IL 7 1.3.2.! :2 vii .( v :5? . “$4.14.: 1: .. ."II? 1.." «3.. b e..- £va 5 . 4r.:.|...tk\. . L I I..- [0413‘ 0’5 1| Icl .I v.lnl.r. . . 15!. .3 rt . In. I) II v‘rlniyoa .L 4. , 1.3.1.! 1h .- : Au. . . , , , :2EZF ll?! 3.2.5....- :. :II 2. , ‘ A: 7| n. !.c3. . ‘ . . , fifimag .. I. 5&9? ‘ . . 5.. . . .v-.. .. v . , .dJl-ifi.).(..1 ..... . . Av.” jmfiiJ’un-vlzzu .I’I .I" . . . v ‘ snowmangipbw : 3|: 1‘» 31‘: ‘ .u. a onto! .04 o { . ‘ w l. 'r " l 5’ [II I l ' FE fi. ~ “3 STATE WNERSITY UBRARES l2! Ill/lI/llll/I/l/Ill/ll!{Ill/IIHIIII/ll ll ‘1! \ E‘— ‘— x \— 1’ HI I . 1 93 00794 4956 w This is to certify that the thesis entitled INVESTIGATIONS OF TRIPHENYLPHOSPHONIUM DERIVATIZATION OF PEPTIDES FOR IMPROVED ANALYSIS BY FAST ATOM BOMBARDMENT MASS SPECTROMETRY presented by Timothy Jay Nieuwenhuis has been accepted towards fulfillment of the requirements for Chemistry M.S ' degree in Major rofessor _ P ofessor 0 Chemistry and Biochemistry Date October 2, 1992 0-7639 MS U is an Affirmative Action/Equal Opportunity Institution LIBRARY Michigan State University cMma-m PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE MSU Is An Affirmative Action/Equal Opportunity Institution cWMS-e: INVESTIGATIONS OF TRIPHENYLPHOSPHONIUM DERIVATIZATION OF PEPTIDES FOR IMPROVED ANALYSIS BY FAST ATOM BOMBARDMENT MASS SPECTROMETRY By Timothy Jay Nieuwenhuis A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Chemistry 1992 ABSTRACT INVESTIGATIONS OF TRIPHENYLPHOSPHONIUM DERIVATIZATION OF PEPTIDES FOR IMPROVED ANALYSIS BY FAST ATOM BOMBARDMENT MASS SPECTROMETRY By Timothy Jay Nieuwenhuis Fast atom bombardment (FAB) is a popular ionization technique in mass spectrometry (MS) because this method can desorb highly polar and thermally labile analytes dissolved in a liquid matrix. Because the surface activity of analyte in the hydrophilic matrices used for FAB affects their ionization emciency, derivatization techniques to increase the hydrophobic character of peptides are investigated. Reagents to selectively attach the triphenylphosphonium (TPP) group to either terminus of a peptide are synthesized or are commercially available. The hydrophobicity of the TPP group increases the ionization efficiency of the derivatized peptides, and its localized charge directs fragmentation from the terminus of derivatization. The FAB-MS response of leucine enkephalin (YGGFL) modified with various TPP-containing reagents is examined and the fragmentation pattern of derivatized VGVAPG is analyzed by FAB-CAD-B/E. Investigation of a series of chemical reactions proposed for the ethyl-TPP derivatization of peptides in a solid phase reactor indicates that this reaction scheme is feasible. Procedures for the isolation and future TPP- derivatization of glycopeptides are described. Dedicated to my parents Henry and Arlene Nieuwenhuis iii ACKNOWLEDGMENTS First, I wish to thank my co-advisors Dr. Rawle Hollingsworth and Dr. J. Throck Watson for ideas, insights, support, and guidance that were crucial to the completion my degree. I am grateful to Drs. John Allison and J efl' Ledford for their time and advice as members of my guidance committee. I acknowledge Dr. Z.-H. Huang for input on the project simulating the solid phase derivatization of peptides and Dr. J. Leykam for assistance and the use of equipment necessary for glycopeptide isolation. The helpful advice of Drs. David Wagner and Yoon-Seok Chang also is appreciated. I would be remiss if I did not acknowledge the members of the Watson group. I thank Kuruppu Dharmasiri, Jennifer Johnson, Jian Wang, Ken Roth, and Ben Gardner for their moral support and laboratory assistance. Thanks also are in order to the Mass Spectrometry Facility gang; the assistance of Dr. Doug Gage, Bev Chamberlin, Mike Davenport, and Mel Micke is appreciated heartily. I am endebted and grateful to my parents for providing me the opportunity to pursue a college education. The love and support and of , my dear friend, Carrie Blok, entitle you to share in my achievement. I thank my siblings, Rob and Alyce, for the family activities and relaxation that made my studies at MSU enjoyable. Finally, but most importantly, "I thank Christ Jesus our Lord, who hath helped me." (I Tim. 1:12 KJV). iv TABLE OF CONTENTS List of Tables ............................................................................... xi List of Figures ............................................................................ xii Chapter]. InhoductionandOlfiectivee A. Introduction ................................................................ .1 B. Fast atom bombardment mass spectrometry 1. Historical background ............................................... 2 2. FAB-MS of underivatized peptides and CAD ................ 3 3. Fragmentation of underivatized peptides by FAB-CAD- M S / MS .................................................................... 5 4. Liabilities of FAB ionization and CAD-MS / MS ............ 11 C. Derivatization in conjunction with FAB-MS analysis 1. Derivatization schemes of other investigators ............. 14 2. Strategies of peptide derivatization ............................ 15 a) Enhancement of surface activity ........................... 16 b) Peptide degradation ............................................ 18 i) Hydrolysis .................................................... 18 ii) Methanolysis ................................................ 18 c) Modification of fragmentation .............................. Z) ' i) Fixed charge ................................................ Z) ii) Adduct formation .......................................... 24 iii) Dialkylphosphorylation .................................. 25 (1) Determination of functional groups ...................... % i) Acetylation ................................................... % ii) Esterification ................................................ % iii) Amino acid detection ..................................... 27 e) Miscellaneous .................................................... so D. Research objectives ...................................................... 33 E. References .................................................................. 34 Chapter II. 'h'iphemdphosphonium Derivatization and the SeparationofComponemaoftheReacfionMixhne A. Introduction ................................................................ 39 B. Synthesis of reagents containing the TPP group 1. (2-Bromoethyl)TPP a) Synthesis by Method I i) Experimental ............................................... 41 ii) Results and discussion .................................. 41 b) Thin-layer chromatography i) Experimental ............................................... Q ii) Results and discussion .................................. 45 c) Flash column chromatography i) Experimental ............................................... 47 ii) Results and discussion .................................. 47 vi d) Synthesis by Method II i) Experimental ............................................... 48 ii) Results and discussion .................................. 49 2. (2-Aminoethyl)TPP a) Experimental ..................................................... 49 b) Results and discussion ........................................ 49 3. (4-Carboxybenzyl)TPP a) Experimental ..................................................... 52 b) Structural investigations 1') Analysis by FAB-MS ...................................... 55 ii) Analysis by NMR .......................................... 55 . Peptide derivatization procedures 1. Bromoalkyl-TPP reagents and vinyl-TPP a) Experimental ..................................................... (i) b) Results and discussion i) (2-Bromoethyl)TPP and vinyl-TPP ................... E) ii) (3-Bromopropyl)TPP and (4-bromobutyl)TPP ..... 65 2. The “Kunz” reagent a) Introduction ....................................................... 68 b) Experimental ..................................................... w c) Results and discussion ........................................ w 3. (4-Carboxybenzyl)TPP a) Experimental ..................................................... 71 b) Results and discussion ........................................ 72 vii 5. (2-Aminoethyl)TPP a) Experimental ..................................................... 74 b) Results and discussion ........................................ 74 Summary of derivatizing reagents ............................ 7'7 . Separation of the C-terminal derivatization mixture 1. Experimental .......................................................... 78 2. Results and discussion ............................................ 79 . Quantitation of the C-terminal reaction yield .................. 82 1. Determination of the absorption coefficient a) Experimental ..................................................... 83 b) Results and discussion ........................................ 83 Direct quantitation of the reaction yield a) Experimental ..................................................... 84 b) Results and discussion ........................................ 84 . FAB-CAD-B/E fragmentation of derivatized VGVAPG 1. 2. Introduction ........................................................... 85 Mass spectrometry a) Instrumentation ................................................. 87 b) Operating principles ........................................... 87 Analysis of underivatized VGVAPG by FAB-CAD-B/E. .................................................... 89 (2-Bromoethyl)TPP and vinyl-TPP derivatized VGVAPG .............................................. 92 (2-Aminoethyl)TPP derivatized VGVAPG .................. 94 (3-Bromopropyl)TPP derivatized VGVAPG ................. % viii 7. (4-Bromobutyl)TPP derivatized VGVAPG ................... % 8. ‘Kunz’ reagent derivatized VGVAPG ........................ Q 9. (4-Carboxybenzyl)TPP derivatized VGVAPG ............. 101 G. References ................................................................. 103 Chapterlll. DevelopmentofOn-line Whiphmwlphosplmnium A. Introduction .............................................................. 105 B. Advantages of an on-line derivatization reactor ............. 107 C. Proposed setup ........................................................... 109 D. Chemical methodology ................................................ 111 E. Separation of a peptide mixture by LC-MS 1. Experimental ........................................................ 113 2. Results and discussion ........................................... 114 F. Reaction in homogeneous solution ............................... 114 1. Formation of the ‘Kunz” reagent o-nitmphenolate a) Experimental .................................................... 119 b) Results and discussion ...................................... 119 2. Peptide derivatization a) Experimental .................................................... 123 b) Results and discussion ...................................... 124 G. Conclusions ............................................................... 132 H. Future work .............................................................. 132 I. References ................................................................. 133 ix Chaptu'N. FutureDezivatizaflonofaGb'copepfide A. Introduction .............................................................. 135 B. Procedures for glycopeptide isolation ............................ 137 C. Glycopeptide analysis ................................................. 139 D. The analysis of glycoproteins by FAB-MS ...................... 140 E. Derivatization of carbohydrates and analysis by FAB-MS ............................................. 142 F. Advantages of ethyl-TPP derivatization of a glycopeptide ......................................................... 144 G. References ................................................................. 145 Table2J. Table2.2. LIST OF TABLES Investigated derivatizing containing the TPP group. Ratio of signal intensity of derivatized leucine enkephalin to that of underivatized, protonated peptide contained in the reaction mixture during analysis by FAB-MS. xi FigureLL FigumLZ. Scheme Ll. FigureL3. FigureL4. Figure 2.]. LIST OF FIGURES Representation of the processes occurring with FAB. Protonated (M-I-H)+ and deprotonated (M-H)' molecules of analyte are produced as well as many neutrals, M, and ions of glycerol, G. If a species exists as a “preformed ion”, it is desorbed as (M)+. Structures of the common ions encountered in FAB- CAD-MS/MS and their nomenclature as revised by Biemann. Adapted from reference 1 2. Fragmentation of the peptide backbone and nomenclature proposed by Roepstorfl' and Fohlmann [11]. Example of a FAB-CAD-B/E product ion spectrum displaying a complete series of y. fragment ions and an overlapping h, series, which permits complete amino acid sequence determination of the pentapeptide Y-G-G-F-M. FAB-CAD-BIE product ion spectrum of the heptapeptide R-V-Y-V-H-P-F that does not provide a complete or overlapping series of peaks for complete amino acid sequence elucidation. The FAB-MS spectrum of the (2-bromoethyl)TPP product synthesized by reaction of 1 ,2-dibromoethane with triphenylphosphine. The molecular ion (M)+ of (2-bromoethyl)TPP is represented by the doublet at m/z 369 and 371; the vinyl-TPP and (2- diphenylphosphinoethyl)TPP impurities appear at m/z 289 and m/z 475, respectively. Nitrobenzyl alcohol (NBA) was used as the FAB matrix. xii 13 Human Figure 2.4. HPLC chromatogram of the (2-bromoethyl)TPP product mixture synthesized by reaction of TPP with 1 ,2-dibromoethane. For this separation, approximately 100 nmol of product mixture was injected. The parameters used in the separation are: mobile phase, 45:55 (v/v) acetonitrile/water; flow rate, 1 mL/min; column, 4.6 x 250 mm; stationary phase, 5 pm diameter, C13 bonded phase; detector, 214 nm, 1.0 AUFS; chart speed, 20 cm/hr. The FAB-MS spectra of the components of the (2- bromo-ethyl)TPP reaction mixture prepared by the method of Fredrich and Henning as separated by flash column chromatography using a 14:100 (v/v) methanol/chloroform mobile phase: a) (2- methoxyethyl) TPP corresponding to the lower TLC spot and last-eluting compound when separated by flash column chromatography. b) vinyl-TPP and (2- bromoethyl)TPP that were the first-eluting compounds during separation by flash column chromatography and was represented by the upper spot after separation by TLC . NBA was the FAB matrix used for both spectra. HPLC separation of the (2-bromoethyl)TPP product synthes-ized by reaction of (2-hydroxyethyl)TPP with PBr3. Approximately 10 nmol of product was injected for separation. The parameters used in the separation are: mobile phase, 45:55 (v/v) acetonitrile/water; flow rate, 1 mL/min; column, 4.6 x 250 mm; stationary phase, 5 pm diameter, C13 bonded phase; detector, 214 nm, 0.5 AUFS; chart speed, 20 cm/hr. Analysis of the product produced by reaction of (2- hydroxy-ethyl)TPP with PBr3 by FAB-MS. The peaks representing the molecular ion (M+) of (2- bromoethyl)TPP dominate the FAB-MS spectrum (m/z 369, 371). Vinyl-TPP impurity appears at m/z 289. The FAB matrix was NBA. xiii 51 F‘isumZ-B. Figure 2.7. W23. Figure 2.9. Figure2.10. Figure 2.11. The chromatogram resulting from the separation of the (2-aminoethyl)TPP product by HPLC. The first on- scale peak represents the (2-aminoethyl)TPP product; the second peak represents vinyl-TPP impurity in the product. For this separation, approximately 200 nmol of product was injected and the parameters used in the separation are: mobile phase, 45:55 (v/v) acetonitrile/water; flow rate, 2 mL/min; column, 10 x 250 mm; stationary phase, 5 pm diameter, 013 bonded phase; detector, 214 nm, 0.5 AUFS; chart speed, 20 cm/ hr. The spectrum resulting from the analysis of the (2- aminoethyl) TPP product mixture by FAB-MS. The molecular ion (M+) of (2—aminoethyl)TPP dominates the spectrum (m/z 306), but some vinyl-TPP is observed due to impurity and fragmentation (m/z 289). The peaks at m/z 136 and 154 are characteristic of the NBA matrix used. Analysis of the (4-carboxybenzyl)TPP product by FAB- MS that was produced by reaction of TPP with 4- (chloromethyl)ibenzoic acid. The peak representing the molecular ion (M+) of the (4-carboxybenzyl)TPP product is at m/z 397. The base peak in this spectrum represents protonated triphenylphosphine oxide and the peak at m/z 201 represents diphenylphosphine oxide. The FAB matrix was glycerol. The spectrum resulting from analysis of the (4- carboxybenzyl) TPP product by FAB-CAD-B/E. The fragmentation observed is relawd to the structure of the product in the inset. The FAB matrix was NBA. Analysis of the (4-carboxybenzyl)TPP product by NMR when dissolved in a) deuterated chloroform and in b) d4-methanol. The FAB-MS spectrum of 1 nmol of YGGFL N- terminal derivatization mixture. Leucine enkephalin was reacted with a 20-fold excess of a) vinyl-TPP and b) (2-bromoethyl)TPP in an acetonitrile/pyridine 1:1 (v/v) reaction mixture. A higher response resulting from the molecular ion (M+) of derivatized peptide (m/z 844) was observed for reaction with vinyl-TPP than (2-bromoethyl)TPP. xiv Figurem Figure2.13. Figure2.l4. Figure2.l5. Figure 2.16. The FAB-MS spectrum corresponding to 1 nmol of YGGFL reacted in a potassium phosphate/sodium borate buffer solution (pH 9.0) with a 20-fold molar excess of a) vinyl-TPP and b) (2-bromoethyl)TPP. The spectrum represents 1 nmol of YGGFL reaction mixture as isolated by the Sep-Pak® procedure. The molecular ion of derivatized peptide (M+) appears at m/z 844, and protonated underivatized peptide at m/z 556. The FAB-MS spectra representing 1 nmol of YGGFL reacted with a) (3-bromopropyl)TPP and b) (4- bromobutyl)TPP. The reaction was performed with a 20-fold molar excess of reagent in an acetonitrile/pyridine 1:1 (v/v) solvent mixture. The molecular ion (M+) of propyl-TPP derivatized YGGFL appears at m/z 858 and butyl-TPP at m/z 872. The FAB-MS spectrum respresenting 1 nmol of YGGFL reacted with a 20-fold molar excess of “Kunz’ reagent in an acetonitrile/pyridine solvent mixture. The peak representing PEOC derivatized peptide (M+) appears at m/z 888, decarboxylated, ethyl-TPP derivatized peptide is at m/z 844, and underivatized peptide appears at m/z 556. The FAB-MS response of 1 nmol of YGGFL when reacted with a 20-fold molar excess of (4- carboxybenzyl)TPP and EDC in an aqueous solution (pH 5.0) buffered with TFA. The peak representing the molecular ion (M+) of derivatized YGGFL appears at m/z 934, the protonated, underivatized peptide appears at m/z 556. The FAB-MS spectrum representing 1 nmol of YGGFL reacted with a 20-fold molar excess of (2- aminoethyl)TPP and EDC in an aqueous solution (pH 5.0). The peak representing derivatized peptide (M+) appears at m/z 843 and underivatized peptide is at m/z 556. XV 76 Figure 2.17. Figure2.18. Figure 2.19. Figurem Figure2.2l. Flgure2.22. Figure2.23. Figure2.24. Figume2.25. HPLC separation of 20 nmol of the C-terminal derivatization mixture of YGGFL. The reaction mixture contained a twenty-fold excess of (2- aminoethyl)TPP and EDC. The compounds that the peaks represent are identified in the chromatogram. The conditions of the separation are: flow rate 2mL/min; column, 10 x 250 mm; stationary phase, 10 um diameter, 013 bonded phase; detector, 214 nm, 2.0 AUFS; chart speed, 30 cm/hr. Analysis of the peak representing separated, C- terminally derivatized YGGFL by FAB-MS. The molecular ion (M+) of derivatized YGGFL dominates the spectrum (m/z 843), although signal representing vinyl-TPP and (2-aminoethyl)TPP (m/z 306) is observable. The FAB-CAD-B/E spectrum of the molecular ion (M+) of C-terminally derivatized YGGFL with (2- aminoethyl)TPP appearing at m/z 843. The FAB-CAD-B/E spectrum of underivatized VGVAPG. The protonated molecule (M+H)+ is at m/z 499. The FAB-CAD-B/E spectra of the molecular ion (M+) of N -terminally, ethyl-TPP