 12. Scheraga, H. A., Konishi, Y., 001, T., Adv. Biophys, 18, 21 (1984). 13. Jacobson, G. R., Schaffer, M. H., Stark, G. R., and vanaman, T. C., J. Biol. Chem, 248, 6583 (1973). 14. Wu, J., Gage, D., and Watson, J. T., Anal. Biochem, 235, 161 (1996). 15. Wu, J ., and Watson, J. T., Protein Sci. 6, 391 (1997). 16. Roder, H., Elove, G., and Englander, S. W., Nature, 335, 700 (1988). 17. Balbach, J., Forge, V., Van Nuland, N. A. J., Winder, S. L., Hore, P. J ., and Dobson, C. M., Nat. Struct. Biol., 2, 865 (1995). 18. Udgaonkar, J. B., and Baldwin, R. L., Nature, 335, 694 (1988). 19. Jennings, P. A., and Wright, P. B., Science, 262, 892 (1993). 47 20. Freedman, R. B., Trends Biochem. Sci., 9, 438 (1984). 21. Creighton, T. E., Biochem. J., 270, l (1990). 22. Haber, E., and Anfinsen, C. B., J. Biol. Chem, 237, 1839 (1962). 23. Creighton, T. E., Methods Enzymal., 131, 83 (1986). 24. Gilbert, H. F ., Adv. Enzymal., 63, 69 (1990). 25. Honk, J ., and Whitesides, G. M., J. Am. Chem. Sac., 109, 6825 (1987). 26. Wilson, J. M., Bayer, R. J ., and Hupe, D. J ., J. Am. Chem. Sac., 99, 7922 (1977). 27. Szajewski, R. P., and Whitesides, G. M., J. Am. Chem. Sac., 102, 2011 (1980). 28. Goldenberg, D., Trends Biochem Sci., 17, 992 (1992). 29. Barron, E. S. G., Adv. Enzymal., 11, 201 (1951). 30. Larsson, A., Orrenius, S., Holrngren, A., and Mannervik, B., in Functions of Glutathione: Biochemical, Physiological, Toxicological, and Clinical Aspects, Raven, New, York, 1983. 31. Iyer, K. S., and Klee, W. A., J. Biol. Chem, 248, 707 (1973). 32. Creighton, T. B., Methods in Enzymolagy, 107, 305 (1984). 33. Misra, H. P., J. Biol. Chem. 249, 2151 (1974). 34. Anfinsen, C. B., Haber, B., Sela, M., and White, F. H., Proc. Natl. Acad. Sci. USA, 47, 1309 (1961). 35. Creighton, T. B., Prog. Biophys. Mol. Biol., 33, 231 (1978). 36. Goldberger, R. F., Epstein, C. J ., and Anfinsen, C. B., J. Biol. Chem, 238, 628 (1963). 37. Venetianer,.P., and Straub, F. B., Biochim. Biophys. Acta, 67, 166 (1963). 38. Bergman, L. W., and Kuehl, W. M., J. Biol. Chem, 254, 8869 (1979). 48 39. Darby, N. J., Norin, P. B., Talbo, G., and Creighton, T. B., J. Mol. Biol., 249, 463 (1995) 40. Weissman, J. S., and Kim, P. S., Proc. Natl. Acad. Sci. USA, 89, 9900 (1992). 41. Gilbert, H. F., in Mechanisms of Protein Folding, Pain, R. H. (Editor), Oxford Press, Oxford, pl 12 (1994).. 42. Chang, J. Y., Biochemistry, 35, 11702 (1996). 43. Chatrenet, B., and Chang, J. Y., J. Biol. Chem, 268, 20988 (1993). 44. Chang, J. Y., J. Biol. Chem, 270, 25661 (1995). 45. Chang, I. Y., Canals, F., Schindler, P., Querol, B., and Aviles, F. X., J. Biol. Chem, 269,22087(1994) 46. Hua, Q. X., Narhi, L., Jia, W., Arakawa, T., Rosenfeld, R., Hawkins, N., Miller, J. A., and Weiss, M. A., J. Mol. Biol., 259, 297 (1996). 47. Chatrenet, B., and Chang, J. Y., J. Biol. Chem, 268, 20988 (1993). 48. Wilkins, J. A., Cone, J ., Randhawa, 2., Wood, D., Warren, M. K., and Witkowska, H. 13., Protein Science, 2, 244 (1993). 49. Hober, S., Forsberg, G., Palm, 6., Hartmanis, M., and Nilsson, B., Biochemistry, 31, 1749 (1992). 50. Glocker, M. 0., Arbogast, B., Milley, R., Cowgill, C., and Deinzer, M. L., Proc. Natl. Acad. Sci. USA, 91, 5868 (1994). 51. Creighton, T. B., Proc. Natl. Acad. Sci., USA, 85, 5082 (1988). 52. Rothwarf, D. M., Scheraga, H. A., J. Am Chem. Sac., 113, 6293 (1991). 53. Konishi, Y., Ooi, T., and Scheraga, H. A., Biochemistry, 21, 4734 (1992). 54. Goto, Y., and hamaguchi, J. Mol. Biol., 146, 321 (1981). 49 55. Torella, C., Ruoppolo, M., Marino, G., Pueci, P., FEBS Letters, 352, 301 (1994). 56. Rothwarf, D. M., Scheraga, H. A., J. Am. Chem. Soc., 113, 6293 (1991). 57. Rothwarf, D. M., and Scheraga, H. A., Biochemistry, 32, 2671 (1993). 58. Bruice, T., and Kenyon, G. L., J. Protein Chem, 1, 47 (1982). 59. Roberts, D. D., Lewis, S. D., Ballou, D. P., Olseon, S. T., and Shafer, J. A., Biochemistry, 25, 5595 (1986). 60. Thannhauser, T. W., Sherwood, R. W., and Scheraga, H. A., J. Protein Chem. 17, 37 (1998) 61. Happersberger, P., Cowgill, C., Deinzer, M. L., and Glocker, M. 0., Proceedings of the 44'" ASMS Conference on Mass Spectrometry and Allied Topics, Portland, Oregon, May, 1996, p1050. 62. Makinen, A. L., and Nowak, T., J. Biol. Chem, 264, 12148 (1989). 63. Wang, Z. X., Preiss, B., and Tsou, C. L., Biochemistry, 27, 5095 (1988). 64. Mak, A. S., and Jones, B. L., Anal. Biochem, 84, 432 (1978). 65. Andrews P. C., and Dixon, J. B., Anal. Biochem, 161, 524 (1987). 66. Krufl, V., Kapp U., and Wittmann-Liebold, B., Anal. Biochem, 193, 306 (1991). 67. Lee, Y.,Kinoshita, H., Radke, G., Weilder, S., Barranger, J. A., and Tomich, J. M., J. Protein Chem, 14, 127 (1995). 68. Krieglstein, K. G., DasGupta, B. R., and henschen, A. H., J. Protein Chem, 13, 49 (1994). 69. Inc, R. A., and Hale, J. B., Anal. Biochem, 221, 374 (1994). 70. Krutzsch, H., and Inman, J. K., Anal. Biochem, 209, 109 (1993). 71. Brune, D., Anal. Biochem, 207, 285 (1992). 50 72. Inc, R. A., and Hale, J. E., Anal. Biochem, 210, 39 (1993). 73. Hale, J. B., Beidler, D. B., and Jue, R. A., Anal. Biochem, 216, 61 (1994). 74. Seifi'ied, S. B., Wang, Y., and Von Hippel, P. H., J. Biol. Chem, 263, 13511 (1988). 75. Pardo, J. P., and Slayman, C. W., J. Biol. Chem, 264, 9373 (1989). 76. Miyanishi, T., and Borejdo, J., Biochemistry, 28, 1287 (1989). 77. First, E. A., and Taylor, S. S., Biochemistry, 28, 3598 (1989). 78. Bishop, J. B., Squier, T. C., Bigelow, D. J., and Inesi, 6., Biochemistry, 27, 5233 (1988) 79. Sturrock, E. D., Yu, X. C., Wu, Z., Biemann, K., and Riordan, J. F., Biochemistry, 35, 9560 (1996). 80. Toyooka, T., and Imai, K., Anal. Chem, 56, 2461 (1984). 81. Kirley, T., J. Biol. Chem, 264, 7185 (1989). 82. Kirley, T., Anal. Biochem, 180, 231 (1989). 83. Chin, C. C. Q., and Wold F., Anal. Biochem, 214, 128-134 (1993). 84. Jason, J ., and Ryden, L., Protein Purification: Principles, High Resolution methods, and Applications, VCH Press, (1993), p252. 85. Kalef, B., Walfish, P. G., and Gitler, C., Anal. Biochem, 212, 325 (1993). 86. Zaluzec, E. J ., Gage, D. A., and Watson, J. T., in Proceeding of the 42th ASMS Conference on Mass Spectrometry and Allied Topics, Chicago, (1994), p20. 87. Catsimpoolas, N., and Wood, J. L., J. Biol. Chem, 241, 1790 (1966). 88. Jacobson, G. R., Schaffer, M, H., Stark, G. R., and Venaman, T. C., J. Biol. Chem, 248, 6583, (1973). 89. Stark, G. R., Methods in Enzymology, vol. 47, 129 (1977). 51 90. Wakselman, M., and Guibe-Jampel, B., J. C. S. Chem. Comm, 21 (1976). 91. Mikami, B., and Morita, Y., J. Biochem., 93, 777 (1983). 92. Nika, H., and Hultin, T., Eur. J. Biochem. 142, 521 (1984). 93. Minard, B., Desmadril, M., Ballery, N., Perahia, D., Mouawad, L., Hall, L., and Yon, J. M., J. Biochem. 185, 419 (1989). 94. Darbre, A., Practical protein Chemistry—A Handbook, John Wiley & Sons, (1985), p149. 95. Daniel, R, Caminade, B., Martel, A., Goflic, F ., canosa, D., Carrascal, M., and Abian, J ., J. Biol. Chem, 272, 26934 (1997). 96. Wu, J., Gage, D. A., and Watson, J. T., Anal. Biochem., 235, 161 (1996). 97. Ryle, A. P., and Sanger, F., Biochem. J., 60, 535 (1955). 98. Biemann, K., in Methods in Protein Sequence Analysis, Plenum Press, New York, (1993), p119. 99. Staswick, P. B., Hermodson, M. A., and Nielsen, N. C., J. Biol. Chem, 259, 13431 (1984). 100. Haefiher—Gormlcy, L., Parente, L., and Wetlaufer, D. B., Int. J. Pept. Protein Res., 26, 83 (1985). 101.Yazdanparast, R., Andrews, P. C., Smith, D. L., and Dixon, J. E., J. Biol. Chem, 262, 2507 (1987). 102. Sun, Y., Bauer, M. D., Keough, T. W., and Lacey, M. P., Methods in Molecular Biology, Vol. 6] : Protein and peptide Analysis by Mass Spectrometry, edited by Chapman, J. R., Humana press Inc., Totowa, NJ, (1996). 103. Smith, D. L., and Zhou, Z., Methods in Ezymol., Vol. 193, 374 (1990). 52 104. Sun, Y., Zhou, Z., and Smith, D. L., in Techniques in Protein Chemistry, (hugli, T. B., ed.), Academic, New, York, p176 (1989). 105. Zhou, Z., and Smith, D. L., J. Protein Chem, 9, 523 (1990). 106. Wetlaufer, D. B., Parente, L., and Haeffner-Gormley, L., Int. J. Prpt. Protein Res., 26, 83 (1985). 107. Thannhauser, T. W., McWherter, C. A., and Scheraga, H. A., Anal. Biochem., 149, 322 (1985). 108. Morris, H. R. and Pucci, P., Biochem. Biophys. Res. Commun., 126, 1122, 1985. 3109. Takao, T., Yoshida, M., Hong, Y., Aimoto, S., and Shimonishi, Y., Biomed. Mass Spectrom., 11, 549, 1984. 110. Yazdanparast, R., Andrews, P. C., Smith, D. L., and Dixon, J. B., J. Biol. Chem, 262, 2507, 1987. 111. Stults, J. T., Bourell, J. H., Canova—Davis, B., Ling, V. T., Laramee, G. R., Winslow, J. W., Griffin, P. R., Rinderknecht, B., and Vandlen, R. L., Biochem. Environ. Mass Spectrom., 11, 549-556, 1984. 112. Vaccaro, A. M., Salvioli, R., Barca, A., Tatti, M., and Pucci P., J. Biol. Chem, 270, 9953-9960, 1995. 113. Yoshida, S., Takamatsu, T., Kanda, F., Rapid commun. Mass Spectrom. 10, 641- 648, 1996. 114. Beam, M. F., and Carr, S. A. Anal. Biochem. 201, 216,1992. 115. Sun, Y., and Smith, D. L. Anal. Biochem. 172, 130, 1988. 116. Hunt, D.F., Yates, J. R., Shabanowitz, J., Winston, S., and Hauser, C. R., Proc. Natl. Acad. Sci. USA, 83,6233, 1986. 53 117. Bauer, M., Sun, Y., Degenhardt, C., and Kozikoski, B., J. Protein Chem. 12, 759,1993. 118.Krishnamurthy, T., and Hauer, C. C., Biol. Mass spetrom 23, 719, 1994. 119. Bayer, M., Sun, Y., Degenhardt, C., and Kozikowsiki, B., J. Protein Chem. 12, 759, 1993. 120. Stewart, A. B., Raffioni, S., and Chandhery, T., Protein Science, 1, 777, 1992. 121. Patterson, S., and Katta, V., Anal. Chem. 66, 3727, 1994. 122. Jones, M. D., Patterson, S. D., and Lu, H. S., Anal. Chem. 70, 136 (1998). 123.Stults, J. T., Bourell, J. H., Canova-Davis, B., Ling, V. T., Laramee, G. R., Winslow, J. W., Griffin, P. R., Rinderknecht, E., and Vandlen, R. L., Biomed. Environ. Mass Spectrom., 19, 655 (1990). 124. Fox, J. W., Elzinga, M., and Tu, A. T., Biochemistry, 18, 678 (1979). 125. Serensen, H. H., Thomsen, J ., and Bayne, S., Biomed. Environ. Mass Spectrom. 19, 713 (1990). 126. Nokihara, K, Morita, N., Yamaguchi, M., and Watanabe, M., Anal. Lett., 25, 513 (1992). 127. Zhang, D., and Liang, S., J. Protein Chem, 12, 735 (1993). 128. Krishnamurthy, T., Hauer, C. R., Prabhakaran, M., Freedy, J. G., and Hayashi, K., Biol. Mass Spectrom., 23, 719 (1994). 129. Tang, Y., and Selsted, M. E., J. Biol. Chem, 268, 6649 (1993). 130.Gray, W. R., Protein Science, 2, 1749 (1993). 131. Haniu, M., Hsieh, P., Rohde, M. F ., and Kenney, W. C., J. Biochem. Biophys., 310, 433 (1994). 54 132. Yamashita, H., Nakatsuka, T., and Hirose, M., J. Biol. Chem, 270, 29806 (1995). 133. Fairlie, W. D., Stanton, P. G., and Hearn, M. T. W., Biochem. J., 314, 449 (1996). 134. Gray, W. R., Luque, F. A., Galyean, R., Atherton, B., Sheppard, R. C., Stone, B. L., Reyes, A., Alford, J., Mcintosh, M., Olivera, B. M., Cruz, L. J ., and rivier, J., Biochemistry, 23, 2796 (1984). 