.3... i i .. .. .333; .1311: . a I... - 334 . c 1 um... 1.5%.... , .... 3....» 5 35 a!" 3&2... ‘1‘ iqltyinn... (to. tu Q .1 9. . tin; 4.715.) flu ”an .1 . VF. Uri ...(. an)... .lI IQ? Ar IJ... . ,-!\.A~.. .bx This is to certify that the dissertation entitled STUDIES OF THE OXIDATIVE REFOLDING OF CYSTINYL PROTEINS BY HYDROGEN/DEUTERIUM EXCHANGE-MASS SPECTROMETRY presented by Xue Li has been accepted towards fulfillment of the requirements for the Doctoral degree in 1 Chemistry Wm Major Professor’s Signature Mt WJM Date MSU is an Affirmative Action/Equal Opportunity Institution LIBRARY Michigan State . University 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 JUL 0 8 2009 071103 2/05 p:lCIRC/DateDue.indd-p.1 STUDIES OF THE OXIDATIVE REFOLDING OF CYSTINYL PROTEINS BY HYDROGEN/DEUTERIUM EXCHANGE-MASS SPECTROMETRY By Xue Li A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2005 ABSTRACT STUDIES OF THE OXIDATIVE REF OLDING OF CYSTINYL PROTEINS BY HYDROGEN/DEUTERIUM EXCHANGE-MASS SPECTROMETRY By Xue Li Hydrogen/deuterium exchange (HDX) with accurate mass measurement provides important information about the conformation of a polypeptide chain. When a protein is dissolved in D20, hydrogens located at peptide amide linkages can exchange with deuteriums fiom the solvent at easily measurable rates. The rates depend on features directly related to protein structure such as whether the amide hydrogens are participating in intramolecular hydrogen bonding, and the extent to which they are shielded from the solvent. Thus, conformational information about a protein can be obtained by assessing the extent of deuterium incorporation during HDX. Various parameters for HDX experiments were investigated to minimize back- exchange and to accurately measure the mass of the deuterium-labeled protein by matrix assisted laser desorption/ionization time of flight mass spectrometry (MALDI-TOF-MS). Optimized protocols for different types of HDX experiments were then proposed. Disulfide bonds, an important post-translational modification of proteins, play a key role in stabilizing a protein’s 3-dimensional structure and controlling its activity. The novel cyanylation/cleavage methodology, developed previously in this laboratory, has been applied in several studies of protein folding because of its effectiveness in trapping, identifying, and preserving the disulfide structure of a given folding intermediate of a cystinyl protein. This methodology was integrated with HDX-MS to study the conformational changes of a mode] cystinyl protein, long-R3 insulin-like growth factor-I (LR3IGF-I), as a function of disulfide bond formation during its oxidative refolding. The integrated approaches establish a novel strategy for studying the refolding of cystinyl proteins. Conventionally, HDX-MS has been used to characterize large proteins. We investigated the applicability and advantages of HDX-MS in the conformational study of small cystinyl peptides. The conformations of five peptide models derived from bovine pancreatic trypsin inhibitor (BPTI) were characterized to study the effect of long-range interactions (induced by the formation of disulfide bonds) in stabilizing their local folded structure. Collisionally induced dissociation-tandem mass spectrometry (CID-MS/MS) was employed to assess the extent of deuterium incorporation at the single amino acid residue level. During the oxidative refolding of LR3IGF-I, only one of the fifteen possible one- disulfide intermediates (the one containing the disulfide bond Cys31—Cys74) accumulates to a significant extent. A structural explanation is proposed for this phenomenon: during the folding of LR3IGF-I, long-range interactions induced by the formation of the Cy531-CYS74 disulfide bond may stabilize the local folded structure around it; the local folded structure, in turn, protects this disulfide bond against attack from reducing reagents in the solution such as free sulfltydryl groups. This hypothesis was tested by probing the structural features of synthetic peptide models that mimic those of the intact Cys3l—Cys74 l-S intermediate. ACKNOWLEDGMENTS I would like to express my gratitude to my advisor, Dr. J. Throck Watson, for his support, advice, and encouragement. I would also like to thank Dr. William J. Wedemeyer for his guidance and help; Dr. Victoria L. McGuffin and Dr. David P. Weliky for helpful discussions and for serving on my committee. I want to acknowledge Dr. Honggao Yan and Dr. Leslie A. Kuhn for helpful discussions; Dr. Robin J. Hood, Dr. Yi-Te Chou, Dr. Rhonda Husain and Dr. Jiong Yu for assistance and discussion; Dr. Steven P. Gygi for access to the LCQDeca XP ion-trap mass spectrometer at Harvard Medical School and Dr. Babak Borhan for access to the CD spectrometer at Michigan State University. To all the Watson group members, past and present, thank you for your fi'iendship and cooperation. I also owe many thanks to the members of the Mass Spectrometry Facility: Dr. Daniel A. Jones, Lijun Chen, Beverly Chamberlin and Susanne Hoffmann-Benning. I especially would like to thank my mother, Shuying Xue, for her love, patience, and support, without which I would not have been able to finish this Ph.D. project. At last, but not the least, I would like to thank my husband, Gabriel D. Musk, and my Chinese and American families for their love and support. iv TABLE OF CONTENTS LIST OF TABLES ............................................................................................................ vii LIST OF FIGURES ......................................................................................................... viii ABBREVIATIONS .......................................................................................................... xii CHAPTER 1 INTRODUCTION .............................................................................................................. l I. Introduction ................................................................................................................... 1 II. Disulfide Mapping of Cystinyl/Cysteinyl Proteins ....................................................... 3 A. Conventional Proteolytic Fragmentation Strategy .................................................. 4 B. Novel Cyanylation/ Cleavage Methodology ........................................................... 9 HI. Oxidative Folding of Cystinyl/Cysteinyl Proteins and Trapping of Folding Intermediates ............................................................................................................... 14 A. Disulfide Formation in Oxidative Folding of Cystinyl/Cysteinyl Proteins .......... 15 B. Trapping of the Folding Intermediates ................................................................. 18 IV. Conformational Study of Proteins and Peptides ......................................................... 24 A. Conventional Technologies in Protein/Peptide Conformational Studies .............. 24 B. HDX—MS ............................................................................................................... 31 V. Desorption/Ionization Mass Spectrometry ................................................................. 34 A. Matrix-Assisted Laser Desorption/Ionization Time of Flight Mass Spectrometry (MALDI-TOF-MS) ..................................................................................................... 35 B. Electrospray Ionization Ion Trap Mass Spectrometry (ESI-Ion Trap-MS) .......... 45 C. The Comparison of MALDI and E81 .................................................................... 51 D. ESI-CID-MS/MS .................................................................................................. 53 VI. References ................................................................................................................... 60 CHAPTER 2 OPTIMIZATION OF CONDITIONS FOR MONITORING HYDROGEN/DEUTERIUM EXCHANGE IN PROTEINS BY MALDI-TOF-MS ....................................................... 73 I Introduction ................................................................................................................. 73 11. Experimental Section .................................................................................................. 75 III. Results and Discussion ............................................................................................... 79 A. Complete Reduction of RNase A .......................................................................... 79 B. Effect of Deuterated and Non-deuterated Matrix ................................................. 83 C. Use of a Deuterium-labeled Internal Standard ...................................................... 88 D. Ablation of Sodium Ion Adducts Early in the Analysis at a Given Spot on the MALDI Target ............................................................................................................ 89 E. Effectiveness of Quenching In-Exchange at pD 2.5 ............................................. 91 IV. Conclusion .................................................................................................................. 92 V. References ................................................................................................................... 93 CHAPTER 3 INTEGRATION OF HYDROGEN/DEUTERIUM EXCHANGE AND CYANYLATION-BASED METHODOLOGY FOR CONFORMATIONAL STUDIES OF CYSTINYL PROTEINS ............................................................................................. 96 I. Introduction ................................................................................................................. 96 11. Experimental Section .................................................................................................. 99 III. Results and Discussion ............................................................................................. 103 A. Experimental Design for the Application of HDX-MS to Cystinyl/Cysteinyl Proteins .................................................................................................................... 103 B. HDX Behavior of Different Segments of LR3IGF-I ........................................... 106 C. Comparison of HDX-MS Results with NMR Data on the Native Structure of LR3IGF-I ................................................................................................................... 111 D. Some Figures of Merit for the Integrated Methodologies .................................. 114 IV. Conclusion ................................................................................................................ 121 V. Reference .................................................................................................................. 122 CHAPTER 4 CHARACTERIZATION OF PEPTIDE FOLDING NUCLEI BY HYDROGEN/DEUTERIUM EXCHANGE-MASS SPECTROMETRY AND CIRCULAR DICHROISM ............................................................................................. 126 I. Introduction ............................................................................................................... 126 H. Materials and Methods .............................................................................................. 129 1]]. Results and Discussion ............................................................................................. 132 A. Criteria for Selection of Peptide Models ............................................................ 132 B. Conformational Study of Selected Peptide Models by CD Spectroscopy .......... 137 C. Study of the Conformational Features of Selected Peptide Models by HDX- MALDI-MS .............................................................................................................. 143 IV. Conclusions ............................................................................................................... 157 V. References ................................................................................................................. 159 CHAPTER 5 IDENTIFICATION AND CHARACTERIZATION OF AN AUTONOMOUS FOLDING UNIT FOR LR3IGF-I ...................................................................................................... 163 I. Introduction ............................................................................................................... 163 H. Materials and Methods .............................................................................................. 166 HI. Results and Discussion ............................................................................................. 169 A. Design of the peptide models .............................................................................. 169 B. CD results ............................................................................................................ 172 C. HDX results ........................................................................................................ 177 IV. Conclusion ................................................................................................................ 199 V. Future Work .............................................................................................................. 200 VI. Reference .................................................................................................................. 201 vi Table 1. Table 2. Table 2. Table 2. Table 2. Table 3. Table 3. Table 4. Table 4. Table 5. Table 5. 2 l 2 LIST OF TABLES Comparison of ESI—MS and MALDI-MS [147] ............................................ 52 Calculated and detected m/z value (average) of protonated cleavage products from RN ase A following complete reduction, cyanylation, and cleavage. . 81 Comparison of observed deuterium incorporation level in reduced (unfolded) RNAse A using deuterated and non-deuterated matrix. .............................. 84 Comparison of the HDX results using internal and external calibration. ...... 88 Deuterium incorporation in native (folded) RNase A after “quenching”. ..... 91 Summary of HDX results on identified peptic fi'agments. .......................... 110 Summary of mass shifts for two peptic fragments from different red-ox species after 10—sec HDX. (Results are the average from three independent experiments; plus/minus terms indicate the 95% confidence interval). 115 HDX results for oc- and B-subunits as assessed by MALDI-TOF-MS. ....... 144 HDX results for the Pp subunit as assessed by CID-MS/MS. ...................... 153 HDX results of each intact peptide model. .................................................. 178 HDX-CID-MS results for each peptide model. ........................................... 196 vii LIST OF FIGURES Figure 1. 1 Cysteine and cystine. ...................................................................................... 3 Figure 1. 2 Conventional proteolytic fiagmentation strategy for disulfide bond assignment of a protein. ................................................................................. 4 Figure 1. 3 Disulfide/sulflrydryl exchange in proteins ...................................................... 5 Figure 1. 4 Conceptual representation of conventional protocol for disulfide-bond assignment based on mass spectrometry ........................................................ 8 Figure 1. 5 Chemical overview of the cyanylation/cleavage strategy. Partial reduction means that the protein of interest is transformed into a mixture of isofonns, each of which corresponds to reduction of only one of its disulfide bonds; for a protein containing 11 cysteines, n isomers of the singly reduced protein will result. Itz stands for iminothiazolidinyl carboxyl residue at the amino terminus [redrawn from [36]]. ..................................................................... 11 Figure 1. 6 Chemistry involved in the cyanylation and cleavage reactions .................... 13 Figure 1. 7 The formation of a disulfide bond upon adding a disulfide reagent to a reduced protein ............................................................................................. 16 Figure 1. 8 Trapping folding intermediates by acidification. ......................................... 19 Figure 1. 9 Chemical trapping of folding intermediates by alkylation. .......................... 20 Figure 1. 10 Reversible trapping of folding intermediates by AEMTS ........................... 22 Figure 1. 11 Representative CD spectra of a-helical, B-sheet, and random coil structure in proteins (redrawn from [89]). .................................................................. 31 Figure 1. 12 Highly charged clusters are formed during disintegration of the analyte/matrix solid [117]. M stands for the neutral analytes, H” for protons, and B' for anions (i.e., trifluoroacetate). ...................................................... 37 Figure l. 13 Chemical structure of compounds commonly used as matrices for MALDI. ................................................................................................................................... 39 Figure 1. l4 MALDI-TOF-MS instrument [123]. ........................................................... 42 Figure 1. 15 Schematic diagram of ion trajectories in a reflectron-TOF m/z analyzer [129]. ............................................................................................................ 44 viii Figure 1. Figure 1. Figure 1. Figure 1. Figure 1. Figure 1. Figure 1. Figure 2. Figure 2. Figure 2. Figure 2. Figure 2. Figure 3. Figure 3. Figure 3. 16 17 18 19 20 21 22 Schematic of the electrospray ionization process [136]. ............................. 47 ESI-MS instrument interface. ...................................................................... 48 Schematic diagram of ion-trap mass spectrometer [146]. ........................... 50 Conceptual representation of the MS/MS technique (redrawn from [149]).54 Segment of a peptide showing the residues of alanine and lysine linked by an amide bond, the peptide bond. ................................................................ 56 Conceptual pathway of forming the y-ions by expulsion of residues from the amino terminus, thereby leaving the charge on detectable fragment ions retaining the C-terminus (redrawn from [156]). .......................................... 57 Conceptual pathway of forming the b-ions by expulsion of protonated residues plus the hydroxyl group at the C- terminus, leaving the charge on N-terminal fragment ions (redrawn from [156]); for convenience, the b-ion here is shown as an acylium ion, however, the actual structure is believed to be an oxazolone species ([157, 158]) ........................................................... 58 Conceptual chemical scheme for complete reduction of bovine pancreatic RNase A, and confirmation of complete reduction by subsequent cyanylation, cleavage,and analysis of cleavage reaction mixture ................ 80 MALDI mass spectrum of the cleavage products of RNase A. ................... 82 Mass spectra of completely reduced RNase A after incubation with D20 and then mixed with (A) deuterated matrix or (B) non-deuterated matrix before analysis by MALDI-TOF-MS ...................................................................... 85 Time course of HDX on reduced RNase A. ................................................. 86 Mass spectra of fully deuterated myglobin (dzéo—Mb) obtained with deuterated sinapinic acid as matrix (A) Cumulative result from the first 50 laser shots; (B) Cumulatlast 50 laser shots. ................................................. 90 Amino acid sequence of IGF-I and LR3IGF-I; the sequence of IGF-I starts at residue 14 of LR3IGF-I with a Glu at position 16. ...................................... 98 MALDI-TOF mass spectrum of peptic digest of LR3IGF-I. ...................... 108 Average NMR structure of LR3IGF—I [8]. N-terminus and C-tenninus of the protein are labeled by N and C, respectively. Three helixes are labeled according to inclusive residues. Three disulfide bonds are represented by dotted lines between the designated cysteines. The dark-gray region is the ix Figure 3. 4 Figure 4. 1 Figure 4. 2 structurally ordered part of the protein; the light-gray segments are highly flexible parts of the protein (in which the heavy atom r.m.s.d. is larger than the average r.m.s.d. of the ordered part of the protein). The figure was drawn using the RasMol software [33]. ..................................................... 112 Mass spectra of (A) the segment 67 — 72 (LRRLEM, 6-mer) from different redox species and (B) the segment 38 - 55 (FNKPTGYGSSSRRAPQTG, 18-mer) from different redox species. The first row shows the mass spectra of the peptides before deuteration; the second row shows the mass spectra after sustained HDX to achieve a high degree of deuteration of the peptides. The remaining four rows are mass spectra of indicated redox species (R: reduced; l-S: one—disulfide intermediate; 2-S: two-disulfide intermediate, N: native) after a 10-s exposure to D20 at pD 6.8. ......................................... 118 Amino acid sequence of BPTI. ................................................................... 134 Ribbon diagram of BPTI [35], with peptides Pm and Pp depicted in dark and light gray, respectively. The three native disulfide bonds are also shown: Cyss-Cy855 (left), Cysl4-Cys33 (right), and Cy53o-Cy551 (middle). The figure was drawn using the MOLMOL software [36]. ......................................... 135 Figure 4. 3 (A) CD spectra of Pap at 0°C, Par; at 60°C and the summed spectrum of PO! Figure 4. 4 Figure 4. 5 Figure 4. 6 Figure 4. 7 and Pp at 0°C. (B) CD spectra of Pa at 0°C, P,m at 0°C, P,m at 60°C, and the difference spectrum of P,m and Pa at 0°C. (C) CD spectra of P5 at 0°C, Ppp at 0°C, PM; at 60°C, and the difference spectrum of P33 and PB at 0°C. (D) Temperature dependence of the CD signal at 217 nm for Pap. .................. 139 Mass spectra of (A) Pug before HDX, (B) Pap after HDX, (C) Pa after HDX, (D) PB after HDX, (E) P.m after HDX, (F) Ppp after HDX. The disulfide bond in each dimer was present during HDX, but was reduced before the mass measurement. The cluster of peaks corresponding to the Pa-chain is on the left, while that for the PB-chain is on the right in (A) and (B). Likewise, P.m and P55 were exposed to HDX as dimers, but reduced to the monomer prior to analysis by MALDI-MS. .............................................. 145 CID-MS/MS spectrum of the PB sequence. ................................................ 151 CID-MS/MS spectrum of the P,x sequence. ................................................ 152 Ribbon diagram of the natively structured PB peptide; Asn24 is indicated in light gray and relevant hydrogen bonds are shown in dark gray. The mainchain—mainchain H-bond donor/acceptor pairs from bottom: Phe33/Argzo, P hezz/ Gln31, Gln31/Phe22, AS1124/L81129, Leuzg/Asn24, and Gly23/Asn24. The H-bonds between the amide hydrogens of Ly57 and A183 Figure 5. 1 Figure 5.2 Figure 5. 3 to the sidechain atom 0°l of Asns are also shown. The figure was drawn using the MOLMOL software [36] ............................................................. 155 HPLC chromatogram showing the time-dependent distribution of CDAP- trapped intermediates during the refolding of LR3IGF-I [13]. The disulfide structure of each species is: I’ (Cysn — Cys74), II’ (Cysn — Cys74, Cyslg - Cy36l)a III’ (Cys31 - CYSM, CY519 - Cysw, Cysts: - Cyses), N (Cy331 - CyS74, CYS19 — Cysm , Cys6o — Cysés), R (completely reduced). (experimental details on page 99 in Chapter 3) ................................................................ 164 The natively structured P1 and P11 [24]. P. is depicted in dark gray, with its initial residue, Argn, shown at the far left and its final residue, Ser39, shown in the back on the right. Similarly, P11 is depicted in light gray, with its initial residue, Arg54, shown at the bottom and its final residue, Gln75, shown at the far right. The crossing-linking disulfide bond, CyS31-Cys74, is shown in the upper right. The side chains of the substituted residues are also shown in dark gray (in P1) and light gray (in P"). The figure was drawn using the MOLMOL software [25]. The side-chains were drawn using SCWRL software [26]. ............................................................................................. 171 (A) CD spectra of PIP" at 0 °C, 60 °C, 80 °C, and the summed spectra of P1 and P11 at 0 °C. (B) CD spectra of P1P] at 0 °C, 60 °C, 80 °C, and the spectrum of P; at 0 °C. (C) CD spectra of PUP" at 0 °C, 60 °C, 80 °C, and the spectrum of P" at 0 °C. ......................................................................... 174 Figure 5. 4 CID-MS/MS spectra of each peptide model: (A) the P1 monomer, (B) the PIP] homodimer, (C) the P” monomer, (D) the ................................................. 180 Figure 5. 5 CID-MS3 spectra of the two dominant MS/MS ions from P; (A) (B) and Pu (C) (D). The charge state of doubly charged ions is shown as (+2). ........ 186 Figure 5. 6 CID-MS3 spectra of the two dominant MS/MS ions from P1P“ (A) (B), PIP] (C)(D), and PnPu (E) (F). The charge state of doubly charged ions is shown as (+2). ....................................................................................................... 190 xi ABBREVIATIONS a—CHCA: a-Cyano-4-hydroxycinnamic acid l-S: l-disulfide 2, 5-DI—IB: 2,5-dihydroxybenzoic acid 2-S: 2-disulfide 3-D: 3-dimensional ACN: acetonitrile AEMTS: 2-aminoethyl methanethiosulfonate AFU: autonomous folding unit BPTI: Bovine pancreatic trypsin inhibitor BPTI: bovine pancreatic trypsin inhibitor CAD/CID-MS: collisionally induced/activated dissociation mass spectrometry CD: Circular Dichroism CDAP: 1-cyano—4-dimethylamino-pyridinium CI: chemical ionization CRM: charge residue model DTT: dithiothreitol ESI: electrospray ionization FA: formic acid FAB: fast atom bombardment FT -MS: Fourier transform-mass spectrometry Gdn-HCl: Guanidine hydrochloride xii GSSG: Glutathione H-bonds: hydrogen bonds HDX-MS: hydrogen deuterium exchange-mass spectrometry HPLC: high-performance liquid chromatography HX: hydrogen exchange ICR: ion-cyclotron resonance IGF-I: insulin-like grth factor-I LC-ESI-MS: liquid chromatography-electrospray ionization-mass spectrometry LR3IGF-I: Long R3 insulin-like grth factor-I MALDI-MS: matrix assisted laser desorption/ionization-mass spectrometry MS: mass spectrometry N: native NMR: Nuclear Magnetic Resonance NMR: nuclear magnetic resonance NOE: nuclear Overhauser enhancement PID: photo-induced dissociation ppm: parts per million PSD: post-source decay analysis R: completely reduced RNase A: Ribonuclease A SLD: soft laser desorption TCEP: tris(2-carboxyethyl) phosphine TFA: trifluoroacetic acid xiii TOF: time-of-flight xiv CHAPTER 1 INTRODUCTION 1. Introduction Biological macromolecules such as proteins are the “main actors” in the makeup of life [1]. Mass spectrometry (MS) has long been used as an important tool for the study of small and medium-sized biomolecules. The revolutionary breakthrough that makes this technique an indispensable tool for macromolecule analysis is the discovery and development of two ionization techniques — electrospray ionization (E81) [2] and matrix-assisted laser desorption/ionization (MALDI) [3]. ESI uses a strong electric field to facilitate spraying the sample, and, thus, produce free-hovering ions [2]; in contrast, MALDI subjects the sample molecules to an intense laser pulse that releases them as free ions in the form of protonated molecules [3]. These developments partly made chemical biology the “leading science” of our time. ESI-MS and MALDI-MS have become among the most popular tools for the understanding of the basis of life — the identity, 3- dimensional (3-D) structure, biofunction, and interaction of proteins and other biological macromolecules [1 ]. Proteins are macromolecules built of combinations of 20 different amino acids linked together by amide bonds in long chains. They are the major functional molecules in a living system [4]. For a protein chain to achieve its specific bioactivity, it must be folded into a very specific globular form [5, 6]. This phenomenon naturally gives rise to the question, “What is the source of the information responsible for this specific folding of the peptide chain?” Ever since C.B. Anfinsen’s Nobel Prize-winning work that showed that the linear sequence of amino acids in a protein determines its biologically active conformation [7], great efforts have been made to understand the rules that govern the folding (and unfolding) pathway of various proteins. Although different mechanisms for the folding of various proteins have been proposed during the last few decades, a general solution to the protein-folding problem is still unknown [8]. The most popular strategy to study the folding mechanism of a protein is to trap and characterize the folding intermediates that appear during the folding process. However, such efforts are often hampered because many intermediates appear only transiently during the course of folding and are therefore difficult to isolate and characterize [9-11]. In the case of disulfide-containing proteins, some intermediates can be trapped by chemical modification of free sulfhydryl groups (-SH) before their isolation and characterization [12-15]. During the last decade, a cyanylation/cleavage methodology was developed and applied to study the oxidative refolding of a few biologically important cystinyl/cysteinyl proteins [16, 17]. In these studies, a picture of disulfide-bond formation along the oxidative folding pathway was proposed based on the determined disulfide structure of each trapped intermediate. The folding process of cystinyl/cysteinyl proteins involves both disulfide-bond formation and conformational changes. During the last decade, hydrogen/deuterium exchange coupled with mass spectrometry (HDX-MS) has become more and more popular in the study of protein conformation and conformational dynamics [18-20]. The main aim of this dissertation is to describe the application of HDX-MS to study the conformational changes during the oxidative refolding of cystinyl proteins as a firnction of their disulfide-bond formation. In Chapter 2, various experimental conditions will be compared to optimize the conditions for monitoring HDX in proteins by MALDI-MS. In Chapter 3, using Long R3 insulin-like growth factor-I (LR3IGF-I) as a model protein, the feasibility and advantages of integrating the cyanylation/cleavage methodology and HDX-MS to study the oxidative refolding of cystinyl proteins will be demonstrated. In Chapter 4, the application of HDX-MS to the characterization of small peptides will be investigated. In Chapter 5, the identification and characterization of an autonomous folding unit (AFU) for LR3IGF-I will be discussed, and the folding process of LR3IGF-I will be firrther investigated by designing peptide models and then characterizing these peptide models using our established methodologies. II. Disulfide Mapping of Cystinyl/Cysteinyl Proteins Among the 20 different amino acids that compose proteins, cysteine is the one with unique properties. In the reduced form, cysteine contains a free sulfhydryl group; while in its oxidized form, cystine, a disulfide bond is formed by linking two sulflrydryl groups together (Figure 1.1). C02“ . , H020 _ c0211 2 HS oxrdatron > Y S S < . reductlon NHz NHz NH2 cysteine cys tine Figure l. l Cysteine and cystine. The oxidation status of cysteine residues is very important to the bioactivity of a protein. For example, the free sulfhydryl groups in cysteine proteases usually work as the active sites for enzyme catalysis; in addition, they are often chelating sites for metals in proteins [21]. Disulfide bonds, on the other hand, usually help to maintain the proper 3-D structure of a protein, and, therefore, its bioactivity. For instance, although the two isomers produced during the refolding of insulin-like growth factor-I (IGF-I) bear exactly the same amino acid sequence and are only different in their disulfide structure, they show significant difference in their 3-D structure and bioactivity [22]. Due to the important roles that free cysteine sulflrydryls and cystine disulfide bonds play, the unambiguous determination of disulfide structure for a protein is a very important aspect in protein analysis. A. Conventional Proteolytic Fragmentation Strategy Modification of fiee sulflrydryl groups Enzymatic cleaiiage of a protein Fractionation of peptides by HPLC Reduction of disulfide bonds Modification of nascent free sulfhydryl groups Separation of the resulting peptides Sequencing by Edrnan degradation/mass spectrometry Figure l. 2 Conventional proteolytic fragmentation strategy for disulfide bond assignment of a protein. Figure 1.2 illustrates a well-established proteolytic fragmentation strategy for disulfide-bond assignment in cysteine-containing proteins [23, 24]. First, a protein is cleaved by enzymes between half-cystinyl residues under conditions that minimize disulfide scrambling (or disulfide bond exchange, Figure 1.3). The resulting digests are then separated using reversed-phase high performance liquid chromatography (rp- HPLC). Finally, the amino acid sequence of the resulting peptides is determined by Edman degradation or mass spectrometry, and the identified peptides are related to specific segments of the protein. The characteristic aspects of the strategy are discussed below. Figure 1. 