derivatized VGVAPG by a) vinyl-TPP and b) (2-bromoethyl)TPP (m/z 787). The FAB-CAD-B/E spectrum of the molecular ion (M‘i') of C-terminally, ethyl-TPP derivatized VGVAPG as modified by (2-aminoethleI‘PP (m/z 786). The FAB-CAD-B/E spectrum of the molecular ion (M+) of N-terminally, propyl-TPP derivatized VGVAPG after derivatization with (3- bromopropyl)TPP (m/z 801 ). The FAB-CAD-B/E spectrum of the molecular ion (M+) of VGVAPG modified at the N-terminus with a butyl-TPP group as yielded from reaction with (4- bromobutyl)TPP (m/z 815). The FAB-CAD-B/E spectrum of the molecular ion (M+) of N-terminally derivatized VGVAPG as derivatized with the “Kunz” reagent to yield PEOC derivatized peptide (m/z 831 ). xvi 100 mm Figure 3.1. Scheme 3.1. Figure3.2. Figure 3.4. Scheme3.2. Figure 3.5. The FAB-CAD-B/E spectrum of the molecular ion (M+) of N-terminally modified VGVAPG as modified by the (4-carboxybenzyl)TPP group (m/z 877). Proposed on-line, ethyl-TPP derivatization setup. The most effective design probably will perform solid phase, precolumn derivatization prior to LC separation and analysis by FAB-MS Chemical reaction scheme for solid phase, ethyl-TPP. derivatization. ‘Kunz" reagent (1) reacts with immobilized o-nitrophenol to form the “Kunz”-o- nitrophenolate (II). Peptide reacts with the covalently bound reagent, derivatizing the peptide. The column will be regenerated with the ‘Kunz’ reagent. UV—Vis trace of the four-component peptide mixture containing 100 pmol of VGVAPG, GLA, YGGFL, and VQAADYING following separation with a microbore column, prior to analysis by frit-FAB-MS. Reconstructed total ion current (TIC) plot (a) obtained when the peptide mixture entered the mass spectrometer. Scanning was commenced when the first-running peptide was detected by the UV-Vis detector positioned prior to the mass spectrometer. Figure 3.3 (b) contains the reconstructed mass chromatograms of the individual components of the peptide mixture. The background-subtracted mass spectra of individual peptides (a—d) eluted during LC-MS. The scan rate was one scan every 10 seconds. Homogenous-phase chemistry simulating the on-line reaction sequence. Nonmobilized o-nitrophenol reacts with the ‘Kunz’ reagent (I) to produce the “Kunz”-o- nitrophenolate (II). This product (II), in turn, derivatizes the peptide to yield the “Kunz” modified peptide (III). FAB-MS spectrum of reaction mixture resulting from mixing a five-fold molar excess of o-nitrophenol and 0.5 uL 2,4,6-trimethylpyridine with 1 mg of “Kunz” reagent for 10 sec. (b) Linked-scan spectrum at constant B/E of the o-nitrophenolate of the “Kunz” reagent. xvii 102 110 112 115 116 117 118 12) Fianna”. Figure 3.7. Scheme 3.3. Figure 3.8. Figure 3.10. Investigation to determine the kinetics of the reaction yielding the o-nitrophenolate of the “Km” reagent. Reaction yield was defined as the ratio of peak intensity at m/z 472 (11, Scheme 3.2) divided by peak intensity at m/z 307 [(2-hydroxyethyl)TPP]. These results show the product forms quickly, but is unstable in a solution with water-containing atmosphere above it. Reaction of 1.25 nmol/uL AAA methyl ester with a theoretical 25-fold molar exess of the “Kunz”-o- nitrophenolate. The following spectra represent 1 1.1L of the reaction medium containing 1.25 nmol of AAA methyl ester after 32 minutes of reaction (a) at room temperature, (b) at 37°C, and (c) in the presence of TEA. Suggested reaction pathway for decarboxylation of the peptide derivatized with the “Kunz” reagent to form an ethyl-TPP derivatized product. FAB-CAD-B/E spectrum (a) of 1 nmol of AAA methyl ester derivatized with vinyl-TPP and (b) FAB-CAD-B/E spectrum of 1 nmol of AAA methyl ester derivatized with the “Kunz”-o-nitrophenolate. Spectra of the reaction mixtures theoretically containing 1 nmol of AAA methyl ester after derivatization: (a) ~8 hrs in a pyridine/acetonitrile (1 :1 v/v) solvent mixture at 37°C and a 20-fold molar excess of vinyl-TPP, (b) 32 min in acetonitrile/TEA with a approximate 25-fold molar excess of “Kunz”-o- nitrophenolate. Investigation of the speed of peptide derivatization with the “Kunz”-o-nitrophenolate. The yield of reaction was determined by the ratio of decarboxylated, ethyl-TPP, derivatized AAA methyl ester (m/z 534) to underivatized and protonated AAAmethyl ester (m/z 246) peak intensity. This reaction was performed at room temperature, 37°C, and in the presence of TEA, as the legend indicates. For these studies , 1.25 nmol of AAA methyl ester was reacted with a 25-fold excess of the “Kunz”-o- nitrophenolate. xviii 122 125 1% 128 129 13) Figure 3.11. (a) FAB-MS of reaction mixture containing 1.25 nmol of VGVAPG after reaction 32 min with a 25-fold molar excess of “Kunz”-o-nitrophenolate with 0.5 uL of TEA present as a base catalyst. Decarboxylated, ethyl-TPP, derivatized product appears at m/z 871, while protonated, underivatized peptide appears at m/z 499. (b) FAB-CAD-B/E spectrum of 1.25 nmol of decarboxylated, derivatized VGVAPG. xix 131 Chapter I INTRODUCTION AND OBJECTIVES A. Introduction Although mass spectrometrists have endeavored to elucidate the structure of biopolymers for over thirty years, the most significant contributions of mass spectrometry (MS) to the analysis of important, biological compounds have occurred in the past decade. Two major developments in instrumentation revolutionized the analysis of biopolymers by MS: i) high-field magnetic mass spectrometers and ii) fast atom bombardment (FAB) ionization. These advances made it possible to obtain mass spectra of intact polypeptides, oligosaccharides, oligonucleotides, and other small biopolymers without the need for prior derivatization or degradation [1]. However, sample derivatization also can be utilized to augment the amount of information available from analysis by FAB-MS. At this time, FAB-MS is the most popular of the desorption ionization techniques and has proved to be a powerful and cost-effective method for the structural determination of natural and recombinant peptides [2]. The tandem FAB-MS technique has become the method of choice to confirm the amino acid sequence of proteins predicted by DNA sequencing [3]. Structural modifications of a biopolymer (posttranslational modifications or alteration of amino acids) can be identified by FAB-MS, as these cannot be deduced from DNA analysis. Even more importantly, FAB- MS can analyze peptide mixtures without costly purification procedures. This is an aspect foreign to Edman degradation. 1 2 There are many schemes for biopolymer modification that serve to complement their characterization by FAB-MS. Many peptides, especially those of higher molecular weight, show poor or incomplete sequence data, even with the use of collisionally-activated dissociation (CAD) in combination with tandem mass spectrometry (MS/MS). This necessitates the use of chemical reagents to modify specific amino acid side chains, or either terminus, to aid in identifying the amino acid sequence of a peptide. Peptide modification is used to increase peptide ionization emciency and direct fragmentation to simplify the elucidation of the amino acid sequence. The objectives of this first chapter are to i) introduce the FAB ionization and CAD techniques utilized for determining the amino acid sequence of underivatized peptides, ii) discuss the strategies, procedures, and efi‘ectiveness of chemical derivatization methods employed to enhance the analysis of peptide biopolymers by FAB, and iii) detail the objectives and criteria used for evaluating the derivatization schemes investigated in this research project. B. Fast atom bombardment mass spectrometry 1. Historical background Before the introduction of FAB-MS by Barber et al., [4,5] chemical derivatization was necessary to increase peptide volatility and thermal stability for sequencing by gas chromatography-mass spectrometry (GC- MS). In the late 19508, Biemann investigated polyamino alcohols derived from peptides with chemical acetylation and reduction by LiAlH4 for GO- MS analysis [6]. The derivatized peptides displayed improved volatility, but also the specific fragmentation of specific bonds in the peptide backbone. Later, this derivatization procedure was modified to yield N- trifluoroethyl-O-trimethylsilyl polyamino alcohols. Peptides modified to N- 3 acetyl-N,O-permethylated derivatives were introduced into the mass spectrometer by a direct probe for electron ionization [7]. Even with these peptide modification schemes, mass spectrometry contributed little to protein primary structure elucidation in the 1 9608 and 1 9708, however. 2. FAB-MS of underivatized peptides and CAD During FAB-MS, a beam of xenon or argon atoms (or Xe'i' ions for liquid secondary ionization mass spectrometry, LSIMS) is focused on a viscous, liquid matrix (commonly glycerol) to desorb and ionize highly polar, thermally labile analyte molecules dissolved in the matrix (Figure 1.1). This "soft" ionization technique has expanded the scope of mass spectrometric analysis to include a large number of biologically important compounds. The secondary ion beam produced is long-lasting and stable, which is another advantage of FAB ionization. Intact, high mass molecules are desorbed into the gas phase. Little excess energy is imparted to the molecule with FAB ionization when an ionic species is produced by the addition or removal of a proton to form stable, even-electron ions, (M+H)+, or (M-H)‘. These ions have a low tendency to fragment, which is useful for the unambiguous determination of molecular weights, but is a liability if fragmentation is desired for structural elucidation. When FAB-MS was first implemented, researchers already realized its potential for sequence determination of high-mass peptides [8]. However, further investigation showed FAB usually does not give enough energy to the peptide to cause sufficient fragmentation for complete structural determination. Tandem mass spectrometry (MS/MS) is the most commonly used technique to increase the internal energy of a peptide and induce fragmentation [9]. With MS/MS, the precursor ion fragments when undergoing high energ CAD with a collision gas (e.g., helium, 10'3 torr). Fast atom beam \5 §3 "Selvedge" Figure 1.1. Representation of the processes occurring with FAB. Protonated (M+H)+ and deprotonated (M-H)' molecules of analyte are produced as well as many neutrals, M, and ions of glycerol, G. If a species exists as a “preformed ion”, it is desorbed a8 (M)+. S Peptide fragmentation is thought to occur at bonds near and remote to the site of protonation [10]. The degree of charge localization and the site of protonation influences the relative intensity and structure of ions produced with CAD fragmentation. With FAB ionization, the site of protonation usually is at basic amino acid side chains; consequently, the positive charge tends to localize on lysine or arginine residues rather than on the less basic amide nitrogens and oxygens on the peptide backbone. 3. Fragmentation of underivatized peptides by FAB-CAD—MS / MS Positive ion FAB-MS has sensitivity advantages for peptide analysis because the basic nature of the polyamide backbone causes protonated molecules (M+H)+ to form more readily than deprotonated molecules (M- H)‘. The repeating units of a peptide biopolymer contain the twenty common amino acids, which have a repeating structure —HN—CH(R)—CO—, where R is the side chain characteristic of each common amino acid. When the peptide bonds are cleaved with CAD after protonation, charge retention on the carbonyl group generates the b" ion series (Figure 1.2). The mass difference of the fragment ions reveals the identity of consecutive amino acids with the exception of isometric leucine and isoleucine or isobaric glutamine and lysine. Another set of peaks representing y" product ions is due to a second fragmentation pathway involving the transfer of a hydrogen atom from a nitrogen atom on the N-terminal side of the amide bond and proton retention at the C-terminal portion of the peptide after amide bond cleavage. Ifno basic amino acids are present in the peptide, mainly hp. and some yn ions occur in the mass spectrum. The first nomenclature for peptide fragmentation (Scheme 1.1) was proposed by Roepstorfi' and Fohlmann [11]. Since then, the nomenclature has undergone revision; Biemann labeled fragment ions by lower-case N -terminal Ions + 0.: H—(NH—CHR—CO),-,—NH=CHR, or Ho CknuRnb H—iNH—CHR—Coi,-.—NH—CH H. R, a, + l: H—‘(NH——CHR-—c0)‘.-.——NH—CH- 1...: H—(NH—CHR—CO).-n—NH—CHR.—CEO‘ HQ c,: H-‘(NH—CHR—CO)‘.—NH3 b H. CR. 4,: H—INH—CHR—coi.-.—NH—cn C-terminal Ion Types H. u,: HN=CH—co—iNH—CHR—c07._.—0H CR,‘ H. w,: CH—CO—iNH—CHR-COIrr—OH x,: +OEC—NH—«CHR,,—CO---(NH—CHR—--CO),.. |—OH or H 0 o=C=N—'CHR.—CO—(NH—CHR—C(R.-.—0H H? y,: H-‘(NH—CHR—Coirou H+ y, - 2: 'HN=CR.—co—(NH—CHR—coi,-.—0H CK‘K. H4 z.: CH—co—iNH—CHR—coi.-.—ou H4 2, + I: -CHR.—co-«'NH—CHR—cofi._.—0H Figure 1.2. Structures of the common ions encountered in FAB-CAD- MS/MS and their nomenclature as revised by Biemann. Adapted from reference 12. XaYszéYzzleYIZ. F— F l— l— T T l— T (Ii lie u 1‘3 (ll li‘ ii HzN—(ll-"C"N"(|3- c- N--(l3- c- N--C—C—OH H H H H H H H _J .i__._i_iJ B. A.B..A.B. 9 fl Scheme 1.1. Fragmentation of the peptide backbone and nomenclature proposed by Roepstorfi‘ and Fohlmann [11]. 8 letters to differentiate them from immonium ions, which are identified with upper-case letters. The structure of the common ions formed with CAD of peptides are outlined in Figure 1 .2 [12]. Interpreting mass spectra usually requires more effort than acquiring the actual data. Deducing the amino acid sequence from the mass spectrum is made easier if other information about the peptide is known; the source of the sample (tryptic fragments contain lysine or arginine at the C-terminus), possible homology with other peptides, or amino acid composition aid in interpreting the mass spectrum. Ideally, the amino acid sequence of a peptide is determined by a complete series (e.g. , bu) or the overlap of two ion series, one derived from the N-terminal sequence of amino acids (an, bu, or cn) and the other derived from the C- terminal sequence of amino acids (x... y", or z"). The FAB-CAD-B/E mass spectrum outlined in Figure 1.3 from 1 nanomole of a peptide is an example of two overlapping ion series unmistakably indicating the amino acid sequence of a peptide. Examining the low-mass end of the spectrum, immonium ions are observed for phenylalanine (m/z 120) and tyrosine (m/z 136), providing a partial amino acid composition. Examination of the high-mass region reveals peaks corresponding to the loss of water and the side chains of methionine, phenylalanine, and tyrosine residues. To deduce the amino acid sequence, the mass difference between (M+H)+ and a fragment ion that corresponds to the loss of mass characteristic of an amino acid residue would represent the yn.1 ion, n being the number of amino acids in the peptide. The peak at m/z 411 represents the loss of 163a from (M+H)+, the mass of a tyrosine residue. Therefore, the tyrosine residue resides at the N-terminus. The next reasonable loss in the y—series is to m/z 354 (57u), which indicates that cum- +8.36 .zacds cosmonflaoa. .05 we 55835233 85:53 can Sada 33988 Egon £033 .958 an $233196 do can meow unofimaum .3» me 858 32988 a mafia—9mm“. 83.8on :3 853.5 Em-Qm Hdawdmwuh lO glycine is the amino acid adjacent to tyrosine in the N -terminal series. Another loss of 57u, represented by the peak at m/z 297, indicates that another glycine is next in the series. A reasonable loss represented by the y“ peak is 147u and this corresponds to a phenylalanine residue, bringing us to the peak at m/z 150. Because two hydrogens are added to the y, fragment ion, the final residue at the C-terminus i8 methionine. The y, ion series indicates the peptide sequence is YGGFM. This sequence can be confirmed by the ion series generated by cleavage at the peptide bond and retention of charge at the N—terminus (bu). The fragment ion represented by m/z 556 corresponds to the loss of water from the (M+H)+. The peak at m/z 425 represents loss of 131u from (M+H-H20)+, which corresponds to a methionine residue. Neutral losses of 147u (to m/z 278) and 57u (to m/z 221) indicate the sequence from the C- terminus is GFM, confirming the N-terminal sequence. The an ion series represented by peaks at m/z 193, 250, and 397 confirms the assignment of the 1),. ion series. Informative fragmentation in this example reveals the primary structure of the peptide as YGGFM, named methionine enkephalin. However, most peptides do not fragment completely because the amide groups are not the most basic functionalities in the peptide. For example, the amine group of lysine and and guanidine group of arginine are more basic than amide groups and the CAD spectra of peptides containing these residues are changed dramatically. Peptides obtained from tryptic digestion, which contain a basic arginine or lysine residue at the C- terminus, display a series of ions (v... w", x", y", 2,.) retaining the charged, C-terminal residue, which effectively competes with amide groups for the proton in (M+H)+ [13]. Wn ions are products of y" ions, resulting from 11 cleavage of the B,y-bond of the amino acid residue. The formation of w" ions complicates the mass spectrum and decreases overall signal intensity, but these unique ions difi‘erentiate leucine and isoleucine; the wn ion of leucine is always 14 mass units lower than that of isoleucine [14]. Wn ions are not produced from glycine, the aromatic amino acids, or alanine; therefore, wn ions rarely represent a continuous series. Y" and w" ion series predominate the FAB-CAD-MS/MS of tryptic peptide fragments with a basic C-terminal amino acid, producing simple, informative spectra. When basic residues are present near the N-terminus of a peptide, the proton is retained at the N-terminus and series of fragment ions retaining the charge at the N-terminus (an, bu, cu, (in) dominate the mass spectrum. When the charge is retained at the protonated guanidino group of arginine, an ions are produced by charge-remote fragmentation, and (1;. ions result from cleavage of the 8,)‘bond of the amino acid side chain. These ion series (an, (1,.) confirm the assignment of peaks caused by amide bond cleavage. The relative abundance of an, bu, cu, x", y", and 2,, ions indicates the position of basic residues along the peptide chain. 4. Liabilities of FAB ionization and CAD-MS / MS A serious liability of FAB-MS is the dependence of peptide ionization efficiency on its surface activity in the liquid matrix. Low surface activity causes poor sensitivity for hydrophilic peptides and discrimination against hydrophilic peptides in the analysis of peptide mixtures by FAB-MS. Often the poor ionization efficiency of hydrophilic peptides prevents their detection in the mass spectrum. This is a handicap for peptide mapping, because it is desirable to obtain a peak of discernible intensity in the mass spectrum with this technique, regardless of the degree of hydrophilicity of the peptide. Although the lack of peaks representing key fragments with 1 2 analysis by FAB-CAD-MS/MS is not restricted to hydrophilic peptides, the low signal abundance from such peptides can increase the difficulties of deducing their complete amino acid sequence. FAB-CAD-MS/MS relies on the capacity of the protonated peptide to fragment eficiently and informatively during FAB ionization or with CAD, prior to analysis by MS/MS. Unfortunately, not all underivatized peptides fragment completely and the product ions required for structural elucidation may be absent or in low abundance. Because the major factor directing the fragmentation of underivatized peptides with FAB-CAD- MS/MS is the site of proton attachment, the presence of a basic residue in the center of a peptide chain complicates mass spectral interpretation. In the absence of strongly basic residues, a mixture of N- and C-terminal ions are observed. These ion series often are incomplete and result in ambiguities when deducing the amino acid sequence. Immonium ions and ions resulting from internal cleavage of the peptide further add to the complexity of the mass spectrum, although they are beneficial for the identification of amino acids near the N-terminus. Another liability of MS/MS is that the efficiency of fragmentation caused by CAD decreases with increasing molecular weight. Therefore, the complete amino acid sequence is diflicult to obtain from the FAB-CAD-MS/MS of a peptide with molecular weight above 1500-1700u. For example, the mass spectrum in Figure 1.4 of 1 nmol of the heptapeptide [val4]-angiotensin III (RVYVHPI) only allows partial determination of the amino acid sequence. Peaks representing the immonium ions of histidine (m/z 110) and tyrosine (m/z 136) are observed in the low-mass end of the spectrum. The side chain losses of valine (m/z 874), histidine (m/z 836), and tyrosine (m/z 810) residues from (M+H)+ are 13 .doflovmoflo 88:qu Soc 888 32988 you $.on mo moron 98893.85 .8 32888 a 035.8 88 noon 85 m-m-H.H->.