135. Gray, W. R, Protein Sci., 2, 1732 (1993). 136. Wu, J ., and Watson, J. T., Protein Science, 6, 391 (1997). 137. Thorgersen, D. F., Skowronski, R P., and Macfarlane, R. D., Biochem. Biophys. Res. Commun., 60, 616 (1974). 138. Barber, M., Bordoli, R. S., Sedgwick, R. D., and Tyler, A. N., Chem. Commun., 325 (1981). 139. Meng, C. K., Mann, M., and F enn, J. B., Proc. 36'” ASMS Conference on Mass Spectrometry and Allied Topics, San Francisco, CA, (1988), p771. 140. Posthumus, M. A., Kistcmaker, P. G., Meuzelaar, H. L. C., deBrauw, M. C., Anal. Chem, 50, 985 (1978). 141. Karas, M., Bachmann, D. B., Hillenkamp, F., Int. J. Mass Spectrom. Ion. Proc., 78, 53 (1987). 142. Hillenkamp, F., Karas, M., Beavis, R. C., Chait, B. T., Anal. Chem, 63, 1193A (1991) 143. Ehring, H., Karas, M., and Hillenkamp, F., Org. Mass Spectrom., 27, 472 (1992). 144. Fitzgerald, M. C., Parr, G. R., and Smith, L. M., Anal. Chem, 65, 3204 (1993). 145. Beavis, R. C., and Chait, B. T., Rapid Commun. Mass Spectrome., 3, 233 (1989). 55 146. Juhasz, P., Costello, C. B., and Biemann, K., J. Am. Soc. Mass Spectrom., 4, 399 (1993). 147. Vorm, 0., Roepstorff, P., and Mann, M., Anal. Chem, 66, 3281 (1994). 148. Xiang, F., and Beavis, R. C., Org. Mass Spectrom., 28, 1424 (1993). 149. Xiang, F., and Beavis, R. C., Rapid Commun. Mass Spectrom., 8, 199 (1994). 150. Cohen, S. L., and Chait, B. T., Anal. Chem, 68, 31 (1996). 151. Preston, L. M., Murray, K. K., and Russsell, D. H., Biol. Mass Spectrom., 22, 544(1993). 152.Vorm, O., and mann, M., J. Am. Soc. Mass Spectrom., 5, 955 (1994). 153. Beavis, R. C., and Chait, B. T., Proc. Natl. Acad. Sci. USA, 87, 6873 (1990). 154. Mamyrin, B. A., Karateev, V. I., Shmikk, D. V., Zagulin, V. A., Sov. Phys. JETP, 37, 45 (1973). 155. Anderegg, R. J ., Wagner, D. S., Stevenson, C. J ., Borchardt, R. T., J. Am. Soc. Mass Spectrom., 5, 425 ( 1994). 156. Bose, G. M., Recherches sur le cause et sur la veritable theorie de l'electricite, Wittenberg, 1745.. 157. Zeleny, J., J. Phys. Rev., 10, 1 (1917). 158. Dole, M., Mack, L. L., Hines, R. L., Mobley, R. C., Ferguson, L. D., Alice, M. B., J. Chem. Phys. 49, 2240 (1968). 159. Mack, L. L., Kralik, P., Rheude, A., Dole, M., J. Chem. Phys., 52, 4977 (1970). 160. Clegg, G. A., Dole, M., Biopolymers, 10, 821 (1971). 161. Gieniec, J ., Mack, L. L., Nakama, K., Gupta, C., Kumar, V., Dole, M., Biomed. Mass Spectrom., 11, 259 (1984). 56 162. Yamashita, M., Fenn, J. B., J. Phys. Chem, 88, 4451 (1984). 163. Aleksandrov, M. L., Gall, L. N., Krasnov, V. N., Nikolaev, V. I., Pavlenko, V. A., Shkurov, V. A., Dokl., Akad. Nauk. SSSR, 277, 379 (1984). 164. Aleksandrov, M. L., Gall, L. N., Krasnov, V. N., Nikolaev, V. I., Pavlenko, V. A., Shkurov, V. A., Baram, G. I., Grancher, M. A., Knorre, V. D., Kusner, Y. S., Bioorg. Kim, 10, 710 (1984). 165. Meng, C. K., Mann, M., Fenn, J. B. Z., Phys. D., 10, 361 (1988). 166. Gomez, A., Tang, K., Phys. Fluids, 6, 404 (1994). 167. Rayleigh, J. W., Philos. Mag, 14, 182 (1882). 168. Siuzdak, G., Proc. Natl. Acad. Sci. USA, 91, 11290 (1994). 169. Gale, D. C., Smith, R. D., Rapid Comm. Mass Spectrom., 7, 1017 (1993). 170. Smith, R. D., Wahl, J. H., Goodlett, D. R., Hofstadler, S. A., Anal. Chem, 65, A574 (1993). 171. Griffin, P. R., Coffinan, J. A., Hood, L. B., Yates, J. R, Int. J. Mass Spectrom. Ion. Proc., 111, 131 (1991). 172. Mann, M., Meng, C. K., Fenn, J. B., Anal. Chem, 61, 1702 (1989). 57 CHAPTER 2 A COMPARISON OF MALDI AND ESI FOR ANALYSIS OF IMINOTHIAZOLIDINE-BLOCKED PEPTIDES IN DISULFIDE MAPPING I. Introduction Cleavage at cysteine residues of peptide chains under alkaline conditions after cyanolysis of disulfide bonds was first observed by Catsimpoolas and Wood (1). However, because of the 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. (2) showed that 2-nitro-5- thiocyanobenzoic acid (NT CB) specifically cyanylates cysteine thiols under alkaline conditions. 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-iminothioazolidine (itz)-4—carboxyl peptides. Although the original paper claimed that the cleavage reaction can come to completion with only minor side reactions for most of the peptides and proteins tested, Digani and Patchornik (3) ‘ found B- elimination, occurring also under alkaline conditions, competes with the cleavage as an adverse reaction. 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 (3). Wu & Watson (4) conducted a systematic study to elucidate the effects of peptide structure and reaction conditions on the kinetics of the cleavage reactions of cyanylated 58 peptides and proteins and the yields of cleavage products. The optimal results were obtained in 1M ammonium hydroxide solution in which cleavage is complete within an hour at ambient temperature (4). Ammonia is a stronger small nucleophile but has smaller proton affinity than hydroxyl ion so that nucleophilic attack of ammonia on the carbonyl carbon facilitates the cleavage process giving an a-amidated N-terminal peptide but low capacity for catalyzing the B-elirnination reaction. Additionally, the excess volatile ammonia can easily be removed from the mixture alter cleavage, an advantage for subsequent analysis of the products by MALDI or ESI-MS (4). Other reagents for cyanylation of SH groups have been suggested (5-8). Among them l-cyano-4—dimethylamino-pyridinium (CDAP) tetrafluoroborate (8) has the advantage of the minimization of thiol/disulfide exchange because of the acidic condition of cyanylation reaction of sulfhydryl groups. Cyanylation and cleavage reactions of the cysteine residue have the capacity be used for its structural analysis. If a protein contains 11 cysteine residues, the cleavage reaction will result in the formation of n+1 peptide fragments; mass analysis of the fragments indicates the number and location of cysteine residues. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) has been used to mass map the cleavage products (9-28). However, peptide assignments using this approach are often complicated by its poor mass accuracy (error>5%). Papayannopoulos and Biemann (29) used CID tandem mass spectrometry to sequence the NTCB cleavage reaction products of a protease inhibitor isolated fi'om Sarcophaga bullata. Their work demonstrated that mass spectrometry could be used to sequence peptides from the NTCB cleavage reaction in spite of the blocked N-terminus. However, tandem MS is practically limited to low- 59 mass peptides produced from cysteine-rich proteins, because of the effective mass limit of CID (<3000 Da). Recently, Wu et al. (30) developed a methodology to recognize the number and location of both cystines and free sulfliydryls in peptides and proteins using the specific chemical cleavage reaction described above, followed by mass mapping of the resulting peptides by MALDI-MS. This approach provides the advantages of fast analysis, easy operation, high mass accuracy, and high sensitivity. Both electrospray ionization (ESI) and matrix-assisted laser desorption/ionization (MALDI) mass spectrometry allow for the analysis of large biomolecules through “mild” desorption and ionization methods (31, 32), each having unique capabilities, as well as some fundamental similarities. They are truly complementary in nature, almost a mirror image being projected between the disadvantages of one technique and the advantages of the other (33). During our original work, analyses of resulting mixtures of iminothiazolidine-blocked peptides from the cleavage of cyanylated protein by MALDI showed fi'equent, and sometimes complete, suppression of a signal for some fragments. Since such a phenomenon can compromise the structural analysis of cysteine residues by the methodology based on the cyanaylation chemistry, an alternative mass analysis technique should be used. ESI-MS will be a reasonable choice because of its capability of analysis of large molecules and its complementary nature for MALDI.' In this chapter, we investigated the use of ESI for the analysis of iminothiazolidine- blocked peptides when MALDI had the signal suppression. We used off-line HPLC to separate the mixture of iminothiazolidine-blocked peptides and then analyzed the HPLC fractions by ESI-MS. Two proteins, B-lactoglobulin A and ovalbumin, were used as 60 model proteins to investigate the complementary nature of MALDI and ESI in separate experiments. The results indicate that the responses of blocked peptides to ESI and MALDI are frequently complementary. II. Analytical Methodology A. Cyanylation and Chemical Cleavage Reaction Cyanylation of cysteine residue employs a one-step selective chemical reaction between free sulfliydryl and NTCB (3, 30, 34) or CDAP (8). Afier selectively cyanylating the sulfhydryls, the cysteinyl peptide bonds can be cleaved under alkaline condition to form an amino-terminal peptide and a series of iminothiazolidine-blocked peptides. The mechanism of cyanylation and cleavage reactions is illustrated in Figure 2.1. The side reaction is that the cyanylated cysteine undergoes a base-catalyzed B- elimination to form thiocyanate and dehydroalanine. This side reaction results in lower yield of cleavage. The cyanylation of sulfhydryl groups was traditionally accomplished by 2-nitro-5- thiocyanobenzoic acid (NT CB) 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 (3, 35, 36). Another reagent, 1-cyano-4—dimethylamino-pyridinium (CDAP) tetrafluoroborate, was proposed for the cyanylation of SH groups under slightly acidic conditions (pH 3-7) (5). CDAP is advantageous over NTCB for specific cyanylation of protein SH groups in the presence of disulfide bonds, because the acidic condition can effectively minimize sulfhydryl/disulfide exchange. Furthermore, the cyanylated proteins 61 SH O CH2 — — -NH-CH-CO — - cyanylation CDAP MezH/_ >1: CN pH3-7; rt, 10' N=C-§ ii ‘iH2 - -NH-CH-CO - — .NH3KO cleavage 1’“ M44OH B elimination ~1h, It 7° NH2 NH loo- - - -CO- NH-CI-CO- - (itz-peptide) (side reaction) Figure 2.1. Mechanisms of cyanylation and cleavage reactions 62 or peptides are stable under acidic conditions. While complete cyanylation by CDAP can be carried out using three to five-fold molar excess of the reagent over fi'ee SH groups under very mild conditions at room temperature (5, 37, 38), the optimal condition should be 10 to 20-fold molar excess of CDAP (over sulthydyl groups) in pH 3.0 buffer for 15 min at ambient temperature (4). A large excess of CDAP (~50 fold over SH group) and excessive incubation time (>2 hours) could result in the formation of an unidentified side reaction product (4). Therefore, like cyanylation by NTCB, the cyanylation by CDAP needs to be performed under controlled conditions to minimize the side reaction. The another characteristic of the reaction is that the resulting peptide fiagments may vary greatly in size because of the widely variable frequency of occurrence of cysteine in