3 Disulfide/sulflrydryl exchange in proteins. Ideally, the chosen enzymes for cleavage of cystinyl/cysteinyl proteins should have strict specificity to produce well-defined fragments, thus simplifying the identification process. Second, the chosen enzymes should have maximum activity under conditions favoring minimal disulfide scrambling. Third, the chosen enzymes should have the capability to cleave between every half-cystinyl residue to obtain peptides containing no more than one disulfide bond. In practice, however, compromise usually has to be made among these three requirements. Trypsin has been the most frequently used enzyme so far due to its high specificity for Arg and Lys residues. Unfortunately, the activity of this enzyme is maximized at pH 8.3, and thus, disulfide scrambling often occurs during the digestion [25, 26]. If a protein cannot be cleaved by specific enzymes, it must be cleaved by non-specific enzymes. Pepsin (with maximum activity around pH 3.0), for example, is one of the most popular non-specific enzymes used for these studies [27-30]. However, when using non—specific enzymes, it is difficult to relate the disulfide- bridged peptides to specific segments of the protein simply by determining their molecular masses by MS; in this case, more complicated and tedious procedures such as Edman degradation have to be used to identify the proteolytic fragments. Following the digestion of a protein, the next step is to purify and to identify each disulfide-bonded peptide. Rp-HPLC under acidic conditions is usually used for this purpose. This is because this method is rapid and employs conditions where disulfide scrambling is minimized (i.e., a gradient of acetonitrile (ACN) in 0.1% trifluoroacetic acid (TFA)). The purification of digests from large proteins to homogeneity may still be challenging because many peptide fragments may be produced during digestion. Conventionally disulfide-containing peptides can be identified using HPLC by comparing chromatograms of digests before and after reduction of the disulfide bonds. A peak corresponding to a disulfide-containing proteolytic fragment in the chromatogram of the unreduced digests should be replaced by two or more new peaks in the chromatogram of the reduced digest. Each new peak corresponds to a component peptide of the original unreduced proteolytic fragment. In the case that a proteolytic fragment has an internal disulfide bond, the hope is that there will be at least a small shift in the elution time upon reduction. The amino acid sequence of the peptides can be confirmed by conventional methods such as amino acid analysis [31]. This procedure, however, not only is tedious and time consuming, but also can be unreliable because it depends too much on the changes of peptide hydrophobicity upon reduction of the disulfide bonds. Modern mass spectrometric techniques such as MALDI-MS and ESI-MS have made the task of disulfide-bond assignment in proteins more tractable [32, 33]. As shown in Figure 1.4, disulfide-linked peptides are identified based on their molecular masses, and by comparing the mass spectra of the sample before and after reduction. Specifically, in the spectrum of the reduced sample, the signal corresponding to the disulfide-linked peptide(s) is replaced by two signals corresponding to each of the component cysteine-containing peptides; if a proteolytic fragment contains an intra— molecular disulfide bond, a mass shift of two mass units should be observed. A major advantage of mass spectrometry in locating disulfide bonds in proteins is that it does not require rigorous purification of peptides before identification. Moreover, modern mass spectrometric techniques such as tandem mass spectrometry and post-source decay analysis (PSD) permit fast and simple sequence identification that can confirm a specific peptide segment in a protein. Disulfide-bonded protein 1 Enzymatic digestion Mixture of peptides 8“ x/ A /;\ \‘S:S7/ s . s \ ' \\ l + other peptides \\ \ / \ \ Mass spectrum of unreduced digest / \ l // / + other peptides Mass spectrum of reduced digest SH X/ / A \‘SH H31 SH. SH c ' I \ l I / \ Figure 1. 4 Conceptual representation of conventional protocol for disulfide-bond assignment based on mass spectrometry. B. Novel Cyanylation/ Cleavage Methodology There are two major problems in the conventional proteolytic fragmentation strategy for locating disulfide bond linkages. One is disulfide scrambling that often occurs in mild alkaline solutions. Because in practice it is often necessary to use a specific enzyme to produce recognizable fragments, and because most specific enzymes only work under alkaline conditions, the scrambling problem is usually difficult to avoid [25, 26]. The second problem is that peptides from which the locations of disulfide bonds can be deduced are not always produced after digestion. This is especially a problem for proteins containing adjacent cysteines [25, 34]. For instance, trypsin cleaves specifically at the C-terminal side of Lys or Arg unless followed by Pro. Therefore, if two half-cystinyl residues are not separated by Lys or Arg, tryptic fi'agments containing more than one disulfide bond will be produced, and thus unambiguous disulfide assignment is very difficult. In addition, the -Cys-Cys- structure in a protein is usually resistant to any enzyme [33]. Due to these two problems, disulfide structure assignment by the conventional approach remains a challenge for protein chemists. The cyanylation/cleavage strategy is a novel analytical approach for the assignment of disulfide bond linkages in proteins, and it is much simpler and more efficient than conventional methodologies. All steps before blocking of free sulflrydryl groups (and thus, “freezing” the disulfide structure of the protein) are under acidic conditions (pH 3), and, therefore, minimize, if not completely avoid, the problem of disulfide scrambling. In addition, this method is especially useful for highly knotted, cysteine-rich peptides such as sillucin [35], which are not readily analyzed by conventional methodology. Figure 1.5 illustrates a chemical overview of the methodology [36]. The basic idea of the cyanylation/cleavage analytical strategy is as follows: First, the protein of known sequence is partially reduced by a reducing reagent such as tris (2-carboxyethyl) phosphine (TCEP) hydrochloride. The nascent partially reduced protein isomers are immediately cyanylated by 1-cyano-4-dimethylamino- pyridinium (CDAP) tetrafluoroborate. The cyanylation products are then separated and collected using HPLC. Finally, the peptide is cleaved on the N-terminal side of the cyanylated cysteines, and any residual disulfide bonds that may connect some of the cleavage products are reduced. A specific partially reduced isoform usually produces a specific set of cleavage products; therefore, the disulfide structure of the isoform can be deduced by identifying its cleavage products, which usually can be achieved simply by measuring their molecular weights. The chemistries of cyanylation and cleavage reactions are illustrated in Figure 1.6. In the last several years, the disulfide-bond structure of various peptides and proteins has been determined by this strategy [35-37]. This, in turn, demonstrated the feasibility of the cyanylation/cleavage approach under various conditions. 10 Figure 1. 5 Chemical overview of the cyanylation/cleavage strategy. Partial reduction means that the protein of interest is transformed into a mixture of isoforms, each of which corresponds to reduction of only one of its disulfide bonds; for a protein containing n cysteines, n isomers of the singly reduced protein will result. Itz stands for iminothiazolidinyl carboxyl residue at the amino terminus [redrawn from [36]]. 11 40 1 10 30 I 50 1" S ¢ Partial reduction; TCEP at pH 3 H H F 20 F 40 F 20 F 40 1 10 I 30 I 50 1 10 I 30 50 S S SH SH 1 Cyanylation; CDAP at pH 3 CN CN 20 40 20 40 1 10 30 I so 1 10 I 30 I 50 n S SCN SCN * HPLC fractionation + MALDI-MS * Cleavage; 1 M NH4OH 1 20 9 11240 1 0 50 I I 11210 S\ 29 1 S\ 19 S S I I 11230 40 50 11220 30 39 Reduction; TCEP I 20 9 11240 20 50 Itle 'SH 29 1 [H 19 I“ I” 112 30 40 50 11220 30 39 (Cysro “ Cys3o) (Cyszo ‘ CYS40) Figure 1. 5 12 - / Cyanylatlon N C y<:\>—NMe2 pH 3; 25°C, 10' _ _ BF4 CDAP N=C_S Cleavage 1 M NH4OH, 1hr, 25°C /0 “Y S o C\ II NH2 H truncated peptide iminothiazolidine (itz)-blocked peptide Figure 1. 6 Chemistry involved in the cyanylation and cleavage reactions. 13 III. Oxidative Folding of Cystinyl/Cysteinyl Proteins and Trapping of Folding Intermediates Protein folding is a subject that has been under investigation for several decades [8]. Although a few folding mechanisms have been proposed and each of them has some supporting experimental evidence, none can yet explain all the experimental observations. Thus, no single folding theory can be regarded as authoritative [8]. At the current stage, research in this area is still focused on the folding study of individual proteins, in the hope that investigation on the folding pathway of various proteins will lead to the final establishment of a general folding mechanism. The major difficulty in studies of protein folding is that many folding intermediates are short-lived and/or exist as a distribution of possible conformers, which usually makes the isolation and characterization of folding intermediates a challenge [9-11]. Protein folding is the process by which a polypeptide chain acquires the proper 3- D structure to achieve its biologically active native state [5], and the folding of a cystinyl/cysteinyl protein involves both changes in its 3-D structure and evolution in its disulfide bond connections. Observations in numerous experiments have illustrated that during the folding process of cystinyl/cysteinyl proteins, the formation of disulfide bonds is usually coordinated with protein conformational changes [38-42]. In other words, disulfides reshuffle (break and reform) readily in order that a protein (or a segment of the polypeptide chain) can adopt conformations that are more energetically favorable. Disulfrde folding intermediates have been proven to be amenable to chromatographic purification, which makes possible the further characterization of these intermediates for information about the folding pathway [12, 16, 17]. To prevent further disulfide 14 rearrangement after the folding process is quenched, folding intermediates are usually trapped by chemical modification of its free sulflrydryl groups before separation and purification. A. Disulfide Formation in Oxidative Folding of Cystinyl/Cysteinyl Proteins Even when two free sulfliydryl groups are in close proximity, they do not form a disulfide bond spontaneously without an appropriate oxidant. Two types of oxidants are usually used for this purpose — oxygen (with the presence of trace amount of heavy metals) [43, 44] or a low molecular weight disulfide [17, 45]. Both types of oxidants are commonly used in practice, although oxygen-assisted disulfide-bond formation is usually much slower, and its mechanism is still not clear. When a low molecular weight disulfide is used, it is proposed that the disulfide bond formation is a sulfl'rydryl/disulfide exchange process between the oxidant and the protein [46]; Figure 1.7 illustrates this process. First, one sulfliydryl group in the reduced protein is deprotonated to produce a nucleophilic thiolate anion group. This group (Sf) then attacks one of the sulfurs of the disulfide bond in the low molecular weight disulfide. In the transition state, there is approximately equal bond formation to the nucleophilic sulfur (S 1') and the leaving sulfur (RLS'). In the next step, a mixed disulfide bond forms between the attacking thiolate anion group (8") and the central sulfur (RCS), and the leaving group (RLS') leaves the complex as a thiolate anion. The third step is an intra-molecular disulfide rearrangement. In this step, a second sulflrydryl group in the reduced protein is deprotonated to form a new attacking nucleophilic group (S 2'); it attacks sulfur S I in the mixed diSulfide bond. 15 + ?— 2 SH RC - ll 88’ s I 1 I 55- RC 2 SH 1 S—S RL I + _| RC 3 2 SH ¢T , "‘1 RC 2 S- - II -. I _$ ..... f5 S RC _ 2 5‘ J I S " | + I 2 S C Figure 1. 7 The formation of a disulfide bond upon adding a disulfide reagent to a reduced protein. 16 Then a disulfide bond between sulfurs S2 and S2 forms through a transition state analogous to that in the second step [46-48]. The observed rate of disulfide formation depends on several factors. The rate increases with pH because sulfhydryl/disulfide exchange starts with the attack of a nucleophilic thiolate anion to an existing disulfide bond, and more sulflrydryl groups exist as reactive thiolate anions at higher pH. The rate constant of the nucleophilic substitution reaction increases as the basicity of the attacking thiolate anion increases or that of the leaving group decreases. In addition, the stability of the mixed-disulfide intermediate and the tendency of thiolate anions of different cysteine residues to come into proximity of the mixed disulfide (which is mainly controlled by the conformation adopted by the segment of the polypeptide chain) are also important factors that control the observed rate of disulfide formation [46-48]. Two types of low molecular weight disulfides are usually used, linear disulfide and cyclic disulfide. Glutathione (GSSG) and dithiothreitol (DTT) are representative of each type. The GSSG/GSH (reduced glutathione) system plays an important role in disulfide bond formation in proteins in vivo [49], and they distribute universally in nature (i.e., in tissues of animals, plants, and microorganisms) [50]. Glutathione is often used in disulfide bond formation in vitro; however, the formation of significant amounts of non- native stable mixed disulfide species usually complicate the experiments [17, 45]. DTT, on the other hand, produces energetically unfavorable mixed disulfides that dissociate rapidly to produce disulfide bonds between cysteine residues in the polypeptide chain. DTT, therefore, is the preferred oxidant in some oxidative folding experiments [45]. 17 Once a disulfide bond is formed during the folding process, it can undergo intra- molecular rearrangements via a process analogous to the intramolecular step illustrated in Figure 1.7. The rate at which such rearrangements transpire is determined by the extent to which the conformation of the polypeptide chain favors a nucleophilic attack of a thiolate anion on the existing disulfide bond (Figure 1.3). Therefore, this step provides very useful information about protein conformation and folding. B. Trapping of the Folding Intermediates ' To study the oxidative refolding mechanism of a cystinyl/cysteinyl protein, the folding process is quenched at pre-determined time intervals, and various folding intermediates are trapped in chemically stable forms that can be separated, purified and characterized. Both inter- and intra- molecular disulfide rearrangements can take place only if there are attacking nucleophilic thiolate anions in the system. The general idea behind all kinds of trapping strategies is to deactivate all the reactive thiolate anions in the system by modifying/blocking them chemically so that they cannot attack any existing disulfide bonds in the system. This prevents any further disulfide rearrangement. A good trapping reagent should be able to block fiee sulflrydryl groups rapidly and completely without modifying other sites in the protein [51, 52]. Several common trapping strategies and reagents are discussed below. 1. Acid Trapping The pKa of the typical sulflrydryl group is near 8.7 [12]. Therefore, sulfhydryl/disulfide exchange can be significantly slowed by protonating reactive thiolate anions, which is usually done by adding acid to the folding solution until its pH value reaches 3 or even less (Figure 1.8). 18 HS pH<3 S » Figure 1. 8 Trapping folding intermediates by acidification. Acid trapping has been applied to the folding study of several proteins [12, 17, 41, 43, 53, 54]. This method is rapid and occurs at the diffusion control rate (109 M'IS'I) [55]. Its major advantage is that it does not require the modification of the free sulflrydryl groups in folding intermediates, thus conformational changes induced by any blocking group are not a concern. Another advantage of acid quenching is its reversibility. In other words, an acid-quenched intermediate can be purified, and then further disulfide rearrangement or folding of this isolated intermediate can be resumed after appropriate adjustment of pH. This is an invaluable tool for protein folding study because it can provide some detailed insight into the folding mechanism that is difficult to obtain otherwise. This strategy has been applied to a few folding studies [12, 17, 54, 56]. The rate constant for disulfide fonnation/rearrangement decreases ten times for each one-unit decrease in the pH value of the system. Therefore, the major disadvantage of this method is that it only slows down, but does not completely stop, the dynamics of disulfide bond formation in a cystinyl/cysteinyl protein. Thus, samples should be analyzed in a reasonably timely manner after the folding process is quenched by 19 acidification [55]. In addition, if the disulfide structure of an acid-trapped species needs to be determined, then its free sulflrydryl groups must be chemically modified. 2. Trapping by Alkylation Alkylating all free thiol groups in the system is by far the most widely used trapping strategy. Many alkylation reagents, such as iodoacetate, iodoacetarnide, and 4- vinyl pyridine have been used to block reactive thiolate anions [57-60]. Iodoacetate is most commonly used. No further sulfhydryl/disulfide exchange can occur after all the free thiol groups in the folding intermediates and the reduced protein are alkylated (Figure 1.9). ICH2COO' ———-> Figure 1. 9 Chemical trapping of folding intermediates by alkylation. The half-life for the reaction between 1 mM GSSG and a sterically accessible sulflrydryl group is about 30 s at pH 8.7 [61]. Using 0.2 M iodoacetate, the half-life of the alkylation reaction of an exposed thiol group is about 0.3 s [62]. The alkylation reaction is usually much faster than many inter- and some intra- molecular thiol/disulfide exchange reactions. However, in the folding study of some proteins such as bovine pancreatic trypsin inhibitor (BPTI) and ribonuclease, rearrangement of intermediates during trapping with iodoacetate has been reported [12, 52]. In these cases, the rate of 20 alkylation is significantly slower than that of some intramolecular disulfide rearrangements. In quantitative analysis, this will cause a problem of underestimation of the intermediates that react comparatively slowly with the alkylation reagent (or overestimation of those intermediates that react comparatively fast with the alkylation reagent). A much higher concentration of alkylation reagents can be used to accelerate the alkylation reaction with the risk that functional groups other than thiols may also be modified [63]. 3. Trapping by AEMTS In order to solve the problems of the conventional alkylation reagents mentioned above, a novel reagent, 2-aminoethyl methanethiosulfonate [(NH2)C2H5SSO2CH3] (AEMTS) was proposed [51, 64]. This reagent reacts with free sulfliydryls rapidly, and one or two minutes are usually long enough to completely alkylate all the free thiol groups without modifying other sites in a protein. Similar to the case of acid quenching, this reagent blocks thiol groups reversibly, and thus, the folding process can be resumed for an isolated intermediate. In addition, AEMTS brings one positive charge to each free thiol group in an intermediate (Figure 1.10); therefore, the more free thiol groups a disulfide species has, the more positive charges it bears after modification. This has been exploited in the separation of complicated folding intermediates by ion-exchange chromatography [65, 66]. 21 +£2|£ol£olmlm .mHEmZ .3 8860::25 wcfiom mo wEaEmb 03533” A: A PSwE o __ .mmzlmmolmmol m1 m lame + L 22 AEMTS has been widely used in the folding study of cystinyl proteins in recent years. Characterization of the 3-D structure of folding intermediates trapped by this reagent has also been reported [67]. In this case, it seems that this rather large blocking group, the 2-aminoethyl methanethiol group [(NH2)C2H5S-], did not significantly effect the 3-D structure and bioactivity of the protein. However, it should still be a concern whether or not this reagent is the best choice if the study focuses on the 3-D structure of folding intermediates. 4. Trapping by CDAP The cyanylation reagent, CDAP, has also been used to trap various folding intermediates in a few studies [16, 17]. After pre-determined time intervals, the folding process is quenched by adding 20-fold more CDAP reagent to the solution and adjusting the pH value is adjusted to 3 by the hydrochloric acid (HCl) in the CDAP solution. The trapped folding intermediates then can be separated and purified using r'p-I-IPLC, and their disulfide structure can be determined using the cyanylation/cleavage methodology (see section II. B.). The major advantage of this method is that the cyanylation of free sulflrydryl groups in the intermediates not only stops any further disulfide exchange, but is also the first step for disulfide assignment based on the cyanylation/cleavage methodology. The cyano- (CN-) group is rather small compared with most blocking groups, and it has been reported not to interfere with the 3-D structure of the trapped intermediates [68]. Thus, CDAP is a good choice if the study is aimed at investigating conformations of the folding intermediates. 23 Using CDAP, the cyanylation of all free sulflrydryl groups is usually complete within 5 to 10 minutes. Thus, similar to the conventional alkylation reagents (as discussed in section 111.82.), the rather slow blocking reaction could cause bias in quantitative analysis. IV. Conformational Study of Proteins and Peptides A polypeptide chain must take a certain 3-D structure for the protein to have its proper bioactivity [6]. During the folding of a cystinyl/cysteinyl protein, the conformation of the polypeptide chain usually changes cooperatively with its disulfide bond formation [38-42]. Therefore, conformational study of the trapped intermediates is an indispensable aspect in the folding study of cystinyl proteins. Some commonly used tools for protein conformational studies are discussed below. A. Conventional Technologies in Protein/Peptide Conformational Studies 1. The X-ray Diffraction Methods X-ray crystallography is among the most powerful and commonly used tools in the investigation of protein 3-D structure. In X—ray crystallography, a crystal of the molecules to be visualized is exposed to a collirnated beam of monochromatic X-rays, and the consequent diffraction pattern is recorded by a radiation counter or on photographic film. The distribution of the electron density, and, therefore, the structure of the crystal, may then be calculated from the diffraction pattern by Fourier transformation [69, 70]. X-ray crystallography has been the most important method in enzyme studies since it has provided the experimental tools to provide our current knowledge of the 3-D 24 structure of proteins [69]. By May 2002, about 80% of atomic coordinate sets deposited in the Protein Data Bank had been determined by this technique [1]. At low resolution (4 to 6 A), the electron density maps usually can only provide information on the overall shape of the molecule; on the other hand, with a high resolution of about 1.9 A and very— well ordered crystals, the positions of atoms can be fitted with an accuracy of :t 0.2 A [69]. The conformation of numerous proteins has been characterized using this technique, and it is almost always the first choice when only the static 3-D structure of a protein needs to be elucidated [71]. Crystallization of the protein molecules of interest is an uncircumventable requirement of X-ray crystallography; thus, for those proteins whose crystals are still not available, other methods have to be used to characterize their conformation [72]. In addition, the 3-D structure of a protein in the crystalline state could differ significantly from that in solution state [71, 73]. Because both in vitro and in viva reactions usually take place in a solution environment, the appropriateness of explaining observed biological phenomena using the static protein structure obtained from X-ray crystallography is always a concern. Furthermore, X-ray crystallography is not suitable for investigation of various dynamic processes such as those that occur during protein folding [71, 73, 74]. 2. Nuclear Magnetic Resonance (NMR) Spectroscopy Nuclear magnetic resonance (NMR) refers to the absorption of electromagnetic radiation of a specific frequency by certain atomic nuclei placed in a static magnetic field and simultaneously exposed to electromagnetic irradiation with certain frequency [75- 78]. The NMR phenomenon results from the quantum mechanical property of nuclear 25 spin. Only those nuclei with nuclear spin I :e 0 (i.e., the nuclei of 1H or 13C) can absorb electromagnetic radiation under certain conditions, and, thus, produce a NMR signal. Common biological nuclei such as those of 12C, and I°O have I = 0, and, therefore, are not detectable by NMR [77]. NMR spectroscopy is the use of the NMR phenomenon to study physical, chemical, and biological properties of compounds [75]. The development of high-field NMR methods is the most important advance in protein structural determination in recent years [79]. Currently, NMR and X-ray crystallography are the major tools for conformational studies of macromolecules such as proteins [1]. The local magnetic field in a nucleus is sensitive to the types and closeness of its neighboring nuclei. The strategy of solving 3-D structures by NMR is to assign the many resonances in an NMR spectrum to their respective nuclei in the macromolecule and, thereby, determine which atoms are close to each other. The overall 3-D structure of a protein is calculated principally from constraints on the distances between the nuclei under investigation [79]. Some NMR parameters usefirl for this purpose, as well as different NMR methods, are discussed below. In a one-dimensional experiment, the sample is subjected to a constant magnetic field and to irradiation by a radio frequency electromagnetic field to generate an NMR spectrum consisting of a series of resonance lines that vary in many ways, including their positions and fine structure. The location of a resonance line is defined by a “chemical shift” that depends on the chemical environment around the nucleus of interest. This parameter is measured in parts per million (ppm) as a fraction of the static magnetic field by which the signal is displaced from that of a reference compound. The spins of nuclei 26 that are separated by a small number of covalent bonds (5 3) interact and split the resonance lines. These “through-bond” effects are called spin-spin (or J) coupling and are the cause of fine structure in an NMR signal. Nuclear Overhauser effect refers to NMR signal intensity changes produced by dipolar cross-relaxation from neighboring spins by radio frequency perturbation of the spin population [80] (named after American physicist Albert Overhauser, who hypothesized it in the early 19505), and the degree of increase in the signal is known as the nuclear Overhauser enhancement (NOE) [81]. This parameter measures “through space” interaction between nuclei. The effect is caused by the coupling of dipoles, the magnitude of which falls off as r'° (r is the inter-atomic distance between nuclei). Thus, information about inter-atomic distance between nuclei close in space can be obtained by studying this parameter. NOEs are usually classified as “strong”, “intermediate”, or “weak”. The assignments are made through identification of the “through-bond” and “through-space” interactions. Observation and quantification of NOE effects between nuclei in close proximity are the basis of determination of protein conformation by NMR [1, 79]. In multidimensional NMR, the sample is irradiated by two radio frequency fields in various sequences and pulses [1, 74, 79]. The 3-D structure of a protein is deduced from a combination of the restraints on the distances between nuclei and energy calculation. Certain mathematical methods (i.e., a method based on metric matrix distance geometry [82]) are used to calculate the 3-D structure for a protein based on the structural constraints obtained from NMR experiments, and a family of possible structures is usually obtained. In regions where considerable consistent information has 27 been obtained, the different solutions usually give essentially superimposable structures; on the other hand, the regions that seem particularly disordered are often highly mobile. The dynamics of a protein molecule is described by parameters such as the varying mobility, timescales, and amplitudes of motions of the mobile segments along the polypeptide chain. Knowledge of the conformation and the dynamics of a protein are essential for understanding how the protein molecule interacts with other molecules in its environment [ 1 ]. The major advantage of NMR, compared with X-ray crystallography, is its ability to yield insight into the dynamics of a molecule [1, 71, 73, 74]. In addition, NMR can be used to characterize the 3-D structure of proteins that fail to crystallize. However, NMR analysis is usually limited to moderate-sized proteins in solution (typically < 50 kDa) [83]. The quality of NMR data is usually compromised if the sample contains paramagnetic ligands [84]. Moreover, NMR analysis is very time-consuming, especially compared with the techniques discussed below. 3. Circular Dichroism (CD) Spectroscopy Plane-polarized light is composed of two vectorial components that are circularly polarized: one is right-handed (R) and polarized in a clock-wise direction; the other is left-handed (L) and polarized counterclockwise. When plane-polarized light is shone through an optically active sample at a wavelength at which the light is absorbed, the left- and right-handed circularly polarized light may not be equally absorbed by the sample. In other words, the L and R lights have different molar absorptivities, £2 and fig, respectively. This phenomenon is called circular dichroism [85]. 28 The molar absorptivity (a) is defined by Equation 1.1: 5 = — (Equation 1.1) The CD is the difference in molar absorptivity, As. A6: 82 - 63 (Equation 1.2) Thus, _ (AL — AR) cl Ar: (Equation 1.3) Where AL and A R are the absorbance of the left- and right-handed polarized light, respectively; c is the concentration of the solution, and l is the length of the light path. A L — AR is usually in the region of 0.005% to 0.1% of A2 or AR. The unit of As is cm M'l or 103 cmzmol'I. The output of many CD instruments is in ellipticity, 9, measured in degrees. 