> ->.m 338888: on» no 838on com 888.8 Smddofism .v.H arm 38 com com OOH. com ocm OOHV com com 2: _ , , _ o 3 m as V \ m r» mm; as J b o m u m N H at 86 \NE “:0 m8 BB ”en fl 5. The 8.. H8 8“ new we a m , m by >H>| m 08. H g g a "2% 8R :8 +8.85 H a m e a o s .2 .H Mona.H>o HflaoflmoH-plh 14 observed at high mass. The peak at m/z 899 represents the loss of water from the protonated peptide (m/z 917). N 0 peak in the mass spectrum corresponds to the loss of an amino acid residue from (M+H-H20)+, which would represent a b...1 ion. However, loss of 193u from the protonated molecule produces an and ion (m/z 724), corresponding to the loss of a phenylalanine residue plus CO. Subsequent losses of 97u, 137u, 99u, and 163u from the as ion indicate the C-terminal sequence is ?YVHPF. A series of peaks at m/z 761, 662, 499, 400, and 263 representing a complementary series of y" ions should confirm the C-terminal sequence. However, only peaks at m/z 499, 400, and 263 are readily apparent from the mass spectrum. Therefore, the incomplete C-terminal sequence cannot be verified, and there is no direct information for the sequence commencing at the N-terminus. Consequently, the complete amino acid sequence of this peptide cannot be deduced from the mass spectrum, although the mass of the protonated molecule and some peaks representing sequence ions are observable in the FAB-CAD-B/E mass spectrum. C. Derivatization in conjunction with FAB-MS analysis 1. Derivatization schemes of other investigators Investigators [15-18] soon realized that chemical modification of an analyte to produce the unconventional properties of nonvolatility and ionic character enhanced its signal-to-background ratio during FAB analysis. The “preformed” ions produced with analyte modification were thought to be more surface active in the FAB liquid matrix than the unmodified species. Protonation, cationization with alkali or transition metal ions, and chemical modification to produce a “preformed” charge were three derivatization techniques used to enhance the analysis of several large, highly polar, and nonvolatile molecules. 15 A dramatic example of the effectiveness of analyte protonation for an improved signal-to-background signal for desorption techniques was demonstrated with the addition of p-toluene sulfonic acid to glycylphenylalanine on a graphite surface [16]. Before treatment, the peak of the protonated analyte was imperceptible from the background, but addition of p-toluene sulfonic acid caused the (M+H)+ peak to dominate the mass spectrum, with the intensity of (M+H)+ greatly increased relative to the intensities of fragment ions. Chemical modification of specific functional groups was also used for improved analysis by FAB-MS. The carbonyl group of keto steroids was targeted with the Girard reagent to form water-soluble hydrazones with a fixed positive charge [1 9]. Although the protonated molecule could not be observed before derivatization, physiological levels (1 ug/L) of androsterone, corticosterone, and androsterone glucuronide were analyzed following derivatization. Following derivatization, suppression of glycerol ion intensity was observed relative to the signal of the modified species. Chemical modification with reagents containing a “preformed” charge produced improved analysis by FAB-MS for biomolecules, as was demonstrated with the modification of the hydroxyl group of the steroid corticosterone with 2-fluoro-1-methyl- pyridinium p-toluene sulfonate [1 8]. 2. Strategies of peptide derivatization Peptide modification is performed in conjunction with FAB-MS for several reasons: i) enhancement of analyte surface activity; it) degradation of the sample to smaller subunits; iii) enhancement of separability; iv) modification of fragmentation: 8) enhanced formation of molecular weight-related ions and b) enhanced formation of structurally informative 16 ions; and v) determination of functional groups. Useful derivatizations should be specific, modifying only the intended portion of the target molecule, and yield a single product. The modification procedure should be fast, simple, and amenable to small sample sizes. Ideally, the derivatization reagents should be safe and stable, but few derivatization methods meet all of these criteria [20]. Many derivatization schemes for peptides have been developed in the last decade. Reports of these derivatization procedures often have appeared in the scientific literature for known, commercially available peptides without a follow-up report on application of the modification technique to unknown peptides isolated from a biological matrix. a) Enhancement of surface activity When a peptide is placed in a drop of liquid matrix on a FAB probe tip surface, a concentration gradient develops, resulting in the surface enrichment of hydrophobic peptides. Inequalities in surface concentrations are observed for different peptides with identical bulk concentrations. This increase in peptide surface concentration or peptide adsorption at the matrix/gas interface is termed “surface activity”. Physical and chemical parameters such as pH, surface tension, micelle concentration, and the hydrophilicity/hydrophobicity of the matrix, charge, size, stability, and solubility of the peptide affect its surface activity [21]. At low bulk concentrations of peptide, the surface peptide concentration is similar to its bulk concentration. As the bulk concentration increases, the peptide forms a monolayer and dislodges the glycerol molecules at the surface. Higher concentrations of peptide cause the formation of micelles composed of peptide. When the matrix surface is covered by a monolayer of peptide, increasing concentrations of peptide do not affect the magnitude of the 17 signal due to the protonated peptide, (M+H)+. Consequently, there is a direct correlation between the sputtering efficiency of protonated molecules of hydrophobic peptides with their surface concentration, which is a function of bulk peptide concentration. One aim of peptide derivatization is to enhance the surface activity of hydrophilic peptides. Many of the liquid matrices used in FAB-MS are polar in character and adding hydrophobicity to a peptide causes it to be drawn to the surface to desorbed preferentially. To increase the surface activity of several dipeptides, Ligon derivatized them with an equimolar amount of dodecanal [22]. Detectability of all the polar, derivatized peptides increased, but the sensitivity for detection of a hydrophobic derivatized peptide was actually less that of the underivatized peptide. Gas chromatography and FAB-MS analysis indicated the reaction was complete for all the investigated peptides. This suggests that the derivatization procedure does not increase the surface activity of hydrophobic peptides. By increasing the surface activity of all components of a peptide mixture, chemical modification also can decrease the discrimination against certain hydrophilic peptides that is observed with the screening of protein digests by FAB-MS [23]. In the peptide mixtures investigated, the relative ion abundance of some peptides was not proportional to their actual concentration in the FAB liquid matrix. The surface activity of one peptide could cause less surface active components in a mixture to be unobservable [24]. Because derivatization inherently increases the homogeneity of molecular mixtures, this technique can equalize differences in surface activity for individual components of a peptide mixture. 1 8 b) Peptide degradation With Edman sequencing, protein degradation is used to identify amino acid subunits, but molecular size reductions used in conjunction with MS are driven by instrument mass range. In the case of tandem mass spectrometry, the need for degradation to smaller size fragments may be motivated by molecular size limits of ion dissociation technology rather than the analyzer mass range. i) Hydrolysis Kidwell et al. [25] first quaternized, then hydrolyzed, esterified, and acylated peptides to achieve the subnanogram detection limits observed by Cooks and coworkers in the static SIMS analysis of quaternary ammonium salts [16]. This reaction sequence produced N-terminally charged products with molecular weights differing by the mass of the amino acids attached to the N -terminus after hydrolysis. Although uncharged side products result from this procedure, they were not observed in the resulting static SIMS spectra. This derivatization procedure produces detection limits in the low nanogram range and compensates for the lack of fragmentation observed in liquid FAB analysis, but deposition on solid silver is necessary and silver- cationized amino acids were observed. Unfortunately, this derivatization method is nonspecific because basic amino acid side chains would also be derivatized. Another problem is that each peptide bond would be hydrolyzed to an unequal extent in this procedure. ii) Methanolysis Treatment of a peptide with dry methanol containing a small amount of hydrochloric acid (0.02-0.1 0N) converts all the free carboxyl groups into methyl esters, but other functional groups remain intact. Reactions with more concentrated HCl (5N, 37°C, 6 h) in dry methanol causes 19 methanolysis of peptide bonds in addition to esterification of the free carboxylic groups. Consequently, methyl ester fragments of the original peptide are produced by cleavage at the carboxylic end of amino acid residues [26]. Treatment of 0.5 mg of peptide with 150 uL of 5N HCl in dry methanol causes primary methanolysis, that is, cleavage of single peptide bonds. The FAB spectra of the methanolysis products display an abundant (M+H)+ ion corresponding to the completely esterified peptide. Information about the original peptide can be obtained from the protonated peptide methyl ester and the complementary fragment series that contain the N - or C-terminal amino acid residues. The (M+H)+ signals for all the methanolytic fragments are present in the mass spectra, but the intensity of the (M+H)+ of fragments containing the esterified C-terminal residue are much more intense than fragments containing the N-terminal residue. The advantages of this procedure are that it is simple, fast, and does not require separation or transfer of the sample. Only primary methanolysis is observed under these reaction conditions and this derivatization technique can be used in the analysis of N-blocked peptides. However, the position of N- and C-terminal residues cannot be distinguished by this technique, and leucine and isoleucine residues cannot be be distinguished. The esterification of glutamic and aspartic acid side chains and methanolysis of the asparagine amide group are also complicating factors. Peptides containing cysteine require methylation of the sulflrydryl group by methyl iodide to give a stable 5-methylcysteinyl residue prior to methanolysis. Methionine residues undergo a partial methylation 'of sulfur during methanolysis. These side reactions cause a complication of the mass spectra of peptides containing cysteine and methionine [27]. 20 c) Modification of fragmentation i) Fixed charge The advent of FAB ionization has led to the use of chemical derivatization to introduce a fixed charge into a peptide, enhancing the detectability of molecular weight related ions. The localized charge has a stronger effect in directing fragmentation than any basic amino acids that may be present in the peptide, drastically altering the appearance of the CAD spectrum of peptides [28] and facilitating structure elucidation by enhancing the formation of structurally informative ions. When a preformed charge is introduced into a peptide, an and tip; ions are more abundant for N -terminally derivatized peptides, while v", y", and w", ions are formed when peptides are derivatized at the C-terminus. The altered fragmentation pattern produced by derivatization with a fixed charge simplifies mass spectral interpretation and usually allows differentiation of the leucine and isoleucine residues. Quaternization is a straightforward reaction that is used for the chemical derivatization of peptides in FAB-MS. “Preformed” ions, especially tetramethyl ammonium (TMA) ions, exist as preformed ions in solution and are ionized preferentially above glycerol [29]. TMA suppresses the ion signal of Bronsted bases in their protonated or deprotonated form. Small suppression effects of TMA over alkali ions were also observed; although when mixtures of alkali metals are made in glycerol, there is little competition between alkali ions in the FAB matrix and cationized species have similar sensitivities. The preference factors displayed by the TMA ions over alkali ions results from the higher surface activity of the TMA ions. 21 A derivatization procedure to convert the e-NHz group of lysine to a 2,4,6-trimethylpyridinium quaternized group was implemented by Johnson et al. [30,31]. This derivative produced decreased abundance of immonium ions, internal fragment ions, and b" ions when compared to the spectrum of underivatized peptide. Under collisionally activated dissociation (CAD) conditions, the derivatized peptide displayed fragmentation remote to the fixed charge of the 2,4,6-trimethylpyridinium quaternized group. As expected, an and (1,. ions predominated when the derivatized lysine was at the N-terminus; v", w", and y" ions predominated when the derivatized lysine was near the C-terminus. When the lysine was in the middle of the peptide chain, the upper mass range displayed intense an and y 3 peaks, but small peaks in the lower mass range were observed because the ions did not contain the preformed charge of the derivatized lysine. , Stults also fixed a positive charge to the amino terminus by chemical derivatization to increase the amount of information obtainable from CAD tandem MS spectra [32]. The two-step procedure selectively quaternizes the N -terminus, exploiting the pKa differences of the N -terminus (pKa~8) and side chain amino groups of lysine (pKa~10.5) and arginine (pK..~12.0). However, cysteine residues must be protected from derivatization by reduction and alkylation. Other disadvantages of this derivatization procedure include a decreased signal intensity of modified peptide and an intense peak resulting from the loss of the -N(CH3)3 group. The use of excessive reagent and a nonvolatile buffer also necessitates sample cleanup (typically by HPLC). Because only well-separated an and d, ions originating from the N-terminus were observed for CAD-MS/MS of the derivatized peptide, enhanced signal due to the ion intensity spread over fewer channels and the need for lower resolving power was noted. 22 Utilizing a microderivatization apparatus, Vath et al. [33,34] modified peptides by covalently linking a trimethylammonium acetyl (TMA) moiety to the N-terminus, converting the peptide to a quaternary ammonium cation. This procedure reproduced the derivatization method developed by Bennett and Day [35], but was altered to be performed in the gas phase. Because no sample transfer was necessary in the microderivatization procedure, the reaction was carried out at the subnanomole level. Although analysis by HPLC and the FAB mass spectra indicated the derivatization was complete, the ratio of signal to background for the derivatized peptide was only one-half that of the underivatized peptide. With FAB analysis, a peak caused by the replacement of the -N(CH3)3 group by a hydrogen was also observed. Moreover, this peptide derivatization procedure is a lengthy three-step process. Despite these disadvantages, peptides derivatized by this method produced simpler and altered CAD mass spectra, providing more sequence information than the CAD spectra of unmodified peptides. The CAD mass spectrum was dramatically changed in one example, giving a complete series of an sequence ions where few sequence ions were observed before derivatization. Fixing a positive charge at the N-terminus by derivatization directs CAD fragmentation to produce an and (1,. ions. The presence of (1.. ions allows the differentiation of leucine and isoleucine. In this sense, the CAD mass spectrum observed is similar to the CAD spectra of peptides containing a basic arginine residue at the N-terminus. Moreover, derivatized peptides with a basic residue at the C-terminus primarily produce N-terminal fragment ions, demonstrating the strong directing effect of a fixed charge on peptide fragmentation. 