proteins. Large or small sized fiagments may be obtained. Both MALDI and ESI-MS can be used to measure the masses of fragments. While MALDI-MS usually has the suppression phenomenon at low mass range, ESI-MS can be used to determine the mass. Therefore, we can expect that the combination of MALDI-MS and ESI-MS can be promising for cysteine analysis. B. Localization of Cysteines and Cystines by ESI-MS The number and location of cysteines and cystines in a protein can be determined by a specific chemical cleavage process, which involves cyanylation of free sulthydryls by CDAP at pH 3 or NTCB at pH 8.7, then cleavage in 1M ammonia. Two steps can be used to locate cysteines and cystines (30), one is called NTCB or CDAP/TCEP (cyanylation/reduction) procedure which involves first cyanylating the free sulfhydryls in non-reducing proteins, then cleaving the cyanylated proteins following reduction of the remaining disulfide bonds. This step can recognize the location of free cysteines. Another 63 step is called TCEP/NTCB or CDAP (reduction/cyanylation) procedure, which first reduces all the disulfides by tris-2-carboxyethyl)phosphine (TCEP), then cyanylates the reduced protein, and then cleaves the peptide bond on the N-terminal side of cyanylated cysteine. This step can locate total cysteines. The location of cystines can be deduced by comparing the results of the two steps (scheme 2.1). III. Experimental Section ESI-MS ESI mass spectra were obtained on a Fisons VG Platform ESI mass spectrometer. This instrument is equipped with a single quadrupole mass analyzer. The flow rate was set at 10u1/min for infusion injection mode, and the solvent is 50:50 acetonitrile/P120 containing 1 % formic acid. The capillary voltage is set at 3.25KV, cone voltage at 37V, and source temperature at 100°C. MALDI-TOF MS MALDI mass spectra were obtained on a Voyager Elite time-of-flight (TOF) mass spectrometer (PerSeptive Biosystems Inc., Framingham, MA) equipped with delayed extraction and a model VSL-337ND nitrogen laser (Laser Science, Newton, MA). The accelerating voltage in the ion source was set to 20 KV. Grid voltage and guide wire voltages were 93.6% and 0.2% of the accelerating voltage, respectively. Data were acquired in the positive linear DE mode of operation. Time-to-mass conversion was achieved by external and/or internal calibration using standards of bradykinin (m/z 3:sz bum $5033 “.o c0280. 9: m£~Emooom .FN oEmcow .8an .000 mm\ Jug 31a omm>mm_o w 71%|) $sz 2\.\/ lea Siege \l: am: am to 5.2.2 .92 a ..fz 2 couozcom m0:_0«w>0 Amy .38 £28. I 08 QTY 255.3 \lw am: o2“. bacon. As +fz o /\ m_m>_mc< 398.0 w .mm .o 5.25. co_am_>cm>o .000\ ong -0001“: mfilzwjomg A 2.\l| 8N: \lije cowosnom 03} ..fz 65 1061.2), bovine pancreatic insulin (m/z 5734.5), and horse skeletal myoglobin (m/z 16952) obtained from Sigma Chemical Co. (St. Louis, MO). All experiments were performed using a-cyano-4-hydroxycinnamic acid (Aldrich Chemical Co., Milwaukee, WI) as the matrix. Two types of saturated matrix solutions were prepared. Type A contains 70% (v/v) solution of acetonitrile/aqueous 1% TFA. Type B contains 50% (v/v) solution of acetonitrile/aqueous 1% TFA. The protein or peptide samples were applied with a sandwich mode to a stainless-steel sample plate: the first layer was 0.7 ul matrix Solution A, second layer was 0.6 ul B, the third layer was 1 ul sample, and the last layer was 0.5 ul matrix solution A again. The sample sandwich was allowed to air dry before being introduced into the mass spectrometer. HPLC HPLC is equipped with Waters model 6000 pumps controlled by a PC computer. UV detection was at 215nm. The column was a Vydac C 18 (#218TP54, 10pm particle size, 300- pore, 4.6x250mm). Chemicals Tris(2-carboxyethyl)phosphine hydrochloride (TCEP) was purchased fi'om Pierce, Rockford, IL. Guanidine hydrochloride was a product of Boehringer-Mannheim Biochemicals (Indianapolis, IN). Proteins, 2-nitro-5-thiocyanobenzoic acid (NT CB), 1- cyano-4-dimethylamino-pyridinium tetrafluoroborate (CDAP), were purchased from Sigma and used without further purification. Acetonitrile and TFA were HPLC grade. The TCEP solution in 0.1 M citrate buffer at pH 3 or in 0.1 M tris-HCl buffer at pH 8.2 66 was prepared as 0.1 M stock solution and stored under N2 at —20°C for weeks with little deterioration. The 0.01 M solution of NTCB was prepared in 0.1 M tris-HCL buffer and the pH of the solution was adjust to pH 8.7. The 0.1 M CDAP solution in pH 3.0, 0.1 M citrate buffer was freshly prepared prior to use. Analysis of Free Sulfhydryl Groups in Proteins (CDAP or NTCB/TCEP Procedure) About 10nmol protein under study is dissolved in lOpl 0.1N tris-HCl buffer (pH 8.0) for NTCB reaction or in 0.1N citrate buffer (pH 3.0) for CDAP reaction, containing 6N guanidine-HC] as denaturing reagent. To the solution, add 5-10 fold molar excess of NTCB or 10-20 fold CDAP (over the free sulfliydryls). Allow the reaction to proceed at 37°C in a water bath for 30min. Adjust pH to 12 by 1N NH3, let it stay in room temperature for 1 hour. Add 5-10 fold molar excess TCEP (over the disulfide bond). The mixture is put in a water bath at 37°C for 30min to reduce the disulfide bond. The masses of the cleavage fragments were analyzed by MALDI-MS. Comparatively, the mixture of the cleavage product was separated by reversed-phase HPLC, the peaks were collected from HPLC and analyzed by ESI-MS. Confirmation of Total Cysteines (TCEP/CDAP or NTCB Procedure) The protein solution is allowed to react at pH 8.7 with 5-10 fold NTCB (over free sulflrydryl groups) or at pH 3.0 with 10-20 fold CDAP (over free sulfhydrl groups) at 37°C for 30min after it is reduced with 5-10 fold TCEP (over disulfide bonds) at 37°C for 30min, and then adjust pH to 12 by 1N NH3, following reaction at room temperature for 1 hours, analyzed by MALDI-MS or off-line HPLC/ESI-MS. 67 IV. Results and Discussion A. Recognizing the Location of Cysteine and Cystines of B-Lactoglobulin A B—lactoglobulin A has 162 amino acids and molecular weight 18368 Da. It contains two disulfide bonds and one free cysteine at Cys 121 (Figure 2.2). The expected masses of cleavage peptide fragments afier two procedures are shown in Table 2.1. Figure 2.2. Structure of B-Lactoglobulin A The expected peaks for the fragments 1-65(m/z 7247.4), itle6-118(m/z 1461.6), itzl 19-120(m/z 274.3), and itzl60-162(m/z 396.5) for the procedure reduction/cyanylation (TCEP/NT CB) are undetected in the MALDI spectrum (Figure 2.3 A) which clearly indicates the signal suppression in MALDI-MS. The abbreviation “itz” represents “iminothiazolidine” derivative at the specified residue. Only two peaks of itz66-105 and it2121-159 were identified by MALDI-MS. Because these two peptide fragments indicate the cyanylation and cleavage only at Cys 66 and Cys 121, the numbers and location of other cysteine residues couldn’t be recognized. Therefore, MALDI-MS gave the incomplete mass mapping of cysteine residues of B-lactoglobulin A. 68 Table 2.1. Expected and Observed Masses of Cleavage Product of B-Lactoglobulin A Reduction/Cyanylation Fragments Expected Mass (Da) ESI(Da) MALDI(Da) 1-65 7247.4 7243.4 no itz66-105 4651.6 4651.1 4653.5 itle6-l 18 1461.6 no no itzl 19-120 274.3 no no itz121-159 4551.2 4550.4 4548.1 it2160-162 396.5 397.0 no *1-105 11822.0 11823.0 no itz*121-162 4870.7 4870.7 no Cyanylation/Reduction Fragments Expected Mass(Da) ESI(Da) MALDI(Da) 1-120 13506.7 no no it2121-162 4905.7 4904.7 4901.1 * B-elimination itz = iminothiazolidine derivative. 69 Instead of the expected two peaks, one peak at m/z 4901.1 (1-120) was identified by MALDI-MS for the cleavage product resulting from the procedure cyanylation/reduction (NT CB/CDAP) (Figure2.3B). Based on the above results, we used off-line HPLC separation/ESI-MS to analyze the cleavage products. The HPLC chromatogram of the mixture of cleavage products resulting from procedure TCEP/CDAP of B-lactoglobulin A is shown in Figure 2.4. The ESI spectra corresponding to the HPLC peaks in Figure 2.4 are illustrated in Figure 2.5. ESI-MS identified peptide fragments of 1-65, itz66-105, it2121-159, and itzl60-162, indicating that the location of cysteine residues should be at Cys 66, Cys 106, Cys 121, and Cys 160. The location of one cysteine residue at Cys 119 cannot be recognized because of two unidentified fragment peaks of itle6-ll8 and itzll9-120, which is probably due to the difficulty of cleaving the peptide bond at the N-terminal side of cysteine residue at Cys119 because of the closeness of two cysteines at Cys 119 and Cys 121. The HPLC chromatogram and corresponding ESI spectra of the cleavage product resulting fiom procedure CDAP/TCEP of B-lactoglobulin A is shown in Figure 2.6 and Figure 2.7, respectively. One fragment of itz121-162 was identified, indicating that the cleavage occurred at Cys 121, thus Cys 121 involves in the pairing of the disulfide bond. However, the cysteine residue paired with Cys 121 cannot be located because of the unidentified fi’agment of 1-120. Comparing the results between MALDI and ESI of B-lactoglobulin A, we can conclude: 70 A TCEP/NTCB .g itz121-159 _ a; \ /It266-105 '0 cu 66 E 3 m _=_ 1...-__-_L. A. J‘M i... A ' 2000 4000 6000 8000 B itz121-162 NTCB/TCEP a. ,1 -- g, 4 ' 4000 8000 12000 16000 mlz Figure 2.3. MALDI spectra of cleavage products of B-lactoglobulin A following reduction/cyanylationlcleavage (panel A) or cyanylationlcleavagelreduction (panel B). itz = iminothiazolidine derivative. 71 .558 s .85. m sense a sense .5: s 3 9.35:8 e_E_e28e $8 ”m 29.8 uamt. e\e F .0 H< cozow .n_o S mueziwoe 2880.5 0532.8 8:85 898.0 00 8538 00 88008525 ohm: .ofim 050E lob—dude 8.3. 8.3 86 p — — F — r— 1: oo.oom 3083 M 89858. I 8.88 «83 H m _ 1 a l 8.88 m i 11 oo.ooo« mzfia\ m.— g u 80 +3 100 790 1011 1-10 ‘°° 11211-29 10110:: 237mm ,9 % % 13 846 s +2 111 II 1190 o 900 1000 1100 D” 0 000 ‘1000 '1mvaa/e +5 +5 100 95“ itz30-72 1m “”5 itz73-119 4760.209 5426.9oa +6 905 % 11. +6 :1 i .. 08001000121000. o1100100012001400Dale +4211 100 1°15 lntact(A)+ilz120-366(B) 5:; itz367-381 +41A +2313 W275810a 100 1713306 +2 959 v. 409 .11: o 500 700 900 110043“9 Figure 2.11. ESI spectra of cleavage products following procedure TCEP/NTCB of ovalbumin 81 +3 +12 . m . 100 844 Itz*30-119 100 1 Itz*367-385 i 10108.30a m5); 1 1 +9 i 1124 l 9‘ 4 % ‘ I . +2 134 $7 1034 l 794 fl 0 Dale 0 ‘ 900 1100 1!!) 500 700 91) 1100 26 10° ‘ 400 4.. 0 1m .. +29 [12. 1tz*120-381 “2332-385 1 1009 291mm 4m % 412 452 I 468 o "7' 73.2.:1 LATATA: We 8; 0 ‘ 1045 111!) 1211 Figure 2.11. Continue 82 .ofiu 0.59”. 5 8888 mm 088 05 em 80288 E285 new Eczow £23.96 a $0882 9:885 ea 0532.2 3885 31.58.20 232E do ESmBmEoEo 04d... .mfiw 059“. laps 00.0N _ r 3 000.33 00.0 00.0 1. 8.8» a u I .n I. 00.000." I 00.003 83 553.95 so awotmopz 2:885 8. 9.52.2 32.85 898.0 .6 808» .mm new 2:9... RNF 90. Ru 2!. RN. Rx... R0 030 _ooio§p§p8=§8~0§8$8vaowvao v9 _ $9 v+ «3 0» 8m .x. a 0+ 8 3:. 