6= 33(AL — A R) (Equation 1.4) In most scientific papers, CD data are reported in mean residue weight ellipticity, [@IMRWI [@]mw = % (Equation 1.5) c Where N is the number of residues in the protein [85]. The backbone of a protein is optically active in the far ultraviolet range (170 — 250 nm), and specific secondary structures produce characteristic CD spectra. For example, the CD spectrum of or-helical structure is typically characterized by strong negative signals at around 208 nm and 222 nm, and a strong positive signal at around 192 nm. The characteristic CD signal of B-sheet structure is a positive signal near 195 nm 29 and a negative signal near 217 nm [86]. Typical CD spectra of a-helical and B-sheet structures are shown in Figure 1.11. In addition to qualitative analysis, in principle, quantitative results such as the amount of the helical content in a polypeptide chain can be obtained by comparing the ellipticity of the sample to a standard reference at a characteristic wavelength for helical structure. In addition to the contribution from the backbone of a protein, signals from natural chromophores sometimes also can be observed in a CD spectrum. This is because some natural chromophores exhibit idiosyncratic CD behavior when placed in a heterogeneous environment in a folded protein. For instance, tryptophan can give CD signals in the near-ultraviolet region (~270 — 300 nm) [85], and tyrosine in the 270 — 300 nm region [85] as well as in the far-ultraviolet region (~230 nm) [87]. CD spectroscopy has been widely used in conformational study of proteins and peptides [88]. It is a convenient tool for investigation of the 3-D structure of a polypeptide chain, as well as conformational changes of a protein upon binding or during the folding process. The major advantage of this technique is that CD experiments are simple, straightforward, and fast. However, the resolution of this technique is limited by the size of the polypeptide chain under investigation; in other words, there is no way to determine the location of the structure along the entire polypeptide chain solely based on CD results. In addition, although a-helical structure usually produces a strong and distinctive CD signal, the CD technique is not a highly effective tool for the study of B-sheet or other secondary structure. This is because the CD signal of B-sheet and other secondary structure is usually much weaker and less distinctive [85]. Moreover, the accuracy of quantitative CD analysis is usually disappointing. CD spectroscopy is often used for preliminary 30 experiments to obtain a rough idea of a protein’s conformation before starting more detailed and more time-consuming analysis such as NMR analysis. or-helix B—sheet Random coil 190 200 210 220 230 240 250 Wavelength (nm) Figure 1. 11 Representative CD spectra of a-helical, B-sheet, and random coil structure in proteins (redrawn from [89]). B. HDX-MS The so-called hydrogen/deuterium exchange technique has for a long time been used to study the 3-D structure of proteins and peptides, as well as their conformational changes [18-20]. Among all hydrogens in a polypeptide chain, those bound to carbon atoms essentially do not undergo exchange, while those bound to oxygen, sulfur, or side- 31 chain nitrogen undergo isotopic exchange too rapidly to be measured by current analytical techniques. In contrast, hydrogens located on the nitrogen atoms in the peptide amide linkages exchange with deuterated solvent at a rate suitable for analytical measurement. When a protein is dissolved in D20, amide hydrogens are replaced slowly by deuteriums if they are located either in well-defined secondary structures such as or- helixes and B-sheets (where the amide hydrogens are usually involved in intramolecular hydrogen bonds (H-bonds)), or in the hydrophobic core of the protein (where access to solvent is limited). In contrast, amide hydrogens located at unfolded regions exchange much more rapidly due either to weakened H-bonds or to much better access to the deuterated solvent. Thus, conformational information of a polypeptide chain can be obtained by assessing the deuterium incorporation level of the protein after HDX; the fewer deuteriums incorporated by the polypeptide chain (or a segment of the polypeptide chain), the more folded the polypeptide chain (or the segment) is, and vice versa [90]. When an amide hydrogen is replaced by a deuterium, its corresponding NMR signal disappears. Thus, NMR can be used to monitor deuterium incorporation. For more than two decades, HDX has been coupled with NMR in many folding studies to investigate conformational changes of proteins and peptides [91]. The resolution of HDX-NMR is at the single amino acid residue level. This method usually requires milligram-scale samples for analysis; however, in protein folding studies to harvest folding intermediates at that scale is not always possible. Its other disadvantages include molecular mass limitations (usually less than 50 kDa) and difficulties in analyzing complexes containing paramagnetic ligands [92, 93]. 32 The atomic mass of hydrogen is ~1 Da, while that of deuterium is ~2 Da. Therefore, a mass shift of ~1 Da should be observed if an amide hydrogen is substituted by a deuterium. MS has long been applied to measure molecular weights and mass changes of proteins and peptides [94, 95]. In recent years, HDX-MS has been successfully used to study the conformational changes in a number of protein systems [18-20]. Due to the superior detection limit of MS, the sample quantity required for HDX-MS is at least a thousand times less than that for HDX-NMR analysis. In many cases, this feature permits the study of intermediates of very low concentration, which might be very important for refolding kinetics. In addition, the upper mass limit of this approach is yet to be established, and paramagnetic ligands present no difficulty to MS. In HDX-MS, conventionally the deuterated protein is often dissected by proteolytic digestion before the assessment of deuterium incorporation. (Pepsin is usually the enzyme of choice for reasons that will be discussed in Chapter 3). Compared with HDX- NMR, the major disadvantage of HDX-MS is that the resolution of this methodology is normally limited by the size of peptic fragments (5 — 12 residues, on average) [92]. However, in a few cases, the resolution has been improved to the single amino acid residue level with the application of tandem mass spectrometry (i.e., collisionally induced dissociation-mass spectrometry/mass spectrometry (CID-MS/MS), see details in the next section) [96-99]. 33 V. Desorption/Ionization Mass Spectrometry Since J .J . Thompson’s pioneering description of the ability to separate molecules based on different mass and charge in 1912, many different theories were proposed and various types of mass spectrometers were invented. All MS process consists of a chemical preparation of charged molecules in the gas phase and then physical separation of ions in vacuum. Although MS has long been used in the analysis of small and medium-sized molecules, the goal of analyzing large macromolecules remained elusive for over 70 years. The efforts were hampered by the idea that one always had to volatilize the molecules first and then ionize them, and by the fact that many macromolecules such as proteins are polar and thermally labile, and, thus, are not suitable for thermal volatilization. The major problem to be solved was how to produce intact gas-phase ions of a molecule from compounds that are usually solids and that degrade or decompose when heated [1]. The use of chemical ionization (CI) made it possible, for the first time, to ionize thermo-labile biomolecules [100]. In CI, a reagent gas is first ionized to form abundant reagent gas ions, which then collide with sample molecules for proton transfer. Another technique, plasma desorption, uses high-energy ions to desorb and ionize molecules of interest [101]. A milestone was passed when the fast atom bombardment (FAB) technique was reported, where a non-volatile chemical protection environment (the matrix, often glycerol) was used to dissolve the sample molecules to enable thermally labile compounds to survive the ionization process [102]. Despite some success with these techniques in mass determination of small biomolecules (molecular weight < 10 kDa), the challenge of how to analyze high-molecular—weight compounds by mass 34 spectrometry and to make mass spectrometry into a powerful detector for liquid separation techniques was not solved until the late 19805 when two novel ionization techniques, MALDI [3] and ESI [2], were reported. These soft desorption/ionization techniques have overcome the problem of the desorption and ionization of large and thermo-labile biomolecules, and, thus, have become indispensable tools for the identification and characterization of these compounds. The last decade has seen explosive grth in the applications of these two techniques in biochemical sciences [l]. A. Matrix-Assisted Laser Desorption/Ionization Time of Flight Mass Spectrometry (MALDI-TOF-MS) The use of laser light as an energy source to volatilize/ionize macromolecules was the focus of several key research groups in the 19805 [1]. In late 19805, M. Karas and F. Hillenkamp showed that a UV-absorbing matrix could be used to volatilize small analyte molecules [103], and the major breakthrough in the application of laser desorption to large biomolecules was reported by K. Tanaka in 1987 [104, 105]. Tanaka presented the ionization of proteins (12 — 35 kDa) using a low-energy (nitrogen) laser and a matrix of glycerol containing colloidal particles. The technique developed by Tanaka is called soft laser desorption (SLD). The currently predominant MALDI technique is a fast growing version of SLD, and its application to proteins was reported shortly after Tanaka’s initial breakthrough. The key to the MALDI process is to incorporate the analyte macromolecules in a low-molecular-weight crystalline matrix whose absorption maximum matches the wavelength of the laser pulse. The matrix molecules absorb energy from the laser pulse and transfer the energy to the analyte molecules, which 35 results in a soft desorption process of the analyte molecules; in other words, little or no fragmentation/decomposition of the analyte molecule occurs during desorption [106]. The major advantage of the MALDI technique is its ability to transfer intact low- charged macromolecules to the gas phase as protonated species. In addition, it also yields good results with contaminated samples. Although the mechanism of sample desorption and ionization has not been completely understood, MALDI, especially when combined with time-of-flight mass spectrometric detection, has found numerous applications in the study of biological macromolecules. 1. The MALDI Process Despite the widespread of analytical applications of MALDI, the mechanism for ionization and desorption of the analyte is still under investigation [107-116]. A comprehensive and general model for MALDI was proposed in recent years that can successfully explain some of the major phenomena observed in experiments [117]. It is assumed that the charge state of the analytes as acquired in solution due to solvent and pH is roughly conserved within the matrix crystals, and that some water is preserved in the surrounding of the charge site of the analyte, and, thus, counter ions may be only roughly associated with the analyte charge site to give overall electroneutrality. One role of the matrix is to separate the analyte molecules (by dilution) to prevent analyte-analyte molecular (or ionic) interactions during the ionization process. The key role of the matrix is in absorbing the radiation, thereby protecting the analyte from radiation damage. Provided that the energy density achieved is large enough, a sudden and explosive phase transition occurs, which is best described by the formation of a gas jet of matrix neutrals, which entrains analyte molecules. Initial species formed as a result of laser 36 desorption/ablation are clusters. They consist of matrix, analyte and ionic species embedded in the matrix crystal, all held together by H-bonds and coulombic interactions. These clusters are highly excited owing to the laser energy absorbed by the matrix and contain enough excess energy to evaporate/desolvate to molecular and ionic species either directly by evaporation of neutral species or after chemical reactions resulting in more stable neutral or ionic species. The charging and, thus, ionization process is the statistical occurrence of clusters with a deficit/excess of anions or cations (Figure 1.12). 0 Matrix O Analyte Figure 1. 12 Highly charged clusters are formed during disintegration of the analyte/matrix solid [117]. M stands for the neutral analytes, H+ for protons, and B' for anions (i.e., trifluoroacetate). 37 The initially generated charged clusters evaporate on a very short time-scale by loss of neutral matrix molecules or neutral reaction products. Upon matrix evaporation, ion—ion and neutral—ion chemical reactions take place, e. g., proton-transfer reactions and proton neutralization to neutral species; the successive steps of evaporation (e.g., HCl from C1' and xH+) are driven by the reaction enthalpy. 2. Matrix and Sample Preparation The choice of proper matrix is the key to successful analysis of high-molecular- weight macromolecules using the MALDI technique. The chosen matrix should exhibit maximum absorbance at the wavelength of the laser used. The low-energy nitrogen laser (337 nm) is the most widely used in commercially available MALDI mass spectrometers. Most currently available matrix compounds show strong UV absorption around this wavelength. In addition, the chosen matrix must have good solubility in the solvent required for dissolution of the analyte and be inert to the analyte and solvents. It must be miscible with the analyte in the solid phase, and be vacuum compatible (with low vapor pressure), and possess a chemical composition that promotes the ionization of matrix substituents that can donate protons to the analyte. Moreover, the proper matrix should crystallize readily and have a low heat of sublimation [118]. According to these criteria, among the hundreds of organic compounds that have been investigated, only a few are suitable for peptide and protein analysis [119, 120]. The structure of some of the most widely used matrices are shown in Figure 1.13. Empirically a-Cyano-4-hydroxycinnamic acid (or-CHCA) is usually the first choice for proteins/peptides less than 10 kDa, 2,5-dihydroxybenzoic acid (2, S-DHB) for peptide 38 analysis, and 3,5-dimethoxy—4-hydroxycinnamic acid (sinapinic acid) for proteins larger than 10 kDa [121]. OH CO H \CN HO OH or-CHCA 2,5-DHB CH=CH-C02H CH3. OCH3 0H Sinapinic acid Figure 1. 13 Chemical structure of compounds commonly used as matrices for MALDI. Several sample preparation methods have been developed for various analytes. Some of the most widely used methods are discussed below. In the dried-droplet method [103], a droplet of the matrix solution (1 —— 10 mmol) and a droplet of the solution containing the protein analyte (1 -— 10 umol) are dried on the sample plate (usually a 39 piece of metal). A common problem with this method is the aggregation of higher amounts of analyte/matrix crystals in a ring around the edge of the drop; these crystals are usually heterogenous and irregularly distributed. Another problem often observed during crystallization is segregation, which refers to the exclusion of salts and some of the analyte from the matrix as the solvent evaporates and the matrix crystallizes. Component segregation generates a heterogeneous mixture of analyte throughout the solidified sample, which results in highly variable analyte ion production as the laser moves across the sample surface. The “dried-droplet” method is the oldest and simplest sample preparation method. Despite its many disadvantages, it has remained the most common method of sample preparation in the MALDI community. The fast evaporation method is a variant of the “dried droplet” method. This method was developed in 1994 with the main goal of improving the resolution and mass accuracy of MALDI measurements [121]. First, a thin layer of the matrix is placed onto the sample plate. A droplet of the analyte solution is then applied on top of the dried matrix bed. As the droplet evaporates, it dissolves some of the matrix that then crystallizes onto the previous layer of matrix. This method, however, does not provide reproducible sample-to-sample data for peptide and protein mixtures, and, thus, it is not the first choice if the protein/peptide sample contains more than one component. In the year 2000, a novel method called the ultrathin layer method was reported by B. Chait [122]. In this method, saturated matrix solution is first applied over the whole surface of a sample plate. The matrix is then gently wiped with a tissue, leaving only a faint layer of matrix visible as a yellowish reflection when the plate is examined at an angle. The sample is dissolved in saturated matrix solution directly, and a small 40 aliquot (usually 0.5 11L) is spotted onto the pre-coated sample plate. A protein-matrix crystalline film forms readily on the ultrathin layer of previously crystallized matrix, leaving most of the salts, detergents and other components in the excess liquid to be removed by vacuum aspiration. The spot can be further washed with dilute acidic solution to remove residual salts and detergents. This ultrathin-layer method presents a homogeneous matrix-protein co-crystallized surface, allowing for reproducible data acquisition over the entire surface area. Compared with the conventional dried droplet method, this method shows higher sensitivity and accuracy, as well as better resolution. It is a robust, detergent-fiiendly method for a variety of proteins and mixtures including membrane proteins. Until today the choice of matrix and sample preparation method is still an empirical process due to a lack of understanding of the desorption/ionization mechanism. The quality of MALDI analysis results is controlled by many factors such as the choice of matrix and solvent, the additive used, and the crystallization conditions employed. Parallel analyses under different conditions are required for optimization of the MALDI experimental conditions. 3. Instrumentation A MALDI ion source is usually interfaced with a time-of-flight (TOF) mass-to- charge (m/z) analyzer. Figure 1.14 is a schematic drawing of a MALDI-TOP mass spectrometer. The sample consisting of co-crystallized analyte and matrix is irradiated by a pulsed laser that is directed and focused by a prism and optical lens. The intensity of the laser is controlled by an attenuator, and it is increased gradually until the threshold for 41 Beam Splitter QII 55-3:- Laser / | / / . T ' / ransrent ll / Trrgger C: Recorder / // ’/ Sample I/ I I I I I- Probews L 633 @665 603 } I, \ 5 ‘\ Flight Tube Li I 1011‘ K ‘\‘ Detector I SourceI \\ \ Protein \\ Matrix Probe Tip Figure 1. l4 MALDI-TOF-MS instrument [123]. ionization is reached. The interaction of photons with the matrix and analyte results in the desorption and ionization of the co-crystallized sample/matrix from the metal surface of the sample plate. A static electric field is then applied to the generated ions by applying a two-stage high voltage (typically i 25 kV, depending on positive or negative ion detection mode) to the sample probe with regard to a closely spaced accelerating electrode. The ions are thus accelerated to the same kinetic energy (assuming the initial kinetic energy is zero) by the electric field toward a long (1 — 2 m) field-free TOF analyzer [123]. 42 The TOF m/z analyzer is simple, inexpensive, and well suited to the pulsed nature of laser desorption in MALDI [124]. MALDI can produce ions with very high m/z values, and the TOF m/z analyzer has virtually no upper mass limit. Therefore, the combination of these two techniques (MALDI-TOF-MS) is very popular, especially in the analysis of macromolecules such as proteins and peptides. The time required for an ion to traverse the flight tube is dependent on its m/z value and is described by Equation 1.6. l TOF = E = L(—"-‘-eV)2 (Equation 1.6) U 22 Where L is the length of the flight tube, 1) and m are the velocity and the mass of the ion, respectively, and V is the acceleration potential [125]. According to Equation 1.6, ions with lower m/z value should have a shorter flight time than heavier ions. Thus, ions with different m/z are separated into a series of spatially discrete ion packets, each of which travels with a velocity characteristic of its m/z ratio. A detector is positioned at the end of the flight tube and produces a signal as each ion packet strikes it. A TOF spectrum is a record of the detector signal as a firnction of time. The difference between the start time, which is triggered by the laser pulse and is common to all ions, and the arrival time of each ion at the detector is proportional to (m/z)”. This is the basis for the conversion of ion flight time (TOF) into an m/z value along the x-axis in a TOF spectrum [125]. Ions of different m/z values produced by a single laser shot arrive at the detector at different times. Therefore, in contrast to a conventional scanning m/z analyzer, a TOF mass spectrometer can generate an entire spectrum from every single laser shot; thus, at least in theory, no information is lost. However, in practice, the spread of initial 43 conditions of ions with the same m/z value, such as the time/location of ion formation and initial kinetic energy distribution, is usually large, which results in poor resolving power of the linear TOF m/z analyzer [126]. Two techniques were introduced to solve the problem of poor resolving power, reflectron-TOF [127, 128] and delayed extraction of the ions [126]. The reflectron-TOF consists of two linear field-free regions and an ion mirror (Figure 1.15); the ion mirror compensates for the spread of initial kinetic energy of ions with the same m/z value. Ions of the same m/z value, but with a distribution of energies, travel at different velocities. After traversing the field—fi'ee region, they reach a series of grids and ring electrodes of increasing potential (the mirror or reflectron) that create a retarding field that acts as a mirror by deflecting the ions and sending them back through the field-free flight tube. Ions with more kinetic energy penetrate more deeply into the reflector region and spend more time in that area, while ions with less kinetic energy penetrate less deeply and spend less time. Thus, both types of ions arrive at the detector at the same time. Ion source f;::" it:::: % 3%: W A ...... ...... ...... ...... ...... ...... ...... ...... ...... IIIIII IIIIII llllll Detector Reflector region Figure 1. 15 Schematic diagram of ion trajectories in a reflectron-TOF m/z analyzer [129]. 44 The delay time is the interval between the laser pulse, which ionizes the sample, and application of the acceleration potential. During this time (usually nanoseconds), ions disperse owing to their initial velocity in the source region (the acceleration region just above the sample surface). Ions with higher initial kinetic energy move farther into the extraction field than ions having less kinetic energy, and, thus, acquire less kinetic energy when the extraction field is turned on; on the other hand, ions of lower initial kinetic energy, having moved not as far into the extraction region, traverse a longer segment of the electric field after the extraction potential is applied, and, thus, acquire slightly more kinetic energy than the other group, which allow them to “catch up” with the ions of more initial kinetic energy. The “time delay” dramatically improves the resolving power achievable in MALDI-TOF-MS. B. Electrospray Ionization Ion Trap Mass Spectrometry (ESI-Ion Trap-MS) Although the electrospray phenomenon has been studied for over two centuries, the use of electrospray as an ionization method for macromolecules was introduced by M. Dole only about four decades ago [130, 131]. He proposed the first model of the electrospray process, including the charge residue model (CRM) that has survived as a main explanation for the enigmatic ESI process. In 1984, the first combination of electrospray and mass spectrometry was reported by J. Fenn [132]. The well-defined breakthrough of ESI for MS came in 1988, when he presented an identification of intact proteins of 40 kDa with a molecular-weight accuracy of 0.01% [133]. Since then, the application of ESI-MS to macromolecule analysis has expanded explosively. Today, different types of commercial instruments are available, and ESI-MS has become a routine and indispensable tool in bioanalytical laboratories in both academia and industry. 45 1. The Electrospray Process in ESI-MS In ESI-MS, an analyte solution is pumped through a capillary biased at a high potential (2 — 5 kV). This potential provides the electric-field gradient required to produce charge separation at the surface of the liquid. As a result, the liquid protrudes from the capillary tip and forms a “Taylor cone” (Figure 1.16). Finally, droplets containing an excess of positive or negative charge leave the tip when the solution that comprises the Taylor cone reaches the Rayleigh limit, which refers to the point at which coulombic repulsion of the surface charge exceeds the surface tension of the solution [134]. These droplets travel through the electric field at atmospheric pressure toward the entrance to the mass spectrometer; the droplets diminish in size due to solvent evaporation and coulombic explosions with the production of gas phase ions (charged analyte molecules). A few mechanisms have been proposed to describe this process. According to the CRM (also known as coulomb fission mechanism), a charged droplet evaporates during the electrospray process, and the coulombic repulsion among the charges on the surface of the droplet becomes larger and larger until a Rayleigh explosion occurs, and a number of smaller dr0plets whose surface also contains electrostatic charges are produced. This process takes place over and over again, and the droplet is divided into smaller and smaller droplets that eventually consist only of single ions [130]. On the other hand, the ion evaporation theory assumes that the release of ions directly from the surface of the droplet takes place when the coulombic repulsion is large enough to overcome the liquid’s surface tension [135]. 46 Reduction Taylor cone «5 2 col 6? High Voltage Power Supply 5 Electrons * roc Figure 1. 16 Schematic of the electrospray ionization process [136]. Regardless of the mechanism by which ions are produced, the ESI process generates gas-phase ions that can be analyzed for m/z ratio in the mass spectrometer. The transition of ions from the liquid phase to the gas phase is very gentle, and the analyte molecule is likely still surrounded by at least one solvation shell when it occurs. Fragmentation during this event has only rarely been observed, and aspects of the tertiary structure of a protein might survive into the gas phase under certain circumstances [137]. 2. Instrumentation Figure 1.17 is a schematic diagram of a typical electrospray interface to a mass spectrometer. The electrospray capillary is typically a metal or glass capillary with a small i.d. (<100 pm). A high voltage (1 — 5 kV relative to the desolvating capillary) is applied to the capillary. An analyte solution is forced into the electrospray capillary at a flow rate of 1 — 100 uL/min and is nebulized by the resulting electrospray plume. A mixture of ACN and water (or methanol and water) is usually used as solvent, with a small percentage (< 5%) of acetic acid or formic acid (FA). The protein/peptide concentration is usually at 10’5 — 10'7 M. The most important application of ESI is as an interface between a mass spectrometer and separation equipment such as electrophoresis [138] or HPLC [139]. Ion Optics Electrospray 4:”: Sklmmer Capillary $22: ................... . ""51..;EZZZZIIIijj' Heated \ l”\ Capillary + High Voltage m/z Analyzer Figure l. 17 ESI-MS instrument interface. 48 Also worth mention is the growing interest in the use of nanospray in LC-ESI-MS analysis due to the greatly improved sensitivity, a result of the use of very small spraying orifice (l — 2 pm inner diameter) and very low flow rate usually at nl/min scale). Nanospray also achieves more efficient ionization and desolvation than does conventional ESI [140, 141]. Once the ions are formed and sprayed into the heated capillary, they need to be moved to the m/z analyzer. During this progression, the instrument needs to focus the ions of interest through several stages of differential pumping of decreasing vacuum pressure regions to reach the very high vacuum analyzer region. This is achieved in the assembly of heated capillary, Skimmers, and ion optics. The typical internal diameter of the heated capillary is between 400 pm and 1mm; the length is usually 10 — 25 cm. Its purpose is both to complete the desolvation process of the charged analytes and to baffle atmospheric pressure on the source side to about 1 torr. The heated capillary is electrically biased (with the same polarity as the ions of interest) to prevent analyte ions from contacting with the inside walls. In the next step, Skimmers are used to transmit ions from the heated capillary as much as possible while decreasing the pressure in each subsequent vacuum stage of the chamber. The Skimmers are also electrically biased to focus the ion beam and define its average kinetic energy. The use of ion optics can firrther enhance ion transmission and focus the ion beam into the m/z analyzer [142]. ESI is commonly coupled with the ion-trap mass spectrometer [143]. As shown in Figure 1.18, a quadrupole ion trap consists of a ring electrode and two endcap electrodes, between which is a small volume to confine ions. A sinusoidal potential (RF at a fixed frequency) is applied to the ring electrode, while the endcap electrodes is 49 usually maintained at an oscillating (AC) potential. Under a certain storage condition (specific instrumental parameters), only ions within a given range of m/z value can be stored in the ion trap; upon change in the storage conditions, the trapped ions can be expelled from the trap. Ions striking the detector transfer charge that registers as a current that is proportional to ion abundance. The high sensitivity and ability to be used for tandem MS experiments have resulted in the wide use of the ion-trap mass spectrometer [144, 145]. RF generator Entrance Exit endcap endcap Ion Injection .................... ’ .................... > Ion Ejection Ring Electrode L AC generator Figure 1. 18 Schematic diagram of ion-trap mass spectrometer [146]. C. The Comparison of MALDI and ESI As shown in Table 1.1, MALDI and ESI have different and complementary natures [147]. Together, they make a formidable set of tools with high levels of sensitivity, accuracy, and mass range. They are widely used for routine identification and quantitation of synthesized small molecules and proteins, as well as compounds obtained directly from biological matrices. In recent years, these techniques also have been applied to novel fields such as the study of non-covalent interactions, protein folding, in vitro drug analysis, and drug discovery. 51 Table 1. 