23 Recently in this laboratory, Wagner et al. [36] have developed (2- bromoethyl)triphenylphosphonium and (2-aminoethyl)triphenylphosphon- ium derivatives to covalently link the ethyl-triphenylphosphonium (ethyl- TPP) group to the N- or C-terminus of a peptide, respectively. The ethyl- TPP moiety incorporates a fixed positive charge and significant hydrophobic character to the derivatized peptide, increasing its ionization efficiency. With CAD, the fixed charge directs fragmentation from the terminus of derivatization. Consequently, a predictable series of predominantly an and (1,. ions is observed after N-terminal derivatization and v", y", and w" ions are observed upon C-terminal derivatization. Significant ethyl-TPP derivatization is accomplished in three hours with simple laboratory procedures. These characteristics, especially the enhanced signal intensity of low levels of peptides and simplified fragmentation observed with CAD of the derivatized peptide, make this a sought-after derivatization approach. The increased detectability and enhanced fragmentation caused by incorporation of the ethyl-TPP group into the derivatized peptide are features that cause this derivative to be ideal for deducing the amino acid sequence of posttranslationally modified peptides. Often, the nature of modification (glycosylation, phosphorylation, sulfation, etc.) increases the hydrophilicity of peptides. The altered peptides often do not fragment efficiently with CAD and a complete ion series for amino acid sequence determination often is absent. Ethyl-TPP derivatization, however, can minimize these problems [37]. Upon derivatization, N-blocked (pyroglutamic acid), phosphorylated, and peptides containing a disulfide bridge yielded a simplified spectrum of sequence information with ion formation directed from the terminus of ethyl-TPP derivatization. Incorporation of the ethyl-TPP group allowed the differentiation of leucine 24 and isoleucine and identification of the posttranslational modification by increased abundance of d, and w" ions. The detectability of the ethyl-TPP modified peptides was at least an order of magnitude greater than the underivatized peptides. These features warrant an investigation of other derivatizing reagents containing the triphenylphosphonium group. ii) Adduct formation Alkali, alkaline earth, and transition metal ions complex with peptides to form cationized complexes. These metal-ligand interactions have been investigated and have resulted in new approaches for mass spectrometric structure determination of peptides. Addition of a drop of an alkali iodide to a liquid matrix containing peptide often causes formation of a peak representing the peptide-alkali cation (M+Cat)+, which many times is of greater intensity than the (M+H)+ of the untreated peptide mixture. Collisionally-activated dissociation studies of the gas phase complexes (M+Cat)+ provide evidence that b, product ions result from the metal cation binding to the N -terminus of a peptide. However, the alkali metal ion can bind to several sites in the peptide molecule and these may include amide oxygens and chelating side chains [38]. These investigators believe that the gas-phase complexes (M+Cat)+ do not contain the metal ion bonded to a deprotonated carboxylate anion, and consequently reflect binding of the metal cation to basic sites in a neutral peptide. Investigators also have studied H 3PO4 adducts that were produced with the addition of concentrated H3PO4 to a glycerol matrix containing leucine enkephalin [39]. Ion series of the general formula (M+nH3PO4+H)+ due to adduct formation, and (M+H+62)+ and (M+H3PO4+H+62)+ series resulting from the slow decomposition of (M+nH3PO4+H)+ ions were observed under FAB conditions. These 25 investigators also used FAB to investigate the aggregation of polyamines, leucine enkephalin, and phosphoric acid [40]. iii) Dialkylphosphorylation Chai and Zhao have reported the analysis of positive FAB of amino acids, dipeptides, and tripeptides after derivatization with diisopropylphosphite [41 ,42]. Although the derivatization procedure is lengthy and requires several steps of sample manipulation, improved sensitivity and decreased background noise from the glycerol matrix were observed after derivatization of the peptides. Most of the derivatized amino acids produced a signal enhanced by a factor of ten in comparison to underivatized amino acids. N-terminal fragment ions dominated the fragmentation of the N -diisopropyloxy phosphoryl derivatized peptides, giving evidence of directed fragmentation. An intense an series of ions was observed, with an-42 and an-84 ion series representing the loss of one and two propene molecules, respectively. Derivatized peptides also displayed suppressed fragmentation of amino acid side chains. The dialkylphosphite reagent used for derivatization also can be used for peptide synthesis [43] and more studies are underway for this derivative [44]. d) Determination of functional groups The presence of functional groups in a peptide can be identified by measuring the mass shift after reacting the peptide with a site-specific reagent. This procedure also can determine the number of targeted functional groups. i) Acetylation To detect the presence (or absence) of a free amino terminus, a peptide is reacted with acetic anhydride on a probe tip for several minutes. An upward shift of (M+H)+ by 42u indicates the presence of a free N- 26 terminus. Another unique feature of this procedure is that any fragment ions derived from the N-terminus of the acetylated peptide are shifted upwards by 42u when compared to the CAD of the underivatized peptide. Because the e-NH2 group of lysine residues will partially acetylate by this procedure, care must be exercised. However, the 8-NH2 group can be targeted for acetylation to distinguish between lysine and glutamine because they have identical nominal mass. Hunt creatively applied an acetylation procedure to deduce the amino acid sequence of individual components of complex peptide mixtures, employing CAD-MS/MS with a triple quadrupole mass spectrometer [45,46]. Acetylation is carried out using an equimolar mixture of acetic anhydride and hexadeuterated acetic anhydride. The product ions containing an acetyl group appear as a visually conspicuous, equal-intensity doublet of peaks separated by 3 mass units when (N-acetyl-M+H)+ and (N -acetyl-d3-M+H)+ are allowed to pass through a 5u-window. Thus, N-terminal ions are easily distinguished from C-terminal ions. ii) Esterification Although this technique is not widely used, esterification can greatly reduce the detection limits of hydrophilic peptides. Falick and Maltby [47] capitalized on this derivatization technique to decrease detection limits to midpicomole levels, increasing the secondary ion yield produced with FAB by a factor of twenty-five or more. Although methyl [48] and isopropyl [49] esters were formerly used to determine the number of free carboxyl groups in a peptide, these investigators developed a simple, rapid derivatization procedure to improve sensitivity with FAB. Esterification was achieved by reacting peptide with 0.2M acetyl chloride in 5 [L of the desired alcohol for 1 h at 45°C. The resulting solution was lyophilized to remove excess 27 reactants, dissolved in 0.1% aqueous TFA. and added to matrix on a probe tip. The signal-to-background ratio increased for peptide esterification with alcohols of increasing alkyl chain length, although 1-octyl and longer chain alcohols were dificult to remove from the reaction mixture by evaporation. The longer chain alcohols increased the surface activity of peptides, and this characteristic was applied to a mixture of peptides of differing hydrophilicity to eliminate discrimination against hydrophilic peptides. However, peptides that contain more than one carboxyl group are not completely esterified with this derivatization procedure, producing a mixture of esterified products. iii) Amino acid detection Caprioli and Beckner [50] investigated reagents for the specific derivatization of peptide side chain groups. The use of site-specific reagents in combination with FAB-MS is a convenient method to confirm the presence of specific residues in peptides. FAB-MS is an ideal technique to determine the completeness of the derivatization reaction, assess the extent of modification for the group of interest, and detect side reactions in the derivatization procedure. The stability of the derivatized peptides also may be established by FAB-MS. Diethylpyrocarbonate (DEP) reacts with the imidazole nitrogen atoms of histidine to give the corresponding diethoxyformyl derivative [50]. Dissolution of 5 x 10'6 mol of peptide in 1 mL of 0.05M phosphate bufi'er (pH 6.5) and the addition of 1 .25 x 10'5M diethylpyrocarbonate in 95% ethanol causes both nitrogen atoms of the imidazole ring to be carboethoxylated after irreversible cleavage of the imidazole ring. However, DEP also has been used as a site-specific derivatizing reagent of free peptidyl amino groups to produce the corresponding ethoxyl derivative by Foti et al. [51]. 28 Consequently, peptides that contain a histidine residue and a free N- terminus display a triderivatized product. FAB-MS indicated that peptides with more than one amino group also produced multiply derivatized products [50], representing the addition of one or more ethoxyformyl groups to the peptide. Therefore, this derivatization procedure requires separation of N-terminally monoderivatized peptide from reagents and side products by HPLC. Because the reaction conditions required to produce the diethoxyformyl derivative of histidine more effectively modify the N- terminus (90% completion) [50], derivatization with DEP is commonly used for the latter purpose. BNPS-skatole oxidizes tryptophan residues to oxyindole and dioxyindole (or dioxyindolelactone) analogues using a two-fold or ten-fold excess of reagent, respectively [50]. A two-fold excess of 17.2mM reagent in 50% acetic acid reacted for 24 h at room temperature produces the oxyindole analog in a 20-30% yield with no dioxy analog. However, a ten- fold excess of BNPS-skatole causes formation of the dioxyderivatized peptide (25-35%) and 30-50% of oxyindole analogue. Side reactions observed in this derivatization procedure include cleavage of the peptide at the peptide bond. Several schemes to specifically modify tyrosine have been reported [50,51]. Tetranitromethane (TNM) nitrates tyrosine residues to form an o- nitrophenol analog. Ten micromoles of peptide dissolved in 1.4 mL of 0.001M Tris and treated with two equivalents of TNM dissolved in 95% ethanol yield o-nitrophenolic tyrosine in a slow reaction. Only a 15% yield of derivatized product was observed for an overnight reaction. However, formation of the dinitrophenolic species was not observed. Tyrosine- containing peptides treated with iodine rapidly react to form iodotyrosine 29 modified in the position ortho to the hydroxyl group. Five micromoles of peptide and 20 micromoles of ICl dissolved in 0.01M glycine buffer nearly quantitatively yielded the derivatized tyrosine residues in 18 hrs [50]. Non- radioactive iodination of tyrosine is also accomplished [52] with Iodogen® (Pierce Chem. 00.; Rockford, IL). However, diiodotyrosine analogues are observed as impurities in both procedures. Reaction of 101 with the imidazole group of histidine and the sulfhydryl of cysteine also are possible. Another scheme identifying the presence of tyrosine with an upward mass shift of 42u after acetylation of the phenolic group with N-acetylimidazole also has been reported [52]. Good yields of monoacetylated tyrosine- containing peptides were observed in one hour of reaction time. Treatment with hydroxylamine confirmed acetylation, because this reagent only reverses the acetylation of phenolic groups. However, amino and hydroxyl groups react with N-acetylimidazole after longer exposure to the reagent (24 h). Oxidation of methionine and an increase in mass of the protonated molecule by 16u confirms the presence of methionine in peptides. Treatment of a peptide with 40 uL of 1 % hydrogen peroxide solution quickly accomplishes oxidation and the resulting product can be directly analysed by FAB-MS. The thioether can be regenerated by reduction with b- mercaptoethanol or dithiothreitol. A derivatization reaction specifically modifying arginine with 1,2- cyclohexanedione to N7,N3-(1,2-dihydroxycyclohex-1,2-ylene)-L-arginine has been reported [53]. In the derivatization reaction, 20 nmol of peptide is reacted with 20 nmol of 1 ,2-cyclohexanedione in borate buffer (pH 8.5) for 2 h at 37°C. After evaporation of the solvent, approximately 1 nmol of peptide is applied to the probe tip for FAB-MS analysis. In the mass 30 spectrum, the peak representing underivatized peptide has a greater intensity than derivatized peptide in some cases, because the chemical modification is not complete. The peak due to derivatized peptide is shifted upward by 9411 for each arginine contained in the original peptide. The intensity of the peak representing completely derivatized peptide decreases with an increasing number of arginines, indicating the derivatization reaction is not complete or the derivatized peptide has a poorer ionization efficiency than the underivatized peptide. 8) Miscellaneous Banner and Spiteller reported the acylation of N-terminally free peptides to form dansyl derivatives [54,55]. Acylation converts the free N - terminal amine to an amide nitrogen, decreasing its basicity and increasing the tendency for protonation along the peptide backbone. N-terminal fragment ions dominate the mass spectra of dansylated peptides; a complete bu series was observed for a dansylated hexapeptide, while little sequence information was apparent from the underivatized peptide. This derivative greatly increases the sensitivity of ultraviolet detection for HPLC and if 2-bromo-5-(dimethylamino)benzenesulfonyl chloride is used for derivatization; the N-terminal fragments are easily detected because they display the “goalpost” isotope pattern of peaks characteristic of bromine separated by 2u. Soon after the introduction of FAB, this ionization technique was employed to identify and quanitate dansyl amino acids obtained in the N- terminal analysis of proteins. The investigated protein was reacted with dansyl chloride and then hydrolyzed to yield the dansylated amino acid for FAB analysis [56]. The dansylation procedure is an involved, time- consuming process. The protonated N-terminally derivatized amino acid is 31 the primary species observed, although fragments due to cleavage of the dansyl group (m/z 170) and amino acid side chain are also observed. The response of the dansyl amino acids is a linear function of concentration up to 20 nmol/uL, with the limit of detection being approximately 100 pmol. This procedure does not differentiate between isomeric leucine and isoleucine, but isobaric lysine and glutamine can be distinguished because glutamine is hydrolyzed to glutamic acid during peptide hydrolysis. The accuracy of the procedure was substantiated with N-terminal analysis of a series of known peptides. To increase the amount of sequence information from FAB-MS spectra, peptides converted to polyamino alcohols were analyzed by tandem FAB-MS [57]. The peptide YAGFL was reduced using diborane. In this procedure, the amide groups are converted to amines and the carboxylic acid group is reduced to an alcohol. Analysis of the reduced peptide by FABoCAD-MS/MS produced complete b... y... and 2,. ion series, which were incomplete before reduction. Therefore, the simplicity of spectral interpretation was increased markedly with reduction of peptides to polyamino alcohols. The t-butyloxycarbonyl (BOC) protecting group was also investigated as a peptide derivatizing agent [58]. The loss of C4H3 and 002 ofien is observed with FAB-MS analysis of BOC-protected peptides. Neutral loss of the BOC group and water [59] is also observed in peptides containing a BOC protecting group. Moreover, the (M+H)+ signal intensity of BOC- protected peptides is less than the protonated molecule of underivatized peptides. Due to the facile fragmentation of the t-butyloxycarbonyl group, the CAD spectra of the derivatized peptides are more complex and yield less sequence information than the underivatized peptides. Consequently, more 32 descriptive mass spectra result when the BOC promoting group is removed after peptide synthesis. N-benzyloxycarbonyl-protected tripeptide ethyl esters containing proline have been investigated by negative ion FAB-MS. The results implied that conformational differences due to the position and number of proline residues may influence the fragmentation of tripeptides by FAB- CAD-MS/MS in the negative ion mode. The fragmentation patterns of the derivatized peptides varied significantly, particularly the intensities of the fragment ions formed by cleavage of the benzyloxycarbonyl group, depending on the number and position of proline residues in the derivatized molecules. However, the fragmentation patterns of the derivatized peptides were nearly identical in the positive ion mode. From their observations of the fragmentation pattern of the N-benzyloxycarbonyl group in model tripeptides, these investigators [60] have been able to empirically predict the numbers and positions of proline residues in N-benzyloxycarbonyl- protected peptide ethyl esters. To develop a method to detect carcinogenic N-nitroso peptides in gastric contents, Wait et al. [61] nitrosated a number of synthetic peptides for examination by FAB-MS. The peptides were nitrosated with N02 in dichloromethane and the N -nitroso peptides were readily desorbed under positive FAB conditions. The structures of the fully nitrosated peptides (N- terminus and peptide bonds) were unambiguous, but when a single NNO group was introduced to the molecule, the product ion spectra produced by FAB-CAD-B/E displayed a dominant loss of 30u (N 0 group). Consequently, the site of introduction of a single NNO group could not be determined from the spectra. 33 Backbone-modified peptides cannot be sequenced by normal techniques, such as Edman degradation, maln'ng tandem FAB-MS a critical technique for determining the amino acid sequence of these peptides. Several types of backbone-modified peptides were synthesized and analyzed by FAB-CAD-MS/MS [62]. The normal peptide bond in the modified peptides was replaced with thiomethylene ether (Cst), thioamide (CSNH), methyleneamine (CHzNH), and thiomethylene sulfoxide (CstO) groups. With CAD, peptides containing thiomethylene and methyleneamine groups predominantly produced C-terminal ions. The thiomethylene sulfoxide modified peptides, on the other hand, produced both N- and C-terminal fragment ions. However, the thioamide group causes only slight difi‘erences in fragmentation in comparison to peptides containing the unmodified amide bond. These investigations showed the presence and location of peptide backbone modification is quickly determined by the fragmentation pattern produced with tandem FAB-MS. D. Research Objectives The primary goal of this research is to develop techniques for derivatization with the triphenylphosphonium group to enhance the analysis of peptides by FAB and FAB-CAD-MS/MS. Because several reagents are commercially available or have been synthesized in-house to perform TPP derivatization, the effectiveness of these derivatizing reagents will be described. Usually small (low picomole) amounts of peptides are available for analysis, and to eliminate sample loss due to transfer, a second goal of this research is to develop a method to derivatize peptides with the ethyl-TPP group on-line. The feasibility of an on-line, ethyl-TPP derivatization procedure will be investigated and described for a homogeneous solution. Finally, firture projects, including the derivatization 34 of a glycopeptide, will be described. Covalently fixing a charge to the peptide portion of a glycopeptide should cause selective fragmentation of the peptide moiety for increased structural information and decreased fragmentation of the carbohydrate moiety, which normally preferentially fragments. The criteria used to evaluate the effectiveness of the TPP derivatizing reagents are the speed and completeness of the derivatizing reaction, the improvement in ionization efficiency, and the efficiency and simplification of fragmentation that results from derivatization. Other criteria used to evaluate the derivatizing reagents include the ease of the modification reaction and the number and complexity of the manipulations necessary to accomplish the derivatization procedure. E. References 1. Dell, A.; Rogers, M.E., Trends. Anal. Chem., 1989, 800), 375-378. 2. Greer, F., Lab. Pract., 1989, 38(10), 81 -2, 84-5. 3. Gibson, B.W.; Biemann, K., Proc. Natl. Acad. Sci. USA, 1984, 81, 1956-1960. 4. Barber, M.; Bordoli, R.S.; Sedgewick, R.D.; Tyler, A.N., J. Chem. Soc. Chem. Commun., 1981, 7, 325. 5. Morris, H.R.; Panico, M.; Barber, M.; Bardoli, R.S.; Sedgwick, R.D.; Tyler, A., Biochem. Biophys. Res. Comm., 1981, 101, 623-631 . 6. Biemann, K.; Gapp, F.; Seibl, J ., J. Am. Chem. Soc., 1959, 81, 2274- 2275. 7. Biemann, K., Anal. Chem., 1986, 88, 1288A-1300A. 8. Barber, M.; Bordoli, R.S.; Elliott, G.J.; Sedgwick, D.; Tyler, A.N., Anal. Chem., 1982, 54( 2), 645A-657A. 9. Gross, M.L.; Hayes, R.N., Methods Enzymol., 1990, 193, 237-263. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 2‘3 22. 25. 35 Carr, S.A.; Hemling, M.E.; Bean, M.F.; Roberts, G.D., Anal. Chem., 1991, 63, 2802-2824. Roepstorff, P.; Fohlmann, J ., Biomed. Mass Spectrom., 1984, 11, 601. Biemann, K., Methods Enzymol., 1990, 193, Appendix 5, 886-887. Biemann, K., Methods Enzymol., 1990, 193, 455-479. Johnson, R.S.; Martin, S.A.; Biemann, K; Stults, J .T.; Watson, J .T., Anal. Chem., 1987, 59, 2621-2625. Unger, S.E.; Ryan, T.M.; Cooks, R.G., Analytica Chimica Acta, 1980, 118, 169-174. Busch, K.L.; Unger, S.E.; Vineze, A.; Cooks, R.G.; Keough, T., J. Am. Chem. Soc., 1982, 104, 1507-1511. Busch, KL.; Cooks, R.G., Science (Washington, D. C.), 1982, 218, 247- 254. Flurer, R.A.; Busch, K.L., Pmc. Indiana Acad. Sci, 1985 (Pub. 1986). 