0+ 8.8% 88$... 888. 884 0+ 3. 8. 30 . 8p . 8.. 0... NF... e8 o88v.§.89o8oe8o§ o§ 83:88:05.8 a one 8.. ..x 82 8m .x. 8: «8.x. 8 ..x N. (8+ <91 838.8884 80.3: 3500-005? 8.8M. Rafi 8:2 /§ 8.33 Asses. 8.28 e. .8. o: =2 is. «B 8. as. .8. 0... En... Because ovalbumin only contains one disulfide bond, the linkage of disulfide bond must be between Cys 73 and Cys 120. Comparing the results of MALDI and ESI, several aspects could be concluded: I). All the undetected peaks by MALDI-MS were identified by HPLC/ESI-MS (summarized in Table 2.2 ). Therefore, HPLC / ESI-MS gave the complete mass mapping for ovalbumin. 2). Peak of ESI at m/z 37684.7 in procedure NTCB/TCEP and peak of ESI at m/z 27581 in procedure TCEP/NT CB are very noisy which are due to the broad HPLC peaks that include the intact ovalbumin. 3). Compare the two HPLC chromatograms for procedure TCEP/NT CB and NTCB/TCEP. The procedure NTCB/'1‘ CEP also forms the peaks for fi'agments it230-72 and itz73-119 at m/z 4760.8 and m/z 5426.2 which should not appear in the procedure of NTCB/'1‘ CEP, although they have lower intensity. Therefore, the cyanylation reaction with NTCB at pH 8.7 induced the thiol-disulfide exchange of ovalbumin. C. Evaluation of CDAP and NTCB Based on the results of ovalbumin in Section B, NTCB induced the thiol/disulfide exchange of ovalbumin in procedure of cyanylation/reduction. In order to evaluate the performance of NTCB and CDAP. We used CDAP to perform the same procedure of CDAP/TCEP as that of NTCB described in section B. The HPLC chromatograms and corresponding ESI spectra of the cleavage product from procedure CDAP/TCEP of ‘ ovalbumin are illustrated in Figure 2.14 and Figure 2.15, respectively. 85 .05 8:9... 5 contomou mm 058 05 2o; 282260 £0.36 new Eczow 5:50.90 do emote/Bo 23809.8 Eat 8:33 mug—No.0 yo «SEE *0 £832.95 0:: .EN 239". O O I O www-monng O O I O 86 oo.ood oo.ooN m oo.oon oo.oo# mwnémflri. ‘11IIIIIIII1TIIIIIIIIIIT—I . +36A . l "92 |ntact(A)+itz30-366(B) 10° Itz11-29 ‘ +368 2377.60a 1045 1212 W3768408 96 733 Q 2 i it 49° 0 J ‘L we a” 1000 1200 +3 52 _ 100 ‘ 100 ItZ367-381 “0 itz382-385 1714.503 m % % 439 $1; 4& m8 412 J a 0 We 0 . J ..l. 41.. IA IL think-Dale 5(1) 700 900 1100 420440460480 Figure 2.15. ESI spectra of cleavage product from procedure CDAP/TCEP of ovalbumin 87 The results are summarized in Table 2.3. Several aspects can be concluded: 1). No peaks for fragments itz30-72 and itz73-ll9 at m/z 4762.5 and m/z 5427.1 resulting from the procedure CDAP/TCEP were identified. Therefore, CDAP reaction did not induce the thiol-disulfide exchange. 2). CDAP has the lower capacity to cyanylate free cysteines. The fragment 1-10 of ovalbumin cannot be detected by HPLC by using 20-fold excess of CDAP. 3). By comparing the results from NTCB and CDAP, we can obtain clear assignment for cysteines and cystines in ovalbumin. C. Conclusions 0 The use of NTCB or CDAP is unique in that it specifically targets the site being analyzed. 0 Since CDAP reaction is performed at pH 3, it eliminates the thiol-disulfide exchange in the analysis. 0 HPLC separation / ESI-MS overcomes the signal suppression phenomenon of MALDI- MS; it can become a complementary method to MALDI-MS. The ESI method can be firrther used with on-line LC/MS. 88 Table 2.3. Comparison of CDAP and NTCB Fragments CDAP/TCEP(Mass: Da) NTCB/TCEPOVIass: Da) 1-10 no 101 1.0 itzl 1-29 2377.6 2377.1 itz30-366 37684.0 37684.0 it2367-381 1714.5 1714.0 itz382-385 430.0 430.0 #itz‘30-1 19 no 10112.6 #itz73-l 19 no 5426.2 #itz30-72 no 4760.8 #: fragments due to thiol—disulfide exchange. "': B-elimination itz = iminothiazolidine derivative. 89 References: I—l . Catsimpoolas, N., and Wood, J. L., J. Biol. Chem, 241, 1790 (1966). 2. Jacobson, G. R., Schaffer, M. H., Stark, G. R., and Vanaman, T. C., J. Biol. Chem, 248,6583(1973) 3. Degani, Y., and Patchornik, A., Biochemistry, 13, 1 (1974). 4. Wu, J ., and Watson, J. T., to appear in Anal. Biochem. (1998). 5. Wakeselman, M., Guibe-Jarnpel, E., Raoult, A., and Busse, W. D., J. Chem. Soc. Chem. Commun., 1, 21(1976). 6. Brockleehurst, 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). 7. Brocklehurst, K., Malthouse, J. P. G., Biochem. J., 175, 761 (1978). 8. Walkselman, M., and Guibe-Jampel, B., J. C. S. Chem. Comm, 21 (1976). 9. Casey, R., and Lang, A., Biochem. J., 145, 251 (1975). 10. Price, N. C., Biochem. J., 159, 177 (1976). 11. Schaffer, M. H., and Stark, G. R., Biochem. Biophys. Res. Comm, 71, 1040 (1976). 12. Lu, H. S., and Gracy, R. W., Arch. Biochem. Biophys., 212 (2), 347 (1981). 13. Thomas, M. L., Janatova, J ., Gray, W. R., and Tack, B. F., Proc. Natl. Acad. Sci. USA, 79, 1054 (1982). 14. Roberts, D. D., and Goldstein, I. J ., J. Biol. Chem, 259, 909 (1984). 15. Matsudaira, P., Jakes, R., Cameron, L., and Atherton, B., Proc. Natl. Acad. Sci. USA, 82, 6788 (1985). 16. Mahboub, S., Richard, C., Delacourte, A., and Han, K., Anal. Biochem., 154, 171 (1986) 90 17. Bahler, M., Benfenati, F ., Valtorta, F., Czemik, A. J ., and Greengard, P. J., Cell Biol., 108, 1841 (1989). 18. Minard, P., Desmadril, M., Ballery, N., Perahia, D., and Mouawad, L., Eur. J. Biochem., 185, 419 (1989). 19. Nefsky, B., and Bretscher, A., Proc. Natl. Acad. Sci. USA, 86, 3549 (1989). 20. Eshdat, Y., Chapot, M. P., and Strosberg, A. D., FEBS Letters, 1-2, 166 (1989). 21. Katayama, B., J. Biochem., 106, 988 (1989). 22. Sutherland, C., and Walsh, M. P., J. Biol. Chem, 264, 578 (1989). 23. Tang, C., Yuksel, K. U., Jacobson, T. M., and Gracy, R. W., Arch. Biochem. Phys., 283(1), 12 (1990). 24. May, J. M., Buchs, A., and Carter-Sn, 0, Biochem. J., 29, 10393 (1990). 25. Wines, B. D., and Easterbrook-Smith, S. B., Molecular Immunology, 28 (8), 855 (1991). 26. Som, S., and Friedman, S., J. Biol. Chem, 266, 2937 (1991). 27 . Altamirano, M. M., Plumbridge, J. A., and Calcagno, M. L., Biochem. J., 1153 (1992). 28. Goold, R., and Baines, A. J ., Eur. J. Biochem., 224, 229 (1994). 29. Papayannopoulos, I. A., and Biemann, K., Protein Sci., 1, 278 (1992). 30. Wu, J., Gage, D. A., Watson, J. T., Anal. Biochem., 235, 161 (1996). 31. Karas, M., and Hillenkamp, F ., Anal. Chem, 60, 2299 (1988). 32. Fenn, J. B., Mann, M., Meng, C. K., Wong, S. F., and Whitehouse, C. M., Science, 246, 64 (1989). 91 33. Muddiman, D. C., Bakhtiar, R., Hofstadler, S., and Smith, R. D., J. Chem. Edu., 74, 1288(1997) 34. Stark, G. R., Methods in Enzymology, vol. 47, 129 (1977). 35. Denslow, N. D., and Nguyen, H. P., in Techniques in Protein Chemistry VII, Academic Press, Inc., (1996), p.241-248. 36. Price, N. C., Biochem. J., 159, 177 (1976). 37. Nakagawa, S., Tamakashi, Y., Harnana, T., Kawase, M., Taketomi, S., Ishibashi, Y., Nishimura, 0., Fukuda, T., J. Am. Chem. Sac., 116, 5513 (1994). 38. Wu, J., and Watson, J. T., Protein Sci., 6, 391 (1997). 92 CHAPTER 3 DISULFIDE MASS MAPPING IN PROTEINS CONTAINING ADJACENT CYSTEINES WITH CYANYLATION/CLEAVAGE METHODOLOGY 1. Introduction Disulfide bond formation between cysteine residues is a posttranslational event that commonly occurs in proteins synthesized in the endoplasmic reticulum. In many cases, disulfide bond formation contributes to the stability of the tertiary structure of the folded protein molecule (1 ). A full description of the covalent structure of proteins demands that the connectivity of the bridged cysteines be analyzed. 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 be a challenging task for the protein chemists. The general strategy for locating disulfide bonds in proteins by conventional methods involves several steps (2, 3). First, a protein is cleaved by enzymes or chemical reagents between half-cystinyl residues to obtain peptides that contain only one disulfide bend. Second, the amino acid composition, amino terminal sequence or molecular masses of these peptides are determined using an amino acid analyzer, manual or automated Edman degradation, and/or mass spectrometry. Finally, peptides identified using these data are related to specific segments of the protein (3). There are two major problems are 93 related to the methodology. First, if a protein contains a pair of adjacent or closely spaced cysteines in its primary structure, it is impossible to obtain peptides containing a single disulfide bond (3). Second, the conditions most frequently used in specific enzymatic digestions may promote disulfide bond scrambling (4). Although supplementary methodologies such as non-specific fragmentation by partial acid hydrolysis (2) have been proposed to avoid disulfide bond scrambling, it is difficult to deal with the data obtained by these non-specific techniques. Difl‘erent approaches have been made to assign the disulfide linkages of proteins containing adjacent or closely spaced cysteine residues (5-17). Among them, the methodology proposed by Gray (16, 17) has the capacity to analyze proteins containing adjacent or closely spaced cysteine residues. In his experiments, peptides were partially reduced, the nascent free thiols alkylated, and their positions recognized from the results of sequence analysis. However, it is obviously tedious, if not impractical, to sequence an alkylated high-mass peptide or protein using this approach. Recently, Wu and Watson described a novel approach for the assignment of disulfide bonds in proteins of known sequence (18). In this approach, the denatured protein was subjected to limited reduction by tris(2-carboxyethyl)phosphine (TCEP) in pH 3.0 citrate buffer to produce a mixture of partially reduced protein isomers; the nascent sulthydryls were immediately cyanylated by 1-cyano-4-dimethylamino- pyridinium tetrafluoroborate (CDAP) under the same buffered conditions. The cyanylated protein isomers, separated by and collected from reversed-phase HPLC, were subjected to cleavage of the peptide bonds on the N-tenninal side of cyanylated cysteines in aqueous ammonia to form truncated peptides that were still linked by residual disulfide 94 bonds. The remaining disulfide bonds were then completely reduced to give a mixture of peptides that can be mass mapped by MALDI-MS. The masses of the resulting peptide fragments were related to the location of the paired cysteines that had undergone reduction, cyanylation, and cleavage. This strategy minimizes the disulfide bond scrambling and is simple, fast, and sensitive. Furthermore, it is possible to assign the disulfide linkages of proteins containing adjacent cysteines by this new methodology. In this chapter, the applicability of partial reduction, cyanylation, chemical cleavage, and mass mapping approach to the disulfide structure analysis of proteins containing adjacent cysteine residues will be demonstrated. Two model proteins, long R3 insulin-like growth factor-1 and insulin-like growth factor-I, will be used to show the capability of the new approach for the determination