1 Comparison of ESI-MS and MALDI-MS [147]. ESI-MS MALDI-MS Mass limit, Da >1,000,000 Da > 1,000,000 Da Ionization Continuous Pulsed Common mass Quadrupole, quadrupole ion TOF, FTICR, analyzer trap, FTICR, TOF, quadrupole ion trap magnetic sector Advantages HPLC/MS capable Tolerant of mM concentrations Multiple charging of salts Capable of observing Highest mass capability noncovalent complexes directly Tolerant of mixtures from solution Limitation of detection: Detection limit: attomole-to- zeptomole-to-femtomole femtomole Being developed as a tool for sequence analysis Disadvantages Multiple charging can be Typically low resolution confusing with mixtures Typically requires < mM salt concentration for good signals Not tolerant of mixtures (< 500) for linear TOF without delayed extraction Difficult to couple with LC/MS Suitable compounds Peptides Proteins Carbohydrates Nucleotides Oligonucleotides Phosphoproteins Small chargable molecules Glycoproteins Peptides Proteins Carbohydrates Nucleotides Oligonucleotides Phosphoproteins Small chargable molecules Glycoproteins Heterogeneous samples 52 D. ESI-CID-MS/MS 1. Concept and the Dissociation Process MS/MS can be thought of as a means by which to obtain the mass spectrum of a mass spectrum. Among these various modes of ion dissociation, CID is probably most widely used, especially in ion-trap instruments [145]. Figure 1.19 is a conceptual representation of CID-MS/MS. First, a given molecule is ionized and allowed to fragment with mass selection to obtain a conventional mass spectrum (the top spectrum in Figure 1.19). Second, ions of a particular m/z value (ions with m/z 129 in this case) are selected as precursor ions for a dissociation process followed by a second mass analysis to produce the product-ion spectrum (the bottom spectrum in Figure 1.19). The chemical structure of the precursor ion can be characterized by determining the products of the dissociation process. As illustrated in Figure 1.19, the key of a MS/MS process is that a significant population of the precursor ions undergoes some form of dissociation between the two stages of mass selection. In CID, the precursor ion collides with a gas molecule (usually an inert gas such as argon) to obtain internal energy, which leads to its decomposition [ 148]. 53 Ionization J m/z analysis 41 MS 1 129 [- .;.—. 1 / / Collision \\ — Dissociation in collision cell—— Cell ‘m/z analysis l I I MS 2 100 114 129 m/z ’ Figure l. 19 Conceptual representation of the MS/MS technique (redrawn from [149]). 2. Instrumentation There are two types of MS/MS instrumentation, MS/MS-in-space and MS/MS-in- time. The first type of instrument use a mass-selective process to select a given precursor ion at a given set of coordinates in space, transfer the ion to a different set of coordinates for ion dissociation, and transfer the products of that dissociation to yet another set of coordinates for secondary m/z analysis. The triple-quadrupole mass spectrometer is representative of instruments that are arranged in this manner. One disadvantage of MS/MS-in-space is that it requires many different points of ion focus that must be optimized at the same time to achieve a successful experiment [150]. 54 In MS/MS-in-time [151], ions are formed in a given set of coordinates in time, and then at a later time some of the ions are expelled from those coordinates so that only the ions of interest remain. At a later time, the residual ions are exposed to some dissociation process again at the same coordinates, and this is followed by yet a later episode of m/z analysis so that the entire MS/MS process is conducted in the same coordinates of space, but at different times. This kind of MS/MS analysis is adopted by ion-cyclotron resonance (ICR), Fourier-transform (FT) mass spectrometers [152], as well as the ion-trap mass spectrometers [145, 153, 154]. MS/MS-in-time is much simpler to operate than MS/MS-in-space because it only has one set of ion optics that must be optimized and merely controlled sequentially to allow the ions of interest to reside in the ion volume to achieve the desired results [155]. 3. Nomenclature for Classifying N-terminal and C-terminal Fragment Ions of Peptides Most proteins are constructed from peptides consisting of 20 common amino acid residues. As shown in Figure 1.20, the amino acids are condensed into the peptide bond (an amide bond) by connecting the carboxyl group of one amino acid with the amino group of the second amino acid. When a peptide is subjected to CID fi'agmentation, b- ions and/or y-ions are most often observed; these are produced through cleavage at the peptide bond. A y-ion is one that is formed by cleavage at a peptide bond with charge retention on the C-terminal end of the molecule (Figure 1.21), while a b-ion is one that is a fragment of the peptide with charge retention on the amino terminal side of the cleaved 55 amide bond (Figure 1.22). The sequence identification of an unknown peptide mainly depends on recognition of peaks representing a particular series of ions [156]. 1’ l’ ----- N H-(liH-C—NH — CH — C CH3 ((I3H2)4 NH2 peptide bond Figure 1. 20 Segment of a peptide showing the residues of alanine and lysine linked by an amide bond, the peptide bond. 56 .23: Bob 5360b 358.350 2.: wEESE meow EoEwmb 038083 :0 omega 05 wE>w£ 3205 255828 05:3 05 Bow 8538 Ho comm—=98 .3 ESTA 05 wEEom Mo 28358 333280 a A «Eur.— 72% a x N m x m m I _ _ _ \ morolmolmz oraclmz Ulmolmz .__. 2. a 2%.. J .4 -olmolmzrm E. .a . .4 .4 _ EOIUIEUIIZ UIEOIEZ DIEUIIZ- .F MW L +2 \ 57 .awm _ Km 9 86on 820538 Be on 8 33:3 2 oEoEm Race 05 £9626: £2 8338 an we. 850% mm Be: :03 05 62645268 no.“ ”new: Bob 5660b 82 8253 33828-2 no omega 05 mars": HESS -0 05 3 98% 1088.3 2: 33 852%: 33583 no comm—=98 .3 edema of mama—com no .3353 333280 «N A 9.55 _ 1:9 fl inn \ \ ‘ 79% x SAN—H #4 _ -ormolmz 01:0 1:2 01:0 1st 4. 44 4 .4 EOIUIEUIEZ- __ o / + III R 58 3. Application MS/MS and higher stage MSn (n > 2) can be used to great advantage in qualitative and quantitative analysis by supplying addition information on fragments that can be related to the structure of a precursor ion (the ion of interest) [159, 160]. Ion trap instruments have the advantages of high sensitivity and high capacity for fragment ions. In recent years, ESI-ion trap-MS/MS has become an indispensable tool in biomedical research [161-165]. It has been widely used in proteomics, as well as in the investigation of small molecules such as fatty esters and ketones. The coupling of HPLC and ion trap- MS/MS through the ESI interface made possible on-line separation, and, thus, direct analysis of biological samples is allowed. 59 VI. References [1] Mass spectrometry (MS) and nuclear magnetic resonance (NMR) applied to biological macromolecules, http ://nobelp1ize.org/chemistry/lgureates/Z002/chemgdv02.pdf. (accessed 10/ 12 2005) [2] J. B. Fenn, M. Mann, C. K. Meng, S. F. Wong, C. M. Whitehouse, Electrospray ionization for mass spectrometry of large biomolecules, Science 246 (1989) 64- 71. [3] M. Karas, F. Hillenkamp, Laser desorption ionization of proteins with molecular masses exceeding 10,000 daltons, Anal Chem 60 (1988) 2299-301. [4] A. Fersht, in Structure and mechanism in protein science: a guide to enzyme catalysis and protein folding, W. H. Freeman and Company, New York (1999) pl. [5] C. Branden, J. Tooze, in Introduction to Protein Structure, Garland Publishing Inc., New York (1999) p89. [6] T. E. Creighton, in Mechanisms of Protein Folding (Pain, R. H., Ed.), Oxford University Press, Oxford (1994) pl. [7] C. B. Anfinsen, E. Haber, M. Sela, F. H. White, Jr., The kinetics of formation of native ribonuclease during oxidation of the reduced polypeptide chain, Proc Natl Acad Sci USA 47 (1961) 1309-14. [8] V. Daggett, A. R. F ersht, Is there a unifying mechanism for protein folding?, Trends Biochem Sci 28 (2003) 18-25. [9] P. S. Kim, R. L. Baldwin, Intermediates in the folding reactions of small proteins, Annu Rev Biochem 59 (1990) 631-60. [10] C. R. Matthews, Pathways of protein folding, Annu Rev Biochem 62 (1993) 653-83. [11] 0. B. Ptitsyn, Structures of folding intermediates, Curr Opin Struct Biol 5 (1995) 74- 8. [12] J. S. Weissman, P. S. Kim, Reexamination of the folding of BPTI: predominance of native intermediates, Science 253 (1991) 1386-93. [13] T. E. Creighton, Renaturation of the reduced bovine pancreatic trypsin inhibitor, J Mol Biol 87 (1974) 563-77. 60 [14] T. E. Creighton, D. P. Goldenberg, Kinetic role of a meta-stable native-like two- disulphide species in the folding transition of bovine pancreatic trypsin inhibitor, J Mol Biol 179 (1984) 497-526. [15] H. A. Scheraga, Y. Konishi, T. Ooi, Multiple pathways for regenerating ribonuclease A, Adv Biophys 18 (1984) 21-41. [16] J. Wu, Y. Yang, J. T. Watson, Trapping of intermediates during the refolding of recombinant human epidermal growth factor (hEGF) by cyanylation, and subsequent structural elucidation by mass spectrometry, Protein Sci 7 (1998) 1017-28. [17] Y. Yang, J. Wu, J. T. Watson, Probing the folding pathways of long R(3) insulin-like grth factor-I (LR(3)IGF-I) and IGF-1 via capture and identification of disulfide intermediates by cyanylation methodology and mass spectrometry, J Biol Chem 274 (1999) 37598-604. [18] T. E. Wales, J. R. Engen, Hydrogen exchange mass spectrometry for the analysis of protein dynamics, Mass Spectrom Rev (2005). [19] J. R. Engen, D. L. Smith, Investigating protein structure and dynamics by hydrogen exchange MS, Anal Chem 73 (2001) 256A-265A. [20] I. A. Kaltashov, S. J. Eyles, Studies of biomolecular conformations and conformational dynamics by mass spectrometry, Mass Spectrom Rev 21 (2002) 37-71. [21] Y. M. Torchinsky, Surfur in Proteins, Pergamon, Oxford (1981). [22] A. Sato, S. Koyarna, H. Yamada, S. Suzuki, K. Tamura, M. Kobayashi, M. Niwa, T. Yasuda, Y. Kyogoku, Y. Kobayashi, Three-dimensional solution structure of a disulfide bond isomer of the human insulin-like grth factor-I, J Pept Res 56 (2000) 218-30. [23] L. Haeffner-Gonnley, L. Parente, D. B. Wetlaufer, Use of proline-specific endopeptidase in the isolation of all four "native" disulfides of hen egg white lysozyme, Int J Pept Protein Res 26 (1985) 83-91. [24] P. E. Staswick, M. A. Hermodson, N. C. Nielsen, Identification of the cystines which link the acidic and basic components of the glycinin subunits, J Biol Chem 259 (1984) 13431-5. [25] A. P. Ryle, F. Sanger, Disulphide interchange reactions, Biochem J 60 (1955) 535- 40. 61 [26] G. Padron, V. Besada, A. Agraz, Y. Quinones, L. Herrera, Y. Shimonishi, T. Takao, Mass spectrometric analysis of recombinant human a-2 interferon, Analytica Chimica Acta 223 (1989) 361-9. [27] B. S. Antharavally, B. R. DasGupta, Covalent structure of botulinum neurotoxin type B; location of sulflrydryl groups and disulfide bridge and identification of C- terrnini of light and heavy chains, J Protein Chem 17 (1998) 417-28. [28] 0. Markovic, H. J omvall, Disulfide bridges in tomato pectinesterase: variations from pectinesterases of other species; conservation of possible active site segments, Protein Sci 1 (1992) 1288-92. [29] N. F. Ng, C. L. Hew, Structure of an antifreeze polypeptide from the sea raven. Disulfide bonds and similarity to lectin-binding proteins, J Biol Chem 267 (1992) 16069-75. [30] T. A. Valueva, T. A. Revina, V. V. Mosolov, R. Mentele, Primary structure of potato kunitz-type serine proteinase inhibitor, Biol Chem 381 (2000) 1215-21. [31] T. W. Thannhauser, C. A. McWherter, H. A. Scheraga, Peptide mapping of bovine pancreatic ribonuclease A by reverse-phase high-performance liquid chromatography. II. A two-dimensional technique for determination of disulfide pairings using a continuous-flow disulfide-detection system, Anal Biochem 149 (1985) 322-30. [32] A. Aitken, M. Learmonth, in Protein Protocols Handbook, Humana Press Inc., Totowa, N. J. (2002) p585-8. [33] D. L. Smith, Z. R. Zhou, Strategies for locating disulfide bonds in proteins, Methods Enzymol 193 (1990) 374-89. [34] K. Hirayama, S. Akashi, in Biological mass spectrometry: Present and future, Wiley, New York (1994) p299-312. [35] J. Qi, J. Wu, G. A. Somkuti, J. T. Watson, Determination of the disulfide structure of sillucin, a highly knotted, cysteine-rich peptide, by cyanylation/cleavage mass mapping, Biochemistry 40 (2001) 4531-8. [36] J. Wu, J. T. Watson, A novel methodology for assignment of disulfide bond pairings in proteins, Protein Sci 6 (1997) 391-8. [37] Y. Yang, J. Wu, J. T. Watson, Disulfide mass mapping in proteins containing adjacent cysteines is possible with cyanylation/cleavage methodology, J. Am. Chem. Soc. 120 (1998) 5834-5. 62 [3 8] E. Welker, W. J. Wedemeyer, M. Narayan, H. A. Scheraga, Coupling of conformational folding and disulfide-bond reactions in oxidative folding of proteins, Biochemistry 40 (2001) 9059-64. [39] W. J. Wedemeyer, E. Welker, M. Narayan, H. A. Scheraga, Disulfide bonds and protein folding, Biochemistry 39 (2000) 7032. [40] M. Narayan, E. Welker, H. A. Scheraga, Characterizing the unstructured intermediates in oxidative folding, Biochemistry 42 (2003) 6947-55. [41] Q. X. Hua, L. Narhi, W. Jia, T. Arakawa, R. Rosenfeld, N. Hawkins, J. A. Miller, M. A. Weiss, Native and non-native structure in a protein-folding intermediate: spectroscopic studies of partially reduced IGF-I and an engineered alanine model, J Mol Biol 259 (1996) 297-313. [42] L. O. Narhi, Q. X. Hua, T. Arakawa, G. M. Fox, L. Tsai, R. Rosenfeld, P. Holst, J. A. Miller, M. A. Weiss, Role of native disulfide bonds in the structure and activity of insulin-like grth factor 1: genetic models of protein-folding intermediates, Biochemistry 32 (1993) 5214-21. [43] J. Y. Chang, P. Schindler, U. Ramseier, P. H. Lai, The disulfide folding pathway of human epidermal growth factor, J Biol Chem 270 (1995) 9207-16. [44] T. G. Oas, P. S. Kim, A peptide model of a protein folding intermediate, Nature 336 (1988) 42-8. [45] D. M. Rothwarf, H. A. Scheraga, Regeneration of bovine pancreatic ribonuclease A. 3. Dependence on the nature of the redox reagent, Biochemistry 32 (1993) 2690-7. [46] J. Houk, R. Singh, G. M. Whitesides, Measurement of thiol-disulfide interchange reactions and thiol pKa values, Methods Enzymol 143 (1987) 129-40. [47] M. Janet, Wilson, and D. J. Hupe, Structure-reactivity correlations for the thiol- disulfide interchange reaction, Journal of the American Chemical Society 99 (1977) 7922-26. [48] Z. Shaked, R. P. Szajewski, G. M. Whitesides, Rates of thiol-disulfide interchange reactions involving proteins and kinetic measurements of thiol pKa values, Biochemistry 19 (1980) 4156-66. [49] A. Larsson, S. Orrenius, A. Hohngren, B. Mannervik, Functions of Glutathione: Biochemical, Physiological, Toxicological, and Clinical Aspects, Raven, New York (1983). [50] E. Barron, S. Guzman, Thiamine and sulfhydryl compounds, Boletin de la Sociedad Quimica del Peru 6 (1940) 7-32. 63 [51] D. M. Rothwarf, H. A. Scheraga, Regeneration of bovine pancreatic ribonuclease A. 1. Steady-state distribution, Biochemistry 32 (1993) 2671-9. [52] D. M. Rothwarf, H. A. Scheraga, Regeneration and reduction of native bovine pancreatic ribonuclease A with oxidized and reduced dithiothreitol, J. Am. Chem. Soc. 113 (1991) 6293-4. [53] J. Y. Chang, F. Canals, P. Schindler, E. Querol, F. X. Aviles, The disulfide folding pathway of potato carboxypeptidase inhibitor, J Biol Chem 269 (1994) 22087-94. [54] B. Chatrenet, J. Y. Chang, The disulfide folding pathway of hirudin elucidated by stop/go folding experiments, J Biol Chem 268 (1993) 20988-96. [55] H. C. Gilbert, in Mechanisms of Protein Folding, Oxford Press, Oxford (1994) p112. [56] J. Y. Chang, The disulfide folding pathway of tick anticoagulant peptide (TAP), a Kunitz-type inhibitor structurally homologous to BPTI, Biochemistry 35 (1996) 11702-9. [57] J. A. Wilkins, J. Cone, Z. I. Randhawa, D. Wood, M. K. Warren, H. E. Witkowska, A study of intermediates involved in the folding pathway for recombinant human macrophage colony-stimulating factor (M-CSF): evidence for two distinct folding pathways, Protein Sci 2 (1993) 244-54. [58] S. Hober, G. Forsberg, G. Palm, M. Hartmanis, B. Nilsson, Disulfide exchange folding of insulin-like grth factor I, Biochemistry 31 (1992) 1749-56. [59] M. O. Glocker, B. Arbogast, R. Milley, C. Cowgill, M. L. Deinzer, Disulfide linkages in the in vitro refolded intermediates of recombinant human macrophage- colony-stimulating factor: analysis of the sulfhydryl alkylation of free cysteine residues by fast-atom bombardment mass spectrometry, Proc Natl Acad Sci U S A 91 (1994) 5868-72. [60] T. E. Creighton, Toward a better understanding of protein folding pathways, Proc Natl Acad Sci U S A 85 (1988) 5082-6. [61] M. M. Lyles, H. F. Gilbert, Catalysis of the oxidative folding of ribonuclease A by protein disulfide isomerase: dependence of the rate on the composition of the redox buffer, Biochemistry 30 (1991) 613-9. [62] T. E. Creighton, Disulfide bonds as probes of protein folding pathways, Methods Enzymol 131 (1986) 83-106. [63] C. Torella, M. Ruoppolo, G. Marino, P. Pucci, Analysis of RNase A refolding intermediates by electrospray/mass spectrometry, FEBS Lett 352 (1994) 301-6. 64 [64] D. M. Rothwarf, H. A. Scheraga, Regeneration and reduction of native bovine pancreatic ribonuclease A with oxidized and reduced dithiothreitol, Journal of the American Chemical Society 113 (1991) 6294-4. [65] M. J. Volles, X. Xu, H. A. Scheraga, Distribution of disulfide bonds in the two- disulfide intermediates in the regeneration of bovine pancreatic ribonuclease A: further insights into the folding process, Biochemistry 38 (1999) 7284-93. [66] D. M. Rothwarf, Y. J. Li, H. A. Scheraga, Regeneration of bovine pancreatic ribonuclease A: detailed kinetic analysis of two independent folding pathways, Biochemistry 37 (1998) 3767-76. [67] S. Talluri, D. M. Rothwarf, H. A. Scheraga, Structural characterization of a three— disulfide intermediate of ribonuclease A involved in both the folding and unfolding pathways, Biochemistry 33 (1994) 10437-49. [68] Y. H. Zhang, X. Yan, C. S. Maier, M. I. Schimerlik, M. L. Deinzer, Structural comparison of recombinant human macrophage colony stimulating factor beta and a partially reduced derivative using hydrogen deuterium exchange and electrospray ionization mass spectrometry, Protein Sci 10 (2001) 2336-45. [69] A. Fersht, in Structure and mechanism in protein science: a guide to enzyme catalysis and protein folding, W. H. Freeman and Company, New York (1999) p4-6. [70] D. Voet, J. G. Voet, in Biochemistry, John Wiley & Sons, Inc., New York (1995) p162. [71] C. Branden, J. Tooze, in Introduction to protein structure, Garland Publishing, Inc., New York (1999) p390-391. [72] C. Branden, J. Tooze, in Introduction to protein structure, Garland Publishing, Inc., New York (1999) p374. [73] D. Voet, J. G. Voet, in Biochemistry, John Wiley & Sons, Inc., New York (1995) p166. [74] A. F ersht, in Structure and mechanism in protein science: a guide to enzyme catalysis and protein folding, W. H. Freeman and Company, New York (1999) p8. [75] J. P. Homak. The basics of NMR, http://www.cis.rit.edu/htbooks/nmr/insidehtm, (accessed 10/ 14 2005). 65 [76] the Free Dictionary by F arlex, http://www.thefreedictionary.com/nuclea1%20magnetic%20resonang§_, (accessed 10/14 2005). [77] J. Evans, N.S. Jeremy, in Biomolecular NMR spectroscopy, Oxford University Press, Oxford ; New York ( 1995) p5-9. [78] T. W. G. Solomons, C. B. Fryhle, in Organic Chemistry, John Wiley & Sons, Inc., New York (2000) p370. [79] A. Fersht, in Structure and mechanism in protein science: a guide to enzyme catalysis and protein folding, W. H. Freeman and Company, New York (1999) p7. [80] A. Borthakur. Nuclear Overhauser Effect (NOE), http://wd40.mmrrcc.upenn.edu/courses/biophvs601/NOE.pdf, (accessed 10/20 2005) [81] D. L. Pavia, G. M. Lampman, G. S. Kriz, in Introduction to spectroscopy: 0 gudie for students of organic chemistry, Harcout Brace College Publishers, New York (1996) p150-2. [82] T. F. Havel, I. D. Kuntz, G. M. Crippen, Effects of distance constraints on macromolecular conformation. 11. Simulation of experimental results and theoretical predictions, Biopolymers 18 (1979) 73-82. [83] P. Permi, Aspects of coherence transfer in high molecular weight proteins, Annual Reports on NMR Spectroscopy 53 (2004) 245-96. [84] J. D. Satterlee, Proton NMR studies of biological problems involving paramagnetic heme proteins, Metal Ions in Biological Systems 21 (1987) 121-85. [85] A. F ersht, in Structure and mechanism in protein science: a guide to enzyme catalysis and protein folding, W. H. Freeman and Company, New York (1999) p193-5. [86] R. W. Woody, H. Sugeta, T. S. Kodama, [Circular dichroism of proteins: recent developments in analysis and prediction], T anpakushitsu Kakusan Koso 41 (1996) 56-69. [87] N. Sreerama, M. C. Manning, M. E. Powers, J. X. Zhang, D. P. Goldenberg, R. W. Woody, Tyrosine, phenylalanine, and disulflde contributions to the circular dichroism of proteins: circular dichroism spectra of wild-type and mutant bovine pancreatic trypsin inhibitor, Biochemistry 38 (1999) 10814-22. 66 [88] C. P. M. v. Mierlo, H. H. J. (1. Jongh, A. J. W. G. Visser, Circular dichroism of proteins in solution and at interfaces, Applied Spectroscopy Reviews 35 (2000) 277-313. [89] L. A. Compton, W. C. Johnson, Jr., Analysis of protein circular dichroism spectra for secondary structure using a simple matrix multiplication, Anal Biochem 15 5 (1986) 155-67. [90] S. W. Englander, N. R. Kallenbach, Hydrogen exchange and structural dynamics of proteins and nucleic acids, Quarterly Review of Biophysics 16 (1984) 521-655. [91] C. Redfield, NMR studies of partially folded molten-globule states, Methods in Molecular Biology 278 (2004) 233-54. [92] X. Li, Y. T. Chou, R. Husain, J. T. Watson, Integration of hydrogen/deuterium exchange and cyanylation-based methodology for conformational studies of cystinyl proteins, Anal Biochem 331 (2004) 130-7. [93] I. A. Kaltashov, S. J. Eyles, Crossing the phase boundary to study protein dynamics and function: combination of amide hydrogen exchange in solution and ion fragmentation in the gas phase, J Mass Spectrom 37 (2002) 557-65. [94] J. T. Watson, in Introduction to mass spectrometry, Lippincott - Raven Publishers, New York (1997) p312-313. [95] J. T. Watson, in Introcution to mass spectrometry, Lippincott - Raven Publishers, New York (1997) p285-6. [96] X. Li, R. J. Hood, W. J. Wedemeyer, J. T. Watson, Characterization of peptide folding nuclei by hydrogen/deuterium exchange-mass spectrometry, Protein Sci 14 (2005) 1922-8. '[97] Y. Deng, H. Pan, D. L. Smith, That Hydrogen Exchange at Individual Peptide Amide Linkages Can Be Determined by Collision-Induced Dissociation Mass Spectrometry, Journal of the American Chemical Society 121 (1999) 1966-7. [98] M.-Y. M. Kim, Claudia S.; Reed, Donald J .; Deinzer, Max L., Site-specific amide hydrogen/deuterium exchange in E. coli thioredoxins measured by electrospray ionization mass spectrometry, Journal of the American Chemical Society 123 (2001) 9860-6. [99] X. D. Cai, Chhabil, Structural characterization of methionine and leucine enkephalins by hydrogen/deuterium exchange and electrospray ionization tandem mass spectrometry, Rapid Communications in Mass Spectrometry 19 (2005) 1-8. 67 [100] M. S. B. Munson, F. H. Field, Chemical ionisation mass spectrometry. I. General introduction, J. Am. Chem. Soc. 88 (1966) 2621-30. [101] R. D. Macfarlane, D. F. Torgerson, Californium-252 plasma desorption mass spectroscopy, Science 191 (1976) 920-5. [102] M. Barber, R. S. Bordoli, R. D. Sedgwick, A. N. Tyler, Fast atom bombardment of solids (FAB): a new ion source for mass spectrometry, J. Chem. Soc. Chem. Commun. (1981) 325-7. [103] M. B. Karas, D.; Bahr, U.; Hillenkamp, F., Matrix-assisted ultraviolet laser desorption of non-volatile compounds, International Journal of Mass Spectrometry and Ion Processes 78 (1987) 53-68. [104] K. Tanaka, Ido, Y., Akita, S., Yoshida, Y. and Yoshida, T., in Proc. Second Japan- China Joint Symposium on Mass Spectrometry. (Matsuda, H. a. X.-t., L., Ed.), Osaka, Japan (1987) p185-8. [105] K. Tanaka, H. Waki, Y. Ido, S. Akita, Y. Yoshida, T. Yoshida, Protein and polymer analysis up to m/z 100.000 by laser ionisation time-of-flight mass spectrometry, Rapid Commun. Mass Spectrom 2 (1998) 151-3. [106] F. Hillenkamp, M. Karas, R. C. Beavis, B. T. Chait, Matrix-assisted laser desorption/ionization mass spectrometry of biopolymers, Anal Chem 63 (1991) 1 193A-1203A. [107] R. Kruger, M. Karas, Formation and fate of ion pairs during MALDI analysis: anion adduct generation as an indicative tool to determine ionization processes, J Am Soc Mass Spectrom 13 (2002) 1218-26. [108] R. Knochenmuss, A Quantitative Model of Ultraviolet Matrix-Assisted Laser Desorption/Ionization Including Analyte Ion Generation, Analytical Chemistry 75 (2003) 2199-2207. [109] R. Zenobi, R. Knochenmuss, Ion formation in MALDI mass spectrometry, Mass Spectrom. Rev. 17 (1999) 337-366. [110] R. Knochenmuss, A. Stortelder, K. Breuker, R. Zenobi, Secondary ion-molecule reactions in matrix-assisted laser desorption/ionization, J. Mass Spectrom. 35 (2000) 1237-1245. [111] E. Lehmann, F. Diederich, R. Zenobi, B. Salih, M. Gomez-Lopez, Do matrix- assisted laser desorption/ionization mass spectra reflect solution-phase formation of cyclodextrin inclusion complexes?, Analyst (Cambridge, U. K.) 125 (2000) 849-854. 68 [1 12] E. Lehmann, R. Knochenmuss, R. Zenobi, Ionization mechanisms in matrix- assisted laser desorption/ionization mass spectrometry: contribution of pre-formed ions, Rapid Commun. Mass Spectrom. 11 ( 1997) 1483-1492. [1 13] E. Lehmann, R. Zenobi, Investigations of the cationization mechanism in matrix- assisted laser desorption/ionization, Adv. Mass Spectrom. 14 (1998) B063490/1- B063490/1 1. [114] K. Dreisewerd, S. Berkenkamp, A. Leisner, A. Rohlfing, C. Menzel, Fundamentals of matrix-assisted laser desorption/ionization mass spectrometry with pulsed infrared lasers, International Journal of Mass Spectrometry 226 (2003) 189-209. [115] M. Karas, U. Bahr, I. Fournier, M. Gluckmann, A. Pfenninger, The initial-ion velocity as a marker for different desorption-ionization mechanisms in MALDI, International Journal of Mass Spectrometry 226 (2003) 239-248. [116] R. Knochenmuss, R. Zenobi, MALDI ionization: the role of in-plume processes, Chem Rev 103 (2003) 441-52. [117] M. Karas, M. Gluckmann, J. Schafer, Ionization in matrix-assisted laser desorption/ionization: singly charged molecular ions are the lucky survivors, J Mass Spectrom 35 (2000) 1-12. [118] P. Juhasz, C. E. Costello, K. Biemann, Matrix-assisted laser desorption ionization mass spectrometry with 2-(4-hydroxyphenylazo)benzoic acid matrix, Journal of the American Society for Mass Spectrometry 4 (1993) 399-409. [1 19] H. Ehring, M. Karas, F. Hillenkamp, Role of photoionization and photochemistry in ionization processes of organic molecules and relevance for matrix-assisted laser desorption ionization mass spectrometry, Organic Mass Spectrometry 27 (1992) 472-80. [120] M. C. Fitzgerald, G. R. Parr, L. M. Smith, Basic matrices for the matrix-assisted laser desorption/ionization mass spectrometry of proteins and Oligonucleotides, Anal Chem 65 (1993) 3204-11. [1 21] O. Vorm, P. Roepstorff, M. Mann, mproved Resolution and Very High Sensitivity in MALDI TOF of Matrix Surfaces Made by Fast Evaporation, Analytical Chemistry 66 (1994) 3281-7. [122] M. Cadene, B. T. Chait, A robust, detergent-fi'iendly method for mass spectrometric analysis of integral membrane proteins, Anal Chem 72 (2000) 5655- 8. [123] J. T. Watson, in Introduction to mass spectrometry, Lippincott-Raven Publishers, New York (1997) p279. 69 [l 24] R. C. Beavis, B. T. Chait, Rapid, sensitive analysis of protein mixtures by mass spectrometry, Proc Natl Acad Sci U S A 87 (1990) 6873-7. [125] J. T. Watson, in Introduction to mass spectrometry, Lippincott-Raven Publishers, New York (1997) p90. [126] W. Wiley, I. McLaren, Time-of-flight mass spectrometer with improved resolution, The Review of Scientific Instruments 26 (1955) 1150-7. [127] B. Mamyrin, V. Karataev, D. Shmikk, V. Zagulin, Relectron for TOF-MS, Sov Phys JETP 37 (1973) 45-8. [128] B. Mamyrin, D. Shmikk, Linear reflection, Sov Phys JE T P 49 (1979) 762-4. [129] J. L. Gallegos-Perez. Optimization and Study of Experimental Conditions of Primary Amines for Nucleaophilic Cleavage of Cyanylated Cystinyl Proteins in Disulfide Mass Mapping Methodology, Michigan State University, East Lansing, 2005. [130] M. Dole, L. L. Mack, R. L. Hines, R. C. Mobley, L. D. Ferguson, M. B. Alice, Molecular beams of macroions, Journal of Chemical Physics 49 (1968) 2240-9. [131] L. L. Mack, P. Kralik, A. Rheude, M. Dole, Molecular beams of macroions. II., Journal of Chemical Physics 52 (1970) 4977-86. [132] C. M. Whitehouse, R. N. Dreyer, M. Yamashita, J. B. F enn, Electrospray interface for liquid chromatographs and mass spectrometers, Anal Chem 57 (1985) 675-9. [133] J. B. F enn, in 36th Annual Conference, Am. Soc. for Mass Spectrom, San Francisco (1988) p773. [1 34] N. B. Cech, C. G. Enke, Practical implications of some recent studies in electrospray ionization fundamentals, Mass Spectrom Rev 20 (2001) 362-87. [135] J. Iribame, B. Thomson, On the evaporation of charged ions from small droplets, J. Chem. Phys. 64 (1976) 2287-94. [136] J. H. Gross, Springer, Berlin; New York (2004) vaiii, 518. [137] G. Siuzdak, The emergence of mass spectrometry in biochemical research, Proceedings of the National Academy of Sciences of the United States of America 91 (1994) 11290-7. [138] R. D. Smith, J. H. Wahl, D. R. Goodlett, S. A. Hofstadler, Capillary electrophoresis/mass spectrometry, Analytical Chemistry 65 (1993) 547A-84A. 70 [139] P. R. Griffin, J. A. Coffrnan, L. E. Hood, J. R. Yates, 111., Structural analysis of proteins by capillary HPLC electrospray tandem mass spectrometry, International Journal of Mass Spectrometry and Ion Processes l l 1 (1991) 131-49. [140] W. J. Griffiths, Nanospray mass spectrometry in protein and peptide chemistry, Proteomics in Functional Genomics 88 (2000) 69-79. [141] M. Wilm, M. Mann, Analytical Properties of the Nanoelectrospray Ion Source, Anal Chem 68 (1996) 1-8. [142] ThermoFinnigan training institute LCQ biotech course. [143] S. J. Gaskell, Electrospray: Principles and Practice, Journal of Mass Spectrometry 32 (1997) 677-88. [144] J. T. Watson, in Introduction to mass spectrometry, Lippincott-Raven Publishers, New York (1997) p80. [145] R. E. March, An introduction to quadrupole ion trap mass spectrometry, Journal of Mass Spectrometry 32 (1997) 351-69. [146] J. Pappas, Quadrupole Ion-T rap Mass Analyzer, Thermo Electron, Somerset, NJ (2004) [147] D. C. Muddiman, R. Bakhtiar, S. A. Hofstadler, R. D. Smith, Matrix-Assisted Laser Desorption/Ionization Mass Spectrometry, Journal of Chemical Education 74 (1998) 1288-92. [148] H. E. De, Tandem mass spectrometry: A primer, J Mass Spectrom 31 (1996) 129- 37. [149] J. T. Watson, in Introduction to mass spectrometry, Lippincott-Raven Publishers, New York (1997) p110. [150] K. Jennings, G. Dolnikowski, in Methods in Enzymology (McCloskey, J ., Ed.), Vol. 193, Academic Press, New York (1990) p37-61. [151] J. T. Watson, in Introduction to mass spectrometry, Lippincott-Raven Publishers, New York (1997) p120. [152] C. Koster, M. Kahr, J. Castoro, C. Wilkins, FT-MS, Mass Spectrom Rev 11 (1992) 495-512. [153] G. Glish, Multiple stage MS, Analyst 119 (1994) 533-7. 71 [ 1 54] C. Evans, P. Traldi, A. Mele, C. Sottani, Ion trap vs TQMS collision-results, Org Mass Spectrom 26 (1991) 347-50. [ 1 55] J. T. Watson, in Introduction to mass spectrometry, Lippincott-Raven Publishers, New York (1997) p121. [156] J. T. Watson, in Introduction to Mass Spectrometry, Lippincott-Raven Publishers, New York (1997) p256-8. [l 5 7] A. G. Harrison, A. B. Young, Fragmentation of protonated oligoalanines: amide bond cleavage and beyond, J Am Soc Mass Spectrom 15 (2004) 1810-9. [158] B. Paizs, S. Suhai, A. G. Harrison, Experimental and theoretical investigation of the main fragmentation pathways of protonated H-Gly—Gly—Sar-OH and H-Gly-Sar- Sar-OH, J Am Soc Mass Spectrom 14 (2003) 1454-69. [159] S. Kawabata, MSn explore into the next generation of proteome analysis, Kagaku to Kogyo 78 (2004) 203-11. [160] J. W. Hager, Recent trends in mass spectrometer development, Analytical and Bioanalytical Chemistry 378 (2004) 845-50. [161] H. Snijders, The next generation of hybrid linear ion trap mass spectrometers, New Drugs 1 (2005) 21-31. [162] J. W. Hager, Q TRAP mass spectrometer technology for proteomics applications, Drug Discovery Today: Targets 3 (2004) 831-836. [163] T. Kawakami, H. Anyoji, T. Nishimura, Use of nano liquid chromatography/electrospray ionization/ion trap mass spectrometer for proteome quantitation and identification, Journal of the Mass Spectrometry Society of Japan 51 (2003) 79-82. [164] J. Peng, S. P. Gygi, Proteomics: the move to mixtures, Journal of Mass Spectrometry 36 (2001) 1083-91. [165] D. S. Kirkpatrick, S. A. Gerber, S. P. Gygi, The absolute quantification strategy: a general procedure for the quantification of proteins and post-translational modifications, Methods 35 (2005) 265-73. 