95, 171 -5. DiDonato, G.C.; Busch, KL., Biomed. Mass Spectrom., 1985, 12(8), 364-366. Knapp, D.R., Methods Enzymol., 1990, 193, 314-329. Allmaier, G.M., Rapid Commun. Mass Spectrom., 1988, 2(4), 74-76. Ligon, W.V. J r., Anal. Chem., 1986, 58(2). 485-487. Clench, M.R.; Garner, G.V.; Gordon, D.B.; Barber, M., Biomed. Mass Spectrom., 1985, 12(7), 355-357. Ligon, W.V.; Dorn, S.B.; Int. J. Mass Spectrom. Ion Proc. 1984, 57, 75- 90. Kidwell, D.A.; Ross, M.M.; Colton, R.J., J. Am. Chem. Soc., 1984, 106, 2219-2220. 27. 29. 30. 31. 32. 33. 35. 36. 37. 38. 36 Foti, S.; Saletti, R., Biomed. Environ. Mass Spectrom. 1989, 18, 168- 173. Foti, S.; Saletti, R.; Marletta, D., Org. Mass Spectrom., 1991, 26, 903- 907. Johnson, R.S.; Martin, S.A.; Biemann, K., Int. J. Mass Spectrom. Ion Proc., 1988, 86, 137-154. Shiea, J .; Sunner, J ., Int. J. Mass Spectrom. Ion Proc., 1991, 109, 262- 278. Johnson, R.S.; Biemann, K., Proc. 35th ASMS Conf. Mass Spectrom. Allied Topics, 1987, Denver, CO., 556-557. Johnson, R.S.; Martin, S.H.; Biemann, K., Int. J. Mass Spectrom. Ion Proc., 1988, 86, 137-154. Stults, J .T.; Halualani, R.; Wetzel, R., Proc. 37th ASMS Conf. Mass Spectrom. Allied Topics, 1989, Miami Beach, FL., 856-857. Vath, J .E.; Zollinger, M.; Biemann, K, Fresenius Z. Anal. Chem., 1988, 331, 248-252. Vath, J .E.; Biemann, K., Int. J. Mass Spectrom. and Ian Proc., 1990, 100, 287-299. Bennett, B.D.; Day, R.A., Proc. 35th ASMS Conf. Mass Spectrom. Allied Topics, 1987, Denver, CO., 568-569. Wagner, D.S.; Salari, A.; Gage, D.A.; Leykam, J .; Fetter, J .; Hollingsworth, R.; Watson, J .T., Biol. Mass Spectrom., 1991, 20, 419- 425. Watson, J.T.; Wagner, D.S.; Chang, Y.-S.; Strahler, J .R.; Hanash, S.M.; Gage, D.A., Int. J. Mass Spectrom. Ion Proc., 1991, 111, 191 -209. Teesch, L.M.; Orlando, R.C.; Adams, J ., J. Am. Chem. Soc., 1991, 113, 3668-3675. 39. 41 . 42. $3: 47. . Kausler, W.; Schneider, K.; Spiteller, G., Biomed. Environ. Mass 49. 51. 52. 53. 37 J ankowski, K.; Virelizier, H.; Lafontaine, P.; Pare, J .R.J ., Spectrosc. Lett., 1991, 24(1), 35-41. . Jankowski, K.; LeBlanc, J .-F.; Virelizier, H.; Tabet, J .-C.; Pare, J .R.J ., Spectros. Int. J., 1989, 1, 49-56. Chai, W.-G.; Yan, L.; Wang, G.-H.; Liang, X.-Y.; Zhao, Y.-F.; J ., Gai- Jiao, Biomed. Environ. Mass Spectrom., 1987 , 14, 331 -333. Zhao, Y.-F.; Zhang, D.-Q.; Xue, C.-B.; Zeng, J .-N .; J ., Gai-Jiao, Org. Mass Spectrom., 1991, 26, 510-513. Zhao, Y.-F.; Zhang, D.-Q.; Xue, C.-B., Int. J. Protein Res., 1991, 37, 457- 461. Ma, X.-B.; Zhao, Y.-F., Biol. Mass Spectrom., 1991, 20, 498-502. Hunt, D.F.; Bone, W.M.; Shabanowitz, J .; Rhodes, J .; Ballard, J .M., Anal. Chem., 1981, 53, 1706-1708. Hunt, D.F.; Buko, A.M.; Ballard, J .M.; Shabanowitz, J .; Giordani, A.B., Biomed. Mass Spectrom., 1981, 8(9), 397-408. Falick, A.M.; Maltby, D.A., Anal. Biochem., 1989, 182, 165-169. Spectrom., 1988, 17, 15-19. Naylor, S.; Findeis, A.F.; Gibson, B.W.; Williams, D.H., J. Am. Chem. Soc., 1986, 108, 6359-6363. Caprioli, R.M.; Beckner, C.F., GC-MS News, 1984, 12(6), 152-159. Foti, S.; Marletta, D.; Saletti, R.; Petrone, G., Biomed. Environ. Mass Spectrom., 1991, 20, 336-339. Silberring, J .; Nyberg, F., J. Chromatogr., 1991, 562, 459-467. Silberring, J .; Nyberg, F., Biomed. Environ. Mass Spectrom., 1990, 19, 819-821. Renner, D.; Spiteller, G., Angew. Chem., 1985, 97, 408-410. 55. 57. 59. 62. 38 Renner, D.; Spiteller, G., Biomed. Environ. Mass Spectrom., 1986, 13, 405-410. Beckner, C.F.; Caprioli, R.M., Anal. Biochem., 1983, 130, 328-333. Lippstreu-Fisher, D.L.; Gross, M.L., Anal. Chem., 1985, 57, 1174-1180. Bathelt, E.R.; Heerma, W., Biomed. Environ. Mass Spectrom. , 1987 , 14, 53-61. Jankowsln', K.; Gaudin, D.; Pham Van Chuong, P.; Spectrosc. Lett. , 1987, 20(3), 255-284. Tsunematsu, H.; Nakashima, S.; Yoshida, S.; Yamamoto, M.; Isobe, R., Org. Mass Spectrom., 1991, 26, 147-150. . Wait, R.; Leach, S.A.; Hill, M.J.; Thompson, M.H., Biochem. Soc. Trans, 1988, 16, 740-744. Deterding, L.J.; Tomer, KB; Spatola, A.F., J. Am. Soc. Mass Spectrom., 1990, 1, 174-182. 39 Chapter II 'h'iphenylphosphonium Derivatizationand the SeparationotComponentsofflreReactionMixtme A. Introduction Peptide derivatization at the N-terminus with (2-bromoethyl)TPP and at the C-terminus with (2-aminoethyl)TPP has been performed in this laboratory [1 -5]. The ethyl-TPP group significantly increases the hydrophobic character of the derivatized peptides, enhancing their ionization efficiency and increasing their signal intensity during FAB- MS. The localized charge of the quaternized phosphorus atom in the ethyl-TPP group directs fragmentation by remote site fragmentation from the terminus of derivatization [6], simplifying the fragmentation pattern produced by CAD-MS/MS, which otherwise produces fragments from both termini of an underivatized peptide. These benefits of derivatization with the triphenylphosphonium group warranted the derivatization of peptides with other reagents containing the TPP group. The signal enhancement produced with derivatization depends on the extent of derivatization and the properties of the modifying group. Consequently, the amount of signal enhancement due to derivatization can be increased by employing derivatizing reagents with optimal properties for the analytical technique employed and by manipulating the reaction conditions to improve the reaction yield. A number of reagents containing the triphenylphosphonium group are commercially available or have been synthesized in this laboratory. The 4 0 compound (1 -bromomethyl)TPP contains one methylene group in contrast to the two methylene groups of (2-bromoethyl)TPP. Commercially available compounds include (3-bromopropyl)TPP and (4-bromobutyl) TPP, which have an increasing alkyl chain length. The 2-(triphenyl- phosphonio)ethoxycarbonyl chloride reagent (the so-called “Kunz” reagent) is a reactive chloroformate. In addition, (4-carboxybenzyl)TPP was synthesized in-house to modify the N-terminus of peptides with the coupling action of a carbodiimide. The derivatization procedures used to modify peptides with these magenta will be described in this chapter. The derivatization of VGVAPG and the fragmentation patterns observed for the modified peptides by FAB-CAD-MS/MS by linked scanning will be explained in this chapter. The C-terminal derivatization of leucine enkephalin (YGGFL) with (2-aminoethyl)TPP, separation of the components of the reaction medium by HPLC, and ensuing quantitation of the derivatized product by UV-Vis detection in combination with HPLC chromatography will be related. Table 2.1. Investigated derivatizing reagents containing the TPP group. Beam Scum (2-Aminoethyl)TPP In-house synthesis (1 -Bromomethyl)TPP bromide Lancaster Synthesis Ltd. (2-Bromoethyl)TPP bromide In-house synthesis (2-Chloroethyl)TPP bromide Lancaster Synthesis Ltd. Vinyl-TPP bromide Lancaster Synthesis Ltd. “Kunz” reagent FlukaChemie AG (3-Bromopropyl)TPP bromide Lancaster Synthesis (4-Bromobutyl)TPP bromide Lancaster Synthesis (4-Carboxybenzyl)TPP In-house synthesis 4 l B. Synthesis of reagents containing the TPP group 1. (2-Bromoethyl)TPP a) Synthesis by Method I i) Experimental The method of Frederich and Henning [7] was used to synthesize (2-bromoethyl)TPP from 1 ,2-dibromoethane and triphenylphosphine in a simple substitution reaction (Reaction 2-1). To 40 mL of xylene, 1.00 g of (2-1) ©9/B}2—CH2—Br ——>- @PLmz—an—Br '4’ Br. triphenylphosphine was added (3.81 x 10'3 mol) along with a hundred-fold molar excess of 1,2-dibromoethane (33.0 mL). The mixture was heated to 50°C and stirred overnight. The desired product precipitated out of the reaction solution and was collected on filter paper with suction filtration. The white precipitate was washed with diethylether and dried for use as a peptide derivatizing reagent. ii) Results and discussion One should note that peptides derivatized with a TPP group and the derivatizing reagents containing a TPP group do not represent a molecular ion according to a strict definition of the term, but a cation of an ionic compound. Therefore, when a solution containing (2- bromoethyl)TPP is analyzed by FAB-MS in the positive mode, the (2- bromoethyl)TPP cation is desorbed (m/z 369, 371), but the compound also contains a bromide anion, and the proper name of the compound is (2- bromoethyl)TPP bromide. 4 2 Analysis of the reaction product by FAB-MS indicated that the product contained (2-bromoethyl)TPP as represented by the bromine isotope peaks of the molecular ion (M+) at m/z 369 and 371 (Figure 2.1 ). The product contains a significant amount of vinyl-TPP impurity (Om/2 289) and some (2-diphenylphosphinoethyl)TPP (m/z 475) salt as a side product. Direct analysis of the reaction product by HPLC confirmed the presence of a significant amount of vinyl-TPP impurity. The chromatogram of the product mixture (Figure 2.2) indicates two main components are contained in the product. The fractions represented by the peaks were collected and analyzed by FAB-MS. The first fraction contained vinyl-TPP, while the second peak represented (2- bromoethleI‘PP. Because this method of synthesizing (2-bromoethyl)TPP also produced a vinyl-TPP impurity and the unpurified reaction mixture was used for peptide derivatization, some necessary questions arose: which compound was derivatizing the N-terminus of the peptides, and if both reagents could react with a peptide, which reacted more efficiently? To answer these questions, isolation of the individual compounds by flash column chromatography was attempted [8]. b) Thin-layer chromatography i) Experimental The product synthesized by Method I was analyzed by thin-layer chromatography (TLC) prior to separation of the mixture by flash column chromatography. TLC separation was performed with Whatman‘” (Whatman Paper Ltd.; Maidstone, Kent, England) premade TLC plates that consist of a flexible aluminum backing coated with 250 pm of silica gel. The plates were cut into 2.5 x 10 cm strips and spotted by micropipet with solution containing the product to be analyzed. The sample spots 43 1:00: Vinyl-TPP" (2-Bromoethyl)TPP“' e i l 1 a 80‘ t . i I v 4 0 60‘ 1 I n . t 40. e . n . s . i 20* TPP“ y i «i. .. 18 10° 20° 300 460' . Y 500 Figure 2.1. The FAB-MS spectrum of the (2-bromoethyl)TPP product synthesized by reaction of 1 ,2-dibromoethane with triphenylphosphine. The molecular ion (M+) of (2-bromoethyl)TPP is represented by the doublet at m/z 369 and 371; the vinyl-TPP and (2-diphenylphosphinoethyl)TPP impurities appear at m/z 289 and m/z 475, respectively. Nitrobenzyl alcohol (NBA)wasusedastheFABmatr-ix. 44 1.0- a. 0.75- E: ‘8 A 3 ' O i, a E 0.5- «a 8 .2. c3 4. E“ "'8. 0.25- ,5 ' ;> i 0 I' F4 r l Tweet 8 12 16 20 24 23 Time (min) Figure 2.2. HPLC chromatogram of the (2-bromoethyl)TPP product mixture synthesized by reaction of TPP with 1,2-dibromoethane. For this separation, approximately 1 00 nmol of product mixture was injected. The parameters used in the separation are: mobile phase, 45:55 (v/v) acetonitrile/water; flow rate, 1 mL’min; column, 4.6 x 250 mm; stationary phase, 5 um diameter, 013 bonded phase; detector, 214 nm, 1.0 AUFS; chart speed, 20 cm/hr. 4 5 were 0.2 cm in diameter and made 1.0 cm above the bottom of the TLC plate. For sample development, the spotted plate was placed in a solvent chamber containing the mobile phase at a depth of 0.5 cm. The plate was removed from the chamber when the mobile phase migrated to within approximately 1 .0 cm of the top of the plate. The plate was allowed to air- dry in a hood. When dry, the plate was observed in a viewing chamber by fluorescence at 254 nm. The positions of the separated compounds appeared as purple spots on the TLC plate. The positions of the spots were marked and their Rf values were calculated. Several solvent combinations were investigated for optimal separation, and when a mobile phase produced a desired separation of the mixture, all the fractions collected by flash column chromatrography were analyzed by TLC using this solvent combination. ii) Results and discussion Optimal separation by flash column chromatography of a mixture is accomplished in a reasonable length of time when the components have Rfvalues between 0.35 and 0.50 and are separated by an Rfvalue of 0.1 . A methanol/chloroform 14:100 (v/v) mixture provided the desired separation of the reaction product obtained by Method 1. The lower spot resulting from TLC separation represents (2-methoxyethyl)TPP (m/z 321 , Figure 2.3 a)). The upper spot represents vinyl-TPP and (2- bromoethyl)TPP (m/z 289 and 369, 371, respectively; Figure 2.3 b)). The (2- methoxyethyl)TPP is a side product resulting from the Michael addition reaction of methanol to vinyl-TPP [9-11]. Separation of the product synthesized by Method I in a 1:1 (v/v) acetonitrile/chloroform solvent combination produced a single spot with an Rf of 0.42. No side products were observed with use of this mobile phase. However, no solvent 46 1100: (2-Methoxyetr.hyl)’I‘PP+ , . (a) 1 . 880‘ i 1 v ‘ 0 860- Vrnyl-TPP" 1 . n . t 40. ° i n I ‘ ‘ r :20- y TPP" 100 200 ml 300 400 500 2 1%00‘ . . (b) mel-TPP“ 1 . 1180‘ .t 4 r I (2-Bromoethyl)TPP"' v 1 O 60‘ _ I l n . t 40. e . n . s . {so TPP” ’ l l... . in. .8. .. . . .4. 100 200 300 400 500 m/z Figure 2.3. The FAB-MS spectra of the components of the (2-bromo- ethyl)TPP reaction mixture prepared by the method of Fredrich and Henning as separated by flash column chromatography using a 14:100 (v/v) methanol/chloroform mobile phase: a) (2-methoxyethyl) TPP corresponding to the lower TLC spot and last-eluting compound when separated by flash column chromatography. b) vinyl-TPP and (2-bromoethyl)TPP that were the first-eluting compounds during separation by flash column chromatography and was represented by the upper spot after separation by TLC . NBA was the FAB matrix used for both spectra. 4 7 combination was discovered that could resolve vinyl-TPP and (2- bromoethyl)TPP using TLC plates coated with silica gel. c) Flash column chromatography i) Experimental Separation of the product produced by Method I was performed with flash column chromatography. A 2-inch diameter column with a stopcock and flow controller was filled with six inches of 200-300-um diameter silica gel (Sigma Chem. Co.; St. Louis, MO). The column was filled with the 14:100 (v/v) methanol/methylene chloride solvent mixture that produced the TLC separation and pressure was applied to rapidly push all air from the silica gel. One gram of the (2-bromoethyl)TPP product mixture dissolved in this solvent mixture was applied to the column and the flow controller was briefly placed on the column to push the sample into the silica gel. The column was refilled with solvent and eluted at a flow rate of 2 inches per minute (approximately 75 mL/min). A rack containing forty 20 x 150 mm test tubes was used to collect the eluted fractions. Small fractions were collected at the start of the run and larger fractions were collected near the end of the run. Composition of the collected fractions was determined by TLC and the fractions containing the compounds corresponding to the lower and upper TLC spots were concentrated by rotoevaporation. ii) Results and discussion Separation of the product produced by Method I by flash-column chromatography provided two fractions. The first-eluting fraction contained (2-bromoethyl)TPP and vinyl-TPP. The second fraction contained the (2-methoxyethyl)TPP side product. The side product results 4 8 from the Michael addition of methanol to vinyl-TPP impurity in the reaction product. To avoid formation of the (2-methoxyethyl)TPP side product, flash column chromatography was repeated in a 1:1 (v/v) acetonitrile/ chloroform mixture and a single fraction containing a mixture of vinyl- TPP and (2-bromoethyl)TPP resulted, as investigation with FAB-MS and separation by HPLC demonstrated. No Michael addition products were observed when the reagents were isolated with these nonhydroxylic solvents. Although separation of vinyl-TPP and (2-bromoethyl)TPP with combinations of other solvents that do not contain hydroxyl groups was attempted, satisfactory separation of the reaction mixture was not achieved. Consequently, an alternative method to produce pure (2- bromoethyl)TPP was pursued. d) Synthesis by Method 11 i) Experimental Nearly pure (2-bromoethyl)TPP was synthesized by reaction of (2- hydroxyethyl)TPP with phosphonium tribromide (Reaction 2-2) [1 2,1 3]. A as .(Q—g-m’flz—m) . Pan ——> s(©—§-Gz-mz—Br) + o=aoan mixture of 5.00 g (0.0129 mol) of (2-hydroxyethyl)TPP and 5.0 mL (0.053 mol) of tribromophosphine were heated on a steam bath for 2 h with occasional shaking. A viscous orange syrup resulted, which was suspended in 100 mL of H20 and extracted with chloroform. The chloroform fractions were washed with H20 and dried over N a2804. The chloroform was removed by rotoevaporation, yielding yellow crystals. 4 9 These crystals were dissolved in acetonitrile with slight heating and recrystallized to yield clear, white crystals. This procedure yielded 4.89 g (0.0109 mol) of (2-bromoethyl)TPP; this is a 84.5% yield of this product. ii) Results and discussion Separation of the reaction product by HPLC and analysis by FAB- MS indicated nearly pure (2-bromoethyl)TPP was produced by this method. Because vinyl-TPP and (2-bromoethyl)TPP have identical extinction coefficients, HPLC indicates that the product is 95% (2- bromoethyl)TPP and 5% vinyl-TPP (Figure 2.4). FAB-MS also indicated a highly pure product of (2-bromoethyl)TPP with slight vinyl-TPP impurities was produced by this method (Figure 2.5). 2. (2-Aminoethyl)TPP a) Experimental The synthesis of high purity (2-aminoethyl)TPP from (2- bromoethyl)TPP now was possible. To a stirred solution of acetonitrile 1.50 g (0.00333 mol) of (2-bromoethyl)TPP was added. When the (2- bromoethyl)TPP dissolved, 12 mL of 27% ammonium hydroxide (~ 15-fold molar excess) was added dropwise to the reaction mixture. The solution was heated to 40°C and stirred for 1 hr. The acetonitrile and excess ammonium hydroxide were removed by rotoevaporation. White crystals of solid (2-aminoethyl)TPP remained in the round-bottomed flask. This solid was used directly for C-terminal derivatization after being dissolved to a known concentration in methanol. b) Results and discussion The reaction used to produce (2-aminoethyl)TPP is outlined in Reaction 2-3. Separation of (2-aminoethyl)TPP synthesized by this method 50 1.0 - + D-I 0.75 .. E: 3’ A 8 8 u m 93 0.5 - 0.25 - 4.: + c: D-I 3 E? H F! 5 .E‘ .2. > a . 0 I I I *ILIJ I I I I trmec't 6 12 18 24 30 Time (min) Figure 2.4. HPLC separation of the (2-bromoethyl)TPP product synthes- ized by reaction of (2-hydroxyethyl)TPP with PBr3. Approximately 1 0 nmol of product was injected for separation. The parameters used in the separation are: mobile phase, 45:55 (v/v) acetonitrile/water; flow rate, 2 mL/min; column, 10 x 250 mm; stationary phase, 10 um diameter, 013 bonded phase; detector, 214 nm, 0.5 AUFS; chart speed, 20 cm/hr. 51 100‘ 3 i (2-Bromoethyl)’l‘PP+ 1 . a 80' .t . 1 4 v . O 60" 1 1 n . t 40. e . g Vinyl-TPP" ; 20‘ ‘ y ‘ TPP" T» u.+-;~L~TL“:‘LJ.L:J1 , t . in ,4- . .krl , A. . . . 100 200 300 400 500 m/z Figure 2.5. Analysis of the product produced by reaction of (2-hydroxy- ethyl)TPP with PBr3 by FAB-MS. The peaks representing the molecular ion (M+) of (2-bromoethyl)TPP dominate the FAB-MS spectrum (m/z 369, 371). Vinyl-TPP impurity appears at m/z 289. The FAB matrix was NBA. 52 Q + + (2-3) ©P—cr12—CH2_Br + NH. —> @P—CHg—CHg—NH. + HBr by HPLC with subsequent UV-Vis detection, indicated that nearly pure product was produced. The chromatogram of the reaction mixture appears in Figure 2.6. Because the absorption coefficients of (2-amino- ethyl)TPP and vinyl-TPP impurity are nearly identical, the product was calculated to be 92% pure, based on the area of peaks in the chromatogram. The analysis of product by FAB-MS also indicated the product was of high purity (Figure 2.7). 