of disulfide structure involved in adjacent cysteine residues. II. Methodology for Assignment of Disulfide Linkages The methodology utilized to determine disulfide linkages of proteins containing adjacent cysteines are based on the strategy proposed by Wu & Watson (18). The approach is illustrated in scheme 3.1. The disulfide bonds of a denatured protein are partially reduced with TCEP by controlled TCEP concentration and reducing time in a pH 3.0 buffer. Then the resulting nascent sulfl'rydryls will be cyanylated by CDAP in the same buffer solution. After the separation of partially reduced/cyanylated protein isomers, HPLC fractions can be analyzed by MALDI-MS or ESI-MS to determine which isomers are singly reduced and cyanylated. Those shifted by +52 Da correspond to a singly reduced/cyanylated species; +104 Da correspond to doubly reduced/cyanylated 95 Denatured Protein partial TCEP, pH 3, reduction 10', rt Partially Reduced Protein Isoforms cyanylation CDAP. pH 3, 610', rt Cyanylated Protein Isoforms HPLC A Separated Protein lsomer —> Identified by MS 1N NH4OH, 1hr, rt cleavage Cleavage Products reduction TCEP Mass Spectrometry Scheme 3.1. Conceptual overview of the methodology for assignment of disulfide bond pairings in proteins 96 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 either MALDI-MS or LC/ESI- MS. The masses of the resulting peptide fragments are related to the location of the paired cysteines that had undergone reduction, cyanylation, and cleavage. III. Experimental Section MALDI-TOF MS MALDI mass spectra were obtained on a Voyager Elite time-of-flight (TOF) mass spectrometer (PerSeptive Biosystems Inc., Framingham, MA) equipped with delayed extraction and a model VSL-337ND nitrogen laser (Laser Science, Newton, MA). The accelerating voltage in the ion source was set to 20 kV. Grid voltage and guide wire voltages were 93.6% and 0.2% of the accelerating voltage, respectively. Data were acquired in the positive linear DE 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 16952) obtained from Sigma Chemical Co. (St. Louis, MO). All experiments were performed using a-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 1% TFA, and mixed in equal volumes with peptide or protein 97 samples, and applied to a stainless-steel sample plate. The mixture was allowed to air dry before being introduced into the mass spectrometer. LC/ESI-MS The LC system used for LC/ESI-MS analysis was Perkin-Elmer API Applied Biosystems 173. The separation was carried out on a C18 capillary column. Solvent A was 0.085% TFA in water and solvent B was 0.085% TFA in CH3CN. The linear gradient was 5% B to 65% B in 135 minutes at a flow rate of 7 ul/min. ESI mass spectra were obtained on-line on a Micromass Platform-LC mass spectrometer. This instrument was equipped with a single quadrupole mass analyzer. The capillary voltage was set at 3.25KV, cone voltage at 30V, and source temperature at 80°C. HPLC The HPLC is equipped with Waters model 6000 pumps controlled by a personal computer. UV detection was at 215nm. The column was a Vydac C18 (#218TP54, lOum particle size, 300-A pore, 4.6x250mm). Chemicals Tris(2-carboxyethyl)phosphine hydrochloride (TCEP) was purchased from Pierce, Rockford, IL. Guanidine hydrochloride was a product of Boehringer-Mannheim Biochemicals (Indianapolis, IN). Human recombinant long R3 insulin-like growth factor I (LR3IGF-I), 2-nitro-5—thiocyanobenzoic acid (NT CB), l-cyano-4-dimethylamino- pyridinium tetrafluoroborate (CDAP), were purchased from Sigma. Human recombinant 98 insulin-like growth factor I was obtained from Austral Biologicals (San Ramon, CA). The proteins were purified before using. Acetonitrile and TFA were HPLC grade. The TCEP solution in 0.1 M citrate buffer at pH 3 was prepared as 0.1 M stock solution and stored under N2 at —20°C for weeks with little deterioration. The 0.1 M CDAP solution in pH 3.0, 0.1M citrate buffer was freshly prepared prior to use. Partial Reduction of Proteins 10nm01 Protein samples (IGF-I and LR3IGF-I) were solubilized in 4on1 of 0.1M citrate buffer (pH 3.0) containing 6 M guanidine-HCI. Partial reduction of disulfide bonds were carried out by adding 10 molar fold of TCEP for the cystine content of the proteins (300nmol of TCEP is reacted with lOnmol of IGF-I or LR3IGF-I as lnmol of IGF-I or LR3IGF-I contains 3nmol of cystine). The reduction was conducted at room temperature for 10 minutes. Cyanylation of Nascent sulflrydryls To the partially reduced protein mixture was added a 30—fold excess of CDAP over the cysteine content of IGF-I or LR3IGF-I. Cyanylation of the nascent sulflrydryl groups was accomplished by incubation at room temperature for another 10 minutes. HPLC Separation of Partially Reduced and Cyanylated Protein Isomers Partially reduced and cyanylated species were separated by reversed-phase HPLC _with linear gradient of 30% B to 50% B in 45 minutes, where solvent A was 10% CH3CN containing 0.1% TFA and solvent B was 90% CH3CN containing 0.1% TFA. The 99 predominant HPLC fractions were collected manually and the masses of the collected protein isomers were determined by MALDI-MS or ESI-MS. Appropriate fiactions were then dried for further use. Cleavage of Singly Reduced and Cyanylated Protein Isomers To the dried HPLC fractions was added 2ul of 6M guanidine-HCI in 1M NH40H aqueous solution to dissolve the protein residue and then 5 ul of 1M NILOH. Cleavage of the peptide chain was performed at room temperature for one 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 2 ul of 0.1M TCEP solution at pH 3.0, 0.1M citrate buffer at 37°C for 30 minutes. Samples were diluted with 100 pl of a 50% (v/v) CH3CN/ 1% TFA for MALDI-MS analysis or 100 pl of 5% (v/v) CH3CN/0.085% TFA for LC/ESI-MS analysis. IV. Results and Discussion A. Disulfide Mapping of IGF-I Insulin-like growth factor I (IGF-I) (Mr = 7648.6) is a single-chain polypeptide of 70 amino acid residues containing three intramolecular disulfide bonds, two of which involve adjacent cysteines (19) (Figure 3.1). IGF-I is postulated to be the mediator of 100 @0 0 $33 96 <9 0 o o e ‘9 e e e g .9er o o is 479152 ”e 8 8 69 4863 me Q Q © 6 G) 09 61 96 60 1a Figure 3.1. Amino acid sequence and disulfide structure of lGF-l 101 growth hormone action on skeletal tissue as well as mitogenic activity on several cell types (20, 21). While the complete amino acid sequence of IGF-I was determined in 1978 (19), very little experimental data is available describing the secondary and tertiary structure of the molecule. A three-dimensional model of human IGF-I has been suggested first by Blundell et al in 197 8 (22). This model is based on the proposed tertiary structure of porcine insulin as determined by X-ray crystallography (23). From this model, a disulfide bond arrangement between Cys6-Cys48, Cys47-Cy352 and Cysl8-Cys6l was predicted for human IGF-I. However, since IGF-I contains three disulfide bonds, several isomeric forms can theoretically be formed, especially in the region with the two adjacent cysteine residues, i.e., Cys47 and Cys48. Several approaches have been used to identify the disulfide linkages of IGF-I (24- 27). While the conventional method of enzymatic digestion did not completely assign the disulfide linkages because of the two adjacent cysteines (24), chemical synthesis and the combination of enzymatic digestion and fast-bombardment mass spectrometry have been utilized to try to assign the linkages. The method based on the chemical synthesis determined disulfide linkages of IGF-I by comparing the retention time during HPLC chromatography of the enzymatic digested peptide fragments with these of chemically synthesized peptides. Two types of peptides related to two adjacent cysteines, Type I with Cys6-Cys47 and Cys48-Cys52 as well as Type H with Cys6-Cys48 and Cys47- Cy552, have been synthesized (25). The disulfide bond system of IGF-I was determined to be the Type 11 form. This method is tedious and only provided indirect experimental evidences. The other approaches utilized a combination of multiple steps of enzymatic digestion, Edman degradation, and mass spectrometry (26, 27). Raschdorf et al. assigned 102 the disulfide linkages of IGF-I by a three-step mass spectrometric analysis (26). Firstly, the correct molecular weight of the intact protein was determined by fast atom bombardment (FAB) mass spectrometric analysis. Secondly, two-fold enzymatic degradation (chymotrypsin followed by V3 protease, FAB mapping of the cleavage products) was employed to assign the disulfide linkage of CyslS-Cys6l. Thirdly, FAB tandem mass spectrometry and manual Edman degradation were further used to analyze the peptide fragment, which contained two other disulfide bonds and two adjacent cysteines. Axelsson et al. (27) utilized the combination of enzymatic digestion, reversed- phase HPLC, FAB mass spectrometry, and one-cycle Edman degradation. They first digested the protein with V3 protease, then analyzed it by FAB-MS to determine the disulfide linkage unrelated to adjacent cysteine residues. The peptide fiagment involved in adjacent cysteine residues was further degraded by one-cycle Edman degradation, followed by trypsin digestion and FAB-MS analysis. The methodology consisting of multiple steps of enzymatic digestion, Edman degradation, and mass spectrometry is tedious and time consuming. The disulfide linkages of IGF-I can be determined unambiguously by the new approach based on the partial reduction, cyanylation, chemical cleavage, and mass spectrometry (18). The overview of the methodology used to assign the disulfide linkages of IGF-I is illustrated in Figure 3.2. Partial Reduction and Cyanylation of I GF-I The partial reduction of IGF-I was carried out in citrate buffer (pH 3.0) containing 6 M Guanidine-HCl. Guanidine-HCl was utilized to completely denature the protein. 103 1II r—I—il 70 6 18 47 48 52 61 Partial Reduction TCEP, pH=3 3" i“ 1 I l—l—l n 70 18 47 48 52 61 (+other isomers) Cyanylation CDAP, pH=3 SCN lSCN 1 I I I I 1 70 6 18 47 48 52 61 (+other isomers) HPLC Separation Chemical Cleavage 1 M NH30H 1—5 G6 1| IQ‘IBI I 70 8 47 52 61 Reduction TCEP SH I Ind I” 6 8 i 1_5 G 13 47 52 61 7° Figure 3.2. Overview of the methodology used to assign the disulfide linkages in lGF-l 104 Tris(2-carboxyethyl)phosphine hydrochloride (TCEP) was used as a reducing reagent since it has proved to be an excellent reducing agent for disulfide bonds (12, 17, 18, 28). 