72 CHAPTER 2 OPTIMIZATION OF CONDITIONS FOR MONITORING HYDROGEN/DEUTERIUM EXCHANGE IN PROTEINS BY MALDI-TOF—MS I. Introduction The 3-D structure or conformation of a protein is the key to its bioactivity [1]. Many diseases stem from impairment of structural properties of specific proteins (e.g., muscular dystrophy [2] and Creutzfeld-Jakob disease [3]). The ultimate goal of protein science is to be able to predict the 3-D structure and biofunction of a protein de novo and how it will bind to ligands [2]. In spite of considerable efforts over the past 25 years, principles that govern the transitions of biopolymers from totally unstructured to highly ordered states remain one of the greatest mysteries in structural biology. Furthermore, adequate experimental methodology is still being developed to provide a practical and effective way to determine a protein’s structure and dynamics [4, 5]. Several experimental methods have contributed to our current understanding in this area, among which amide hydrogen exchange (HX) has played a particularly important role [6, 7]. About 50 years ago, it was recognized that different amide hydrogens in different protein structures undergo isotope exchange at different rates [8, 9], which was followed by the widespread use of hydrogen exchange as a tool to study protein structure and dynamics. HX has been used to characterize protein folding [IO-12], functionally different forms of - proteins [13-15], and protein-protein interactions [16, 17]. Among the many different kinds of hydrogens in a protein, only those located at peptide amide linkages have exchange rates in a range that can be easily measured, and 73 the rates depend on whether they are participating in intra-molecular H-bonding and on the extent to which they are shielded from the solvent. Because intra-molecular H- bonding and solvent shielding are directly related to protein structure, it follows that amide HX rates can be used to probe protein structure [18]. Historically, the monitoring of HX rates depends on qualitative infrared spectroscopic measurements [19] and quantitative density-gradient measurements [20]. With the new techniques developed in NMR, replacement of amide hydrogens by deuteriums (HDX) as monitored by NMR became one of the most practical and powerful tools in protein conformational studies since the 19708 [21]. During the last two decades, detecting HDX by MS has become a powerful and popular new approach for investigating protein structure and dynamics due to the unmatched detection limit and the ability to analyze large proteins (when coupled with peptic digestion). Furthermore, this approach has the power to decipher heterogeneity within an ensemble of a particular species, an aspect not readily amenable to conventional techniques [21-23]. Both MALDI-MS [24-26] and LC-ESI-MS [27-30] have been used to monitor HDX. LC-ESI-MS provides preliminary separation of mixtures, and, thus, relieves the problem of signal suppression that usually occurs in the analysis of a complicated mixture such as the peptic digest of a large protein. In addition, the tandem mass spectrometry technique (MS", 11 > 1), which is routinely coupled with ESI-CID-MS makes single-amino-acid-residue resolution achievable in HDX-MS. This method, however, requires significant modification of the instrument, and the operation of the instrument requires considerable expertise. HDX-MALDI-MS, on the other hand, provides a simple and rapid approach to obtaining data about the exchange of deuteriums 74 with amide hydrogens in a protein. No modifications of the instrument are required, and the exchange data for many components of a mixture can be obtained from a single spectrum. The experiments do not require complicated handling, and are usually less time consuming than similar experiments with LC-ESI-MS. HDX-MALDI-MS has found broad utility for analyzing HDX of proteins in studies of protein folding and protein-ligand interactions. This chapter describes the results of investigating various conditions for protocol optimization of HDX experiments on proteins. MALDI-MS was used to measure the mass shift of bovine pancreatic ribonuclease A (RNase A) afier HDX under various conditions. This protein is a medium-sized protein that consists of 124 amino acid residues, including eight cysteines in the form of four intra-molecular disulfide bonds. RNase A has considerable secondary structure (three or-helices and six B-sheets) as stabilized by four disulfide bonds [31]; thus, its native and denatured states provide diverse conformations for our assessment of HDX by MALDI-T OF mass spectrometry. 11. Experimental Section Materials RNase A (type I-AS), TCEP hydrochloride, sodium citrate, or-CHCA, CDAP tetrafluoroborate, and TF A (HPLC grade) were obtained from Sigma and used without further purification. Guanidine hydrochloride (Gdn-HCI) and ACN (HPLC grade) were purchased from Invitrogen Corporation (Carlsbad, CA) and EMD Chemicals Inc., respectively. The 0.10 M TCEP solution in 0.10 M citrate buffer at pH 3.0 was freshly prepared before use. 75 Preparation of Model Proteins RNase A (20 nmol) was dissolved with 0.1 M citrate buffer, pH 3.0, containing 6 M Gdn-HCl and 0.1 M TCEP reducing agent to a final concentration of 0.6 nmol/pL. The denaturation and reduction of RNase A were carried out at 37 °C for 2 hr. The reduced protein was purified by HPLC with linear gradient of 10% to 40% B in one hour, where solvent A was H20 containing 0.1% TFA, and solvent B was ACN containing 0.1% TFA. The HPLC fraction was collected, dried under reduced pressure, and stored at —80 °C. The reduced/denatured protein was reconstituted in HzO (pH ~7.0) right before the HDX experiment. Native RNase A was dissolved and equilibrated in H20 (pH ~7.0) at room temperature for about one hour to achieve an equilibrium distribution of the native state in solution. Confirmation of Complete Reduction To the dried and completely reduced RNase A was added 0.2 M CDAP solution (in 0.1 M citrate buffer, pH 3.0) to a final concentration of l nmol/uL protein. The cyanylation of free thiol groups by CDAP proceeded at room temperature for 15 min under nitrogen. The cyanylated protein was then desalted using HPLC and the proper fraction was dried. To cleave the polypeptide chain on the N-terminal side of each cyanylated cysteine residue, 4 [1L of 6 M Gdn-HCl in 1 M NH40H aqueous solution was added to the dried protein to dissolve it, and then 10 uL of 1 M NH.;OH was added. Cleavage of the polypeptide chain was performed at room temperature for one hour. Excess ammonia was removed in a vacuum system, and the cleavage products were analyzed by MS for identification. 76 Mass Spectrometry For the confirmation of complete reduction, MALDI mass spectra were obtained with a Voyager DE-STR TOF mass spectrometer (Applied Biosystems Inc., Frarningham, MA) using an accelerating voltage of 25 kV with the grid voltage at 93%, guide wire at 0.002%, and extraction delay time at 300 ns. The mass spectrometer was equipped with a 337-nm nitrogen laser. Time-of-flight to mass conversion was achieved with the use of internal standards of bradykinin (average m/z ratio for [M + H]+ = 1061.22), bovine pancreatic insulin (average calculated m/z ratio for [M + H]+ 5734.56). A saturated solution of or-CHCA, prepared in ACN / water / TFA (50:50:0.05), was used as the matrix. The sample spots for MALDI were prepared by the “ultrathin layer method” [32] as described in Section V. A. 1. of Chapter 1. For the comparison of deuterated and non-deuterated matrix, instrument settings were the same as described above. The matrix used in these experiments was a solution containing 5 mg/mL of sinapinic acid in 0.1% deuterated TFA (in D20)/ acetonitrile 1:2 by volume (pD 2.5). The matrix was incubated with D20 overnight before use. The conventional non-deuterated matrix was prepared exactly the same way as for the deuterated matrix, except that H20 and non-deuterated TFA were used instead of D20 and deuterated TFA (dl-TFA) (TFA: CF 3CO2H; dl-TFA: CF3CO2D). Time-of-flight to mass conversion was achieved with the use of external calibration of horse skeletal myoglobin (average calculated m/z ratio for [M + H]+ = 16,952.56, [M + 2H]2+ = 8476.78). For all the other experiments using deuterated matrix, a Voyager DE MALDI- TOF mass spectrometer equipped with a 337-nm nitrogen laser (Perseptive Biosystems 77 Inc., Framingham, MA) was used. The accelerating voltage was set at 25 kV, grid voltage at 93%, and guide wire at 0.02%. The extraction delay time was 600 ns. The matrix used in these experiments was in a solution containing 10 mg/mL of sinapinic acid in 0.1% deuterated TFA (in D20)/ acetonitrile 1:2 by volume (pD 2.5). The matrix was incubated with D20 overnight before use. Time-of-flight to mass conversion was achieved with the use of internal or external calibration of fully deuterated myoglobin (d260-Mb) (average calculated m/z ratio for [d26o-Mb + D]+ = 17214.18, [d26o-Mb + 2D]2+ = 8608.09). The fully deuterated myglobin (d260—Mb) was prepared by incubating apomyoglobin (1 nmol/uL) in buffered D20 (pD 7.0) at 45 0C overnight. Ion current from 200 laser shots was accumulated for each spectrum. Deuterium Exchange Experiments The deuterium exchange was initiated by diluting 1 [1L of the protein solution (- 750 pmol/uL) with 20 uL D20 containing 20 mM phosphate buffer (pH 7.0). After different time intervals, the exchange reaction was quenched by adding a 120-uL aliquot of matrix solution (pD or pH 2.5). An aliquot of 0.5 uL of the resulting solution was spotted on the MALDI sample plate and analyzed by MALDI-TOF-MS. All pH or pD values were measured using a Beckman (D 40 pH meter or ColorpHast pH indicator strips, pH 0-14 (EM Science, Gibbstown, NJ). pD values were calculated by adding 0.4 to the corresponding pH meter readings [33]. 78 III. Results and Discussion A. Complete Reduction of RN ase A Previously, reduction of the native RNase A was achieved by adding reduced DTT in pH 8.0 buffer [34]; our strategy was to reduce the protein with TCEP in aqueous solution at pH 3.0. Complete reduction is possible under such conditions and re- oxidation of the reduced protein is minimized due to the low pH. To confirm complete reduction, the reduction product was cyanylated by CDAP followed by cleavage into the nine cleavage products represented at the bottom of Figure 2.1. Table 2.1 lists the calculated m/z values of expected cleavage fragments and the experimentally observed values. The MALDI spectrum is shown in Figure 2.2. Seven peaks corresponding to seven out of the nine possible fragments appear in the mass spectrum, which confirms the reduced state of seven cysteine residues in the completely reduced RNase A. The missing two peaks corresponding to fragments with m/z 774.86 and 789.83 (corresponding to fragment Itz-58-64 and Itz-65-71, respectively) lead to ambiguity in the redox state of Cysés. However, if all the other seven cysteine residues are reduced in RNase A, Cysos also must be in the reduced state because it would have no partner for forming a disulfide bond. The signal corresponding to these two fragments is probably suppressed by ionization of the matrix or obscured by peaks representing matrix ions. The assignment of some peaks in Figure 2.2 (those labeled with “?”) has not yet been determined. They probably correspond to products of unidentified side reactions occurring during the cyanylation and cleavage process. These experimental results confirm the complete reduction of RNase A under our experimental conditions. 79 (D S i i_ls S 1 26 40 58 65 72 84 95 110 124 l TCEP, 0.1M, pH3, 37°C, 2Hrs 1 26 40 58 65 72 84 95 110 124 l CDAP, 0.2M, pH3, rt, 15min I] ll SCN jCN ISCNSCNSCN SCN SCN ISCN 26 40 58 65 72 84 95 110 124 1 lAmmonia, 1M, rt, 1hr 1-25 , ItzZ6-39 Itz40-57 It258-64 Itz65-71 Itz72-83 Itz84-94 Itz95-109 Itzl 10-124 Expected Cleavage Products Itz = iminothiazolidinyl carboxyl residue at the amino terminus of the cleavage products Figure 2. 1 Conceptual chemical scheme for complete reduction of bovine pancreatic RNase A, and confirmation of complete reduction by subsequent cyanylation, cleavage,and analysis of cleavage reaction mixture. 80 388% 8: u dd . .3558“ 0:58 05 “a 2668 5898 Rfiwzoumfiuofig n N: . N_ .002 05.32 mmfiofi mwémi dd Bfid 3.39 36m: afimonm $8 33030 owdmg coewg mmfiomfi mmdNE mwdwh 3.3% vmdwa 3.3: 5.35m NE @800an VNTS TN: 926%: vméwfiu mw-NhN: :Aofifi vc-wm§_ 36$: mmbmwfi. mm; “:2:me .owgfixo use down—>55 £28sz ouoEEoo wEBo=£ < 032% Set 32603 omm>m2o coemcoaoa .«o Aomfiogv 02$, N\E @8033 EB vows—3:3 — .N 0.5:. 81 .< 0872 .3 8260.5 ow~>w2o 05 no 82.38% 32: 5‘22 N .N 95$...— N\E 8% 88 $3 82 3: 2: _ 2.32 . , s 33%: $.28 :82 mm Ami s . . em 28 e 3.0:; m 3.2% was»: 2.32 \, 2.8.: \1 % Intensity 62 82 B. Effect of Deuterated and Non-deuterated Matrix As described in most current papers, a matrix solution with H20 and organic reagents (i.e., ACN) is usually used for MALDI-MS analysis of deuterated proteins [24- 26]. However, H20 is introduced to the deuteration reaction system when the non- deuterated matrix solution is mixed with the deuterated protein. As a result, a large portion of the incorporated deuteriums on the protein will be back-exchanged by hydrogens from H2O or carboxyl hydrogens of the matrix molecules (most matrix molecules used in MALDI-MS are carboxylic acids). It is assumed that back-exchange only happens to deuteriums attached to oxygen, sulfirr, or side-chain nitrogen, while all the amide deuteriums incorporated are retained. 0n the other hand, the use of deuterated matrix, as prepared with deuterated solvents and matrix that has been incubated in D20 overnight, has been reported [35]. The use of conventional non-deuterated matrix solution and deuterated matrix solution was compared using the firlly reduced RNase A as a model protein; the experimental results are discussed below. A comparison of observed deuterium incorporation using the deuterated and non- deuterated matrix is shown in Table 2.2. The spectra are shown in Figure 2.3. As shown in Figure 2.3 (A), using the deuterated matrix, after incubating with D20 (pH 7.0) for 10 s, the fully reduced RNase A showed an exchange level of over 95%. Longer incubation time with D20 to even 90 min does not improve the deuterium incorporation level (Figure 2.4). This implies that under the conditions described above, an equilibrium of H/D exchange can be achieved within 10 s, which is in agreement with the result obtained by theoretical calculation [36]. On the other hand, using conventional non- deuterated matrix, an exchange level of 50% (Table 2.2) was detected as shown in Figure 83 Table 2. 2 Comparison of observed deuterium incorporation level in reduced (unfolded) RNAse A using deuterated and non-deuterated matrix. RNase A Proteins (Completely reduced) Number of amino acid residues 124 (4 Pro) Average molecular weight (Da) 13690 Number of amide bonds 123 Number of amide hydrogens 119 (Expected mass shift if the amide hydrogens (120) are replaced by deuteriums) Number of side-chain hydrogens 126 Total number of hydrogens 245 (Expected mass shift if all hydrogens (246) are replaced by deuteriums) Observed mass shift with deuterated matrix 13927 — 2 - 13690 (Da) = 235 (Percent deuteration) ( _2_3_5 = 95.5%) 246 Observed mass shift with non-deuterated matrix (Da) 13814 - 1 - 13690 = 123 (Percent deuteration) (5% = 50.0%) 84 .mEiOHAQg as $9228 0.8.33 552: c88833é0: Amv no xEmE 33833 GE «23 858 :05 Ba ONO £3, comma—65 “can < omwzm 3268 bone—9:8 mo $8on 332 n .N ns.—arm Amv N\E coom~ oowom oowmz ooec_ z. - - r. mm woae m m , w . .m vzwm_ m con 3 N\E ooomm - oomom oowmz ocean comm ooozo 8% % m. .w. Aw nmom_ OS 85 .< 632M @0268 so XQE .«o 358 025. v .N ousumm 00? E25 9:: 20:33:. ow 00 CV _ h p $0 O O LO -cor .omw + 0 L0 N parerodroou! surngrarnep to '0" 86 2.3 (B). This corresponds to an extent of back-exchange of about 50%, which is also in agreement with previously reported results in similar experiments [24-26]. The extent of back-exchange is greatly reduced by employing deuterated matrix. In addition, the use of deuterated matrix facilitates the use of a fully deuterated internal standard to improve the accuracy and precision of mass measurement (see details in the following section). The use of non-deuterated matrix with fully deuterated internal standard is not feasible because it is difficult to control the back-exchange of the fully deuterated standard once it is mixed with non-deuterated matrix solution. Therefore, deuterated matrix could be a good choice in the study of intact proteins in systems with a low concentration of salts. However, if the polypeptide chain is dissected by peptic digestion before mass measurements, then it is probably more practical to use a non- deuterated matrix because, otherwise, all the reagents and chemicals involved in sample treatment after quenching HDX, including pepsin and desalting reagents, should be completely deuterated to match the use of the deuterated matrix. In practice, it is not always easy to satisfy this requirement. In the following sections of this chapter, a deuterated matrix was used in all experiments because all of them are focused on intact RNase A. In the other chapters, however, the conventional non-deuterated matrix was used when assessing deuterium incorporation of various peptic fragments of a protein. 87 C. Use of a Deuterium-labeled Internal Standard In TOF-MS measurements, time-of-flight to mass conversion is usually achieved with the use of external or internal standards, and the latter approach normally provides the highest accuracy and precision. In HDX-MS experiments, a fully deuterated internal standard is preferred over non-deuterated internal standards. This is because the non- deuterated species may be partly deuterated when it is mixed with the deuterated sample, and, thus, unpredictable change in the molecular weight of the non-deuterated standard will result in inaccuracy in mass determination. In this section, we compared the use of fully deuterated myglobin (d26o-Mb) for internal and external calibration with deuterated matrix. The results are summarized in Table 2.3. Native RNase A was used as the model protein in the experiments; it was exposed in D20 for various periods before mass measurement. Table 2. 3 Comparison of the HDX results using internal and external calibration. Incubation time 10 15 30 (S) Observed m/z with internal calibration 13879 13883 13883 (standard deviation ) (s = 3) (s = 3) (s = 2) Observed m/z with external calibration 13851 13864 13862 (standard deviation ) (s = 10) (s = 5) (s = 7) . Each value is the average of five repeated experiments 88 As expected, the results obtained by using an internal standard have much better precision than the corresponding results from the use of an external standard. In addition, the measured m/z value is lower using an external standard rather than an internal standard, which could be a result of systematic error caused by the use of an external standard. D. Ablation of Sodium Ion Adducts Early in the Analysis at a Given Spot on the MALDI Target When analyzing relatively large proteins (> 10 kDa) in the low-resolution linear mode, each isotope peak cannot be resolved, and a single broad peak is usually observed that consists of all the unresolved isotope peaks. A symmetric bell-shaped peak is usually required to determine the average mass of the molecule accurately and precisely. However, in our experiments, a tail in the distribution is usually observed early in the analysis at a given spot on the MALDI target (Figure 2.5 (A)). The tail might be caused by Na+-adducts of the protein under investigation (d26o-Mb in this case). Due to the low resolution of this type of analysis, the Na+-adducts peak cannot be resolved from the analyte peak. The sodium ions may come from various sources. For example, the H20 used in the experiments was stored in a glass bottle, and a trace amount of sodium ions could be dissolved from the glass into the water. In addition, the matrix and other chemicals used may contain a trace amount of salts including some sodium salts. The sodium ion adducts may cause overestimation of the m/z value of the analyte or poor reproducibility for repeated experiments. 89 .305 Home. on “3:2:an Am: ”macaw .69: cm raw on“ Bob :32 0333850 A2m§nmo$803205;mm359.92an3985355ficofimmoifimfimfifi: _I__ 98 Here, we use our cyanylation/cleavage methodology to trap the early forming 1- disulfide (l-S) intermediate and a later forming 2-disulfide (2-S) intermediate of LR3IGF-I as we reported previously [17]. The cyanylated l-S intermediates (Cys31-Cys74) and 2-S intermediates (Cys31—Cys74, Cyslg—Cysm) (disulfide structure was determined by our cyanylation/cleavage methodology) are isolated by rp-HPLC and then exposed to D20 at pD 6.8 for HDX. The intermediate is then digested with pepsin and analyzed by MALDI-MS to determine the extent of deuterium incorporation at approximate locations (limited by the size of the peptic fragments) of various amides along the polypeptide backbone. In this way, we can indirectly monitor both the progress of the refolding process as a fimction of disulfide bond formation and the conformational status of these trapped species by the degree of exposure of various amide hydrogens as assessed by HDX. Progressive “snap shot” results during these studies provide insight to the dynamic properties of the protein. II. Experimental Section The HDX experiments were carried out generally according to procedures described previously for assessment of deuterium incorporation by LC-MS [2], but in our work modified to allow removal of Gdn-HCl and other salts prior to analyses by MALDI- MS as described below. Our application also involves reduction of disulfide bonds, a step that complicates the sample processing relative to those previously reported applications of HDX as monitored by MALDI for proteins that do not contain disulfide bonds [23]. 99 Materials All reagents were of the highest purity available from commercial sources; they were used without further purification unless described specifically. NaHzPO4 was purchased from Mallinckrodt Baker (Phillipsburg, NJ). Gdn-HCl, tris (hydroxymethyl) aminomethane-hydrochloride (Tris-HCl), and sodium citrate were obtained from Invitrogen (Carlsbad, CA) and Spectrum Quality Products (Gardena, CA), respectively. ACN (HPLC grade) was from EMD Chemicals (Gibbstown, NJ). Immobilized pepsin was obtained from Pierce (Rockford, IL). C18 ZipTip® was obtained from Millipore (Bedford, MA). All other chemicals were purchased from Sigma-Aldrich. The TCEP solutions were freshly prepared before use. All pH or pD values were measured using a Beckman CD 40 pH meter or ColorpHast pH indicator strips, pH 0 — 14 (EM Science, Gibbstown, NJ). pD values were calculated by adding 0.4 to the corresponding pH meter readings [24]. Harvest of folding intermediates The oxidative folding, trapping, and separation of intermediates and the determination of their disulfide structure are done essentially as described previously [17]. LR3IGF-I (0.1 mg) was dissolved in 0.5 m1 of citrate buffer, pH 3.0, containing 6 M Gdn-HCl and 0.1 M TCEP. Reduction of the proteins was carried out at 37 °C for 2 hr; the reduced/denatured LR3IGF-I was purified by rp-HPLC, dried under reduced pressure, and stored at -70 °C. The refolding of reduced protein was initiated by diluting the reduced/unfolded protein sample with 0.10 M Tris-HCl buffer (pH 8.7), containing 1 mM oxidized Glutathione (GSSG), 10 mM GSH, 0.2 M KCl, and 1 mM EDTA, to a final protein 100 concentration of 0.1 mg/ml. The refolding intermediates were trapped in a time-course manner by removing aliquots (0.1 m1) of protein solution and mixing with 1.0 M HCl containing freshly prepared 0.2 M CDAP to give a solution of pH 2 — 3. Cyanylation of free thiol groups by the CDAP proceeded at room temperature for 10 min. The trapped intermediates were immediately separated and purified by HPLC. The column was a Vydac C13 (catalog no. 218TP54; S-um particle size, 300-A pore, 4.6 x 250 mm). Solvent A was 0.1% aqueous TFA. Solvent B was ACN containing 0.1% TFA. The linear gradient was 30 — 50% solvent B in 45 min at a flow rate of 1 ml/min. The HPLC fractions were collected manually, dried under reduced pressure, and stored at —70 °C for fisrther use. The identification of each intermediate was accomplished previously using the cyanylation/cleavage methodology [17]. HDX Each protein sample was equilibrated in aqueous phosphate buffer (20 mM, pH 7.0) for about an hour before the HDX experiment. HDX was initiated by diluting 1 uL of aqueous protein solution (100 pmol/uL) with 20 [LL of 20 mM phosphate buffer/D20, pD 6.8 at room temperature. After 10 8, exchange was “quenched” by decreasing pH and temperature of the reacting system to about pH 2.5 at 4 °C by adding 19 uL of aqueous TCEP solution (0.5 M, in 0.1 M citrate buffer containing 2 M Gdn-HCl, pH 2.5, 4 0C) to reduce the residual disulfide bonds. The proteins were exposed to D20 for 10 s, a period long enough to exchange deuteriums for “free” amide hydrogens [25], but short enough to avoid significant exchange with amide hydrogens involved in H-bonding or 101 “protected” by interior folds. Labeling time of less than 10 s requires the use of fast mixing and quenching instruments. Pepsin digestion and reduction of residual disulfide bonds Immobilized pepsin slurry (80 uL) was washed three times with 400 uL of chilled 0.1% TFA (pH 2.5) in a 0.65-mL siliconized tube and stored on ice. “Quenched” protein samples (40 uL) were added to the pepsin slurry and digested for l min on ice with gentle - mixing. The immobilized pepsin was removed by a 30-s centrifugation. Analysis by MALDI-TOF-MS Mass spectra were acquired on a PerSeptive Biosystems Voyager DE STR instrument (Applied Biosystems Inc., Framingham, MA). Data were acquired in the positive reflectron mode of operation. Accelerating voltage was 20 kV. Grid voltage and guide wire voltage were 76% and 0.04% of the accelerating voltage, respectively. The mirror voltage ratio was 1.12. Extraction delay time was 150 ns. Matrix was 5 mg/mL a-CHCA in a solution containing ACN, ethanol, and H20 (1:1 :1, v:v:v); the final pH was adjusted to 2.5 with TFA. During the analysis to obtain the data reported herein, the instrument was internally calibrated using the monoisotopic mass of two well- characterized peptic fragments (sequences LRRLEM and LSSLFVNGPRTLCGAELVDALQF; monoisotopic masses are 816.4640 and 2449.2675 Da, respectively). Typically, ion current from 256 laser shots was accumulated for one spectrum. The average mass of a peptide was calculated by determining the centroid of its distribution of isotope peak intensities using Data Explorer TM 4.0 installed by the manufacturer of the MALDI instrument. 102 The sample was desalted using a C13 ZipTip® (Millipore, MA) pipette tip before being eluted with matrix solution and spotted on a chilled MALDI target. The target was immediately placed in a desiccator under a moderate vacuum action of a mechanical vacuum pump such that the spots would dry within 1 — 2 min. Then, the target was transferred to the MALDI mass spectrometer as soon as possible. The completely deuterated protein standard was prepared by dissolving LR3IGF-I in DzO at pD 7.0 containing 8 M d4-urea and incubating at room temperature for over 48 hr; this fully deuterated species was treated by the same sample-handling protocol as all the other proteins to determine the extent of back-exchange during sample processing after HDX. The extent of back-exchange during all the experimental procedures described ranged from 38% to 47%. III. Results and Discussion A. Experimental Design for the Application of HDX-MS to Cystinyl/Cysteinyl Proteins The key procedure leading to optimal HDX results is to minimize back-exchange during sample treatment after deuterium incorporation is “quenched”. If all (or most of) the incorporated deuterimns are substituted by hydrogens from aqueous solution before mass measurement, then the HDX results will be meaningless. For the most part, HDX- MS has been widely used in the conformational study of proteins containing no disulfide bonds [1, 23, 26]. Few researchers, if any, have applied this approach to cystinyl proteins. This is probably because the special features of cystinyl proteins require special 103 sample treatment before analysis by MS, where significant back-exchange could easily occur. To investigate the extent of deuterium incorporation in different segments along the polypeptide chain of the protein, deuterated proteins are usually digested before mass measurement. Proteolytic digestion of cystinyl proteins is usually very slow, if not impossible [27, 28], unless the protein is denatured and its disulfide bonds are reduced. Therefore, the disulfide bonds in a deuterated cystinyl protein should be reduced to facilitate digestion and peptide identification. This leads to two problems that may cause significant back-exchange. The first is that only reducing reagents effective under the conditions required for minimum back-exchange (pH ~3.0 and low temperature) can be used, and even then, the reaction has to be rapid enough to avoid significant back- exchange. The other problem is that salts introduced to the system have to be removed after reduction to facilitate analysis by MS. Significant back-exchange could occur during this desalting step both because it is time consuming and large amounts of aqueous solvents are used. This is less of a problem in HDX-LC-MS experiments because most LC-MS instruments include an automatic sample-desalting step prior to injection of the sample into the mass spectrometer. After the sample is loaded, it is usually flushed with copious inorganic solvents to remove salts and other inorganic contaminants, which normally takes only 1 or 2 min. When pre-cooled solvent of proper pH (~ 3.0) is used, the extent of back-exchange can be held to less than 20% [29, 30]. In the case of HDX-MALDI-MS, however, even for non-cystinyl proteins that usually do not require desalting, the extent of back-exchange is usually around 50% [23, 31]. The desalting has to be done manually before loading the MALDI plate to the mass 104 spectrometer, and, thus, special sample treatment protocols have been developed to minimize possible back-exchange during this step. To solve the first problem, TCEP, which is active at pH ~3.0, was used for the reduction of disulfide bonds. Its concentration was high (0.5 M) to accelerate reduction (within a few minutes) at low temperature (4 oC). Gdn-HCl was added to the system to denature the protein, thus facilitating the reduction of disulfide bonds buried in the native protein. To shorten digestion and reduction time, TCEP solution (0.5 M, in 0.1 M citrate buffer containing 2 M Gdn-HCl, pH 2.5, 4° C) was added to the HDX system at appropriate time intervals to “quench” deuterium incorporation by decreasing the pH of the system to ~30. The reduction of disulfide bonds starts immediately after adding TCEP. The mixture is then promptly transferred to the pepsin slurry so that the digestion and reduction occur simultaneously. A large amount of pepsin (Enzyme : Substrate = 30 : 1) was used to achieve significant digestion within a short time. Under our experimental conditions, complete reduction and digestion are achieved within 1 min, even shorter than the reported time required for most non-cystinyl proteins (usually between five and ten minutes [23, 29]). Efficient reduction/digestion minimizes back- exchange during this step. TCEP, Gdn-HCl and Tris-HCI are salts introduced to the system during reduction and digestion; they must be removed before the sample is mixed with matrix for MALDI- MS analysis. The Zip-Tip® is a novel desalting tool developed by Millipore (Billerica, MA), and can be thought of as a HPLC column at the tip of a pipette tip. The Zip-Tip® pipette tip used in our experiments was a lO-uL pipette tip with a bed of chromatography media (C13 material) embedded at its end. To use a Zip-Tip® pipette tip for desalting, 105 first the pipette tip is wetted and equilibrated using first organic and then aqueous solvent. Peptides/proteins are bound to the C13 bed by aspirating and dispensing the sample solution for a few cycles. The analytes bound to the tip are then washed by aspirating water into the tip and then dispensing the water. Finally, desalted proteins/peptides are eluted from the tip by flushing the C13 bed with an elution solution that usually contains 50% organic solvent (i.e., ACN). To minimize back-exchange, the Zip-Tip® and all the desalting solvents used were pre-cooled to 4 °C, and the desalting solvents were adjusted to pH ~3.0. Afier some practice, the whole desalting procedure can be performed within one minute. The designed procedures serve to minimize back- exchange. The extent of back-exchange during the entire experimental procedure described ranged from 38% to 47%, typical in HDX-MALDI-MS experiment [23, 31]. The variation of back-exchange among different peptides is presumably due to side-chain effects, which can alter the amide hydrogen exchange rate under quenching conditions by as much as 10-fold [32]. Because the back-exchange rate of different peptic fragments varies slightly [2, 32], the adjustment of back-exchange for each peptide was based on its specific back-exchange value (between 38% and 47%) as determined during the “control” experiment with the completely deuterated LR3IGF-I. B. HDX Behavior of Different Segments of when Ten peptic fragments of LR3IGF-I (corresponding to 78% sequence coverage) were identified by molecular weight after analysis of the peptic digest mixture by MALDI-MS, using the results of an earlier report on the peptic digestion of IGF-I as a guide [2]. The mass spectrum of the digest is shown in Figure 3.2. The sequences of the 106 ten peptic fragments examined by MS following HDX are represented by the arrows below the following sequence of LR3IGF-I, which also shows the three disulfide bonds of the native structure (Figure 3.1). It should be pointed out that some of the sites for peptic cleavage are different in the oxidized and reduced forms of LR3IGF-I; this is because the selectivity of pepsin is conformation dependent. Of the ten peptic fragments that were common to most of the oxidized and reduced forms of LR3IGF-I, seven were identified by having a calculated mass that was within i10 ppm of the m/z value for the corresponding observed mass spectral peak; further, no other possible peptic fragment had a calculated mass within i500 ppm of the experimentally observed m/z values for these seven peptides. Each of the other three peptic fragments had alternative candidates having a calculated mass within :50 ppm of the experimentally observed m/z value. In each of these three cases, the peptic fragment chosen was the one that agreed with the results of a published study of the peptic digest of IGF-I [2], and each had a calculated mass within 115 ppm of the observed values. As discussed in the following sections, the HDX behavior of each identified peptic fragment is consistent with its known conformational features, which, in turn, confirms the correct identification of these peptic fragments. 