3. (4-Carboxybenzyl)TPP a) Experimental To prepare the (4-carboxybenzyl)TPP derivatizing reagent, 8 1 :1 molar ratio of triphenylphosphine was reacted with 4-(chloromethyl) benzoic acid. The reaction was performed in 300 mL of xylene; 1.54 g (1.18 x 10’2 mol) of triphenylphosphine and 1.00 g (1.18 x 10'2 mol) of 4- (chloromethyl)benzoic acid were dissolved in this solvent with stirring. The reaction was performed with gentle reflux at 144°C for 24 hours. Analysis of the reaction mixture by TLC after reflux demonstrated that a high yield of (4-carboxybenzyl)TPP was present and a small amount of 4- (chloromethyl)benzoic acid and some TPP impurity remained in the reaction mixture. The TLC separation was achieved in a 4:1 (v/v) acetone/chloroform solvent mixture. After cooling, xylene was removed by rotoevaporation until a light brown syrup remained. Further drying of 53 0.5- r 4. E5 ‘5. fl 4) 0.375- g a A £3 '55 w 8 8 . N a l E 0.25- 93 4.3 a 0.125— ‘E + n. . 5 E: ,2, . r3 '8 S 0 r HI I. r I meet 8 12 16 2o 24 Time (min) Figure 2.6. The chromatogram resulting from the separation of the (2- aminoethyl)TPP product by HPLC. The first on-scale peak represents the (2-aminoethyl)TPP product; the second peak represents vinyl-TPP impurity in the product. For this separation, approximately 200 nmol of product was injected and the parameters used in the separation are: mobile phase, 45:55 (v/v) acetonitrile/water; flow rate, 2 mL/min; column, 10 x 250 mm; stationary phase, 5 um diameter, 013 bonded phase; detector, 214 nm, 0.5 AUFS; chart speed, 20 cm/hr. 54 -8 (2-Aminoethyl)TPP+ 1 R e 1 l , a 80‘ .t 4 1 1 V e sol I i n J t 40. 0 w n 4 I . :20. Vinyl-TPP y 3 ‘ ‘ A} A , . . . A. i, . . q) 100 200 ml 300 400 600 z Figure 2.7. The spectrum resulting from the analysis of the (2-aminoethyl) TPP product mixture by FAB-MS. The molecular ion (M+) of (2- aminoethyl)TPP dominates the spectrum (m/z 306), but some vinyl-TPP is observed due to impurity and fragmentation (m/z 289). The peaks at m/z 136 and 154 are characteristic of the NBA matrix used. 5 5 the product was performed under vacuum until a solid product formed. The yield of product was 3.09 g (60.6 %). b) Structural investigations i) Analysis by FAB-MS The reaction used to produce (4-carboxybenzyl)TPP is described in Reaction 2-4. Although analysis by FAB-MS revealed the expected © . o ‘2'" Gig + a—Cflz‘Q’g"°" —_> ©6CH2-<>—CI=I—0H + cf (<5 molecular ion (M+) for (4-carboxybenzyl)TPP at m/z 397 with ~ 10% relative intensity, a base peak at m/z 279 was observed (Figure 2.8). The peak at m/z 279 represents the protonated triphenylphosphine oxide molecule. The ion observed at m/z 279 probably is formed in the FAB selvedge region due to the high-energy conditions present in and above the FAB-MS matrix. Analysis of the (4ecarboxybenzyl)TPP product by FAB-CAD-B/E produced the fragmentation pattern appearing in Figure 2.9. The ion series at m/z 108, 183, 185, and 262 is characteristic of a compound containing triphenylphosphine. The peaks at m/z 307, 320, 352, 369, and 381 support the identity of the product and fragment at the positions described in the inset of Figure 2.9. ii) Analysis by NMR The identity of the (4-carboxybenzyl)TPP product was investigated by NMR, which indicated the product was pure (4-carboxybenzyl)TPP. Deuterated chloroform and d4-methanol were used for analysis by NMR and the resulting spectra can be observed in Figures 2.10 a) and b), respectively. 56 -3 ('l‘l:'Poxlde+I-I)+ A l A A A J A A A l A A %fifl'flflOHuI-U Q<flofipma ”H '° 8 O I l I ‘ H (It-carbixybenzyll'EPP" (i) 4“ A 1* *fl . , 4. . Tr . 100 200 300 . 400 500 m/z Figure 2.8. Analysis of the (4-carboxybenzyl)TPP product by FAB-MS that was produced by reaction of TPP with 4-(chloromethyl)benzoic acid. The peak representing the molecular ion (M+) of the (4-carboxybenzy1)TPP product is at m/z 397. The base peak in this spectrum represents protonated triphenylphosphine oxide and the peak at m/z 201 represents diphenylphosphine oxide. The FAB matrix was glycerol. ‘8 (30" w H . 9." I-I °--. 9'--- $“DIBG95H O<""¢*fl"‘0 (9+ - 50 100 150 200 250 300 350 400 m/z Figure 2.9. The spectrum resulting from analysis of the (4-carboxybenzyl) TPP product by FAB-CAD-B/E. The fragmentation observed is related to the structure of the product in the inset. The FAB matrix was NBA. 57 (a) JJLU, .JJ L V V I I T V I I V Y Y ' ' V ' ' ‘ ' Y V V I U ‘ ‘ 4 d J d ‘ 8 7 6 5 3 2 I 0 DO. l 1“ JL M [J.n A 44L l r 1 U j I ' ‘ 1 l ' ' ' ' I I I I 7 6 5 4 3 2 I on. Figure 2.10. Analysis of the (4-carboxybenzyl)TPP product by NMR when dissolved in a) deuterated chloroform and in b) d4-methanol. 5 8 The NMR spectra for (4-carboxybenzyl)TPP dissolved in the two unlike solvents differ markedly. The analysis by NMR of the sample dissolved in deuterated chloroform appears in Figure 2.10 a). The three signals furthest downfield (7.85, 7.80, and 7.70 ppm) represent the hydrogens contained in the phenyl rings of the triphenylphosphonium moiety. The peak at 7.24 ppm is caused by non-deuterated chloroform impurities in the solvent. The phenyl hydrogens ortho to the phosphorus atom are expected to cause the signal farthest downfield because they are deshielded by resonance with the charged phosphorus atom (Scheme 2-1). A resonance structure also can be drawn for the hydrogens para to the phosphorus atom, (Scheme 22), but their signal is not shifted downfield I \ Scheme 2-1 R-+P—R <-—> R _ P‘R I l R R Scheme 2'2 R-+P—R <-——> R—P—R l l R R as much as that of the hydrogens in the ortho position because the loss of electron density by induction to the positively charged phosphorus atom is not as strong. The peak at 7.70 ppm represents the hydrogens meta to the phosphorus atom. These hydrogens have the least shift downfield because no electron density is shifted to the charged phosphorous by resonance. The quintet at 4.7 ppm and the twin triplets centered at 3.8 5 9 ppm are due to the geminal methylene hydrogens. The large coupling constant observed for these signals is thought to be caused by the dimerization of two (4-carboxybenzyl)TPP molecules to neutralize the positive electrostatic charge of this molecule when it is dissolved in relatively nonpolar methylene chloride. The dimer probably results from the charged TPP group of (4-carboxybenzyl)TPP complexing with the carboxy group of another (4-carboxybenzyl)TPP molecule. Both the signal at 4.7 ppm and 3.8 ppm appear to be doublets of triplets. The splitting observed is caused by interaction of the geminal methylene hydrogens with one another and the triplet signals observed are caused by each geminal hydrogen interacting with the hydrogens ortho to the methylene group on the adjacent phenyl ring. Further splitting of the signal is caused by hydrogens on the phenyl rings of the TPP moiety. The peak at 1.6 ppm is caused by water impurities in the sample. The NMR spectrum of (4-carboxybenzyl)TPP dissolved in d4- methanol (Figure 2.10 b)). on the other hand, displays absorbances at 7.0- 7.4, 7 .5-7.8, and 8.0 ppm due to hydrogens of the phenyl rings in the TPP group. The peak at 4.9 ppm is due to water, and the peak at 3.3 ppm is due to undeuterated methanol impurity contained in the solvent. The peak (4.7 ppm) representing the methylene hydrogens a to the quaternized phosphorus atom is a singlet because methanol is a more polar solvent and can accommodate the electrostatic charge of (4- carboxybenzyl)TPP more readily. Therefore, dimerization does not occur. The environment of the methylene hydrogens is similar and a single signal results. Conseqently, this spectrum indicates the product is (4- carboxybenzyl)TPP. 6 0 C. Peptidedaivatimfionpncedmu I. Bromoalkyl-TPP reagents and vinyl-TPP a) Experimental The peptide to be derivatized was reacted with a twenty-fold molar excess of the bromoalkyl-TPP reagent or vinyl-TPP. First, the derivatizing reagent was dissolved in 1 mL of the derivatizing medium contained in Eppendorf tube. The derivatizing medium was either an aqueous potassium phosphate/sodium borate buffer of pH 9.0 (Micro Essential Laboratory; Brooklyn, NY) or a 1:1 (v/v) acetonitrile/pyridine mixture (pr pyridine = 8.83 [14]). The desired amount of peptide to be derivatized was placed in the reaction medium and the solution was vortexed. The Eppendorf tube was placed in a water bath (37°C) overnight. If the N-terminal derivatization was performed in the acetonitrile/pyridine solvent mixture, the volatile solvents were evaporated under a nitrogen gas stream from a N-EVAPQ (Organomation Associates, Inc.; N orthborough, MA). The remaining solid residue was redissolved in water and lyophilized to remove any remaining pyridine. The solid residue was dissolved in 1-1.5 “L of 1:1 (v/v) acetonitrile/water mixture and added to 1 uL of glycerol positioned on the FAB probe tip. b) Results and discussion i) (2-Bromoethyl)TPP and vinyl-TPP All the bromoalkyl-TPP reagents derivatized the N-terminus of peptides at an appreciable level, with the exception of (1- bromomethyl)TPP. When the (1 -bromomethyl)TPP reaction mixture was analyzed by FAB-MS, a signal representing the expected product was barely perceptible. The (2-bromoethyl)TPP (Reaction 2-5) and vinyl-TPP 61 0 ll (2-5) @PtCHZ—CHZ—Br + H—(lil—(IZ—C)fi-OH —-> H R ’I‘ (ll HBr + @PLCHZ-CHz—(If—CIJ-C)n—OH H R reagents produced high yields of derivatized peptides modified with the ethyl-TPP group at the N-terminus. Under non-aqueous conditions, vinyl-TPP can under Michael addition (Reaction 2-6) with primary amines and proteins, as reported by Swan and Wright [1 5]. W (2-6) @p’f—CIFCHZ + H—(lil—C'Z—C)n-'OH ——> H R (C? 'r 0 + II ©— —CH2— CHz—(rlI—ci—C);0H H R A volatile buffer system of 1 :1 (v/v) acetonitrile/pyridine reduced sample loss when used for bromoalkyl-TPP derivatization. Initially, an aqueous bufl'er system composed of potassium carbonate/sodium borate salts was employed. Use of this buffer system necessitated separation of the N-terminally derivatized peptide from the bufl'er salts by a Sep-Pak® procedure. The Sep-Pak® procedure requires several sample transfer steps contributing to sample loss. A study comparing the reaction yield produced by vinyl-TPP and (2-bromoethyl)TPP when reacted with 1 nanomole of leucine enkephalin (YGGFL) in a 1:1 (v/v) acetonitrile/pyridine solvent mixture overnight (~ 8 h) was made. A higher FAB-MS response of N-terminally derivatized YGGFL was observed for the reaction with vinyl-TPP than (2-bromoethyl)TPP (Figures 2.11 a) and b)). Also, a larger signal representing underivatized peptide 62 gm: (a) - . TPP+“" Vinyl-T133 e . X 1 . / - a80‘ .t < l V 060 I . n 4 1:40. 3 I +TPP-Ethyl-YGGFL I 1 ' . :20: t c - . 1 . .o y ‘ ll I II C ' 100 200 300 400 500 600 700 800 900 - m/z 1"" ‘ TPP" 'Vin l-TPP“ R . y 1° , Cb) ‘ x / £180 A t 1 i . v . 060‘ I I ‘ n . 1140- o . n . .I . l . § 2°. . . *"TPP-Ethyl-YGGFL ‘ lA $an “9* ' 11' . .A' l- (I, 100 200 300 400 500 600 700 300 900 ml: Figure 2.11. The FAB-MS spectrum of 1 nmol of YGGFL N-terminal derivatization mixture. Leucine enkephalin was reacted with a 20-fold excess of a) vinyl-TPP and b) (2-bromoethyl)TPP in an acetonitrile/pyridine 1 :1 (v/v) reaction mixture. A higher response resulting from the molecular ion (M+) of derivatized peptide (m/z 844) was observed for reaction with vinyl-TPP than (2—bromoethyl)TPP. 6 3 (m/z 556) is observed in the (2-bromoethyl)TPP reaction medium than in the vinyl-TPP reaction medium, indicating that vinyl-TPP is a more efficient derivatizing reagent. Because the reaction product for peptide modification with these reagents is identical, the difference in the FAB- MS response for ethyl-TPP derivatized peptide can be related to the reaction yield. The results of this study indicated that vinyl-TPP is a more effective derivatizing reagent than (2-bromoethyl)TPP when the acetonitrile/pyridine reaction mixture is utilized. When (2-bromoethyl)TPP and vinyl-TPP are reacted with YGGFL under aqueous conditions (pH 9.0), vinyl-TPP produces a higher yield of N-terminally derivatized peptide with these reaction conditions (Figure 2.12). Again, a larger amount unreacted YGGFL is present in the (2- bromoethyl)TPP derivatizing medium than in the vinyl-TPP derivatizing medium, indicating vinyl-TPP is a more effective derivatizing agent. The FAB-MS spectra of 1 nmol of YGGFL reacted with vinyl-TPP and (2- bromoethyl)TPP is presented in Figures 2.12 a) and b), respectively. The majority of vinyl-TPP is converted to (2-hydroxyethyl)TPP (m/z 307) during the course of the reaction, as only a small amount of vinyl-TPP (m/z 289) is observable after 8 h of reaction time. However, the FAB-MS response due to N-terminally, ethyl-TPP derivatized peptide produced in the potassium carbonate/sodium borate aqueous bufl‘er (pH 9.0) is lower than the FAB-MS response observed for the acetonitrile/pyridine buffer system. The FARMS response results from the loss of sample incurred with the transfer of sample during the Sep-Pak® procedure used to separate the derivatized product from the buffer salts. Lower reaction yield also is caused by the addition reaction of water and vinyl-TPP to form unreactive (2-hydroxyethyl)TPP; the large excess of water present in the reaction 64 100- . a “M 3 1 (a) (zHydmxyet-hylfl‘ +TPP-Ethyl-YGGFL 1 4 . 9.80 t 4 i J v 4 e 60- . I 4 n 4 l t 40. i e 4 n . l .8 4 l t 2". i . y ‘ ‘ . f ‘ q. i . . 4 llil I 'if ltl‘lllll ill lij|l!..1[“‘ . All-IAA_L-J.-L4.-.L. L ‘1) 100 200 300 400 500 600 700 8 -‘ (1)) AAA 1AA +TPP-Ethyl-YGGFL O A E A A A O wen-unenu—t (semen-mg“ Whig 1-3.‘1 11., I, 100 200300400 500600 700'860'900 A A Figure 2.12. The FAB-MS spectrum corresponding to 1 nmol of YGGFL reacted in a potassium phosphate/sodium borate buffer solution (pH 9.0) with a 20-fold molar excess of a) vinyl-TPP and b) (2-bromoethyl)TPP. The spectrum represents 1 nmol of YGGFL reaction mixture as isolated by the Sep-Pak® procedure. The molecular ion of derivatized peptide (M+) appears at m/z 844, and protonated underivatized peptide at m/z 556. 6 5 medium makes this reaction likely. Although (2-bromoethyl)TPP may react directly with the amine group of a peptide to form the ethyl-TPP derivatized product through a substitution reaction, elimination of HBr from (2-bromoethyl)TPP may occur, producing vinyl-TPP, which, in turn, may react with the peptide (or with water). Consequently, the addition of water to vinyl-TPP may lower the derivatization yield for both reagents in the aqueous derivatizing medium. Reaction of vinyl-TPP and (2-bromoethyl)TPP with peptides to form the N-terminally, ethyl-TPP derivatized products could theoretically derivatize lysine and arginine residues because they contain primary amine groups on their side chains. Buffering the aqueous medium at pH 9.0, and using pyridine (pr = 8.83) as a component of the non-aqueous reaction medium, maintains the amino groups on the side chains of lysine and arginine predominantly in their protonated form (pKa’s 10.8 and 12.5, respectively). However, at this pH, the N -terminal amino group predominantly is unprotonated. Because the derivatization with (2- bromoethyl)TPP liberates HBr, this also requires a bufl'ered system to keep the reaction medium basic. Because a buffered system is necessary for N -terminal derivatization, the use of the vinyl-TPP derivatizing agent in an acetonitrile/pyridine 1:1 (v/v) volatile, buffered system is an effective, convenient, and time-saving procedure. ii) (3-Bromopropyl)TPP and (4-bromobutyl)TPP Derivatization of leucine enkephalin with (3-bromopropyl)TPP (Reaction 2-7) and (4-bromobutyl)TPP to form propyl-TPP and butyl-TPP 66 H I ll (2-7) Q—PL(CH2)3 —Br + H—(lY—C—C);1—OH——-> (C? it H R HBr + ©PL(CH2)3—(hll -C)n—0H © 31': (Reaction 2-8) modified peptides did not produce as high a FAB-MS O l M (2-8) ©PL(CH2)4—Br + H—(If—c—C)g-0H-—-> H R 0 . HBr + ©PL(CH2)4—(li1—$—(ll)n—OH H R response (Figures 2.13 a) and b), respectively) as derivatization with vinyl- TPP. A higher FAB-MS response caused by underivatized YGGFL (m/z 556) was observed with the (3-bromopropyl)TPP and (4-bromobutyl)TPP reagents than with vinyl-TPP, indicating the TPP reagents with a longer alkyl chain have a lower reaction efficiency when reacted under identical reaction conditions. The reaction efficiency for derivatization with (3- bromopropyl)TPP and (4-bromobutyl)TPP cannot be related directly to the FAB-MS response of the derivatized peptides because the propyl-TPP and butyl-TPP modified peptides presumably have a higher hydrophobicity and ionization eficiency than the ethyl-TPP derivatized peptides. The lower derivatization efficiency observed with (3-bromopropyl)TPP and (4- bromobutyl)TPP reagents may be caused by their lower solubility in the reaction medium or perhaps because elimination of HBr to form a reactive alkene-TPP intermediate (analogous to vinyl-TPP) is not favored. 67 8 1 5, 4 (a) " " Allyl-TPP“ 1 4 a 80-- t . 1 V e 604 I I n 1 t 40- 0 4 n 4 ' ‘ . i 20 ; . *TPP-Propyl-YGGFL 4, «U 200 400 ml 600 300 1000 I 100- , , ‘1: : (b) TPP+ Vinyl-TPP" 1 4 a; 80 H20=CH(CH2)2-TPP+ i 4 ' V e 60- 1 i n 4 f, 404 n j 4 ‘ 0140+ : . YGGFL t 20: c y W . +TPP-Butyl-YGGFL ‘ LA -AAA.-.A A A A .A1 A ‘1. 200 400 m" (700 800 1000 Figure 2.13. The FAB-MS spectra representing 1 nmol of YGGFL reacted with a) (3-bromopropyl)TPP and b) (4-bromobutyl)TPP. The reaction was performed with a 20-fold molar excess of reagent in an acetonitrile/pyridine 1 :1 (v/v) solvent mixture. The molecular ion (M+) of propyl-TPP derivatized YGGFL appears at m/z 858 and butyl-TPP at m/z 872. 6 8 2. The “Kunz” reagent a) Introduction Horst Kunz has introduced a protective functionality for amino acids, which is useful for peptide synthesis. Because amino acids derivatized with the 2-(triphenylphosphonio)ethoxycarbonyl (PEOC) group show enhanced solubility in aqueous solutions [1 6], they are efl‘ective in the synthesis of peptides. An N-terminally PEOC-protected amino acid was coupled with a tert-butylesterified amino acid by dicyclohexylcarbodiimide (DOC) to form a dipeptide which was reacted with TFA to deprotect the C-terminus. The free carboxy group can be reacted with another esterified amino acid to lengthen the peptide as desired. The protecting group is extremely stable under acidic conditions, but its removal is possible under mildly basic conditions because the methylene group located a to the phosphorus is acidic, making B-elimination possible. During peptide synthesis, the PEOC peptide esters precipitate as an oil from solution, making extraction with dichloromethane possible [17]. The PEOC moiety has also been used to aid in the synthesis of other naturally-occurring compounds, including depsipeptides [18] and O-linked glycopeptides [19,20]. However, there are several disadvantages with using the “Kunz” reagent: i) the PEOC group is a bulky residue that hinders the condensation reaction in peptide synthesis, ii) the high lability of the PEOC group in presence of base, which is the basis of its removal, demands special care in its synthesis, and iii) the PEOC protective group is a cationic substituent whose halogen counter ion leads to problems with glycosylation with mercury or silver-containing reagents [21]. 