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 (17 ). After the partial reduction of the protein, the nascent sulfhydryls can be cyanylated by CDAP in the same citrate buffer (pH 3.0). There are several advantages of using CDAP as the cyanylation reagent (18, 28). Firstly, cyanylation by CDAP can be carried out at pH 3.0, which is compatible with the partial reduction by TCEP. It is therefore unnecessary to remove excess TCEP, change buffer, or readjust the pH prior to using CDAP. Secondly, CDAP reacts instantly with TCEP even at pH 3.0, and the excess of TCEP in the reaction system will be consumed upon addition of CDAP, and no further reduction during cyanylation is possible. This is definitely an advantage for the control of reduction in an unknown protein. While a larger amount of CDAP has to be applied in case more TCEP is required for partial reduction of proteins with a tight structure, a large excess of CDAP (>50 fold molar excess over peptide sulflrydryls) and excessive incubation time (> 2 hours at room temperature) will result in some side reactions. Finally, CDAP is water soluble, which is compatible to the reaction system. Because IGF-I consists of three disulfide bonds, use of the partial reduction technique should result in the production of three isoforrns, each of which consists of a singly reduced isomer emanating from a difi‘erent disulfide bond. However, as shown in the chromatogram (Figure 3.3) obtained after partial reduction and cyanylation, besides a major peak for the intact unreduced protein there are only two discernible peaks (peak 1 105 Intact _J - 1'5 2'5 35 45 minutes Figure 3.3. HPLC separation of denatured IGF-I 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 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. Peak 3 represent doubly reduced/cyanylated species, as determined by MALDI-TOP analysis. 106 V and peak 2) each of which represents a singly reduced isofonn as determined by mass analysis showing that the molecular weight of each species shifted by 52 Da from that of the intact protein (a 2-Da shift for conversion of a cystine to two cysteines followed by a 50-Da shift for the replacement of a hydrogen on each of two free sulflrydryls with a cyano group). Peak 3 resulted from the reduction of two disulfide bonds as indicated by a mass shift of 104 Da from the intact protein. This result is not unexpected, as the some disulfide bonds are much more stable than others as indicated in a recent study of the stability of the disulfide bond of Cysl8 to Cys6l in IGF-I (29). This CyslS-Cys61 disulfide bond is preferentially formed in the folding process, a feature that may explain its resistance to reduction under our reaction conditions. In any case, in dealing with a protein known to contain three disulfide bonds, the determination of the pairing of two of the disulfide bonds allows the remaining one to be determined by default. Cleavage of Peptide Chains After the reversed-phase HPLC separation of partial reduced/cyanylated protein isomers, the peptide bonds at N-terrninal side of cyanylated cysteine residues can be cleaved in 1N NHiOH at room temperature for 1 hour in the presence of guanidine-HCl (18). These modified cleavage conditions have the advantage over the conventional cleavage method by using the mildly alkaline conditions (pH~9) (18, 30). Ammonia can be easily removed from the reaction system and the cleavage reaction is fast (in one hour). 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, 107 the peptide mixture is diluted to minimize the adverse effect of guanidine on the subsequent analysis by MALDI-MS. Interpretation of MALDI Data Because cleavage of the peptide chain takes place only at cyanylated cysteinyl sites, there will be three fi'agments for each singly reduced/cyanylated protein isomer (and sometimes two overlapped fragments corresponding to 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. Table 3.1 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 3.4A and B are two MALDI spectra of peptide mixtures resulting fi'om cleavage of isomers of singly reduced/cyanylated IGF-I corresponding to HPLC peaks 1-2 in Figure 3.3, respectively. HPLC peak 3 is due to the reduction of disulfide bonds of Cys47-Cy852 and Cys6-Cys48 according to the MALDI spectra of the cleavage mixture of HPLC peak 3. The mass spectrum in Figure 3.4A corresponds to the cleavage products represented by HPLC peak 1 in Figure 3.3. Two peaks at m/z 4910.4 and m/z 2192.1 are due to fragments 1-46 and it252-70, respectively (expected m/z 4909.3 and 2189.6). From these data, one can deduce that peptide chain cleavages occur at Cys47 and Cy352. The 108 Table 3.1. Calculated and observed mlz values for possible fragments resulting from the cleavage reaction of IGF-I chains at sites of designated cysteine pairs Reduction of Disulfid e Fragment Calculated mlz Observed mlz Cys47-Cys52 1-46 4909.3 4910.4 47-51 638.7 nd 52-70 2189.6 2192.1 CysG-Cys48 1—5 516.5 nd 6-47 4539.0 4539.2 48-70 2682.1 2533.3 *1-47 4976.5 4977.4 Cys15-cys61 Unreduced 109 A /.t252-70 Peak 2 Cys47-Cy352 I (D .15 ,r 1-46 Ex ._ JL. .JIAAA __L_A e Itz48-70 Peak 3 Cys6-Cys48 itz6-47-103 \itz/647 *147 JL . i _ 1000 2000 3000 4000 5000 6000 7000 8000 Figure 3.4. The MALDI mass spectra of peptide mixtures resulting from the cleavage of the two singly reduced/cyanylated lGF—l isomers, corresponding to the HPLC peaks 2 and 3 in Figure 3.2, respectively. The symbols "itz'" and * represent the iminothiazolidine derivatives and protonated B—elimination products, respectively. 110 MALDI peak marked itz47-70-SH in Figure 3.4A indicates the cleavage of the peptide bond at Cys47, and the cleavage of SH group from Cys52. This further confirms the cleavage site at Cys47. The fragment itz47-51 (m/z 638.7) was not determined by MALDI, which is due to the signal suppression at low-mass range of MALDI. Overall, a disulfide bond linkage between Cys47-Cy552 can be unambiguously deduced. With similar strategy, the other disulfide-bond linkage, Cys6-Cys48 also can be recognized from Figure 3.4B. The MALDI spectrum in Figure 3.4B, corresponding to the cleavage product represented by HPLC peak 2 (Figure 3.3), shows peaks at m/z 2683.8 and m/z 4539.2, corresponding to fi'agment itz48-70 and itz6-47, respectively. Although another expected fragment 1-5 (m/z 516.5) is missing, it is still possible to deduce Cys6- Cys48 fiom these two fragments because no other combination gives such masses. One minor B-elimination product, residues 1-47 (m/z 4977.4), is particularly informative for confirmation of the assignment in this case. It should be pointed out that the question- marked peak in Figure 3.4B corresponds arithmetically to itz6-47 minus 103 Da; we have no rational explanation for its origin at this time. Whether this peak corresponds to the expulsion of a cysteine residue from itz6-47 awaits further study with other peptides or proteins containing adjacent cysteines. After the two disulfide linkages of Cys47-Cys52 and Cys6-Cys48 are assigned unambiguously, the third one can be reasonably deduced as Cysl 8-Cys6l because there are three disulfide bonds in IGF-I, even though the singly reduced/cyanylated isomer could not obtained for this disulfide bond. lll B. Disulfide Mapping of LR’IGF-I Recombinant LR3IGF-I is a variant of human IGF-I that contains arginine replacing glutamate-3; it also has an amino terminal extension of 13 amino acids (30). LR3IGF-I shows higher biological activities than its analog, IGF-I. Like IGF-I, LR3IGF-I contains adjacent cysteines at positions 60 and 61. The disulfide-bonding scheme, assigned on the basis of homology to the insulin (or IGF-I) sequence and shown in Figure 3.5, 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 (31). Figure 3.6 shows the HPLC separation of LR3IGF-I and isomers of its partially reduced/cyanylated species. The chromatogram shows similar results as those of obtained for IGF-I. There are only two discernible peaks, each of which represents a singly reduced isoforrn as determined by mass analysis. Stronger reducing conditions (increase the ratio of reducing reagent, apply higher temperature, and prolong the reduction time) result in the formation of doubly reduced/cyanylated protein isomers, but the third disulfide bond still refused to reduce, indicating the third disulfide bond is very stable. Both MALDI-MS and on-line LC/ESI-MS were used to analyze the cleavage products of singly reduced/cyanylated isomers. The results are shown in Table 3.2. 112 Figure 3.5. Amino acid sequence and disulfide structure of LR3IGF-l The bold letters at N-terminal site indicate the 13 amino acid extension. The bold R represents the replacement of E3 of lGF-l. 113 Intact A] Figure 3.6. 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 mllmin with a linear gradient 30-50% B in 45 l . 10 1'5 20 2'5 3'0 3'5 4'0 45 min 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. 114 0.88 0.88 .88 8-8 0.084 0.88 0.88 8-8 0.58; 0.22 4.82 0: 880-280 e: .88 00:0 8-85 588 0.88 0.88 8-8 ammo 0: Esme voéo ~88 0.88 «:8 8-. 880-808 m2-_mm\o.. >0 5.2.2 B 005305 00 NE. 0020000 NE. 0020000 N)». 0003200 2050.000 0000:00m 0:00 050005 0000:9000 .6 00:0 00 050.0 700.0”... 00 000000. 0m0>00_0 05 E0: 000.0000 00:09.00: 0.90000 00.. 0029 N0: 0020000 000 0000.30.00 NM 030... 115 Interpretation of MALDI Data The two HPLC fractions (HPLC peak 1 and 2 in Figure 3.6) were subjected to cleavage under the described conditions. The corresponding MALDI mass spectra are shown in Figure 3.7A and B. The peaks at m/z 2191.6 and 6373.3 are due to fragments itz65-83 and 1-59, respectively. From these data, one can deduce that the cleavage sites are at positions 60 and 65. The fragment itz60-64 was missing from the MALDI spectrum due to the low-mass range signal suppression of MALDI-MS. However, the peaks at m/z 2755.1, representing overlapped peptide fragment itz60-83, where B- elimination occurs at Cys60, confirm that position 60 is the cleavage site. Therefore, disulfide linkage between Cys60 and Cys65 can be recognized. The cleavage products from HPLC peak 2 show m/z values attributable to fragrnentsl-l8, itz19-60, and itz61-83, respectively, suggesting Cysl9 and Cys6l are cleavage sites (Figure 3.7B). The peak with a single question mark showed a mass increase of 103 Da fiom 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 as occurrence in IGF-I. Like IGF-I, further study should be conducted to address the problem. On the other hand, the third disulfide bond in LR3IGF-I can be easily assigned to Cys31-Cys74 based on the linkages of other two disulfide bonds. Therefore, unambiguous assignment of the disulfide structure is evident even if such undesired fragments are present, because the fragments from specific cleavage at the native disulfide bonds are sufficient to make a positive conclusion. 