107 Eur”: do some 2:5 do 838% use 865?: N .m 2:2... N\E 82 . on: 1 8th _ om: 5: . one 33E} r -1 1 4:- - 1. a... on 22 \ $3: 8.0% . 6.32 8 8mm 8.82 8%: 8.22 $62: $4.th C 52 Kuwaiti} % 108 A summary of the mass shifts during HDX (corrected for back-exchange) for these ten peptic fragments (listed according to residue number in the sequence of LR3IGF-I) is presented in Table 3.1 for the completely reduced (R), l-S intermediate (1S: Cys31—Cys74), 2-S intermediate (28: Cys31—Cys74, Cyslg—Cysm), and native forms (N: Cys31—Cys74, CYS19—CYS61, Cys6o—Cys65) of LR3IGF-I. After HDX of the completely reduced protein, the mass shift of most peptic peptides is approximately the same as the number of exchangeable amide hydrogens, suggesting that these amide hydrogens are solvent accessible and not engaged in H-bonds. Additionally, the sequential formation of the three disulfide bonds in LR3IGF-I appears to afford progressively increasing global protection against hydrogen exchange as reflected in the corresponding diminution in mass shift for most of the peptic fragments; these coordinated trends indicate a mode of folding dependent on the formation of disulfide bonds. At the stage of having formed only one disulfide bond (Cy531—Cys74), the mass shift of each peptic fragment is slightly decreased, except in the fragment consisting of residues 38 - 55. Thus, the connection of Cys31 and Cysu may initiate partial secondary structural conformations that involve H-bonds or reduce solvent accessibility as deduced from only slightly decreased mass shifts. The mass shift data from analysis of peptic fragments of the 2-S intermediate shows that the number of exchangeable amide hydrogens declined significantly after the formation of disulfide bonds Cys31—Cy574 and Cyslg—Cysm, but the amide hydrogens in fragment 38 — 55 are still freely exchangeable. Finally, peptic fragments of the native protein (three disulfide bonds) after HDX show mass shifts similar to those in the results for the 2-S intermediate. Certain fragments that 109 murmufim warn—Eu :2 G8» do.“ :83 02 .mnwmm u 3 .o .3235 ooaovcaon. $3 05 88%.: 3:8 :H: as: E3532: 8:: Bob 8:58 he owns; 05 2a 829, vouched d doufioEmaoo 35 5x8 35 ONO wfivvu “can ESQ? XO: 2: E ON: 9&4. ~§2m2 05 633058.33 no.“ dogma?“ soon 96: mafia v.32 .m dd H 06 Nd H dd Nd H E. md H ms w mmén _dHN._ _de.~ N.de.m fideé m 2mg; NBS md H Nd_ vd H wdfi md H v.3 e B M: omEHOHogmmmO>OH§ oodv fid H ”E fl H 2 md H N.m_ md H WE 2 OHOmgmmmoCGHmMzL mmém Nd HoN vd H dN vd H Nd e 3 h ”#5500?" emdN Nd H dN Nd H 0N _d H wd e 3 m >momaco>mo wwwwN vd H dd Nd H mm dd H Th e3 3 >mOMDOU>mOA RNN Nd H dd md H ed Nd H Nd Nd H dd 2 mm NN-: $535525 8; d Nd H cs cd H a.» Nd H dd e3 Z Eni \NNA: ea $0-8 . 5 28-2 2823: 5-2 .42sz «Em ”ms 5 3-: Hm: 3 a 8:8 as $5 mmmz as dEm 932 A“ E5 334 3 cu? $32 do 59:52 oososvom :oEmom .mEoEmfim 2an 85:52 :0 33%: KO: no bdgm w .n 935,—. 110 have their amide hydrogens further protected may consist of residues involved in the formation of secondary and tertiary structure as discussed in the following section. C. Comparison of HDX-MS Results with NMR Data on the Native Structure of LR3IGF-I The HDX results reported in Table 3.1 are consistent with NMR data on the native structure of LR3IGF-I (Figure 3.3) and IGF-1, including those from studies using HDX, which provide an alternate means of studying the dynamics of cystinyl proteins [7- 10]. LR3IGF-I is an or-helical orthogonal bundle protein composed of three or helixes and three disulfide bonds. The l-S intermediate in our study has a disulfide bond within Cys31-Cys74 that connects the C termini of helix 1 (21 - 31) and helix 3 (67 — 73) as illustrated in Figure 3.3. The differences observed in our MS study between the mass shift of peptic fragments from the l-S intermediate and the number of constituent amide hydrogens occur mostly for fragments that are in the vicinity of either Cys31, Cys74, or the helixes in the NMR structure. This observation suggests that the Cys31—Cys74 disulfide bond induces the early formation of secondary structures involving H-bonds and/or stabilizes the local conformational fluctuations that allow transient solvent penetration. The additional disulfide bond Cyslg—Cysm in 2-S seems to close up the hydrophobic core within the three helixes of the NMR structure (Figure 3.3), further reducing the accessibility of solvent; this development apparently promotes the development of secondary and tertiary structure and stabilizes the conformation of the protein. In this regard, the NMR data are in reasonable agreement with our MS HDX results. The mass- shift data from MS analysis of peptic fragments 67 — 72, 73 — 83, and 10 - 22/11 — 23 from the native protein suggest that the CYSGO‘CySQSS disulfide bond reinforces the 111 coupled conformation of helix 2 and helix 3 and increases their contact with helix 1, further diminishing access of the solvent to amide hydrogens. Helix 2 '5 (56-63) Helix 3 (67-73) Figure 3. 3 Average NMR structure of LR3IGF-I [8]. N-terminus and C-terminus of the protein are labeled by N and C, respectively. Three helixes are labeled according to inclusive residues. Three disulfide bonds are represented by dotted lines between the designated cysteines. The dark-gray region is the structurally ordered part of the protein; the light-gray segments are highly flexible parts of the protein (in which the heavy atom r.m.s.d. is larger than the average r.m.s.d. of the ordered part of the protein). The figure was drawn using the RasMoI software [33]. 112 In the HDX-NMR studies [7, 9, 10], all the helixes in native LR3IGF-I and IGF-1 were found to contain slowly exchanging amide hydrogens, indicating some local structure; these NMR data are similar to our MS results. However, the HDX-NMR study of LR3IGF-I [7] also found slowly exchanging amide hydrogens among the first ten residues of the N-terrninal tail and in the segment consisting of residues 43 — 51; this indication of some structure among residues 43 — 51 contradicts the results from the study of IGF—I [9, 10], although results from another study by the same group [8] indicate that this segment is highly flexible as indicated by a much higher than average heavy atom root mean square deviation (r.m.s.d.) of LR3IGF-I and an increase in T 2 values for several residues (Figure 3.3), suggesting that solvent was accessible to this region. Surprisingly, our MS data from analyses of peptic fragments 40 —- 60 and 38 - 55 indicate that both of the above-described scenarios may occur simultaneously. The mass shifi of the segment 40 — 60 is smaller than that of segment 38 — 55 as listed in Table 3.1; this disparity seems strange at first glance because the latter segment is almost completely included in the former. However, a reasonable explanation could be that these two peptic fragments are actually fiom two different conformational states of a given redox state and that pepsin can distinguish between them. This dual-conforrnation explanation is supported by NMR results [7, 8], where two distinct conformations of Gly55 were detected, which indicate that the backbone between Lys4o and Cyséo has two different conformations in native LR3IGF-I, which might be induced by isomerization of Pro.“ and/or Prosz. Therefore, pepsin may be able to attack cleavage sites in the Lys4o/Cy560 segment in one conformational state of LR3IGF-I, but not in the other. The backbone, including the Phe3g/G1Y55 segment, is probably much less structured in one of 113 the two conformational states. In contrast, at least some of the Phe3g/GIYS5 segment is more structured in the other conformational state, a condition precluding access of pepsin, and the resulting missed cleavage gives rise to the peptic fragment 40 — 60. This sort of evidence for the putative dynamic conformational equilibrium observed in the NMR studies [7, 8] and here in our HDX-MS study of LR3IGF-I has also been reported in an NMR study of highly mobile regions of an ATPase protein [34]. D. Some Figures of Merit for the Integrated Methodologies Some figures of merit for our HDX methodology can be gleaned from a detailed examination of our results for two of the ten peptic fragments of LR3IGF-I listed in Table 3.2. Analyses of one of these peptic fragments, a 6-mer (LRRLEM), from triplicate HDX experiments shows considerable slow amide hydrogen exchange in one region of various redox states of LR3IGF-I, while results from analyses of an 18-mer (FNKPTGYGSSSRRAPQTG) indicate fast amide hydrogen exchange in a different region; these results are in agreement with an NMR study of the conformation of native IGF—I and LR3IGF-I [8-10]. The extent of deuterium incorporation into the two peptic fragments of interest (the 6-mer and the 18-mer described earlier) is summarized in Table 3.2. For each peptic fragment, the listed constituent residues are identified according to their residue position in the sequence of LR3IGF-I. The number of amide hydrogens in the fragment (in the longer peptic fragment, the number of amide hydrogens is less than the number of residues because of constituent prolines, which do not have an amide hydrogen) together with the experimentally determined number of deuterium atoms incorporated (not adjusted for back-exchange) is also listed. 114 Table 3. 2 Summary of mass shifts for two peptic fragments from different red-ox species after 10-sec HDX. (Results are the average from three independent experiments; plus/minus terms indicate the 95% confidence interval). (A) 6-mer (Sequence = LRRLEM; number of amide hydrogens = 5) Disulfide Species R 1S 2S N (31-74) (31-74, 19-61) (31-74, 19-61, 60-65) Experimental mass shift (Da) 2.3 :t 0.05 1.7 i 0.1 0.77 i 0.05 0.59 :t 0.05 Adjusted 4.5 3.3 1.5 1.2 mass shift (Da) Percent deuteration 90% 66% 30% 24% 4.5/5.0 3.3/5.0 1.5/5.0 1.2/5.0 Degree of “protection” 10% 34% 70% 76% (5.0-4.5)/5.0 (5.0-3.3)/5.0 (5.0-1.5)/5.0 (5.0-1.2)/5.0 (B) 18-mer (Sequence = FNKPTGYGSSSRRAPQTG; number of amide hydrogens = 15) Disulfide Species R 15 2S N (31-74) (31-74, 19-61) (31-74, 19-61, 60-65) Experimental mass shift 8.5 :t 0.2 8.8 i 0.2 8.5 i 0.7 8.5 :t 0.1 (Da) Adjusted 14.8 15.2 15 14.8 mass shift (Da) Percent deuteration 98% 100.7% 99.3% 98.7% 14.8/15.1 15.2/15.1 15/15.l 14.8/15.1 Degree of “protection” 2% -0.7% 0.7% 2% (15.1-14.8)/15.1 (15.1-15.2)/15.1 (15.1-15)/15.1 (15.1-14.8)/15.1 115 In particular, examine the HDX results for the peptic fragment 67 — 72. In consideration of back-exchange, the listed 2.3-deuterium atom incorporation (or 4.5 Da after adjustment for back-exchange) in Table 3.2 (A) means that essentially all five amide hydrogens were exchanged for deuterium atoms during the HDX experiment; this result indicates that little protection against exposure to D20 is afforded these five amide hydrogens in the reduced (R) form of LR3IGF-I. In Table 3.2 (A), the listed 1.7- deuterium atom incorporation means that 3.3, on average, of the five amide hydrogens were exchanged for deuterium atoms during the HDX experiment with l-S. This reduced level of deuteration of 66% (3.3/5.0) indicates that some degree of secondary structure appears in the l-S intermediate, which partially protects the five amide hydrogens from complete exposure to the D20 during HDX. We numerically indicate this degree of protection as the difference between the maximum possible shift and the observed mass shift relative to the maximum possible shift (in this case, (5 - 3.3)/5 x 100% = 34%). The results for the 2-S intermediate show a substantially smaller mass shift of 0.77 Da, which gives an adjusted mass shift of 1.5 Da; this corresponds to 30% deuteration or 70% protection, meaning that there is significantly more secondary structure in this segment after the second disulfide bond has formed. Finally, the HDX results for the native structure show the smallest mass shift of 0.59 Da, which gives an adjusted mass shift of 1.2 Da; this corresponds to 24% deuteration or 76% protection, which means that the native structure contains the greatest amount of secondary structure in this segment, as expected. The results for HDX of residues 67 — 72 in the native structure of LR3IGF-I are in good agreement with HDX assessment of the conformation of these structures using NMR [7, 9, 10]. 116 Abbreviated segments of the MALDI mass spectra of the peptic fragment 67 - 72 after HDX of the reduced, l-S intermediate, 2-S intermediate, and native forms of LR3IGF-I are shown in Figure 3.4 (A). The mass spectrum of the normal (no HDX) peptic fragment (6-mer) of LR3IGF-I shown in the top panel of Figure 3.4 (A) serves as a reference point from which to compare the mass shift associated with this peptic fragment after HDX of the polypeptide in the various redox states indicated in the lower four panels of Figure 3.4 (A). The mass shift is assessed by determining the centroid of the multiplet of isotope peak intensities representing the peptic fragment (monoisotopic mass with no HDX = 816.4640 Da). The centroid of the multiplet of peaks in the third panel of Figure 3.4 (A) indicates a mass shift of 2.3 Da from the reference point in the top panel of Figure 3.4 (A); this result, when adjusted for back-exchange, gives the shift of 4.5 Da for the reduced (R) protein reported in the second column, second row of Table 3.2 (A). The centroid of the multiplet of peaks in the fourth panel of Figure 3.4 (A) indicates a mass shift of 1.7 Da from the reference point in the top panel of Figure 3.4 (A); this result, when adjusted for back-exchange, gives the shift of 3.3 Da for the l-S intermediate reported in Table 3.2 (A). Similarly, the centroid in the fifth panel of Figure 3.4 (A) indicates a mass shift of 0.77 Da, which corresponds to the adjusted value of 1.5 Da in Table 3.2 (A) for the 2-8 intermediate. Finally, the centroid in the sixth panel of Figure 3.4 (A) indicates a mass shift of 0.59 D3, which corresponds to the adjusted value of 1.2 Da in Table 3.2 (A) for the native form of LR3IGF-I. The mass spectrum in the second panel of Figure 3.4 (A) (corresponding to the firlly deuterated species) served as the basis for determining the degree of back-exchange under the sample-handling conditions described above. 117 Figure 3. 4 Mass spectra of (A) the segment 67 - 72 (LRRLEM, 6-mer) from different redox species and (B) the segment 38 — 55 (FNKPTGYGSSSRRAPQTG, l8-mer) from different redox species. The first row shows the mass spectra of the peptides before deuteration; the second row shows the mass spectra after sustained HDX to achieve a high degree of deuteration of the peptides. The remaining four rows are mass spectra of indicated redox species (R: reduced; 1-S: one-disulfide intermediate; 2-S: two-disulfide intermediate, N: native) after a 10-s exposure to D20 at pD 6.8. 118 v .m an; a: g N\E N\E ode ”32 v.32 v.53 N422 odoE odmw NNNw dem oANw Wm; 0.05 I. ’ @252 7323 t 2: 2e} 2 1.53 3.3 73%; nooseom_ E d , CO—HEOuH-Oo 11 1 Wage-1.4 é. , 1-311 11 - 9.039.939“. : 1 111 02 119 For comparison, examine the HDX results in Table 3.2 (B) for the peptic fragment 38 — 55. In consideration of back-exchange, the 8.5-deuterium atom incorporation listed in Table 3.2 (B) means that essentially all fifteen available amide hydrogens were exchanged for deuterium atoms during the HDX experiment. This result indicates that no protection against exposure to D20 is afforded these fifteen amide hydrogens in the reduced (R) form of LR3IGFI. Similarly, the HDX results for the fifteen hydrogens in this eighteen-residue peptic fragment from all other red-ox states indicate that no protection is ever apparent, and thus little protection from solvent is available to this segment of the sequence in any of the redox states of LR3IGF-I. The abbreviated mass spectra of the peptic fragment 38 — 55 (lS-mer) from which the data in Table 3.2 (B) were gathered for different redox states of LR3IGF-I are shown in Figure 3.4 (B). As is apparent from a cursory examination of the mass spectra, a substantial and nearly constant mass shift was detected in all cases, indicating that none of the amide hydrogens is afforded protection from exposure to DZO. These results indicate that there is no significant conformational structure among residues 38 — 55 in any of the indicated redox states of LR3IGF-I. Further examination of the mass spectra in Figure 3.4 (B) might suggest a lesser mass shift in the third panel than in the lower three; however, the mass shift in all five panels is essentially the same as indicated by the centroid calculation on the cluster of peaks in each panel. The major factor in the lower three panels is apparently the substantial leading edge to the most intense peaks; because the magnitude of the ion current in the lower three panels is two to three times that in the third panel, the leading edge in the spectra is real with regard to ion counting statistics, and it causes the centroid to be approximately the same as that in the third panel. 120 The thiocyanocyanate group on the modified cysteines is not expected to influence the conformation of the intermediates in any significant way, as reasoned by analogy to NMR work by Talluri et a1. [35], who used the (2-aminoethyl) methanethiosulfonate derivative of cysteine, a blocking group much larger than the thiocyanate group. IV. Conclusion These results demonstrate that cyanylation-based disulfide mapping methodology can be combined successfully with HDX methodology for a more comprehensive assessment of the conformation of cystinyl proteins as a function of cysteine status. The cyanylation-based methodology is effective in determining and controlling the cysteine status in cystinyl proteins during assessment of its conformation by HDX. The results reported here show that these integrated methodologies can monitor a clear and distinct dependence of the conformation of a cystinyl protein on the progressive formation of its constituent disulfide bonds during oxidative refolding. Our results of HDX on LR3IGF-I and its 2-disulfide folding intermediate as monitored by MALDI-MS provide evidence for two dynamic conformations of a segment of the polypeptide backbone; this finding is consistent with results from several NMR studies of LR3IGF-I and IGF-1 [7-10]. In this way, the integrated methodologies will help elucidate the detailed molecular mechanisms of protein dynamics and protein folding. 121 V. Reference [1] J. R. Engen, D. L. Smith, Investigating protein structure and dynamics by hydrogen exchange MS, Anal Chem 73 (2001) 256A-265A. [2] H. Ehring, Hydrogen exchange/electrospray ionization mass spectrometry studies of structural features of proteins and protein/protein interactions, Anal Biochem 267 (1999) 252-9. [3] T. E. Wales, J. R. Engen, Hydrogen exchange mass spectrometry for the analysis of protein dynamics, Mass Spectrom Rev (2005). [4] I. A. Kaltashov, S. J. Eyles, Studies of biomolecular conformations and conformational dynamics by mass spectrometry, Mass Spectrom Rev 21 (2002) 37-71. [5] Y. H. Zhang, X. Yan, C. S. Maier, M. I. Schimerlik, M. L. Deinzer, Conformational analysis of intermediates involved in the in vitro folding pathways of recombinant human macrophage colony stimulating factor beta by sulfhydryl group trapping and hydrogen/deuterium pulsed labeling, Biochemistry 41 (2002) 15495-504. [6] Y. Bai, T. R. Sosnick, L. Mayne, S. W. Englander, Protein folding intermediates: native-state hydrogen exchange, Science 269 (1995) 192-7. [7] L. G. Laajoki, E. Le Breton, G. K. Shooter, J. C. Wallace, G. L. Francis, J. A. Carver, M. A. Keniry, Secondary structure determination of lSN-labelled human Long- [Arg-3]-insulin-like grth factor 1 by multidimensional NMR spectroscopy, FEBS Lett 420 (1997) 97-102. [8] L. G. Laajoki, G. L. Francis, J. C. Wallace, J. A. Carver, M. A. Keniry, Solution structure and backbone dynamics of long-[Arg(3)]insulin-like growth factor-I, J Biol Chem 275 (2000) 10009-15. [9] A. Sato, S. Nishimura, T. Ohkubo, Y. Kyogoku, S. Koyarna, M. Kobayashi, T. Yasuda, Y. Kobayashi, lH-NMR assignment and secondary structure of human insulin-like growth factor-1(IGF-I) in solution, J Biochem (Tokyo) 111 (1992) 529-36. [10] A. Sato, S. Nishimura, T. Ohkubo, Y. Kyogoku, S. Koyarna, M. Kobayashi, T. Yasuda, Y. Kobayashi, Three-dimensional structure of human insulin-like growth factor-I (IGF-I) determined by lH-NMR and distance geometry, Int J Pept Protein Res 41 (1993) 433-40. [11] J. S. Weissman, P. S. Kim, Reexamination of the folding of BPTI: predominance of native intermediates, Science 253 (1991) 1386-93. 122 [12] T. E. Creighton, Protein folding coupled to disulphide bond formation, Biol Chem 378 (1997) 731-44. [13] E. Welker, W. J. Wedemeyer, M. Narayan, H. A. Scheraga, Coupling of conformational folding and disulfide—bond reactions in oxidative folding of proteins, Biochemistry 40 (2001) 9059-64. [14] Y. Yang, J. Wu, J. T. Watson, Disulfide mass mapping in proteins containing adjacent cysteines is possible with cyanylation/cleavage methodology, J. Am. Chem. Soc. 120 (1998) S834-5. [15] J. Wu, J. T. Watson, A novel methodology for assignment of disulfide bond pairings in proteins, Protein Sci 6 (1997) 391-8. [ 16] J. Wu, Y. Yang, J. T. Watson, Trapping of intermediates during the refolding of recombinant human epidermal growth factor (hEGF) by cyanylation, and subsequent structural elucidation by mass spectrometry, Protein Sci 7 (1998) 1017-28. [17] Y. Yang, J. Wu, J. T. Watson, Probing the folding pathways of long R(3) insulin-like growth factor-I (LR(3)IGF-I) and IGF-I via capture and identification of disulfide intermediates by cyanylation methodology and mass spectrometry, J Biol Chem 274 (1999) 37598-604. [18] R. Gill, P. De Meyts, J. Pitts, C. Verma, A. Wollrner, S. Wood, Structure and function of human insulin-like grth factor-1, Recent Research Developments in Protein Engineering 2 (2002) 105-34. [19] O. G. Isaksson, A. Lindahl, A. Nilsson, J. Isgaard, Mechanism of the stimulatory effect of grth hormone on longitudinal bone growth, Endocr Rev 8 (1987) 426- 38. [20] W. H. Daughaday, P. Rotwein, Insulin-like growth factors I and II. Peptide, messenger ribonucleic acid and gene structures, serum, and tissue concentrations, Endocr Rev 10 (1989) 68-91. [21] R. King, J. R. Wells, P. Krieg, M. Snoswell, J. Brazier, C. J. Bagley, J. C. Wallace, F. J. Ballard, M. Ross, G. L. Francis, Production and characterization of recombinant insulin-like grth factor-1(IGF-I) and potent analogues of IGF-I, with Gly or Arg substituted for Glu3, following their expression in Escherichia coli as fusion proteins, J Mol Endocrinol 8 (1992) 29-41. [22] G. L. Francis, M. Ross, F. J. Ballard, 8. J. Milner, C. Senn, K. A. McNeil, J. C. Wallace, R. King, J. R. Wells, Novel recombinant fusion protein analogues of insulin-like grth factor (IGF)-I indicate the relative importance of IGF-binding 123 protein and receptor binding for enhanced biological potency, J Mol Endocrinol 8 (1992) 213-23. [23] J. G. Mandel], A. M. Falick, E. A. Komives, Measurement of amide hydrogen exchange by MALDI-TOF mass spectrometry, Anal Chem 70 (1998) 3987-95. [24] P. K. Glasoe, F. Long, Use of glass electrodes to measure acidities in deuterium oxide, Notes 64 (1960) 188-90. [25] Y. Bai, J. S. Milne, L. Mayne, S. W. Englander, Primary structure effects on peptide group hydrogen exchange, Proteins 17 (1993) 75-86. [26] Z. Zhang, D. L. Smith, Thennal-induced unfolding domains in aldolase identified by amide hydrogen exchange and mass spectrometry, Protein Sci 5 (1996) 1282-9. [27] D. L. Smith, Z. R. Zhou, Strategies for locating disulfide bonds in proteins, Methods Enzymol 193 (1990) 374-89. [28] J. Qi, J. Wu, G. A. Somkuti, J. T. Watson, Determination of the disulfide structure of sillucin, a highly knotted, cysteine-rich peptide, by cyanylation/cleavage mass mapping, Biochemistry 40 (2001) 4531-8. [29] J. R. Engen, B. E. Morton, X. Chen, Using Stable-Isotope-Labeled Proteins for Hydrogen Exchange Studies in Complex Mixtures, Analytical Chemistry 74 (2002) 1680-6. [30] Y. Deng, H. Pan, D. L. Smith, That Hydrogen Exchange at Individual Peptide Amide Linkages Can Be Determined by Collision-Induced Dissociation Mass Spectrometry, Journal of the American Chemical Society 121 (1999) 1966-7. [31] G. S. Anand, C. A. Hughes, J. M. Jones, S. S. Taylor, E. A. Komives, Amide H/2H exchange reveals communication between the CAMP and catalytic subunit- binding sites in the R(Dalpha subunit of protein kinase A, J Mol Biol 323 (2002) 377-86. [32] D. L. Smith, Y. Deng, Z. Zhang, Probing the non-covalent structure of proteins by amide hydrogen exchange and mass spectrometry, J Mass Spectrom 32 (1997) 135-46. [33] R. A. Sayle, E. J. Milner-White, RASMOL: biomolecular graphics for all, Trends Biochem Sci 20 (1995) 374. [34] Y. T. Chou, J. F. Swain, L. M. Gierasch, Functionally significant mobile regions of Escherichia coli SecA ATPase identified by NMR, J Biol Chem 277 (2002) 50985-90. 124 [35] S. Talluri, D. M. Rothwarf, H. A. Scheraga, Structural characterization of a three- disulfide intermediate of ribonuclease A involved in both the folding and unfolding pathways, Biochemistry 33 (1994) 10437-49. 