6 9 b) Experimental N -terminal peptide derivatization with the commercially available “Kunz” reagent (Fluka Chemie AG; Buchs, Switzerland) was performed with a twenty-fold excess of this reagent. The derivatizing agent was dissolved in 1 mL of an acetonitrile/pyridine 1 :1 (v/v) mixture, the desired amount of peptide was added to the reaction mixture, the solution was vortexed, and the reaction solution was heated at 37°C for 3 h in a water bath. The volatile solvents then were evaporated under a N-EVAP®, the remaining solid was dissolved in H20, and the water was lyophilized in a Speed Vac‘D. The remaining solid residue was dissolved. in a water/acetonitrile 1 :1 (v/v) solvent mixture and placed on a FAB-MS probe tip. Alternatively, 10 uL of 2,4,6 trimethylpyridine (TMP) were added to 1 mL of acetonitrile containing a 20-fold molar excess of the “Kunz” reagent. The desired amount of peptide was added to this solution and the solution was heated for 1 h at 37°C. The volatile components of this solution were evaporated under a N -EVAP®, dissolved in a 1:1 (v/v) acetonitrile/water solution and placed on a FAB—MS probe tip for analysis. c) Results and discussion As demonstrated in Figure 2.14, the N -terminal derivatization with the “Kunz” reagent (Reaction 2-9) is an eficient reaction, when judged by 70 1%001 0 . TPP+ o . 6 J ‘ ___———V'lnyl-TPP+ 1 4 / - - . + :80? / A (2 Hydroxyethylyrg i 4 V t e 60‘ I 1 n . z 40 VGGFL ' n 4 (M +H)"' +TPP-OCOEthyl-YGG‘FL i 204 “mama-Yea y 4 , L \ . W All ulm “14* l 1- ltni (b 100 200 300 400 ml: 500 600 700 800 900 Figure 2.14. The FAB-MS spectrum respresenting 1 nmol of YGGFL reacted with a 20-fold molar excess of “Kunz” reagent in an acetonitrile/pyridine solvent mixture. The peak representing PEOC derivatized peptide (M+) appears at m/z 888, decarboxylated, ethyl-TPP 'derivatiwd peptide is at m/z 844, and underivatized peptide appears at m/z 556. 7 1 the amount of underivatized peptide remaining in the reaction medium. In all the cases investigated, only a small amount of unreacted peptide remained in the reaction solution. This indicates derivatization with the “Kunz” reagent is an efficient reaction that is probably over 75% complete. However, as Figure 2.14 demonstrates, there is a large amount of signal at m/z 787, which corresponds to the peptide derivatized with the ethyl- TPP group, rather than the ethoxycarboxyl-TPP group expected with acylation of the N-terminus of the peptide with the “Kunz” reagent. The ethyl-TPP derivatized peptide arises from decarboxylation of the 2- (triphenylphosphonio)ethoxycarbonyl (PEOC) group as described in Section F.2.b) of Chapter III. This side reaction causes formation of a mixed product, which decreases the signal produced if a single product formed, making this process a disadvantage. Nonetheless, the FAB response due to peptide derivatized with the triphenylphosphine(2- ethoxycarbonyl) group is significant, which makes this a quick, effective derivatization procedure. On-line derivatization with the “Kunz” reagent as described in Chapter III could be adapted as an effective method to produce ethyl-TPP derivatized peptides in a matter of minutes. 3. (4-Carboxybenzyl)TPP a) Experimental N -terminal derivatization with (4-carb0xybenzyl)TPP was per- formed in 1 mL of a water/TFA buffer (pH 5.0). A twenty-fold molar excess of (4-carboxybenzyl)TPP and N-ethyl-N’-(3-dimethylamino- propyl)carbodiimide-HCI (EDC) (Aldrich Chem. Co.; Milwaukee, WI) as a coupling agent were added to the buffer before the desired amount of peptide. An Eppendorf tube containing the reaction medium was placed in a water bath 3 h at 37°C. The water/TFA buffer was lyophilized in a 72 Speed Vac® and the solid residue remaining was dissolved in 1-1.5 11L of an acetonitrile/water mixture 1 :1 (v/v) before being applied to the FAB probe tip containing 1 uL of glycerol. b) Results and discussion Derivatization of the N-terminus with (4-carboxybenzyl)TPP was performed by acylation of the N-terminus with benzoic acid (Reaction 2- 1 0). The reagent contains a methyl-TPP group rather than a vinyl-TPP o H 0 + I I I EDC (210) ©p—m,—©—c—OH + a—(If—(lz—o—on ———> a n Q—E- CH2 ‘Q-E -(liT-(%—E) —OH + 1120 H B group. The vinyl-TPP group is the most abundant fragment with CAD- MS/MS presumably because formation of this fragment involves cleavage of a bond near to localized charge on the TPP group and because of the stability of the vinyl-TPP species. Replacing the ethyl-TPP group with a methyl-TPP group decreases the likelihood of vinyl-TPP elimination. However, the molecular weight of the derivatizing group is significant (380u), increasing the molecular weight of the derivatized peptide. The eficiency of - this derivatizing reaction was low, because underivatized peptide was observed in the reaction mixture (mlz 499) and the signal due to derivatized peptide (mlz 87 7) was not very intense when the reaction medium for (4-carboxybenzyl)TPP derivatization was analyzed by FAB- MS (Figure 2.15). Improvement of the reaction efi'iciency was attempted by converting the peptide into a methyl ester in a 2% (v/v) HCl methanol solution, but little improvement in the FAB-MS response due to derivatized peptide was observed probably due to the additional sample 73 R004 "'71“ ‘ ' (2-Hydroxyethyl)TPP+ b0 10 j (4-Car xybenzylyrgllrtm a 80‘ YGGFL P t l 1 X sol +TPP(4-benzylcarboxy)-YGGFL I 4 n . t 40. l 8 4 n . i8 20.1 - ‘2 .0 t 1 4 y 4 4 4 . l II 4 ‘ 4] 'I' H 4 ‘ ‘ ‘ (1) 200 400 m/z 600 800 1000 Figure 2.15. The FAB-MS response of 1 nmol of YGGFL when reacted with a 20-fold molar excess of (4-carboxybenzyl)TPP and EDC in an aqueous solution (pH 5.0) buffered with TFA. The peak representing the molecular ion (M+) of derivatized YGGFL appears at m/z 934, the protonated, underivatized peptide appears at m/z 556. 7 4 handling required for this procedure. Derivatization under more acidic conditions (pH 2.0 and 3.0) was also attempted, but little increase in the derivatization efficiency was observed. In fact, no (4-carboxybenzyl)TPP derivatized peptide product was observed when the reaction was performed at pH 2.0. 4. (2-Aminoethyl)TPP a) Experimental The C-terminus of peptides were derivatized by (2-aminoethyl)TPP. A twenty-fold molar excess of (2-aminoethyl)TPP and EDC were dissolved in an aqueous solution (pH 5.0) buffered with trifluoroacetic acid. The desired amount of peptide then was added to the solution, the solution vortexed, and placed in a water bath (37°C) for 3 h. The water and TFA bufl‘er were lyophilized before the reaction mixture was analyzed by FAB- MS or HPLC. This procedure effectively coupled the aminoethyl-TPP group to the peptide through an amide bond. b) Results and discussion The coupling reaction utilizing the water-soluble coupling reagent EDC effectively linked the carboxy terminus of the peptide with (2- aminoethyl)TPP [1]. Carbodiimides are routinely used for for peptide synthesis and also convert carboxylic acids to active acylating agents. This reagent also has also been used for high-sensitivity solid phase sequence analysis [23]. The EDC is used to couple peptides to solid supports with the favorable activation of the carboxyl group. The acidic conditions used in the derivatization procedure with (2-aminoethyl)TPP protonate one of the nitrogens of EDC and deprotonate the carboxyl group of the peptide. Addition of the carboxylic acid to a C=N bond of the carbodiimide generates an O-acylated derivative. The (2-aminoethyl)TPP 7 5 amine reacts with this agent because there is a strong driving force for elimination of the urea unit, with formation of a stable amide carbonyl group (Reaction 2-11). The (2-amin0ethyl)TPP reacted with the C- '11] (2-11) H-(N—C —C) —OH _ _ + EDC I l n 4' HaN CHz—Cl-Ig P~© ——>- H R 'i' u H—(N—ci—C) firearm—Ag» . 44,0 H R H terminal carboxy group in high yield, producing a high FAB response (Figure 2.16). Although (2-aminoethyl)TPP is expected to derivatize the side chains of aspartic acid and glutamic acid residues (pKa’s 3.9 and 4.3, respectively), the double-derivatized or the side-chain derivatized peptides were not observed. This reaction probably does not occur because the formation of the O-acylated derivative after addition of the carboxylic acid to the C=N bond of EDC is sterically hindered when the side chain is in the middle of the peptide chain. If this complex is formed, attack by (2- aminoethyl)TPP is very disfavored due to steric hindrance. Because a 20- fold molar excess of (2-aminoethyl)TPP was used, the possible side reaction involving the coupling of two peptides through their carboxy groups was not observed. In these investigations, only decamers and shorter peptides were derivatized by (2-aminoethyl)TPP. However, derivatization efficiencies would be expected to decrease for larger peptides because secondary structure would be expected to produce further steric hindrance. For these smaller peptides, derivatization efficiencies of 95% yield were observed, making this an effective derivatization procedure. 76 1i°°4 ‘ i / (2-Aminoethyl)TPP+ _ i 4 /p Vinyl-TPP :80 l .L—w/ Cyclic YGGFL x I v .. 1. I n . t . e 40‘ «4140+ YGGFL-NHEthyl-TPP" n 4 YGGFL : J ° t 2°] 3’ 4 ° ° 200 400 m/z 600 800 1000 Figure 2.16. The FAB-MS spectrum representing 1 nmol of YGGFL reacted with a 20-fold molar excess of (2-aminoethyl)TPP and EDC in an aqueous solution (pH 5.0). The peak representing derivatized peptide (M+) appears at m/z 843 and underivatized peptide is at m/z 556. 7 7 5. Summary of derivatizing reagents All the reagents discussed yielded derivatized peptides containing a triphenylphosphonium group. One measure of the effectiveness of a derivatization method is to evaluate the signal intensity representing the derivatized product when analyzed by an instrumental method. To compare the yield of TPP-derivatized product produced by these derivatization methods, 1 nanomole of leucine enkephalin was reacted with a twenty-fold molar excess of the derivatizing reagents listed in Table 2.1 for 8 h at 37°C in the appropriate reaction medium. Derivatization with the bromoalkyl-TPP reagents, vinyl-TPP, and the “Kunz” reagent were performed in an acetonitrile/pyridine 1:1 (v/v) solvent mixture. Derivatization of leucine enkephalin with (4- carboxybenzyl)TPP and (2-aminoethyl)TPP was performed with a twenty- fold molar excess of EDC in an aqueous solution (pH 5.0) buffered with TFA. The signal representing derivatized leucine enkephalin after derivatization with the various reagents observed during analysis of the reaction mixture by FAB-MS is largely dependent on the efiiciency of the derivatization reaction. However, the signal produced during analysis by FAB-MS is not a direct measure of the derivatization yield because the derivatized leucine enkephalin produced with the various reagents has diverse structures, causing the surface activity of modified leucine enkephalin to vary with the reagent used. However, the surface activity of TPP-derivatized leucine enkephalin probably differs minimally because the hydrophobic effects of the TPP group dominate the surface activity of the derivatized peptide. The FAB-MS response was calculated as the ratio of signal intensity representing derivatized leucine enkephalin to the signal produced by underivatized, protonated YGGFL (mlz 556) during 7 8 analysis of the reaction mixture. Table 2.1 (below) indicates that the highest FAB-MS response for N-terminally derivatized peptide was produced by modification with vinyl-TPP and that the C-terminus was efficiently derivatized by (2-aminoethyl)TPP. Table 2.1. Ratio of signal intensity of derivatized leucine enkephalin to that of underivatized, protonated peptide contained in the reaction mixture during analysis by FAB-MS. D'I” B | Bl' [5' III 'I Vinyl-TPP bromide 16 (2-Bromoethyl)TPP bromide 3.1 (3-Bromopropyl)TPP bromide 2.4 (4-Bromobutyl)TPP bromide 1 .2 2-(tri1phenylphosphonio)ethyl chloroformate chloride 0-15 (the “Kunz” reagent) (0.25 for decarboxylated prod.) (4—carboxybenzyl)TPP chloride 2.0 (2-aminoethyl)TPP bromide 15 When evaluated by the ratio of signal intensity produced by modified peptide to the signal intensity due to underivatized peptide, vinyl-TPP is the most effective TPP-containing derivatizing reagent. Only the (2- aminoethyl)TPP reagent modifies the C-terminus of peptides, but fortunately produces a high ratio of signal representing derivatized peptide when compared to the signal of unmodified peptide. D. SeparationoftheC-tenninalderivatizationmixtm'e 1. Experimental The derivatization mixture was separated with a gradient of 30% to 70% acetonitrile in 25 mins and holding the acetonitrile concentration constant for 15 mins. The separation was accomplished with a Beckman 7 9 342 Gradient Liquid Chromatograph system (1 1 2 Solvent Delivery Module, 420 Controller, and 340 Organizer; Beckman Instruments, Inc.; Fullerton, CA). The HPLC column used was an Alltech EconosilO 013 column (Alltech Associates, Inc.; Deerfield, IL) with dimensions of 10 mm x 250 mm. The stationary phase was contained on 10-um spherical particles. The flow rate during chromatographic analysis was 2 mL per minute. The mobile phase exiting the column was monitored by a UV-Vis detector at 214 nm (Kratos Spectroflow 757 Absorbance Detector, Kratos Analytical; Ramsey, NJ). The flow cell had an 8-mm pathlength and 12- pL volume. Alternatively, smaller amounts of the peptide derivatization mixture were separated on an Alltech Econosphere® C13 column. This column has dimensions 4.6 x 250 mm with a 5-um diameter particle size. The flow rate used for this column was 1 mL/min. 2. Results and discussion Separation of the (2-aminoethyl)TPP reaction mixture by HPLC effectively isolated C-terminally derivatized leucine enkephalin from the derivatizing reagent and resulted in a chromatogram which contains five major peaks (Figure 2.17). Comparison of the FAB-MS spectrum of 1 nmol of the unseparated reaction mixture (Figure 2.16) with 1 nmol of the product separated by HPLC (Figure 2.18) demonstrates that the signal due to (2-aminoethyl)TPP and EDC are decreased with isolation of the derivatized product. The solvent fractions represented by these peaks were collected and analyzed by FAB-MS. The first peak is caused by the difference in refractive indices of the solvent that the reaction mixture was dissolved in and the mobile phase, which commonly is called the solvent front. The coupling agent EDC is represented by the second peak, and (2-aminoethyl)TPP was the primary compound in the fraction 80 2.0— E. 1.5- *5 FE + _ ,2 >1 0.. A ‘3 5 E .2. a .:>~ g :3 ,’ g \ 2 r: -— E _ 1.0- 8 5* El 1 z s ' S ' >4 o 4 J D “I: .44 ‘43, 0.5- f 0 l I I l I l l I f f TIniect 4 '8 12 16 20 Time(min) Figure 2.17. HPLC separation of 20 nmol of the C-terminal derivatization mixture of YGGFL. The reaction mixture contained a twenty-fold excess of (2-aminoethyl)TPP and EDC. The compounds that the peaks represent are identified in the chromatogram. The conditions of the separation are: flow rate 2mL/min; column, 10 x 250 mm; stationary phase, 10 um diameter, 018 bonded phase; detector, 214 nm, 2.0 AUFS; chart speed, 30 cm/hr. 81 1004—0” . R e 4 1 4 a 80- P ‘ . l I Z 60 Vinyl-TPP“ 1 ' ((2 Amino: 3 40 V ”ethyl)TPP“ f; 1 / / YGGFL-NHEthyl-TPF‘ 8 t 204 / . . Y 4 a ‘ 1 4 4 I i” 1 4 . . wi 4 ' 1 44 ‘ . 4 I , - . 4 4 I , . 4 300 400 500 600 700 300 900 m/z Figure 2.18. Analysis of the peak representing separated, C-terminally derivatized YGGFL by FAB-MS. The molecular ion (M+) of derivatized YGGFL dominates the spectrum (m/z 843), although signal representing vinyl-TPP and (2-aminoethyl)TPP (mlz 306) is observable. R10: n. 4 s 2 1 (M); e ‘ "n3 474 am 1' l ‘ x": m 46 '3 z 8: Y": 7 as: as an 275 an i 13' an) a -Y v e 6? fl . I v c c r 1. c 1Iv (cm, (on), r-@ n J I"; I" © 1: 4. 0 4 n 4 4 9 Ya I t 21 06332 a 537 a y 1 1131 507 l: ‘8 ‘ 1*.1 l 4. . l 44 .4‘ l . l . . ' : '1 , ‘4‘ H 400 500 600 700 800 Figure 2.19. The FAB-CAD-B/E spectrum of the molecular ion (M+) of C- terminally derivatized YGGFL with (2-aminoethyl)TPP appearing at m/z 843. 8 2 collected when the third peak eluted. Underivatized YGGFL is represented by the fourth peak, and cyclized YGGFL appears in the main component of the fifth peak. The final peak represents C-terminally ethyl-TPP derivatized leucine enkephalin. In the FAB-CAD-B/E spectrum of derivatized leucine enkephalin (Figure 2.19), directed fragmentation from the localized charge of the ethyl-TPP group attached to the C-terminus is observed. All the fragments observed in the spectrum are derived from the C-terminus. No ion series dominates the spectrum and the x, and y" ion series form a complete series, which are complemented by y 1 — y3 fragment ions. Thus, the complete amino acid sequence of the peptide can be elucidated with ease. The ion at m/z 680 represents the loss of a tyrosine residue from the N-terminus. Losses of 57u, 57u, and 147u indicate the amino acid sequence from the N-terminus is Y—G—G—F—L. The ion at m/z 417 (y 1) is characteristic of a (2- aminoethyl)TPP derivatized leucine residue. E. Quantitationof the C-terminal mactionyield Quantitation of the reaction yield was the ultimate goal after separation of the component of the reaction mixture was achieved by HPLC chromatography. Once the reaction mixture is separated into individual components, UV-Vis absorption measurements with a detector on-line can be performed separately for each of the individual components, provided the compounds are entirely separated with baseline resolution. Other requirements of this procedure include the necessity that no other components of the reaction mixture can co-elute with the compound of interest when the absorbance is being measured and the identity of the peak of interest must be positively confirmed. The detected amount of absorbance caused by each separated component then 8 3 may be used to quanitate the amount of that component, if the absorption coefficient for the compound is known. Separation of the (2- aminoethyl)TPP derivatization mixture of leucine enkephalin (Y GGFL) by HPLC with on-line UV-Vis detection was used to quantitate the reaction yield. 