116 A Peak 1 CysGO-Cys65 /it265-83 bum-83 l i - /rt260-83 /1_59 Peak 2 Cys1 9-Cys61 /itz61-83 Q ‘- .L 7 / 7? itz19-60 \ / ..ll-J. n 2000 £00 4000 5000 6000 7000 8000 9000 Mass (mlz) Figure 3.7. The MALDI mass spectra of peptide mixtures resulting from the cleavage of the two singly reduced/cyanylated LR3IGF-l isomers, corresponding to the HPLC peaks 1 and 2 in Figure 3.4, respectively. The symbol "itz" and * represent iminothiazolidine derivatives and B—elimination products, respectively. The peaks with question marks are discussed in the text. 117 Interpretation of L C/ESI-MS Data In order to confirm the disulfide structure of LRJIGF-I assigned by MALDI-MS, especially for the low-mass range fragments and question marked peaks, LC/ESI-MS was utilized to analyze the cleavage products of singly reduced/cyanylated isomers. Figure 3.8A and B as well as Figure 3.9A and B show the results. The first isomer (peak 1 in Figure 3.6) was analyzed by LC/ESI-MS to give the reconstructed total ion current (RTIC) chromatogram in Figure 3.8A and the corresponding mass spectra in Figure 3.8B. The peaks in the RTIC in Figure 3.8A can be assigned to specific fragments 1-59, itz60-64, itz65-83, itz65-83 minus the mass of H20, and the uncleaved protein (IP in Figure 3.8A) according to the corresponding masses from ESI spectra (Figure 3.83). The ESI spectra in Figure 3.8B give masses 6372.7Da, 639 Da (protonated peptide), and 2189.7 Da which are due to fragments 1-59, itz60-64, and itz65-83, respectively. The missing peak (mlz 638.0 corresponding to fiagment itz60- 64) in MALDI spectrum was identified by LC/ESI-MS. From these data, the cleavage sites can be easily deduced, which are at positions 60 and 65. Therefore, the disulfide linkage should be between Cys60 and Cys65. The peaks in the RTIC in Figure 3.9A resulting from analysis of cleavage products of the second singly reduced/cyanylated isomer (peak 2 in Figure 3.6) can be assigned to fragments 1-18 (1977.8 Da), itz19-60 (4540.98 Da), and itz6l-83 (2684.93 Da) according to corresponding ESI spectra in Figure 3.9B, which indicate Cysl9 was linked to Cys 61. While the peak corresponding to 1-18 plus 103 Da that was identified by MALDI-MS was not detected by LC/ESI-MS, the ‘redundant’ peak B in the BSI spectrum in the middle panel of Figure 3.9B that corresponds arithmetically to itzl9—60 minus 103 Da 118 002002.00 050__0N0_£o5:._ €000.00. N... 60.50 0.... <00. 0.090 5 00.000 0:. 05 0. 05000000000 0000000 05 .0 0.0.0000 000E _8 e... 0. 8.0 0.8.. 08:85 8880 0...... 0E. 0. <00 0.8.. 0.0 0.8.. e. . 0.8. 0.5: e. 8.88088 000.05 _-00_0m._ 00003005000000. >650 0 .0 00.0800 05 So... 0.35:. 000000 .0 00:00. 020090.. .0.0 0.390 NE... . 8.0.. . 8.0,. 08 08 8.. 8: 8m. 80. 8... 08 80 80. 8-0. 8.. 8. 08 08 a: i 0 NE .................. 1. o NEV ......... 1.14.1 .13....- . . L . L —O 0.. 0 4 A .81. 000. 00+.- ~+ 08 0mm. ..x. ..+ 0.. ..x. + 000. 0+ . . . 0 .. 038.88. 2. cm 008 «.8 .< . 8.. . a 0.. .8. 8.8 u. .2 .8. 8.8 u. 08 8. m h... . 0+ 88 88 800 88 08. 08. 08 cmowptlil-tt-.-.--b--.»--l-r--i-f-[-i.ti-rl-r-- ---.-i-_° / ._. - /.._ one M 00 A“ 6 ... .x. 0. m a a 8 o 0.0 6 .0 . < . uoow 119 .<0.m 0.090 5 00.000 0.... 0.0 00 05000000000 0000000 000 00 0.00000 0000. .ww 0.0 0. 00.0 0.090 08:00.0 000>00.0 0.0 00 0...”. 0. 4.0.0 0.090 0.0 0.090 5 N 0.000 040... 00 05000000000 000.00. 700.00.. 00003003000000. 2050 0 00 000>00.0 000 E00 0.02:. 000000 00 00.000. ms.-.wmm2o+ oz ov on cm or w 20% 20w + co ov on em or u _. 20% 20W «ofiulqyzozmicg Sensuom REE on S. on cm 2 al.. an am: also 8'30 anomflml mlfiPbWIMh v Bi V Ill O I” I o a o Q _. tOtOBbmm cm 2 84:0 mmlqmc mudlv + 8% Elev mIF «usuISoumSmoG ov on ow or on _. 23, 5% _L + co CV omoN or u w 7 o.._ 5% a .mutscofixcgo counsnmm Ream om ov. on em 3... \ x \ omév .mnucn .muur .cmuou .232. .m-.. L 123 LC/ESI-MS might be used to solve the problem, disulfide linkages may still be identified even though the peptide fragments cannot be completely determined by mass spectrometry. The peptide fragments indicated in bold numbers in Figure 3.10 for each case can be used to identify the disulfide linkages because they have unique masses from the set of peptide fragments that result from the other cases. Alternatively, the combination of the masses of several not complete peptide fragments can also be used to determine the disulfide linkages. For example, the combination of fiagments 1-9, 10-19, and 20—50 in case I of Figure 3.10 can be used to determine the disulfide linkage pattern of Cyle-Cys20 and Cys30-Cys40 because the same combination cannot be obtained for the other two cases (11 and III). An effective way to do the calculation and matching is by the computer sofiware. The combination of partial reduction, cyanylation, mass spectrometry and computer software matching would provide a powerful technique to assign the disulfide linkages in proteins. 124 VII. References l. Creighton, T. B., Bioessays, 8, 57 (1988). 2. Smith, D. L., Zhou, Z., Methods Enzymal., 193, 374 (1990). 3. Hirayama, K., Akashi, S., in Biological Mass Spectrometry: Present and Future, Matsuo T., Caprioli, R. M., Gross, M. L., and Seyama, Y., eds., pp 299 (1994). 4. Ryle, A. P., Sanger, F ., Biochem. J., 60, 535 (1955). 5. Stults, J. T., Bourell, J. H., Canova-Davis, B., Ling, V. T., Laramee, G. R, Winslow, J. W., Griffin, P. R., Rinderknecht, B., and Vandlen, R. L., Biomed. Environ. Mass Spectrom., 19, 655 (1990). 6. Fox, J. W., Elzinga, M., and Tu, A. T., Biochemistry, 18, 678 (1979). 7. Serensen, H. H., Thomsen, J ., and Bayne, S., Biomed. Environ. Mass Spectrom. 19, 713 (1990). 8. Nokihara, K, Morita, N., Yamaguchi, M., and Watanabe, M., Anal. Lett., 25, 513 (1992) 9. Zhang, D., and Liang, S., J. Protein Chem, 12, 735 (1993). 10. Krishnamurthy, T., Hauer, C. R., Prabhakaran, M., Freedy, J. G., and Hayashi, K., Biol. Mass Spectrom., 23, 719 (1994). 11. Tang, Y., and Selsted, M. B., J. Biol. Chem, 268, 6649 (1993). 12. Gray, W. R., Protein Science, 2, 1749 (1993). 13. Haniu, M., Hsieh, P., Rohde, M. F., and Kenney, W. C., J. Biochem. Biophys., 310, 433 (1994). 14. Yamashita, H., Nakatsuka, T., and Hirose, M., J. Biol. Chem, 270, 29806 (1995). 15. Fairlie, W. D., Stanton, P. G., and Hearn, M. T. W., Biochem. J., 314, 449 (1996). 125 16. Gray, W. R., Luque, F. A., Galyean, R., Atherton, B., 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). 17. Gray, W. R., Protein Sci., 2, 1732 (1993). 18. Wu, J ., and Watson, J. T., Protein Science, 6, 391 (1997). 19. Rinderknetcht, B., and Humbel, R. B., J. Biol. Chem, 253, 2769 (1978). 20. Froesch, E. R., Burgi, H., Ramseier, E. B., Bally, P., and Labhart, A., J. Clin. Invest., 42, 1816 (1963). 21. Rinderknecht, B., and Humbel, R. B., Proc. Natl. Acad. Sci., USA, 73, 2365 (1976). 22. Blundell, T. L., Bedarkar, S., Rinderknecht, B., and Humbel, R. B., Proc. Natl. Acad. Sci. USA, 75, 180 (1978). 23. Blundell, T. L., Dodson, G. G., Hodgkin, D. C., and Mercola, D. A, Adv. Protein Chem, 26, 279 (1972). 24. Forsberg, G, Palm, G., Ekebacke, A., Josephson, S., and Hartmanis, M., Biochem. J., 271, 357 (1990). 25. Iwai, M., Kobayashi, M., Tamura, K., Ishii, Y., Yamada, H., and Niwa, M., J. Biochem., 106, 949 (1989). 26. Raschdorf, F., Dahinden, R., Maerki, W., and Richter, W. J ., Biomed. Environm. Mass Spectrom., 16, 3(1988). 27. Axelsson, K., Johansson, S., Eketorp, G., Zazzi, H., Hemmendorf, B., and Gellerfors, P., Eur. J. Biochem., 206, 987 (1992). 28. Wu, J., Gage, D. A., and Watson, J. T., Anal. Biochem., 235, 161 (1996). 126 29. Hober, S., Forsberg, G., Palm, G., Hartmanis, M., Nilsson, B., Biochemistry, 31, 1749 (1992) 30. Tomas, F. M., Knowles, S. B., Chandler, C. S., Francis, G. L., Owen, P. C., and Ballard, F. J ., J. Endocrinology, 137, 413 (1993). 31. Rinderknetcht, B., Humbel, R. B., FEBS Lettt., 89, 283 (1978). 127 CHAPTER 4 TRAPPING AND IDENTIFICATION OF INTERMEDIATES DURING THE REFOLDING OF LR3IGF-I AND IGF-I I. Introduction Considerable insight to the folding and unfolding pathways of a protein can be obtained from trapping and characterizing intermediate structures involved in the dynamic process (1-4). A particularly difficult aspect in the study of protein folding is the fact that intermediates may be short-lived and therefore difficult to isolate and analyze structurally and functionally. Disulfide-containing proteins provide an opportunity to capture sulfhydryl-containing intermediates by chemical reaction during the time course of folding or unfolding (5-8). The folding pathways of several proteins (9-11) have been studied in this way; among them, bovine pancreatic trypsin inhibitor (BPTI) (6, 8, 12, 13), and ribonuclease A (14-18) have been extensively characterized. In order to isolate and characterize the intermediates that are involved in folding or unfolding of proteins containing disulfide bonds, it is necessary to stop both the inter- and intramolecular thiol/disulfide exchange reactions that convert one intermediate to another. Two traditional trapping methods have been used. One method traps all sulthydryls irreversibly by alkylation with iodoacetate or iodoacetamide. However, rearrangement of intermediates during trapping with iodoacetate has been observed for both BPTI (8) and ribonuclease A (19), since quenching by 0.2M iodoacetate is comparable to or, in some cases, slower than the rate of some of the intramolecular rearrangements that interconvert intermediates of the same oxidation state (20). Another 128 method quenches thiol/disulfide exchange by lowering the pH to < 2 by adding acid. An advantage of acid trapping is its reversibility; intermediates trapped at low pH can be transferred to high pH to allow further rearrangement or folding in experiments designed to more completely characterize particular pathways (8, 10). While quenching by acidification is rapid and occurs at the diffusion-controlled rate, it does not completely stop thiol/disulfide exchange (21). Therefore, structural characterization of a given trapped intermediate must be done promptly. We have developed methodology to trap and identify folding intermediates based on cyanylation methodology and mass spectrometry (22). Our approach has two advantages. Firstly, acidification quenches the thiol/disulfide exchange rapidly, and cyanylation of free sufl'iydryls precludes further exchange because of irreversible chemical modification. Secondly, trapping of sulfhydryls is already part of our procedure for structural elucidation of the intermediates. After trapping and isolation of folding or unfolding intermediates, the disulfide structure must be determined. Conventional methodology for determining disulfide structures of folding intermediates involves cleavage between every cysteine residue in the intermediates by enzymatic/chemical digestion, which is tedious, cumbersome, and may risk artifact formation due to disulfide exchange. Furthermore, conventional approaches may have difficulty in analyzing intermediates containing adjacent cysteines since proteolytic and/or chemical degradation may not achieve cleavage between cysteine residues (23). 