125 CHAPTER 4 CHARACTERIZATION OF PEPTIDE FOLDING NUCLEI BY HYDROGEN/DEUTERIUM EXCHANGE-MASS SPECTROMETRY AND CIRCULAR DICHROISM I. Introduction Protein folding is a central problem of molecular biophysics. An early model of folding hypothesized that proteins adopt “flickering” native structure in high-propensity segments, and that protein folding proceeds by the gradual association and mutual stabilization of such marginally stable segments [1, 2]. Consistent with this hypothesis, mixing synthetic peptide fragments of a protein in sufficiently high concentrations can produce folded species analogous to those in the corresponding segments of the intact protein [3]. On the other hand, numerous experiments show that most peptides corresponding to secondary-structure elements (e.g., tit-helices, B-hairpins, B-meanders, etc.) from naturally occurring proteins exhibit little secondary structure in aqueous solution [4], probably owing to competition between water-amide and amide-amide hydrogen bonds (H-bonds) [5, 6]. Therefore, long-range interactions are usually invoked to account for the stabilization of short-range structure during protein folding. However, the mechanisms by which interactions between peptides stabilize the structure within peptides have not been studied systematically. It would be useful, for example, to assess the correlations between secondary, side-chain, and tertiary structure in pairs of covalently linked peptides and to characterize quantitatively the stabilization of local structure conferred by various factors, e.g., amino-acid sequence, specific side-chain 126 packing, steric, and/or hydrophobic exclusion of competing water molecules, etc. For clarity, we distinguish three types of structure. The degree of secondary structure is defined by the average number of specific H-bonds between amide groups of the protein backbone, which imposes local (i.e., short-range in the polymer sense) conformational order on the backbone. The average is taken over the equilibrium conformational ensemble under the experimental conditions (e.g., the thermodynamic average). A polypeptide system has stable secondary structure if, in the thermodynamic average, the majority of its amide groups are involved in H-bonds with specific amide groups. The degree of tertiary structure is defined by global conformational order, i.e., the ordering of the protein backbone in space as measured by the mean CQRMSD within the conformational ensemble. Thus, a polypeptide system has stable tertiary structure if the majority of conformations in its equilibrium ensemble are within 6 A CaRMSD of a single, physically plausible conformation. Finally, the degree of side-chain structure is defined by the ordering in the dihedral angles of the protein side chains. Such structure is often associated with tertiary structure (side chains are constrained because of dense packing), but may also arise from local secondary structure. For example, a specific side- chain rotarner may predominate because of steric interactions or H-bonding to the backbone (e. g., a-helix capping) or other side-chains interactions (e. g., salt bridges or or- helices). Thus, a polypeptide system has stable side-chain structure if, on the average, its side-chain x dihedral angles are within 30° of a specified rotarner. HDX, as monitored by MS and NMR spectroscopy, has been used widely to study conformations and conformational dynamics of proteins in various conditions ranging from isolated states to non-covalent interactions [7, 8]. The cru'rent understanding of 127 protein folding suggests three connected hypotheses that can be tested by HDX on pairs of covalently linked peptides. (H1) The association of any two peptides should lower the rate of HDX (even in the absence of significantly increased secondary structure) because steric hindrance alone will inhibit water from encountering the amide groups. (HZ) The association of any two peptides should increase the secondary structure of both peptides because amide-amide H-bonds will have less competition from amide-water H-bonds. (H3) The association of any two sufficiently nonpolar peptides may provoke stable secondary, side-chain, and/or tertiary structure. In this chapter, the three hypotheses (Hl—H3) are tested by comparing the structure in monomeric and dimeric peptides with HDX-MS and CD. HDX has been rarely applied to study oligopeptide systems [9], nor has HDX-MS been widely applied to the analysis of polypeptides containing disulfide bonds [10]. We chose to measure HDX using MS because it requires significantly less material than NMR and tolerates poorly soluble peptides. The major advantage of HDX-NMR, as compared to HDX-MS, is its resolution at the single amino acid residue level; while the resolution of HDX-MS analysis is conventionally limited by the size of the peptic fragments of deuterated polypeptides (usually 5 — 12 resiudes). However, in recent years it was demonstrated that collision-induced dissociation tandem mass spectrometry (CID-MS/MS) can assess the deuterium incorporation of individual amides from the b-series of CID fragment ions [11- 13]. In 2004, HDX-CID-MS/MS results were reported in a study of two oligopeptides and their homodimers formed by non-covalent interactions using ESI with an ion-trap mass spectrometer [14]. The main purpose of this chapter is: (1) To test the three hypotheses of protein 128 folding; (2) To investigate the feasibility of characterizing protein folding nuclei by the combined approaches of HDX-MS and CD. II. Materials and Methods Materials The model peptides (PO. and PB, see definition in Section HI. A.) were synthesized by the Genomics Technology Support Facility (GTSF) at Michigan State University and purified by rp-HPLC using a Vydac C13 analytical column (catalog #218TP54). All other reagents were of the highest purity commercially available and used without further purification. Dibasic sodium phosphate (NaZI-IP04) and sodium sulfate (Na2804) were purchased fi'om Spectrum Chemical Mfg. Corp. (Gardena, CA) and Columbus Chemical Industries, Inc. (Columbus, WI), respectively. Gdn-HCl was purchased from Invitrogen (Carlsbad, CA). ACN (UV grade) and ACN (HPLC grade) were purchased from Honeywell International, Inc. (Muskegon, MI) and EMD Chemicals (Gibbstown, NJ), respectively. The TCEP solutions were freshly prepared before use. All other chemicals were purchased from Sigma. All pH/pD values were measured using a Beckman (D 40 pH meter or ColorpHast pH indicator strips (pH 0 — 14; EM Science, Gibbstown, NJ). pD values were computed by adding 0.4 to the corresponding pH readings [15]. The standard buffer in this study contains 200 mM Na2804, 10 mM NaHzPOa, adjusted to pH 6.0. For the monomeric Pa and PB peptides, the buffer also contained 1 mM reduced dithiothreitol to prevent dimerization of the monomers. No noticeable oxidation was detected after the experiment, as checked by rp-HPLC. 129 Preparation of the Peptide Dimers The P.m and PB]; homodimers and the Pap heterodimer were prepared as described previously [16]. Briefly, roughly equal amounts of the Pa and PB monomers were mixed in a solution containing 5 M Gdn-HCl and 200 mM Tris-HCl at pH 8.0 and were air- oxidized for 48 hr at 25°C. The dimers were separated and purified on a Vydac C13 analytical column using a water/ACN gradient in the presence of 0.1% TFA. Each dimer was collected manually and dried under reduced pressure for further use. CD CD spectra were obtained at 0 °C on a Jasco J-810 CD spectropolarimeter (Jasco, Inc., Easton, MD) using a therrnostatted l-mm pathlength cell. The samples were dissolved in the standard buffer. The concentrations were determined by UV absorbance at 205 nm [17] and were typically 0.15 mg/ml. HDX The HDX experiments were carried out in a cold room, typically at 4 °C. Each peptide sample was equilibrated with the standard buffer (1 nmol/uL) for about half an hour before each experiment. HDX was initiated by diluting 1 uL of aqueous protein solution with 19 uL of standard buffer made with D20 at pD 6.4. After 15 s, 18 11L of aqueous TCEP solution (500 mM in 2 M GdnHCl, pH 2.5, 4 °C) were added to quench the exchange by decreasing the pH to 2.5 and to reduce the disulfide bond. The pH was maintained at 2.5 in all subsequent steps. Incubation of the sample with D20 for 15 5 should be long enough to exchange exposed amide hydrogens with deuteriums [18], but short enough to avoid significant exchange with H-bonded/protected amide hydrogens. 130 Analysis by MALDI-TOF Mass Spectrometry Mass spectra were acquired on a PerSpective Biosystems Voyager DE STR instrument (Applied Biosystems Inc., Framingham, MA) in the positive ion reflectron mode of operation. The accelerating voltage was 20 kV; the grid voltage and guide wire voltage were 76% and 0.04% of the accelerating voltage, respectively. The mirror voltage ratio was 1.12, whereas the extraction delay time was 150 ns. The matrix was on- CHCA in a 5 mg/mL solution containing ACN, ethanol and H20 (1:1:1, v:v:v); the final pH was adjusted to 2.5 with TFA. The measurement was done using next-spot external calibration with the Pa and Pp monomers as the calibrants, for which the calculated monoisotopic [M + H]+ masses are 1671.7219 Da and 1681.8160 Da, respectively. The average mass of each peptide was measured experimentally by determining the centroid of the distribution of its isotope peak intensities using Data Explorer TM 4.0. Ion current from 256 laser shots was accumulated for each spectrum. Immediately after quenching the HDX process, the sample was desalted with cold water containing 0.1% TFA (pH 2.5) using a C13 ZipTip® (Millipore, Bedford, MA) pipette tip, eluted with matrix solution, and spotted on a chilled MALDI target. The target was immediately placed in a desiccator under moderate vacuum to dry the spot within 1 — 2 min. The target was immediately transferred to the MALDI instrument. A highly deuterated standard was prepared by dissolving the sample in D2O at pH 7.0 with 8 M d4-urea and incubating at room temperature for over 48 hr. Treating the highly deuterated standard by the same protocol indicated that the extent of back-exchange during the experimental procedures was 35%. 131 Analysis by CID Tandem Mass Spectrometry Mass spectra were acquired on a Finnigan LCQDecal ion-trap mass spectrometer (ThermoQuest, San Jose, CA) with an ion-spray voltage of 3 kV, a capillary voltage of 24 V, a tube lens offset of 55 V, and a capillary temperature of 150 °C. Data-dependent MS/MS conditions were set with a default collision energy of 35%, a default charge state of 2, and an isolation width of 5 (m/z). Scans were taken over the range m/z 235 to 1500. Typically, 100 scans were accumulated per spectrum. Immediately after quenching the HDX process, the sample was desalted with cold water containing 0.1% formic acid (FA) (pH 2.5) using a C13 ZipTip® pipette tip and eluted with 100 11L of a water/ACN solution (1:1, v:v) pre-adjusted to pH 2.5 with FA. The processed sample was infused into the LCQDeca immediately with a pre-chilled (ice) lOO-uL syringe at a flow rate of 20 uL/min; the syringe was kept buried in ice throughout the experiment. The extent of back-exchange during the infusion-CID-MS/MS procedures was 25%, as determined from analysis of the highly deuterated standard prepared as described above. Mass spectra were obtained with the operating software Xcalibur, and the centroid values were calculated with the Magtran software [19]. III. Results and Discussion A. Criteria for Selection of Peptide Models To validate our approach for studying protein folding nuclei, the selected model peptides should adopt distinct conformations in different states (e.g., significantly folded in one state and essentially unstructured in another state), and these conformations should have been well characterized using conventional techniques (i.e., NMR); to test the three hypotheses of protein folding, the peptide models should be from a protein that is 132 important and interesting to scientists in the area of protein structure and folding. In addition, to enforce the association of designated peptides, we restrict our study here to ( 1) individual monocysteinyl peptides that are linked by a disulfide bond in the native structure of a folded protein and (2) all possible covalent dimers of these peptides. BPTI is a small single-domain protein that consists of 58 residues including 6 cysteines in the form of three intra-molecular disulfide bonds (Figure 4.1). The native structure of BPTI consists of an amino-terminal 310-helix, a highly twisted anti-parallel B-sheet, and an a-helix near the C-terminus (Figure 4.2). The native structure is stabilized by the three disulfide bonds. This protein is one of the most extensively studied systems in protein sciences. It has served as a model system for the examination of almost all aspects of protein structure including X-ray crystallography and neutron diffraction [20], the development of two-dimensional solution nuclear magnetic resonance (NMR) techniques for protein structure determination [21], measurement of protein dynamics [22, 23], studies of the position and dynamics of water molecules [24]. BPTI is also a popular model in protein folding. It has been used for development of computational approaches to protein folding [25-27], and its oxidative folding pathway has been thoroughly studied in terms of the disulfide-bonded intermediates that accumulate during folding [28-34]. 133 Ham mo oo=o=u8 E8 oEE< m .v 0.5»:— w m m m _ m mm Om E m L ? omN OVN 0mN ONN 2N CON 0% . p b _ p _ . p p _ . _ . WI ., - a , W . a T t - aces-8...; . a in. a . ._ . x . .6 .II .I. Coos a -e- 0. Co . so). 9 rNtaMr . w m. ........ to 31 . m (Z 1N 139 Ed A85 5w:0_0>0>> OmN 3N omN CNN 2N ddN r AUOOvfimrAUOOvddm.... b hr Co 8 8m 1... So 8d sea. 1 1 Go 8 88%|. o o.— ' I lOfil O (\l (zuroIJourp 83p)c_01x MHW[®] 8.508 m .w 03E 140 g A85 saws—2983 0mm ovm 0mm omm 3 N com 02 h p _ L F F _ r p p — p .VI I NI f .. o . I N AUOOvflmnAUOOvQQm....H 1 ADO Ov flaw I..l.. IV Co 08 am. .. .. . Go 9 Emil I o (zonJowp 39P)€-01x MHWI®1 sacs m e an 141 EV Gov ohBEoQEoH ow 8 ow cm 0 . _ . p . _ . P P _ N.MI .. . m.- -o m- [ . W 13cm -3: H v 0. -3..- )9 . w. INN- 8 -o.~- W -3- o... . . w. :07 Z T Il\ FVA: 3.253 m .v “Ema 142 does not exhibit the double minima at 208 nm and 222 nm expected for a-helical structure (Figure 4.3 (8)). Hence, the enhanced protection of P,m against HDX (which rivals that of the native-like Pap) likely stems from the association of the two PCl monomers connected by the disulfide bond, rather than from enhanced secondary structure within each, which supports hypothesis (H1). By contrast, the CD spectrum of the PM; peptide at O 0C (but not at 60 °C) suggests the presence of side-chain structure; specifically, the positive peak at 230 nm likely corresponds to the CD signal of one or more ordered-tyrosine sidechains [39]. This suggests that Ppp may also have stable tertiary structure, similar to that in Pug; further evidence is cited below. C. Study of the Conformational Features of Selected Peptide Models by HDX- MALDI-MS Results from HDX and matrix-assisted laser-desorption/ionization time-of-flight mass spectrometry (MALDI-TOF-MS) of the five peptides are reported as total deuterium incorporation determined by the observed mass shifi (Figure 4.4 and Table 4.1). The monomeric PCl and PB peptides showed average mass shifts of 10.2 Da and 8.6 Da, respectively, from their 15 and 13 exchangeable hydrogens. Hence, Pa and PB show some protection against HDX under our experimental conditions. This protection may reflect “flickering” local structure, as has been observed in other experiments [9, 40] and in our CD and CID-MS/MS experiments reported below. Much smaller mass shifts of 6.8 Da and 3.5 Da were measured for the subunits of the P,m and PM; dimers, respectively; the native-like Pap heterodimer showed slightly smaller mass shifts of 5.5 Da (on-subunit) and 2.9 Da (B-subunit). Hence, all three dimers showed significantly more protection than their component monomers, consistent with hypothesis (H1). 143 Table 4. 1 HDX results for a- and B-subunits as assessed by MALDI-TOF-MS. Peptide model Pa PB P0,, P00. Pm: Mass shift' 10.2 8.6 5.5 (or-chain) 2.9 (IS-chain) 6.8 (each oc-chain) 3.5 (each B-chain) 'Mass shift (adjusted for back-exchange) is the average from triplicate runs. The uncertainty of the reported values is $0.1Da expressed as standard deviation. 144 Figure 4. 4 Mass spectra of (A) PM before HDX, (B) PM afier HDX, (C) Pa after HDX, (D) PB after HDX, (E) PM after HDX, (F) Pm, after HDX. The disulfide bond in each dimer was present during HDX, but was reduced before the mass measurement. The cluster of peaks corresponding to the Pa-chain is on the left, while that for the Pp-chain is on the right in (A) and (B). Likewise, PM and P59 were exposed to HDX as dimers, but reduced to the monomer prior to analysis by MALDI-MS. 145 100- O ' l ' l ' r ' l ' I v F 1665 1670 1675 1680 1685 1690 1695 1700 m/: (A) 100« 80~ l @360- l g 1 8 540.. C,\" 20- O ' l ' 1 f l ' l ' l j I l 1665 1670 1675 1680 1685 1690 1695 1700 m/z (B) Figure4. 4 146 Figure 4. 4 (cont’d) l 100~ 80- .‘E‘6Od m c: 8 .540- o\° . 20. N O'i T I ' I ' I r I 1665 1670 1675 1680 1685 1690 1695 1700 m/: (C) 100~ .4 80- m c: 3 540- ,,\° 20- 0 r I ' I 'r'fi— ' I ' I r u ' I 1665 1670 1675 1680 1685 1690 1695 1700 m/: (D) 147 Figure 4. 4 (cont’d) 100- 80- 360- '53 G 8 .540- ..\° , 0 20- 0 fl I ' I ' I ' I ' I ' I ' I 1665 1670 1675 1680,1/1685 1690 1695 1700 (E) 100~ 80- 3‘60" ’5‘) c: 8 .540- o\° . 204 O ' I ' I"_" I ' I I I ' I 1665 1670 1675 1680 1685 1690 1695 1700 m/z (F) 148 In this study, CD spectroscopy was used for a preliminary investigation of the conformation of each peptide model. Although CD results have been qualitatively related to HDX results in several studies [41, 42] (the more structure detected by CD, the more protection against HDX observed), more research will be necessary to determine whether estimation of secondary structure content based on CD can be quantitatively correlated to protection against HDX. For example, amide hydrogens along a polypeptide chain that adopts 100% specific secondary structure (i.e., oc-helix) may be replaced by deuteriums during HDX [10] depending on various factors and conditions including breathing of the polypeptide chain, the HDX conditions (e.g., pD and labeling time), and the degree of involvement of the amide hydrogens in intra-molecular H-bonds. Protection against HDX, on the other hand, may result from factors other than those associated with well-defined secondary structures that can be detected by CD spectroscopy (e.g., lack of exposure to the deuterated solvent due to tertiary structure). Therefore, it seems that one should hesitate before quantitatively correlating HDX results with secondary structural content of a polypeptide. In addition, a quantitative study of the CD results does not help significantly to validate our strategy of studying small peptides using HDX-MS. The resolution of CD analysis is low, and the accuracy of quantitating secondary structure by CD has been disappointing [40, 43, 44]. To further characterize the 3-D structure of each peptide model (including determining the secondary structural content), NMR or X-ray crystallography should be used. On the other hand, as demonstrated in this study, when the deuterium content of each amide linkage can be deduced from HDX-CID-MS/MS, a qualitative correlation of the extent of deuterium incorporation and intra-molecular H- 149 bonding (or other structural features of the polypeptide chain) is reasonable to a first approximation and can provide a means to monitor changes in local structure of peptides as a fimction of cysteinyl/cystinyl status, for example. D. Assess Deuterium Content at the Single Amino Acid Residue Level by HDX- CID-MS/MS The anti—parallel B-sheet segment is the core element of BPTI [9]. To identify the position(s) of the enhanced HDX protection induced by the dimerization in the B-unit, we used CID-MS/MS to probe the HDX of individual amide groups in P13: P33, and PM. Deca These measurements were done using direct infusion ESI on an LCQ ion trap mass spectrometer. HDX-CID of b-ions has been reported to be reliable for assessing deuterium content at the single amino-acid level, whereas that of y-ions is unreliable [11- 13]. A series of b-ions (b 3 - b1o) was formed during the CID process for the PB subunit (Figure 4.5); however, only y-ions were produced for the Pa subunit (Figure 4.6). The masses of the b 3 — b ,0 ions were used to assess deuterium incorporation in residues Tyr23 — LCUzg of the PB subunit in PB, PBS, and PM (Table 4.2). The results indicate that the amide hydrogens of Ty1'23, Asn24, Ly326, and Alan exchange readily in PB (0.6 - 1.1 Da), but are protected in Pap (0.1 — 0.3 Da) and even more so in PM (0.0 Da for Lys26, 01 Da for Alan). By contrast, the Ala25 amide hydrogen exchanges moderately and equally well (0.4 Da) in the monomer (Pp) and both dimers (PW, and PM). Finally, the GIY28 and LCUzg amide hydrogens exchange weakly in PB (0.2 Da) and less so in PM; and Pap (0.1 and 0.0 Da, respectively). Due to the missing fragment in the series of b1, b2, bu-bn ions, deuterium incorporations for Tyr21, Phezz, Cys3o — Phe33 are not available. 150 .8538 am 05 mo 838on $634.90 m .v 9.33 N\S com? com? com? 003. 002 com? co: 003 com com cow com com oov oom _FFFWLFF;befkulrkwu.rlw*2rwas.Irwlluk rpmwu _m1_._.w._ru .___.a r_-._.a_ h_~ O . . . :3 .1... . t . 3 22.1.... 2.54 .12 m m n ma mwdmo .1. v mom hvéw: mmévn Q n QN m. U Q Q I 9‘ W I m n M. mom w m m flow W 3.32 H m m ME Q 0.35 m “5 m8 mom moo.- 3.3 I 151 .8533 am 2: mo 83.58% mngAHHU w .v «Sum..— N\E com? 003. ooww 000? com com oov .ILI.LZCTFLit_..£L.:..L..ILlr LIZr I..T......T..rl...:_L_.:L. L: CL.ZL.:;.._E_.lr.rTL.;£..._ZC.. a . .1 .J.. .1 . .5... . d 0‘...¢. . u nAJi..o.1.flu‘~J. fi‘4l~¢}4.~fll1.d A—IJqflig .. afiquj‘1—11444 41.7.1111: .‘I‘_flflldfiJ.dq.Iul ovafi 3n 88$ 5% 2 .m : Exam 3.22 8. so . 8 8.... 2 an a.» man 3. 53 2 an 8.9.: I—I—rIIIIIIIIIIIIrITIIIIIIITIIIIIIIIIIIIIITIIIIIIIWI hmdm: N; O O V ('0 0 l0 eouepunqv aAIielaa o o r» (O 152 O (D .aouagou 3353. 3 33895 an: .oH mm 82.3 v3.83» 05 Mo 35850:: 2:. .88 88:95 me owfigm 05 fl Aowuanoxoéoan Sm “628.83 «Em $34. Assamese 303-238 envisiov 3 3 mo 3 2 3w .5 Asa-kc od 3 33$ 8 3 Nag-é No 3 .5 E39 3 $2 939 3 $2 $39 No $2 3 E2$ 3- 3 32$ 3 E 32$ 9o 3 3 33$ 3 ed 33 no 3 3%.: mo 2 s 32$ to 3 m§$ to 3 «22$ .2 3. s 35$ .2 S 35$ no 3 35$ .3 S. 3 A5»: No 3 3233 3 ES: 2 _.m s waitfib finite»: moire»: 3 3 3 3 3 2 3 053mg mo dim mme udEm mmmE 053mg mo fiEm mmme .flEm mama gamma .«0 @Em mmma LflEm mmaa mum an am now $352.96 3 Baas .3 E533 .5 2: 5 33.2 xom N .4 2...:- 153 However, the 2.0-Da mass shift of the b3 ion of PB indicates that Tyrzi and Phezz amide hydrogens exchange readily in the monomer. In addition, the l.O—Da decrease in the mass shift of the b3 ion in P53 and P043 suggests that one of these two amide hydrogens is protected in the dimers. Similarly, segment Cys3o — Phe33 of the [3 sequence shows less protection in PB (2.3 Da) than in Pm; and Pap (0.9 Da and 1.0 Da, respectively). These site-specific HDX results for the PB subunit are generally consistent with the solvent exposure and H-bonding of the native structure (See Figure 4.7). If the [3- chain adopts its native structure, the amide hydrogens of Phezz, Asn24, Glyzg, Gln31, and Phe33 should be well protected because they are all involved in strong backbone (amide- amide) H-bonds; however, the Phe33-Arg20 H—bond is unlikely to be maintained in the peptide model, owing to the well-known flexibility of chain termini. Consistent with these predictions, significant protection for Asn24 and Glyzg is observed in PM; and Pap (Table 4.2). Similarly, the mass shifts of the b3 and b” ions of PBB and P04; are consistent with significant protection of Phezz and Gln31 against HDX, which are mutually H- bonded in the native structure. The amide hydrogen of Leuzg is well protected in PM; and Pap (Table 4.2); in the native structure, the amide hydrogen of Leuzg is protected from exchange by the surrounding residues and is weakly H-bonded to the backbone oxygen of Asn24 (although it competes with the amide hydrogen of Glyzg). Interestingly, the amide hydrogens of Gl)’23 and LCUzg exchange relatively slowly in monomeric PB, suggesting that Pp adopts local structure in its B-hairpin (residues 5 —10). Finally, the Lyszo and Alan amide hydrogens form a bifurcated H-bond with the side-chain oxygen of Asn24 in the native structure of BPTI. These hydrogens are well protected in PM, and even more so in Pap, consistent with a structured Asn24 side-chain. These two hydrogens are chiefly 154 responsible for the differences in the protection of the Pp subunit against HDX in Pap and PM Combined with the CD results, these data suggest significant native secondary and side-chain structure in the P53 homodimer, as well as the Pug heterodimer [16]. Figure 4. 7 Ribbon diagram of the natively structured PB peptide; Asnz4 is indicated in light gray and relevant hydrogen bonds are shown in dark gray. The mainchain- mainchain H—bond donor/acceptor pairs from bottom: Phe33/Arg20, Phe22/Gln31, Gln31/Phe22, Asn24/Leu29, Leu29/Asn24, and Glyzg/Asn24. The H-bonds between the amide hydrogens of Lys7 and Alas to the sidechain atom 051 of Asns are also shown. The figure was drawn using the MOLMOL software [36]. 155 The absence of protection against HDX also can be a cross-check of native structure; the amide hydrogens of Alazs, Cysm, and My should exchange readily because they are solvent-exposed and not H-bonded in the native structure. The increased structure in Pm; and Pug does not provide enhanced protection to Alazs (Table 4.2). Notably, Alazs does not fully exchange even in PB; this effect remains to be explained, although slow exchange at this position has been noted in BPTI variants [45], and may reflect a stable B-turn at residues Asn24-Leu29. The mass shifts of b], and blo from PM; and Pug suggest that at least one residue in the segment Cys3o-Phe33 is unprotected. Finally, the amide hydrogens of Tyr21 and Tyr23 should be protected in the Pap dimer owing to intermolecular H-bonds with Phezz and Asnzo from the Pa subunit. Consistent with this prediction, no protection is observed for Tyrz: and Wm in P5, but modest and roughly equal protection of Tyr23 is observed in both P33 and Pap (Table 4.2). This protection of Tyr23 in the P39 dimer suggests stable intermolecular H-bonding (and, thus, tertiary structure) in Ppp, in addition to the secondary and side-chain structure cited above. 156 IV. Conclusions We have demonstrated the usefulness of HDX-MS, when combined with CD, for studying conformational features of model peptide folding systems. Our protocol uses MALDI-MS to assess overall deuterium incorporation and CID-MS/MS to obtain residue-level HDX data. These data coupled with CD spectra (with thermal transitions) allow us to assay secondary and other structural changes that occur upon dimer formation in model peptides. Our method is general and should be useful in studying the physics of protein folding and disorder. Applied to peptide models from BPTI, our results provide evidence for the three hypotheses outlined in the introduction. First, Pa shows some protection against HDX, but the P,m dimer has almost as much protection as the native-like Pug heterodimer, despite having significantly less secondary structure than Pug (as assessed by CD). This supports hypothesis (H1) that the association of two peptides can inhibit HDX by lowering the local activity of water independently of secondary structure. Second, the P,m dimer exhibits more secondary structure than the Pa monomer, even at 60 °C; analogous results pertain to P3 and Pm}. This supports hypothesis (H2) that the decreased local activity of water stabilizes local secondary structure. Finally, the Pm dimer does not appear to be fully structured, judging from its non-helical CD spectrum; by contrast, the data suggest that the Pap and P33 dimers have stable secondary, side-chain, and even tertiary structure. Thus, our results are consistent with hypothesis (H3) that stable secondary, side-chain, and even nonnative tertiary structure may develop when two sufficiently nonpolar peptides are linked covalently. The difference between the PM and Pap dimers may result from the hydrophobicity of the two peptides; PC, has only two large 157 hydrophobic residues (Phe45 and Metsz) of 16 residues, whereas PB has four (Tyr21, Phezz, Tyrzg, and LCU29) of 14 residues. Therefore, the PM; dimer is better able to form a hydrophobic core around its disulfide bond, further shielding its amide-amide H-bonds from attack by water molecules. This is consistent with the global HDX results; PM; exchanges 27% (3.5/ 13) of its amide hydrogens, whereas P,m exchanges 45% (6.8/ 15). 158 V. References [1] P. N. Lewis, N. Go, M. Go, D. Kotelchuck, H. A. Scheraga, Helix probability profiles of denatured proteins and their correlation with native structures, Proc Natl Acad Sci USA 65 (1970) 810-5. [2] H. A. Scheraga, Theoretical and experimental studies of conformations of polypeptides, Chem Rev 71 (1971) 195-217. [3] H. Taniuchi, C. B. Anfinsen, Simultaneous formation of two alternative enzyrnology active structures by complementation of two overlapping fragments of staphylococcal nuclease, J Biol Chem 246 (1971) 2291-301. [4] G. T. Montelione, H. A. Scheraga, Formation of local structures in protein folding, Accounts of Chemical Research 22 (1989) 70-6. [5] W. F. Harrington, J. A. Schellrnan, Evidence for the instability of hydrogen-bonded peptide structures in water, based on studies of ribonuclease and oxidized ribonuclease, C R T rav Lab Carlsberg [Chim] 30 (1956) 21-43. [6] I. M. Klotz, J. S. Franzen, Hydrogen bonds between model peptide groups in solution, Journal of the American Chemical Society 84 (1962) 3461-6. [7] I. A. Kaltashov, S. J. Eyles, Studies of biomolecular conformations and conformational dynamics by mass spectrometry, Mass Spectrom Rev 21 (2002) 37-71. [8] S. W. Englander, Protein folding intermediates and pathways studied by hydrogen exchange, Annu Rev Biophys Biomol Struct 29 (2000) 213-3 8. [9] N. Carulla, C. Woodward, G. Barany, Synthesis and characterization of a beta-hairpin peptide that represents a 'core module' of bovine pancreatic trypsin inhibitor (BPTI), Biochemistry 39 (2000) 7927-37. [10] X. Li, Y. T. Chou, R. Husain, J. T. Watson, Integration of hydrogen/deuterium exchange and cyanylation-based methodology for conformational studies of cystinyl proteins, Anal Biochem 331 (2004) 130-7. [11] M. Y. Kim, C. S. Maier, D. J. Reed, M. L. Deinzer, Site-specific amide hydrogen/deuterium exchange in E. coli thioredoxins measured by electrospray ionization mass spectrometry, J Am Chem Soc 123 (2001) 9860-6. [12] Y. Deng, H. Pan, D. L. Smith, Selective Isotope Labeling Demonstrates That Hydrogen Exchange at Individual Peptide Amide Linkages Can Be Determined by Collision-Induced Dissociation Mass Spectrometry, Journal of the American Chemical Society 121 (1999) 1966-1967. 159 [13] J . K. Hoemer, H. Xiao, A. Dobo, I. A. Kaltashov, Is there hydrogen scrambling in the gas phase? Energetic and structural determinants of proton mobility within protein ions, J Am Chem Soc 126 (2004) 7709-17. [14] X. Cai, C. Dass, Structural characterization of methionine and leucine enkephalins by hydrogen/deuterium exchange and electrospray ionization tandem mass spectrometry, Rapid Commun Mass Spectrom 19 (2005) 1-8. [15] P. K. Glasoe, F. Long, Use of glass electrodes to measure acidities in deuterium oxide, Notes 64 (1960) 188-90. [16] T. G. Oas, P. S. Kim, A peptide model of a protein folding intermediate, Nature 336 ( 1988) 42-8. [1 7] R. K. Scopes, Measurement of protein by spectrophotometry at 205 nm, Anal Biochem 59 (1974) 277-82. [1 8] Y. Bai, J. S. Milne, L. Mayne, S. W. Englander, Primary structure effects on peptide group hydrogen exchange, Proteins 17 (1993) 75-86. [1 9] Z. Zhang, A. G. Marshall, A universal algorithm for fast and automated charge state deconvolution of electrospray mass-to-charge ratio spectra, J Am Soc Mass Spectrom 9 (1998) 225-33. [20] A. Wlodawer, J. Nachman, G. L. Gilliland, W. Gallagher, C. Woodward, Structure of form III crystals of bovine pancreatic trypsin inhibitor, J Mol Biol 198 (1987) 469-80. [21] K. Wuthrich, G. Wider, G. Wagner, W. Braun, Sequential resonance assignments as a basis for determination of spatial protein structures by high resolution proton nuclear magnetic resonance, J Mol Biol 155 (1982) 311-9. [22] G. Wagner, K. Wuthrich, Amide protein exchange and surface conformation of the basic pancreatic trypsin inhibitor in solution. Studies with two-dimensional nuclear magnetic resonance, J Mol Biol 160 (1982) 343-61. [23] K. S. Kim, C. Woodward, Protein internal flexibility and global stability: effect of urea on hydrogen exchange rates of bovine pancreatic trypsin inhibitor, Biochemistry 32 (1993) 9609-13. [24] G. Otting, E. Liepinsh, K. Wuthrich, Protein hydration in aqueous solution, Science 254 (1991) 974-80. [25] M. Levitt, Effects of proline residues on protein folding, J. Mol. Biol. 145 (1981) 251-63. 160 [26] S. Cusack, J. Smith, J. Finney, B. Tidor, M. Karplus, Inelastic neutron scattering analysis of picosecond internal protein dynamics. Comparison of harmonic theory with experiment, J Mol Biol 202 (1988) 903-8. [27] V. Daggett, M. Levitt, A model of the molten globule state from molecular dynamics simulations, Proc Natl Acad Sci U S A 89 (1992) 5142-6. [28] T. E. Creighton, Experimental studies of protein folding and unfolding, Prog Biophys Mol Biol 33 (1978) 231-97. [29] T. E. Creighton, D. P. Goldenberg, Kinetic role of a meta-stable native-like two- disulphide species in the folding transition of bovine pancreatic trypsin inhibitor, J Mol Biol 179 (1984) 497-526. [30] J. S. Weissman, P. S. Kim, Kinetic role of normative species in the folding of bovine pancreatic trypsin inhibitor, Proc Natl Acad Sci U S A 89 (1992) 9900-4. [31] J. S. Weissman, P. S. Kim, The pro region of BPTI facilitates folding, Cell 71 (1992) 841 -5 1 . [32] J. S. Weissman, P. S. Kim, Efficient catalysis of disulphide bond rearrangements by protein disulphide isomerase, Nature 365 (1993) 185-8. [33] J. S. Weissman, P. S. Kim, Reexamination of the folding of BPTI: predominance of native intermediates, Science 253 (1991) 1386-93. [34] J. S. Weissman, P. S. Kim, A kinetic explanation for the rearrangement pathway of BPTI folding, Nat Struct Biol 2 (1995) 1123-30. [35] S. Parkin, B. Rupp, H. Hope, Structure of bovine pancreatic trypsin inhibitor at 125 K definition of carboxyl-terminal residues Gly57 and Ala58, Acta Crystallogr D Biol Crystallogr 52 (1996) 18-29. [36] R. Koradi, M. Billeter, K. Wuthrich, MOLMOL: a program for display and analysis of macromolecular structures, J Mol Graph 14 (1996) 51-5, 29-32. [37] J. P. Staley, P. S. Kim, Formation of a native-like subdomain in a partially folded intermediate of bovine pancreatic trypsin inhibitor, Protein Sci 3 (1994) 1822-32. [38] R. W. Woody, H. Sugeta, T. S. Kodama, [Circular dichroism of proteins: recent developments in analysis and prediction], T anpakushitsu Kakusan Koso 41 (1996) 56-69. [39] N. Sreerama, M. C. Manning, M. E. Powers, J. X. Zhang, D. P. Goldenberg, R. W. Woody, Tyrosine, phenylalanine, and disulfide contributions to the circular dichroism of proteins: circular dichroism spectra of wild-type and mutant bovine 161 pancreatic trypsin inhibitor, Biochemistry 38 (1999) 10814-22. [40] E. M. Goodman, P. S. Kim, Folding of a peptide corresponding to the alpha-helix in bovine pancreatic trypsin inhibitor, Biochemistry 28 (1989) 4343-7. [41] C. S. Maier, M. I. Schimerlik, M. L. Deinzer, Thermal denaturation of Escherichia coli thioredoxin studied by hydrogen/deuterium exchange and electrospray ionization mass spectrometry: monitoring a two-state protein unfolding transition, Biochemistry 38 (1999) 1 136-43. [42] M. Y. Kim, C. S. Maier, D. J. Reed, M. L. Deinzer, Conformational changes in chemically modified Escherichia coli thioredoxin monitored by H/D exchange and electrospray ionization mass spectrometry, Protein Sci 11 (2002) 1320-9. [43] S. W. Provencher, J. Glockner, Estimation of globular protein secondary structure from circular dichroism, Biochemistry 20 (1981) 33-7. [44] S. Y. Tetin, F. G. Prendergast, S. Y. Venyaminov, Accuracy of protein secondary structure determination from circular dichroism spectra based on immunoglobulin examples, Anal Biochem 321 (2003) 183-7. [45] B. A. Schulman, P. S. Kim, Hydrogen exchange in BPTI variants that do not share a common disulfide bond, Protein Sci 3 (1994) 2226-32. 