1 Our methodology involves partial reduction of disulfide bonds in a protein, cyanylation of suthydryl groups, HPLC separation of partially reduced/cyanylated protein 129 isoforms, chemical cleavage of the peptide bonds at the N-terminal side of cyanylated cysteine residues, and mass analysis of cleavage products by mass spectrometry (24). This new developed methodology has the advantages of preventing disulfide scrambling and cleaving the peptide backbone between adjacent cysteines; furthermore, it is fast, simple, and sensitive (25). Insulin-like growth factor-I (IGF-I) is a single-chain polypeptide of 70 amino acid residues containing three intramolecular disulfide bonds, two of which involve adjacent cysteines (26) (Figure 4.1). IGF-I is postulated to be the mediator of growth hormone action on skeletal tissue as well as mitogenic activity on several cell types (27, 28). Several research groups have studied refolding pathway of IGF-I (29-33). Hober, et al (29, 30) trapped the folding intermediates of IGF-I by pyridylethylation at pH 8.7. Five major forms of IGF-I were detected, which includes a one-disulfide intermediate, two two-disulfide intermediates, and a mismatched three-disulfide intermediate (29). A different folding pattern was obtained by trapping the folding intermediates with acidification (31-33). Instead of five intermediates detected by pyridylethylation, six intermediates were identified by trapping with acidification, which include a mixed disulfide intermediate (3 2, 33). A nonnative two-disulfide intermediate was. identified with acidification, but not identified with pyridylethylation. Furthermore, while a one- disulfide intermediate (Cysl 8-Cysl6) was identified as a major form in the folding process of IGF-I by trapping with pyridylethylation (29), it only appeared as a minor component when trapped by acidification (32, 33). “Therefore, a further study of the folding intermediates of IGF-I needed to be conducted. 130 Figure 4.1. Amino acid sequences and disulfide structures of LR3IGF-l and lGF-l. The numbers in the parentheses indicate the location of cysteine residues of lGF—l. The 13 hold letters at the N-terminus indicate the 13-amino acid extension in LR3IGF-l. The bold R represents the replacement of E3 of lGF-l. 131 Recombinant human long R3 insulin-like growth factor-I (LR3IGF—I) is a variant of human insulin-like growth factor-1 (IGF-I) in which glutamate 3 is replaced by arginine and a 1.3-residue extension appears at the N-terminus (Fig. 4.1). It contains three disulfide bonds and two adjacent cysteines. The disulfide linkages have been determined by partial reduction/cyanylation/mass mapping (25). LR3IGF-I is substantially more potent than IGF-I in affecting carbohydrate metabolism and in stimulating the growth of fetal tissue in animals (34). The results of a folding study of LR3IGF-I (31) were significantly different from those published for IGF-I refolding (29, 30, 32-33). However, the disulfide structure of the folding intermediates of LR3IGF-I has not been studied by the authors (31). In this chapter, we integrate the different structural features of IGF-I and LR3IGF-I into the observed folding and refolding dynamics by using the same methodology to study the behavior of both proteins. Furthermore, our methodology permits direct determination of the disulfide pattern in each of the intermediates, which were trapped without artifact formation. We have trapped the intermediates of refolding of IGF-I and LR3IGF-I and determined disulfide structures of the intermediates by cyanylation methodology and mass spectrometry. The disulfide structure of all the refolding intermediates of both IGF-I and LR’IGF-I were identified by our methodology. It is the first direct experimental evidence for the disulfide structure of the nonnative three- disulfide intermediate, which was only predicted previously by the structural constraints of disulfide bonds in the three-disulfide intermediates (29). The refolding pattern of IGF-I obtained by our methodology was similar to that of obtained by trapping with 132 acidification. Furthermore, the results of our investigation provide integrated information for the folding of insulin-like growth factor families. 11. The Strategy for Trapping Folding Intermediates The procedures for trapping and analyzing refolding and reductive unfolding intermediates of proteins are shown in Figure 4.2. Reduced-unfolded protein was refolded at pH 8.7 in the presence of redox buffer consisting of oxidized glutathione (GSSG) and reduced glutathione (GSH). The folding intermediates were trapped by cyanylating the sulflrydryl groups of the intermediates with 1-cyano-4-dimethylarnino- pyridinium (CDAP) at pH 2~3. The trapped cyanylated intermediates were separated by reversed-phase HPLC. The numbers of disulfide bonds were determined by MALDI-MS based on the mass shift in MW between native protein and that of intermediates. The mass shifl should be 104 Da between native protein and the one-disulfide intermediate, 52 Da between native protein and the two-disulfide intermediates, and no mass shift between the intact protein and the three-disulfide intermediates. The 52-Da mass shift between intact protein and the two—disulfide intermediate arises from 2 Da for reduction of one disulfide bond and 25 Da for replacement of a hydrogen atom on each of the sulflrydryl groups with CN, etc. Afier HPLC separation and purification, the disulfide structure of the intermediates was determined based on the methodology of partial reduction, cyanylation, chemical cleavage, and mass mapping (24). The reverse direction of refolding involving disulfide formation in a protein is reductive unfolding involving the reduction of disulfides by a free-thiol reagent such as cysteine. A study of reductive unfolding can provide insight to protein refolding since 133 2:205 ho 9.5.8:: 9:032 ucm @5282 no $3803 2: .6 263.05 on... .Né 2:9“. . mOwEUOr—tmuc— ho Ghauuzbw 005330 0.3 05559.00; _ 855852. 85.5 _ mecca $535 _ 5:838 Co .353: mos—E030 _ . Dull _mofifioctcg 9.6.8:: cognac...— mofieoctous 9.6.8 339... n_ SH SH I F i i i l 1 6 16 H 47 46 52 61 7° W l T l l‘_l 1 6 16 47 46 52 61 7° . l l .—l—. l 7.... 6 16 47 46 52 61 It should be pointed out that the folding pathways of LR3IGF-I and IGF-1 described in Figure 4.21 and Figure 4.22 are predicted only based on the time dependent appearance and disulfide structure of the folding intermediates. Two problems must be solved before the actual pathway can be defined. Firstly, the well-populated species are not necessarily the productive species that account for the flow of intermediates. One way to identify the productive intermediates is to perform stop/go folding experiments of all species of the intermediates, both major and minor, that present at the same stage of equilibrium. The dilemma faced by this approach is that productive intermediates may exist as minor species. Finding all these minor species will be a daunting task. The predicament can be further complicated by the argument that an undetected species is not necessarily a non-existing species. Secondly, protein folding may occur through multiple pathways and these pathways may not be distinguishable. The other techniques such as mutation of the cysteine residues might solve the problem. The trapping methodology, based on the cyanylation of he thiol groups under acidic conditions, showed similar results to those of acid trapping (31-33), but different fiom those of pyridylethylation trapping at alkaline condition (29) in terms of the distribution of intermediates during the folding processes of LR’IGF-I and IGF-1. The methodology for the determination of disulfide structure by chemical cleavage and mass mapping of the fragments can be used to determine the disulfide linkages of the folding intermediates containing adjacent cysteines. The direct experimental evidences for the disulfide linkages of the nonnative three-disulfide intermediates of LR’IGF-I and IGF-1 were obtained for the first time by our methodology, which cannot be solved by the 184 conventional approach. Furthermore, the cyanylation methodology is much faster and sensitive. 185 VIII. References 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. Kim P. S., Baldwin R. L. (1990) Annu. Rev. Biochem. 59, 631-660. Matthews CR. (1993) Annu. Rev. Biochem. 62, 653-683. Ptitsyn QB. (1995) Curr. Opin. Struct. Biol. 5, 74-78. Li Y.J., Rothwarf D.M., Scheraga H.A. (1995) Nature Struc. Biol. 2, 489-494. Creighton, T.E. (1974). J. Mol. Biol. 87, 563-577. Creighton, T.E., Goldenberg, D.P. (1984) J. Mol. Biol. 179, 497-526. Scheraga, H.A., Konishi, Y., Ooi, T. (1984) Adv. Biophys. 18, 21-41. Weissman, J.S., Kim, RS. (1991) Science 253, 1386-1393. Creighton, T.E., Ewbank 1.]. (1994) Biochemistry 33, 1534-1538. Chang, J .Y. (1996) Biochemistry 35, 11702-11709. Glocker M.O., Arbogast B., Milley R., Cowgill C., Deinzer M.L. ( 1994) Proc. Natl. Acad. Sci. USA 91, 5868-5872. Darby, N.J., Norin P.E., Talbo, G., Creighton, T.E. (1995) J. Mol. Biol. 249, 463- 477. Weissman, J.S., Kim, RS. (1992) Proc. Natl. Acad Sci. USA 89, 9900-9904. Xu, X., Rothwarf, D.M., Scheraga, H.A. (1996) Biochemistry 35, 6406-6417. Rothwarf, D.M., Scheraga, H.A. (19936) Biochemistry 32, 2671-2679. Rothwarf, D.M., Scheraga, H.A. (1993b) Biochemistry 32, 2680-2689. Rothwarf, D.M., Scheraga, H.A. (1993c) Biochemistry 32, 2690-2697. Rothwarf, D.M., Scheraga, H.A. (1993d) Biochemistry 32, 2698-2703. Rothwarf, D. M., Scheraga, H. A. (1991) J. Am. Chem. Soc. 113, 6293-6294. Creighton, T.E. (1992) Bioessays 14, 195. 186 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. Pain, RH. (1994). In mechanisms of protein folding. Oxford University Press, New York, p112. Wu, J., Yang, Y., Watson, J .T. (1998) Accepted by Protein Science. Smith, D.L., Zhou, Z.R. (1990) Methods Enzymol. 193, 374-389. Wu, J., Watson, J.T. (1997) Protein Sci. 6, 391-398. Yang, Y., Wu, J ., Watson, J .T. (1998) Accepted by J. Am. Chem. Soc.. Rinderknetcht, B., and Humbel, R. B., J. Biol. Chem., 253, 2769 (1978). Froesch, E. R., Burgi, H., Ramseier, E. B., Bally, P., and Labhart, A., J. Clin. Invest., 42, 1816 (1963). Rinderknecht, B., and Humbel, R. B., (1976) Proc. Natl. Acad Sci., USA, 73, 2365. Hober, S., Forsberg, G., Palm, G., Hartmanis, M., and Nilsson, B. (1992) Biochemistry, 31, 1749-1756. Hober, S., Forsberg, G., Palm, G., Hartmanis, M., and Nilsson, B. (1992) Biochemistry, 31, 1749-1756. Milner, S.J., Francis, G.L., Wallace, J.C., Magee, B.A., Ballard, F.J. (1995) Biochem J. 308, 865-871. Hua, Q.X., Narhi, L., Jia W., Arakawa, T., Rosenfeld, R., Hawkins, N, Miller, J .A., Weiss, M. (1996) J. Mol. Biol. 259, 297-313. Miller, J.A., Narhi, L.O., Hua, Q.X, Rosenfeld, R., Arakawa, T., Rohde, S.P., Lauren, S., Stoney, K.S., Tsai, L., Weiss, M.A. (1993) Biochemistry 32, 5203-5213. 34. Tomas, F.M., Knowles, S.E., Chandler, C.S., Francis, G.L., Owens, P.C., Ballard, F.J. (1995) J. Endocrinology 137, 413-421. 187 "‘llllllllllllllllllll“