162 CHAPTER 5 IDENTIFICATION AND CHARACTERIZATION OF AN AUTONOMOUS FOLDING UNIT FOR LRJIGF-I I. Introduction To clarify the mechanism of protein folding, great effort has been made to characterize the 3-D structure of intermediates at various stages along the folding pathway [1-4]. However, the cooperativity of protein folding and the limited solubility of many folding intermediates often hinder researchers’ efforts [5, 6]. One way to circumvent these problems is to design and characterize peptide models that can fold into a native-like conformation. The protein subdomains that correspond to these peptide models are called AFU [7]. AFUs from various proteins have been described and characterized in recent years [8-10]. In this chapter, a new AFU for recombinant human LR3IGF-I is described and characterized by HDX-MS combined with CD spectroscopy. Recombinant human long R3 insulin-like grth factor-I (LR3IGF-I) is a highly potent growth factor [11, 12]. Its amino acid sequence is shown in Figure 3.1. Cysteinyl/cystinyl intermediates involved in progressive disulfide-bond formation during the refolding of LR3IGF-I have been trapped and characterized by our cyanylation/cleavage methodology [13]. During this study, it was interesting to observe that only one of the fifteen theoretically possible l-S intermediates, (namely, that containing the disulfide bond Cy531—Cy874) accumulates to a significant degree during the oxidative refolding process (Figure 5.1). One explanation for this phenomenon involves 163 1 min .‘ [#Nfia‘vmv' "V . .u“.A.' .~ ‘ (“x -A-.._-O-'"" ' "’ 2min . “Jr Fm“, ”m,__. a.-. *- -.\-._,..FIJ' Minutes Figure 5. 1 HPLC chromatogram showing the time-dependent distribution of CDAP- trapped intermediates during the refolding of LR3IGF-I [13]. The disulfide structure of each species is: I’ (Cys31 — Cys74), II’ (Cys31 — Cy574, CYS19 - Cysm), III’ (Cys31 — Cy819 - Cysoo, Cysor -' Cysos), N (CYS31 - CyS74, Cy819 - CySm, Cysoo - CySos), R (completely reduced). (experimental details on page 99 in Chapter 3) 164 protective tertiary structure formed via long-range interactions between the residues flanking this particular disulfide bond. According to this hypothesis, the disulfide bond stabilizes the local structure, while the structure protects the disulfide bond, in turn, from attack by reducing reagents in the system (such as the other free sulfhydryl groups in the protein). To test this hypothesis and investigate the relationship between disulfide bond formation and the stabilization of local folded structure, peptide models were designed and characterized that mimic the sequence of LR3IGF-I in the vicinity of Cys31 and Cysu. Specifically, synthetic monomer P1 corresponds to residues 17 — 39 (RLSGAELVDALQFVCGDRGFYFS) in LR3IGF-I, and synthetic monomer Pn corresponds to residues 54 — 75 (RGIVDEASFRSVDLRRLEMYCQ). Segment 17 — 39 corresponds to helix 1 in LR3IGF-I, and segment 54 — 75 corresponds to helix 2 and helix 3. As shown in Figure 3.3, helix 1 and helix 3 are connected by the disulfide bond Cys31—Cys74. Because P1 and Pu each contain one cysteine residue, they can be disulfide— bonded to form the PIP! or the PuPu homodimers or the PIP” heterodimer. The P1P" heterodimer corresponds to the Cys31—Cys74 l-S intermediate that accumulates early during the refolding of LR3IGF-I [13]. The two peptides, P1 and P", were synthesized and purified, and then used as reagents to form the three dimers (PIPI, PnPn, and P1P") by air oxidation. The structure of each of the five peptide models (PI, Pu, P1P], PnPn, and P1P") was probed using CD spectroscopy and HDX-MS. CD reports on the secondary structure (i.e., or-helix and [3- sheet structure) of each peptide model [14], whereas HDX-CID-MS monitors the degree of protection of amide hydrogens against exposure to solvent [15-17]. The experimental 165 results demonstrate that significant 3-D structure is associated with the formation of the disulfide bond in each dimer. This observation supports our hypothesis that P; and P" correspond to an AFU subdomain. 11. Materials and Methods Materials The P. and P" peptides were synthesized by the Genomics Technology Support Facility (GTSF) at Michigan State University. The target peptides (approximately 10% yield in the synthetic mixture) were then purified by rp-HPLC using a Vydac C13 analytical column (catalog #218TP54) in our laboratory. All other reagents were of the highest purity commercially available and used without further purification. Dibasic sodium phosphate (NazHPO4) was purchased from Spectrum Chemical Mfg. Corp. (Gardena, CA). ACN (UV grade) and ACN (HPLC grade) were purchased from Honeywell International, Inc. (Muskegon, MI) and EMD Chemicals (Gibbstown, NJ), respectively. Gdn-HCl was purchased from Invitrogen (Carlsbad, CA). All other chemicals were purchased from Sigma. All pH/pD values were measured using a Beckman (D 40 pH meter or ColorpHast pH indicator strips (pH 0-14; EM Science, Gibbstown, NJ). pD values were computed by adding 0.4 to the corresponding pH readings [18]. The standard buffer in this study contains 0.1 M NaCl and 10 mM NaHzPO4, adjusted to pH 7.0. For the monomeric PI and P" peptides, the buffer also contained 1 mM reduced dithiothreitol to prevent dimerization of the monomers. No noticeable oxidation during the experiment was detected as checked by rp-HPLC. 166 Preparation of the Peptide Dimers The P1P; and PnPu homodimers and the P1P" heterodimer were prepared as described previously [6]. Briefly, roughly equal amounts of the P1 and P" monomers were mixed in a solution containing 5 M Gdn-HCl and 200 mM Tris-HCl at pH 8.0 and were air-oxidized for 48 hr at 25 °C. The dimers were separated and purified on a Vydac C13 analytical column using a water/ACN gradient in the presence of 0.1% TFA. Each dimer was collected manually and dried under reduced pressure for further use. CD Spectroscopy CD spectra were obtained at 0 °C on a J asco J-810 CD spectropolarimeter (J asco, Inc., Easton, MD) using a therrnostatted 1-mm pathlength cell. The samples were dissolved in the standard buffer. The concentrations were determined by UV absorbance at 205 nm [19] and were typically 0.11 mg/ml. HDX The HDX experiments were carried out on ice. Each peptide sample was equilibrated with the standard buffer (1 nmol/uL) for about half an hour before each experiment. HDX was initiated by diluting 1 uL of aqueous protein solution with 19 uL of standard buffer made with D20 at pD 7.4. After 60 s, 80 uL of buffer A solution (containing 0.1% FA and 2.5% ACN) was added to quench the exchange by decreasing the pH to 2.5 and to reduce the disulfide bond. The pH was maintained at 2.5 in all subsequent steps. Incubation of the sample with D20 for 60 3 should be long enough to exchange exposed amide hydrogens with deuteriums [20], but short enough to avoid significant exchange with H-bonded/protected amide hydrogens. 167 Analysis by CID Tandem Mass Spectrometry De“ XP ion—trap mass Mass spectra were acquired on a Finnigan LCQ spectrometer (ThermoQuest, San Jose, CA) with an ion-spray voltage of 2.2 kV, a capillary voltage of 43 V, a tube lens offset of 0 V, and a capillary temperature of 225 0C. Data-dependent MS/MS conditions were set with a default collision energy of 40%, a default charge state of 1, and an isolation width of 10 (m/z units). Scans were taken over the range m/z 235 to 2000. Typically, 10 scans were accumulated per spectrum. Immediately after quenching the HDX process, the sample was loaded onto a C13 trap (Upchurch, C281) using a pre—chilled syringe and flushed with 100 [4L of buffer A (pH 2.5) for 60 °C in for desalting. The processed sample was eluted into the LCQDeca XP by adjusting the gradient to 50% buffer B (97.5% ACN, 0.1% FA) at a flow rate of 100 uL/min; the injector and all the loops were buried in ice. The extent of back- exchange during the procedures was 20%, as determined from analysis of the highly deuterated standard prepared as described above. Mass spectra were obtained with the operating software Xcalibur, and the centroid values were calculated with the Magtran sofiware[21]. 168 III. Results and Discussion A. Design of the peptide models In native LR3IGF-I, the disulfide bond between Cys31 and Cysu connects two of the three (it-helices (helix 1 and helix 3 in Figure 3.3). The core region of the designed peptide model therefore includes residues that correspond to the three helices described earlier. Residues outside this core region also were included if they satisfied either of the following criteria: (1) A residue that is conserved in the IGF family; (2) A residue in one peptide interacts with residues in the second peptide according to the NMR structure of LR3IGF-I [22, 23]. Our goal was to include all residues that satisfy these criteria, while keeping the peptides as short and soluble as possible. P1P" represents approximately half of LR3IGF-I (45 out of 83 residues) and includes most of its hydrophobic core. The amino-acid sequences of P1 and P" are: P; (17-39): RLSGAELVDALQFVCGDRGFYFS Pu (54-75): RGIVDEASFRSVDLRRLEMYCQ Pr contains 23 residues and corresponds to residues 17 — 39; this includes helix 1 (21 — 31) of LR3IGF-I. Pu contains 22 residues corresponding to residues 54 — 75; this includes both helix 2 (56 - 63) and helix 3 (67 — 73). Eight residues of LR3IGF-I were replaced to Optimize the solubility and stability of our peptides, and to encourage the formation of b-ions during CID-fragmentation: in P1, Thrn was replaced by Arg, CYS19 by Ser, and Asn39 by Ser; in P", Thr54 was substituted by Arg, CYSéO by Ala, Cys61 by Ser, CYSfiS by Val, and Ala75 by Gln. To improve solubility, we replaced the terminal residues 169 of our peptides with charged or more polar/flexible residues: Thrn by Arg, Asn39 by Ser, Thr54 by Arg, and Alan by Gln. We also eliminated four of the six cysteines, generally by substituting them with Ala (when the cysteine was buried in the hydrophobic core) or the isosteric residue serine (when the cysteine was exposed). Cys(,5 was mutated to Val (instead of Ala) for two reasons: (1) to help fill the void in the hydrophobic core left by the elimination of the cyS60-‘Cy865 disulfide bond; and (2) Val forces the extended conformation (ct = —80°, w = +139°) of the backbone at position 65. By contrast, the Cysés to Ala substitution might have encouraged the formation of one long a-helix (because Ala is helix-favoring), instead of the two broken helices observed in the native structure (Figure 5.2). Throughout this chapter, the residues are numbered according to their position in native LR3IGF-I (e.g., Cys3r, Va165, etc). 170 Figure 5. 2 The natively structured Pr and P11 [24]. P1 is depicted in dark gray, with its initial residue, Argn, shown at the far left and its final residue, Ser39, shown in the back on the right. Similarly, P" is depicted in light gray, with its initial residue, Arg54, shown at the bottom and its final residue, Gln75, shown at the far right. The crossing-linking disulfide bond, Cys3r—Cys74, is shown in the upper right. The side chains of the substituted residues are also shown in dark gray (in P1) and light gray (in P“). The figure was drawn using the MOLMOL software [25]. The side-chains were drawn using SCWRL software [26]. 171 B. CD results To investigate the structural changes occurring upon disulfide bond formation, the five peptides (P1, P", P1P", P1P], PUP") were first studied using CD spectroscopy. Although the absolute magnitude of the CD signal shown in Figure 4.3 is smaller than expected ideally, the spectral shape is clearly consistent with a-helical structure [14]. As shown in Figure 5.3 (A), the CD spectrum of P1P" at 0 0C shows minima near 208 nm and 222 nm, which are characteristic of helical proteins [14]; at 60 oC and 80 °C, these troughs are replaced by one trough near 204 nm as expected for random-coil structure [14]. The summed CD spectrum of P1 and P11 at 0°C is dominated by contribution from random structure; the minimum near 222 nm suggests that a small population of the monomers probably takes or-helical structure, although the majority of the population does not. The summed spectrum resembles the spectra of P1P" at 60 °C and at 80 °C. These results suggest that significant oc-helical structure is formed in the PIP" heterodimer, consistent with its native structure. At 0 °C, the PIP; homodimer also gives CD signals characteristic of or-helical structure; the two minima disappear as the temperature increases to 60 °C and 80 0C (Figure 5.3 (B)). The CD spectrum of P; at 0 oC resembles that of PP at 60 °C and at 80 °C. The CD spectrum of PuPu at 0 °C is significantly different from that of Pu, and the minimum near 217 nm disappears gradually as the temperature increases (Figure 3 (C)). This observation indicates that, although the majority of the P" monomers adopt a random coil structure, a novel folded structure is formed in the PUP“ homodimer. The changes in 3-D structure are probably caused by long-range interactions induced by the 172 formation of the disulfide bond. Interestingly, the CD spectrum of PuPn resembles that of a pure B-sheet protein (a negative signal around 217 nm [14]), raising the possibility that 173 o. o 8 f 8 588% ea 28 do 3 .oo 8 .6. o 8 is 8 88% 8 Q .o. o 8 5e 888% e5 28 .0. cm .0. 8 .o. o 8 :E 8 88% no a: .6. o 8 f 28 a 8 88% 888% 8 28 .0. cm .0. 8 .0. o 8 is .6 88% no 3 m .m 8.88 $ 3:: sage—38>? 0mm 03 0mm omm o _ N com 03 p p . _ p F p 7 p _ . _ r o—-. 6085191.: . ] e.. ufio/ I WI m G. 08 P? z: .. ... ., . w ..\\ X. z . I or Va 6.. 08 :85- .. r .8 .\ , . m I VI X 6.. 8 =8: . m. I N: Cc . we r O 00 . w I N ..Ol. 1 3 . a. m (7» r o 174 EV A85 Swans—26>» 03 28 0mm 08 Sm com 02 . _ . _ _ _ _ — . _ OMI . m .8 M r X m -o e . m 9 00 6.81828 3m. 6.. 81.3- r r or H H rim 8.8 .ZII . (z -2 3.88 m .m 88E 175 ADV A85 £353.63 0mm 0.8m 0mm 0mm 2m com 03 8m . -8 ./ . M 4-..... z . 0 a. [NIm 9 . 8 . . ,8. .\ r O Ceow 7. In. IN 3 6.8155--- . mz l.—V( 5855' . 9.88 m .m 883 176 PUP" adopts a different secondary structure than observed in the native structure of LR3IGF-I. In summary, the comparison between the CD spectrum of each dimer and its corresponding monomer(s) suggests that formation of the disulfide bond induces the formation of a more folded structure in all three dimers. C. HDX results During the last decade, HDX-MS has become more and more popular in the study of protein conformation and conformational dynamics [15, 27, 28]. Here, HDX was used to further characterize the structural features of each peptide model. HDX results of the five intact peptides are reported as total deuterium incorporation determined by the observed mass shift (Table 5.1). For monomeric P1 and P", 76% (16.7 out of 22) and 86% (18.1 out of 21), respectively, of the total exchangeable amide hydrogens were replaced by deuteriums from the solvent. Under the experimental conditions used in this study, P1 and P11 show some protection against HDX, which may reflect “flickering” local structure in each monomer as indicated by our CD results and which has been observed in other experiments [29, 30]. The three dimers showed much less deuterium incorporation (9% for PIP", 14% for P1P], and 15% for PuPu). In summary, all the dimers exhibit much more protection than their corresponding free monomers. The HDX results are consistent with the observation in our CD experiments that significantly more folded structure is present in each dimer than in the individual constituent monomers. In a few studies, resolution of the loci of incorporated deuteriums has been improved by analyzing the peptides by HDX-CID-MS [16, 17, 31, 32]. When a series of 177 dew—8:68 23:88 8 38298 .80 me“ 8 83.8 venom“: 05 mo b55665. 2E. .83 38:95 mo owfiog 05 a Aowfifioxorxoan co.“ 833.33 «Em 882 . .mcowohin 0288 03803305 me 538:: \ $5 888 n 8983508 .«o 83x? .82 .8: .8 .88 .88 .8888 8 888m 8 3 a. a 8 888.8 888 288.888 4.8 o8 3 3: $2 .82.. .82 is is =8: 5 a 882 e888 .3on 0232“ 685 :80 .«o 3:52 XQZ — .m 038,—. 178 b-ions is produced during fragmentation by CID, the degree of deuterium incorporation of individual amides can be assessed. Here, we applied CID-MS/MS to the five peptides (Pl, Pu, P1P", P1P], and PUP"); the spectrum of each peptide is shown in Figure 5.4. The fragmentation pattern of each dimer is similar to that of its constituent monomer(s). During CID-MS/MS of the P1 monomer, major fragmentation takes place at the C-tcrminal side of Phe36 and Try37, giving rise to b202+ and b212+ (Figure 5.4 (A)). For the PIP] homodimer, major fragmentation takes place at the same sites, but in only one of the two F; chains during the CID-MS/MS process, giving rise to analogous b-ions that are more massive (shified by the mass of the P1 monomer); these ions are quadruply charged as annotated in Figure 5.4 (B). As shown in Figure 5.4 (B), the most abundant ions produced from PIPI are [(bzo of P1) + P1]4+ and [(1)21 of P1) + Pd“; When the P11 monomer was subjected to CID-MS/MS fragmentation, y92+ and b132+ were the major ions produced as a result of fragmentation at the C-terminal side of Asp66 (Figure 5.4 (C)). For the PUP" homodimer, the same b132+ and bmz+ ions from one of the component P" chains are the major products (but mass-shifted by the mass of P") of the CID-MS/MS process (Figure 5.4 (D)). In the case of the PIP" heterodimer, major CID-MS/MS fragmentation takes place at the C-terminal side of Asp“ of the component Pu chain; b- and y-ions are produced that are analogous to those produced from the Pu monomer, but mass shifted by the mass of P; (Figure 5.4 (E)). In the heterodimer, the P1 chain and the P11 chain are held together covalently by the disulfide bond between Cys3l (in P1) and Cys74 (in Pu). Thus, the P1 chain is connected with the yg fragment of the P11 chain through this disulfide bond. As shown in Figure 5.4 (E), the most abundant ions produced from P1P” are b132+ of P" and [09 of P11) + P11“. 179 686882 $5 on. m: 68652 if 05 A9 .8882: um 05 8V .886080: 55 05 Amv £08258 5 05 Ap—>>p—-pp—->—-- _.-—rp——-pp—ppp—>.p—h rh—p-p—pnv—--—-Pp—.—r—-.~——-__-p——-»r—».n—_p’—pnr—-»*n-.—n»»——>-_upu—-p— _ y 1,. n 1 1“ 1. <1 44 la. . 1| . lid II Ill: dun: I: I: 1. .‘.-.1II. 1. . .74 d J iii—.1 O r. VNNE 520 f .6 +ng O V 8 1111‘Ilr1‘11jTjT11111111111TT I 8 aouepunqv emeIaa 111111111 0 N +a? 3% .525. H 91.3 3:53 in 0.55 184 Because each of the five peptides showed essentially only two or three channels of fragmentation during CID-MS/MS, the ion current fi'om the most populated channels of b-ions was subjected to CID-MS/MS/MS in an effort to promote further fragmentation. Representative spectra are shown in Figure 5.5 and Figure 5.6. Ions with reasonable abundance were identified and their corresponding peaks were labeled in the figures. None of the five peptides produces a series of b-ions or y-ions. Both P1 and P” contain more than one basic residue (e.g., Arg) in each sequence, and, thus, the fragmentation pattern of these peptides is complicated and unpredictable [33]. In this study, only b-ions with reasonable abundance were used to assess the level of deuterium incorporation in different sections of each peptide because b-ions of very low abundance, as well as y-ions, are not reliable for assessment of deuterium incorporation [31, 32]. The HDX-CID-MS/MS and HDX-CID-MS/MS/MS results are summarized in Table 5.2. Although single amino-acid residue resolution was not achieved due to the absence of a series of b-ions for each sequence, the b-ions detected for each peptide enable some insight into the deuterium incorporation of a few sections along each polypeptide chain. All the sections along the polypeptide chain of the monomers are significantly deuterated. However, the deuterium incorporation level of each section along the backbone of the dimers is limited. For instance, Argn-Aspzs of the P1 sequence has 8 eligible amide hydrogens; 7 of these 8 amide hydrogens were replaced by deuterium in the P1 monomer (Table 5.2 (A)); on the other hand, only 1.4 of them were deuterated in the PIP; homodimer (Table 5.2 (B)). This observation is consistent with the fact that the dimers adopt a significantly folded structure, while the monomers do not. 185 .AN+V mm 8505 m_ 22 swat $830 as. omega .5 .Ae g f Ea 8V 9: a soc 82 mzaz .5568 92 25o sea... .3490 m .m 2:»: 73 NE ooom com? com: 83 com? 009 com com oov Pru—ber kph twp—pprr—pbpp—bbrr—p>r>_»Pv— :17. .. _._ 17%.»? . _ .1 __ .5 .. _ .i .. :fi .4... .— _ m. 5325:: .4... —LV —>u>h—4>P .— . 57:..:_ :3: .4 1 O O ‘_ O N O (‘0 O V O In 3‘69 r O (D aouepunqv eAnelea O [s O m atxeme $18 a 30050.2 d —» 32a3=_+3343 FF LivylP r?» by: __ :. 333 333:._ _33_._.3.333__:33: or ON om. 93$? 8 on 00 0v. :5 aouepunqv anneleu 1' v1 1.. .I '1 .1. II 1 1| 1| II I» vl .1 7| 3' I: wl f. II .II ‘ 3‘ .I. .II ‘1 l .l. VI I .I 3' r 105 1 338.32 3 2.. 33 . 3395850 .2 3353.33 H 3.5.3 5.8: [83 333.3338 nm 233333.33 187 CV NE 83 003 833 333339 cot. 083. 08 com 003. com com oov com com 3 3 3 3 373 3.3. .313 .3 13133. 33311333 34333331333373 3 313.3313; .3.+3433 33 3.. 3 1O 3 . _ I 3 .33 3 3 3 .— . 3 _ H m? mom mom new - w n w. nomV W m r U 18% .1. U .. m n2 mg 93.3% H wow Q>m¢mm3oqfim30~3 H T . woo3 wmdow mm 3m 8.33300 9m 03333333 191 63 NE comp 83 com? 89 oo: 89 com com con com com oov or cm 8 8 3 8 eouepunqv eAgelea 0 rs O m om ww. 23¢ 3393338230.? 935.8333 9.. 3a 335.38 3 8&5 192 3e N\E com w 83 com 3 com F 83. w coop com com cos com com co... :3....3...33:..3..3«.33: .3...3..._....13:.3:.3:. .3:.353.3.3.3.333..3...373...3:3..3.3 3.3...3...3...3...3....3....3....3:.3....3......3..«+7.13 o . 33333 33 3.33 33333333333333 3333. 33.3 13333333 33333.33 3 . H .18 n mom won .133va 1 B 1 n" 1 on m H w 1 n 1 U 18%. m U . m 3. .1333 wow .18 3.38 3 . m2: 3339585033. 335938333 333. 33 333.3338 3 2333.33 193 003 com .. comp oo 3. 3333333._rh333L3 F7 3. 3 a: fig 000—. com com com com com _ 3.33. : .3 2.3m Q>mdmmm~3n3m3OM oonw oowv cop F ocow «NS: com com on:‘ L 3F333 com 3333333 _ . 3 3 3 3 33.413 13:33.3 4443:4111. ov. Sum 111-343“ 31414 O C O C O O O o N (D ID V CO N 1- eouepunqv ameIea O (I) O O) IIIIIIIITIYTTIITIIIIIIIIIVYVIITjTIjTTITIII1IjTTII [ O O F 333.3338 5 2&3”: 195 Rom gov $0 3 $33 3 3833883033 .30 :3me 3 .o v.33 3 . 3 v. 3 38338030233 8:332:00 N 3 33 w meowepE 02:8 oEMowamnoxm ocomkmonm 53? £23 I o~m3< mam< I Ema cocoom QM ad Wm v.3 a 3.33% "3me Re 53 $3 23 283 3:2:me 3.3.3 38 $03. Ax5333 $3 $3 3353338333 .30 :3me v. 3 o. 3 vs a6 3303382338333 3353303359 N 3 33 w mnemonic 02:8 oBmowSonm ocomamoam SSH £25 I 33¢. «83. I Ewe cocoom 5.03 m.m3 M31 a6 a 333% 833/3 33 E3 353 23 £83 EoEwSm :3 23 .382: 03.33%; 3308 53 3332 3-80-3833 N .m 233 196 .xam 30m $0 fin 3 383338833 3o ngm w. 3 m6 md v.0 35332098333 Escmgofl w 3 w v muowEEE 02:8 oEmowSonm E330 I wow? $33 Sam/33 I 3.530 3539333 I vmwz cocoom NM. #3 3.3 0.0 a £5 @833 NS VS 33 3 W353 335:3me =A—3—am Ae $333 $3 Ax53: 83382833 30 Eaxm ms wd 0% 38338098333 5:338:33 a w v msomegn 2335a oSaownmnoxm 3.3.330 | 3.053 oomwdx | $330 wmam< | Ema Gowuoom 3.w3 33.03 ed a «Em $33 33 33 S 303 3:253me ._.3 63 8.33383 N .m 0:3 197 .8036»: 033335 oBuowgoxo .30 3238:: 3 «Em «mg n 3833885033 .30 Baum—.3 dongs“. c3353 8 338.3298 60 3.3.. m3 8:3? 33.3238 05 .30 3:33.88: 23.3. .83 8.833%: .3 owns: 233 m3 Aowgoxoiumn 8.3 33833333 333333» 8332 . 198 {an {on 38338833033 .30 353$ 1o No 3833809883 83333839 w v 303333.333 033335 o3n~ow§330xm 8333. I .3330 «£32 I «£2 newcom ed N6 .3 333% $52 33.3 .30 as 33m .30 3 £83 EoEwEm 3x33. $03 $3» 383338333 30 EBxM no 3.33 No 8338338333 8:553 N3 3 m mange»: 02:8 oSmowSonm mnh—O I 33> 2:3 320 I Ewe :Omaoom wd md N6 .3 3.33% man: :3 3o 33 3.3 3233 3.3 3o :3 82 35332.3 33.33.. 38 33.33383 N .m oEwH These results support our hypothesis that long-range interactions induced by the formation of a disulfide bond can stabilize surrounding local folded structure. IV. Conclusion The peptide models (P1 and P") were designed to represent the amino acid sequence in the vicinity of the Cy331-Cys74 disulfide bond in the l-S folding intermediate of LR3IGF-I. It was assumed that they would mimic the conformational features promoted by the corresponding environment in the intact protein. Our experimental results (CD and HDX—MS) show that the native-like heterodimer (PxPn) and the two homodimers (P1P; and PUP") adopt significant folded structure as compared with that in the component monomers. During the folding of LR3IGF-I, long-range interactions induced by the formation of the Cy831-Cy574 disulfide bond may stabilize the local folded structure around it. The local folded structure, in turn, protects this disulfide bond against attack from reducing reagents in the solution such as free sulfhydryl groups. This provides a structural explanation for the observation that the Cys31-Cys74 l-S intermediate accumulates at high levels in early folding intermediates of LR3IGF-I. 199 V. Future Work Further experiments (i.e., size-exclusion chromatography) will be carried out to investigate if any of the dimers aggregate to form non-covalently associated multimers under our experimental conditions. If so, then the stacking of the peptide chains in the multimer probably also contributes to the protection against HDX as observed in our experiments. In addition, further interpretation of our experimental results will be correlated with and rationalized by computer simulation of the folding process of each peptide model. NMR and/or X-ray crystallography can be used to obtain detailed conformational information of the peptide models. 200 VI. Reference [1] X. Li, Y. T. Chou, R. Husain, J. T. Watson, Integration of hydrogen/deuterium exchange and cyanylation-based methodology for conformational studies of cystinyl proteins, Anal Biochem 331 (2004) 130-7. [2] O. B. Ptitsyn, Structures of folding intermediates, Curr Opin Struct Biol 5 (1995) 74- 8. [3] E. Welker, W. J. Wedemeyer, M. Narayan, H. A. Scheraga, Coupling of conformational folding and disulfide-bond reactions in oxidative folding of proteins, Biochemistry 40 (2001) 9059-64. [4] D. J. States, T. E. Creighton, C. M. Dobson, M. Karplus, Conformations of intermediates in the folding of the pancreatic trypsin inhibitor, J Mol Biol 195 (1987) 731-9. [5] L. C. Wu, P. B. Laub, G. A. Elove, J. Carey, H. Roder, A noncovalent peptide complex as a model for an early folding intermediate of cytochrome c, Biochemistry 32 (1993) 10271-6. [6] T. G. Oas, P. S. Kim, A peptide model of a protein folding intermediate, Nature 336 (1988) 42-8. [7] Z.-Y. Peng, L. C. Wu, Autonomous protein folding units, Advances in Protein Chemistry 53 (2000) 1-47. [8] M. Nishiyama, M. Vetsch, C. Puorger, I. Jelesarov, R. Glockshuber, Identification and characterization of the chaperone-subunit complex-binding domain from the type 1 pilus assembly platform F imD, J Mol Biol 330 (2003) 513-25. [9] L. Ellgaard, P. Bettendorff, D. Braun, T. Herrmann, F. Fiorito, I. J elesarov, P. Guntert, A. Helenius, K. Wuthrich, NMR structures of 36 and 73-residue fragments of the calreticulin P-domain, J Mol Biol 322 (2002) 773-84. [10] B. Zhang, Z. Peng, A minimum folding unit in the ankyrin repeat protein p16(H\IK4), J Mol Biol 299 (2000) 1121-32. [11] R. King, J. R. Wells, P. Krieg, M. Snoswell, J. Brazier, C. J. Bagley, J. C. Wallace, F. J. Ballard, M. Ross, G. L. Francis, Production and characterization of recombinant insulin-like growth factor-1(IGF-I) and potent analogues of IGF-I, with Gly or Arg substituted for Glu3, following their expression in Escherichia coli as fusion proteins, J Mol Endocrinol 8 (1992) 29-41. [12] G. L. Francis, M. Ross, F. J. Ballard, S. J. Milner, C. Senn, K. A. McNeil, J. C. Wallace, R. King, J. R. Wells, Novel recombinant fusion protein analogues of 201 insulin-like growth factor (IGF)-I indicate the relative importance of IGF-binding protein and receptor binding for enhanced biological potency, J Mol Endocrinol 8 (1992) 213-23. [13] Y. Yang, J. Wu, J. T. Watson, Probing the folding pathways of long R(3) insulin-like growth factor-I (LR(3)IGF-I) and IGF-1 via capture and identification of disulfide intermediates by cyanylation methodology and mass spectrometry, J Biol Chem 274 (1999) 37598-604. [14] R. W. Woody, H. Sugeta, T. S. Kodama, [Circular dichroism of proteins: recent developments in analysis and prediction], T anpakushitsu Kakusan Koso 41 (1996) 56-69. [15] T. E. Wales, J. R. Engen, Hydrogen exchange mass spectrometry for the analysis of protein dynamics, Mass Spectrom Rev (2005). [16] X. D. Cai, Chhabil, Structural characterization of methionine and leucine enkephalins by hydrogen/deuterium exchange and electrospray ionization tandem mass spectrometry, Rapid Communications in Mass Spectrometry 19 (2005) 1-8. [17] X. Li, R. J. Hood, W. J. Wedemeyer, J. T. Watson, Characterization of peptide folding nuclei by hydrogen/deuterium exchange-mass spectrometry, Protein Sci 14 (2005) 1922-8. [18] P. K. Glasoe, F. Long, Use of glass electrodes to measure acidities in deuterium oxide, Notes 64 (1960) 188-90. [19] R. K. Scopes, Measurement of protein by spectrophotometry at 205 nm, Anal Biochem 59 (1974) 277-82. [20] Y. Bai, J. S. Milne, L. Mayne, S. W. Englander, Primary structure effects on peptide group hydrogen exchange, Proteins 17 (1993) 75-86. [21] Z. Zhang, A. G. Marshall, A universal algorithm for fast and automated charge state deconvolution of electrospray mass-to-charge ratio spectra, J Am Soc Mass Spectrom 9 (1998) 225-33. [22] L. G. Laajoki, G. L. Francis, J. C. Wallace, J. A. Carver, M. A. Keniry, Solution structure and backbone dynamics of long-[Arg(3)]insulin-like growth factor-I, J Biol Chem 275 (2000) 10009-15. [23] L. G. Laajoki, E. Le Breton, G. K. Shooter, J. C. Wallace, G. L. Francis, J. A. Carver, M. A. Keniry, Secondary structure determination of 15N-labelled human Long-[Arg—3]-insulin-like growth factor 1 by multidimensional NMR spectroscopy, FEBS Lett 420 (1997) 97-102. 202 [24] F. F. Vajdos, M. Ultsch, M. L. Schaffer, K. D. Deshayes, J. Liu, N. J. Skelton, A. M. de Vos, Crystal structure of human insulin-like growth factor-1: detergent binding inhibits binding protein interactions, Biochemistry 40 (2001) 11022-9. [25] R. Koradi, M. Billeter, K. Wuthrich, MOLMOL: a program for display and analysis of macromolecular structures, J Mol Graph 14 (1996) 51-5, 29-32. [26] A. A. Canutescu, A. A. Shelenkov, R. L. Dunbrack, Jr., A graph-theory algorithm for rapid protein side-chain prediction, Protein Sci 12 (2003) 2001-14. [27] J. R. Engen, D. L. Smith, Investigating protein structure and dynamics by hydrogen exchange MS, Anal Chem 73 (2001) 256A-265A. [28] I. A. Kaltashov, S. J. Eyles, Studies of biomolecular conformations and conformational dynamics by mass spectrometry, Mass Spectrom Rev 21 (2002) 37-71. [29] N. Carulla, C. Woodward, G. Barany, Synthesis and characterization of a beta- hairpin peptide that represents a 'core module' of bovine pancreatic trypsin inhibitor (BPTI), Biochemistry 39 (2000) 7927-37. [30] E. M. Goodman, P. S. Kim, Folding of a peptide corresponding to the alpha-helix in bovine pancreatic trypsin inhibitor, Biochemistry 28 (1989) 4343-7. [31] Y. Deng, H. Pan, D. L. Smith, That Hydrogen Exchange at Individual Peptide Amide Linkages Can Be Determined by Collision-Induced Dissociation Mass Spectrometry, Journal of the American Chemical Society 121 (1999) 1966-7. [32] J. S. Weissman, P. S. Kim, Reexamination of the folding of BPTI: predominance of native intermediates, Science 253 (1991) 1386-93. [33] K. Biemann, Contribution of mass spectrometry to peptide and protein structure, Biomedical and Environmental Mass Spectrometry 16 (1988) 99-1 11. 203 I[III[III‘IIIIIJ’IIIIIIIIJ‘IIII