1. Determination of the absorption coefiicient a) Experimental A large quantity (1.50 mg, 2.70 x 10'6 mol) of YGGFL was reacted to yield an adequate amount of C-terminally derivatized peptide to be weighed with a reasonable amount of accuracy. The peptide was dissolved in 100 mL of H20 (pH adjusted to 5.00 with TFA) with a twenty- fold excess of EDC (0.01035 g, 5.40 x 10'5 mol) and (2-aminoethyl)TPP (0.01035 g, 5.40 x 10'5 mol) added. The solution was stirred overnight at room temperature. The solution was concentrated by rotoevaporation at room temperature and subsequently lyophilized. Transparent crystals were observed when all the water was lyophilized. The crystals were dissolved in 250 uL of acetonitrile and 10 11L portions of the reaction mixture were separated by HPLC under semipreparative conditions. The fractions containing C-terminally derivatized YGGFL were collected, pooled, concentrated by rotoevaporation and lyophilized. b) Results and discussion After large scale derivatization, semipreparative separation of the derivatized peptide from the reaction mixture by HPLC allowed collection of a fraction containing C-terminally derivatized leucine enkephalin. Evaporation of the solvents contained in the collected fraction by lyophilization and a mechanical pump yielded a yellow-brown, oily solid that weighed twice the theoretical yield of C-terminally derivatized 8 4 leucine enkephalin expected. The FAB-MS spectrum of the collected fraction displayed a peak at m/z 149 that is characteristic of phthalate compounds. These observations indicated phthalate impurities were contained in the collected fraction. Due to the presence of impurities with C-terminally derivatized leucine enkephalin, an accurate weight of C- terminally derivatized leucine enkephalin could not be obtained for determination of its absorption coefficient. Thus, direct, on-line quantitation of the reaction yield could not be performed based on the absorbance of the peak corresponding to derivatized peptide in the HPLC chromatogram. Rather, quantitation of the reaction yield was based on the loss of peak area observed for the peak in the HPLC chromatogram representing underivatized leucine enkephalin following derivatization with (2-aminoethyl)TPP. 2. Direct quantitation of the reaction yield a) Experimental About 75 11L of the C-terminal reaction mixture (20 nmol YGGFL and a twenty-fold excess of (2-aminoethyl)TPP and EDC dissolved in a 1:1 (v/v) HzO/ACN solvent mixture after lyophilization of the original H20 (pH 5.0) reaction medium was separated by HPLC under conditions identical to those described in the separation of the C-terminal reaction mixture described above except UV-Vis absorbance was monitored at 260 nm. b) Results and discussion As was already mentioned, phthalate impurities in the collected fraction containing derivatized peptide were detected, consequently direct quantitation of C-terminally derivatized YGGFL on-line was invalid because an accurate extinction coefficient for this product could not be obtained. Separation of the reaction mixture by HPLC produced a 8 5 chromatogram similar to that contained in Figure 2.17. The fourth major peak in the chromatogram represents underivatized leucine enkephalin. Consequently, quantitation of the C-terminal reaction yield was based on the difference in absorbance observed for the peak representing underivatized leucine enkephalin when 20 nanomoles of this peptide were injected alone and the absorbance observed when 20 nanomoles of reaction mixture were injected. Absorbance was monitored at 260 nm because the phenol group of the tyrosine residue and the phenyl group of the phenylalanine residue in YGGFL absorb radiation at this wavelength whereas the other portion of the peptide has little absorbance at this wavelength. Although a side product of cyclized leucine enkephalin is observed by the coupling of the N - and C-terminus by EDC, the amount of side product produced by this reaction can be quantitated accurately at 260 nm because the phenyl and phenol groups remain intact during cyclization and the new structure of the peptide produced after cyclization causes little difference in absorbance. The amount of side product detected then can subtracted from the yield calculated by the loss of signal intensity for the peak representing underivatized leucine enkephalin after derivatization. Taking into account the quantity of side product formed by cyclization of YGGFL, the yield of C-terminally derivatized leucine enkephalin was 59 % after reaction with a twenty-fold molar excess of (2-aminoethyl)TPP 3 h at 37°C. F. FAB-CAD-BIE fragmentation of derivatized VGVAPG 1. Introduction Investigations on the peptide VGVAPG derivatized with various reagents containing the triphenylphosphonium group were performed by FAB-CAD-B/E. These investigations will demonstrate the advantages 8 6 and the complications encountered when utilizing these reagents. Although the TPP-containing derivatizing reagents were used to modify other peptides, only the modification of VGVAPG with these reagents and analysis by FAB-CAD-B/E will be reported here for the sake of brevity. The fragments observed for the TPP-derivatized peptides containing a localized charge likely are produced by remote-site fragmentation, similar to the fiagmentation of fatty acids as described by Gross [24]. One should be aware that the m/z values of the peaks representing fragments of the TPP-derivatized peptides (beyond the mass shift associated with derivatization) differ from the m/z values observed for underivatized peptides, although the same nomenclature is used in symbolizing these fragments. This is because the peptides derivatized with the TPP group contain a “preformed” charge. Consequently, the derivatized peptides are not protonated and fragment by remote-site processes. For N -terminally derivatized peptides, an-H, bn-H, and cn-3H fragments are observed (A-H, B-H, and C-H according to the Roepstorff nomenclature), insted of the a... b... and c" ions observed with underivatized peptides. In the mass spectra of C-terminally, TPP derivatized peptides, xn-H, yn-3H, and zn-H ions are observed (X-l, Y-l, and Z-l according to Roepstorff nomenclature), rather than the x”, y... and 2,. ions observed in the fragmentation of underivatized peptides. Thus, the y" fragments observed are 3 mass units lower in mass than the y, (Y” according to Roepstorff nomenclature) fragment expected for the underivatized peptide. 8 7 2. Mass spectrometry a) Instrumentation FAB has greatly increased the number and variety of compounds that can be analyzed by a mass spectrometer. Particularly because high- mass molecules can be ionized by this technique, large-sized mass spectrometers are necessary to analyze these molecules at high resolution and sensitivity. The JEOL JMS-HX-llO is an ultra high- performance mass spectrometer that meets these demands. A resolution over 100,000 is possible because the JEOL HX-110 employs a perfect second-order double-focusing system with an ion optics system whose second-order aberration is less than one-tenth that of conventional instruments. The velocity dispersion of isobaric ions that does exist in the electric field is compensated for by direction-focusing of the magnetic field. Ions are focused with such accuracy with instruments of the Nier- Johnson design that the exact mass of ions can be measured to six significant figures and ions of the same nominal mass, but different elemental composition are separable. High sensitivity is possible because ion beams of a wide angle distribution and energy width are converged by a combination of quadrupole lenses and a large, uniform, and cylindrical magnetic field. This configuration reduces ion loss in the vertical direction by a factor of ten-fold. The JEOL HX-110 can detect ions with a mass-to-charge ratio (mlz) over 15,000 with a 30-cm radius electromagnet with a maximum strength of 1 8.8 kilogauss. Automatic evacuation of the direct probe inlet and ion source make this instrument ideal for FAB-MS. b) Operating principles The fragment ions produced by collisionally-activated dissociation in the first field-free region of the mass spectrometer can be separated 8 8 according to their mass in a double-focusing instrument with the B/E linked scan. Linked scanning is performed at a constant accelerating voltage. The electric and magnetic fields are scanned simultaneously so that the ratio of their field strengths is constant [25]. These scans . normally are carried out under computer control and prior mass calibration is used to produce accurate mass-to-time correlations based on the scan rate. In the linked scan, the spectrum consists of product ions formed in a collision cell placed close to the source-defining slit from a chosen precursor ion. When the precursor ions (m1+) fragment to produce product ions (m2+ ), these ions require different electric field strengths, E1 and E2, respectively, to be transmitted through the electric field sector, and magnetic fluxes, B1 and B2, to be transmitted through the magnetic sector. Because these scans are run at a constant accelerating voltage, V, the energies of all the ions leaving the source are equal. Ions with mass, m, and velocity, v, pass through the electric and magnetic sectors to be detected when the electric field strength, E, meets the conditions (Equation 2-3), where R, is the radius of the electric sector: (2-3) eE = vaRe For the magnetic sector field strength (Equation 2—4) where Rm is the radius of the magnetic sector: (24) eB = mv/Rm Consequently, the conditions needed to pass m1+ and m2+ in the electric sector are (Equations 2-5,6): (2-5) eE1 = m1v12/Re (2-6) eEz = m2v12/Re 89 The conditions required to pass these ions in the magnetic sector are (Equations 27,8): (27) eB1 = m1V1/Rm (2-8) e32 = m2V1/Rm Combining these equations gives Equation 2-9: (2-9) m1/m2 = E1/Ez = 131/32 Consequently, B1/E1 = B9/E2 = constant. The value of the constant or ratio of magnetic field strength to electric field strength (B/E) is determined by the electric and magnetic field strengths needed to transmit the precursor ion. The resolution needed to separate the fragment ions is dependent on the translational energy released during fragmentation. The resolution normally achieved in linked scanning makes it possible to assign the product ion mass accurately to the nearest integer. The resolution by which the precursor ion can be selected is only from 300 to 400; consequently, m1+ ions containing one or more 13C isotopes will be collected [26]. c) Experimental FAB desorption of protonated and derivatized peptides was performed with a 6-keV beam of xenon atoms. The desorbed ions were analyzed in a JEOL HX-110 double-focusing mass spectrometer. The resolution was set at 1000 and the accelerating voltage remained at 10 kV for all scans. A JEOL DA-5000 data system generated linked scans at constant B/E to produce CAD-MS/MS spectra. Helium was used as the collision gas in the first field-free region to produce CAD. 3. Analysis of underivatized VGVAPG by FAB-CAD-B/ E To compare the fragmentation of TPP-derivatized peptides with underivatized VGVAPG, the FAB-CAD-B/E spectrum of underivatized 9 0 peptide is presented in figure 2.20. The I), and y, fragments resulting from fragmentation of the amide bond [C(O)—N(H)] dominate the mass spectrum. The amide functionality is the most basic functionality in VGVAPG and is the site of predominant protonation. Consequently, fragmentation of this functionality in the peptide backbone is probably caused by charge-initiated chemistry. The 1),. fragments are fragments resulting from cleavage of the peptide bond and retention of charge on the N-terminal side of the molecule. The y, ions are formed by charge retention on the C-terminal portion of the molecule and by hydrogen transfer from the neutral fragment produced by fragmentation at the amide bond. Specifically, protonation of the amine group of the amide functionality occurs, and a hydrogen shift from a N (H)n group of the neutral fragment causes formation of the y” fragments. In the FAB- CAD-B/E spectrum of underivatized VGVAPG, the fragments representing cleavage of the [CH(R)—C(O)] bond (an and x" ions) and the [N(H)—CH(R)] bond (0,. and 2,. ions) are not nearly as abundant as fragmentation of the amide bond. Immediately, one observes that fragments (y, and bu) resulting from cleavage of the peptide bond dominate the spectrum of underivatized VGVAPG (Figure 2.20), because the amide bond represents the most basic site in peptides that do not contain basic amino acid side chains. Examining the low-mass end of the spectrum, a discernible immonium ion is observed at m/z 72, corresponding to the mass of the valine immonium ion, indicating that valine is present in the peptide being analyzed. The high-mass region of the spectrum displays the loss of water and also the side chain of a valine residue. Observing the direct loss of residues (the y, series) from the protonated molecule (M+H)+, a 91 com 001"‘ 1 v.» .3». >1 no on * a“ Cum- +Am+u$ .mmv a): an 2 +Am+2v canoe—o8 33:38AM 25. finger» 833:85 as 5884 Emdééé 23. :8 u 28E as com 3 no 3 ”a an 0A.: in .H m 4N v a N a 3 fl a use "no 4 5 8a a: "an .m a ”an . < > Mpg—I . ”as . a v a n: 4 r m mvfldum>0 Hflumfimwuh 9 2 small peak at m/z 400 (the y5 fragment) represents a reasonable loss, indicating a valine residue resides at the N -terminus. Subsequent losses of 57a, 99u, 71u, and 97u are represented by the y1 - y; fragments appearing at m/z 343, 244, 173, and 76, respectively. The 25 + 1 fragment confirms the assignment of the y5 fragment ion. Surprisingly, a y1 ion can be observed at m/z 76, indicating that glycine resides at the C- terminus. Consequently, the y" ion series identifies the amino acid sequence from the N-terminus as V—G—V-A—P—G. The b" ion series begins at m/z 424; this peak represents the loss of a glycine residue (57u) and a water molecule from (M+H)+. The bu ion series continues at m/z 327, 256, and 157, representing the b4 - I): fragment ions. The bu ion series is confirmed by peaks representing the a4, a5, c3 and c5 fragment ions, which appear at m/z 228, 299, 396, and 272, 441, respectively. The amino acid sequence deduced from the b, series is ?-V-A-P-G. The amino acid sequence generated from the C- terminus (b, ion series) verifies the amino acid sequence generated from the N-terminus (the y" ion series), allowing the amino acid sequence for this peptide to be deduced from these overlapping series of fragment ions. The deduced sequence is VGVAPG. However, the spectrum of the underivatized peptide is complicated by the generation of fragments from both termini. 4. (2-Bromoethyl) TPP and vinyl-TPP derivatized VGVAPG Analysis of the vinyl-TPP and (2-bromoethyl)TPP (Figures 2.21 a) and b) respectively) derivatized VGVAPG revealed similar fragmentation patterns, confirming that similar peptide modification is produced by derivatization with these reagents. The fragmentation pattern for the N - terminally, ethyl-TPP derivatized peptide changed dramatically when 93 ‘4‘ H -..“? wowmnonnm oo> 883:8 $3.38 53553 .0 no ALE com 838—08 23 me 280% Sm.9<0.m<.m SE. dud earn 95 as GE. 02. O8 cow 0mm com and. ...1 ; ... an 4 an «N J 4.» N 4.. a.» v> an a.» u> @ ._. ©I+m «58 «E8 2 m c 41% .> e.,->lz«m 0 © 53 QB v.5 PW ”EN >. «8 Eh g Q» k L RM "ch k a “an o8 ”:3 Bu\ iv “A; «h H a a v a ..c Momaumbo HGH°QMHHH>S 96 was augmented by the formation of v" and w” ions, which result from side chain cleavage of the 2,. + 1 ions. The abundance of the w" and v" ions is enhanced by ethyl-TPP derivatization, from which significant structural information about the amino acid side chains can be obtained. From this mass spectrum, enough information is discernible to provide a complete amino acid sequence and the spectrum is simplified because only fragments for the C-terminus are produced. 6. (3-Bromopropyl)TPP derivatized VGVAPG Although derivatization of VGVAPG by (3-bromopropyl)TPP was not as efficient as derivatization with vinyl-TPP, a sufficient amount of derivatized peptide was present to perform analysis by FAB-CAD-B/E (Figure 2.23). Losses of the valine and alanine side chains from the molecular ion of VGVAPG derivatized with (3-bromopropyl)TPP was observed in the high-mass region of the spectrum. From the N-terminus of derivatization, a complete series of an, b", and c, ions could be observed. However, these ions were not as intense as those produced by ethyl-TPP derivatization. The most intense peaks in the spectrum (m/z 339, 383, and 385) were produced by fragmentations that were not accounted for. The peak at m/z 321 is due to cleavage of the [N(H)1 - CH(R)1] bond and retention of charge on the modifying group. The peaks at m/z 339, 383, 385, 444, 499, and 614 also could not be accounted for. The c: fragment also seems unusually abundant. These results indicate that an increase in the alkyl chain length of the TPP-derivatizing group complicates the mass spectrum, perhaps causing fragmentation by alternative mechanisms. 97 gm +95 NOIOA<>U >1? game 8» as No A an use 8522888; -8 fies 888388 8% cm<>c> 38.328 83.888 5:383 -z .8 Ea 8... 8:528. 2: .s 8888.. Eméééé 2:. dad 28a...— a... cow man wm fir V T H v f m0~w8-~>o Hduoflmmuh 9 8 7. (4-Bromobutyl)TPP derivatized VGVAPG The FAB-CAD-B/E mass spectrum of VGVAPG derivatized with (4- bromobutyl)TPP was similar to the spectrum of VGVAPG derivatized with (3-bromopropyl)TPP, except the fragments were shifted upward in mass by 14u (Figure 2.24). Again, fragments from the modified N- terminus predominated in the spectrum because the butyl-TPP modifying group contains a localized charge. A complete series of an ions was observed in the mass spectrum, which is complemented by a nearly complete series of bu and c" ions. At the high-mass end of the spectrum, the losses of alanine and valine side chains provide a partial amino acid composition. For the butyl-TPP derivatized peptide, the relative abundance of the bu and cu fragment ions have a greater signal intensity relative to that of the an ions. This may indicate that placing a fixed charge more distal from the peptide (increasing the alkyl chain length) causes the likelihoood of backbone fragmentation to be more equal for each bond. Peaks also appear in this spectrum that can not be accounted for. The ions at m/z 397, 399, 499, 560, and 628 can not be accounted for. The peak appearing at m/z 335 is due to cleavage of the [N(H)1—CH(R)1] bond, with charge retention at the derivatized portion of the molecule. Perhaps these fragments are indicative of an alternative mechanism of fragmentation. The high as fragment ion abundance also seems unusual. Although these unknown fragments are observed for butyl-TPP derivatized RVYVHPF, the unidentified peaks were not nearly as intense. 8. “Kunz” reagent derivatized VGVAPG The FAB-CAD-B/E spectrum of VGVAPG derivatized with the “Kunz” reagent displays a simplified fragmentation pattern (Figure 2.25) 99 com +35 :. _ _ 3 .a MOI-G m <4 > 0 ?T game . 5» "no 1 A. _ E at .38 c8 8:55.888 £3 8382 sec 822.1 8 88» 95-38 a £5 8:832 2: a Becca 24>? co £6 :2 ceases 2: co a338,. Em-n0 Hfiuofimwuh 100 .33 say 039% 853:8 comm 20R 3 gauge ..Ear 2: £3 Euuguwu ma 84>? Bungtwu magmas -z .3 £6 :3 3:538— 2: he again Emdééfi 2a. and 25E com com com com cow ” u a u 3 m5 >. H ‘b 0 Q3 Q Iwuo solo. a < > E >l Z—fl: gr ”8 izvmvmnfi 333.23 «a Didi—.dum'ém Hduonmwuh l 01 comparable to that of VGVAPG modified at the N-terminus by ethyl-TPP, except that peaks representing fragment ions were shified upward in mass by 44u. In this spectrum, only N-terminal ion series are observed, with the an ion series dominating the mass spectrum, which is complemented with the ha and c, ion series. From the fragmentation pattern of the “Kunz” derivatized peptide, the amino acid sequence is easily identified to be VGVAPG. However, the peaks at m/z 481 , 649, and 741 cannot be accounted for. The increased formation of d, ions also facilitates elucidation of the amino acid sequence by informative fragmentation of the amino acid side chains. 9. (4-Carboxybenzyl) TPP derivatized VGVAPG The product of derivatization of VGVAPG with (4- carboxybenzyl)TPP also was analyzed by FAB-CAD-B/E (Figure 2.26). Fragments derived from the modified N -terminus dominated the mass spectrum. The (4-carboxybenzyl)TPP modifying group causes predominance of the an ion series, which represents a complete series. In addition, complementary b, and en ion series are observable in the spectrum at lower abundance. The high-mass region of the spectrum displays side chain losses of alanine and valine from the molecular ion, providing a partial amino acid composition. Peaks corresponding to the descriptive (1,. ions are also observed. However, all the peptides derivatized with this reagent displayed a peak at m/z 595, but the source of this peak cannot be accounted for. These results indicate the (4- carboxybenzyl)TPP derivatizing group effectively causes the formation of a... fragment ions for facilitated peptide sequencing. 102 $5 £5 9.2» mmhauamfiuoeswev 2: .3 guess 8 3.33 8528 £3353 .2 no is :3 3:338 2: me $88QO Emdééfi. 2:. .93 25E e8 cow cor com can :sw: ..i ... :.; n.2a . __ ,. n m . . . . ., O «Q «a 0 fl «.0 _. a” 3 a“ v“ a <. >. a . A ._. © . N . mole m < > e.>#ZImIAUY;8.ElflV . . 0 RM mg m u G . 8n 3 a ".5 . fi "co 3 Sic we an ad . 3 Eu mmmflam>fl Hnuonmwum 99°F.“ 10. 11. 12. 103 References . Wagner, D.S.; Salari, A.; Gage, D.A.; Leykam, J .; Fetter, J .; Hollingsworth, R.; Watson, J .T., Biol. Mass Spectrom., 1991, 20, 419- 425. Watson, J.T.; Wagner, D.S.; Chang, Y.-S.; Strahler, J .R.; Hanash, S.M.; Gage, D.A., Int. J. Mass Spectrom. and Ian Process, 111, 1991, 191-209. Wagner, D.S.; Salari, A.; Fetter, J .; Gage, D.A.; Leykam, J.; Hollingsworth, R.; Watson, J .T., Proceedings of the 38th Annual Conference on Mass Spectrometry and Allied Topics, Tucson, AZ, pp. 307-308. Wagner, D.S.; Gage, D.A.; Allison, J .; Watson, J .T., Proceedings of the 39th Annual Conference on Mass Spectrometry and Allied Topics, Nashville, TN, 1991, pp. 1330-1331. Wagner, D.S.; Nieuwenhuis, T.J.; Chang, Y.-S.; Gage, D.A.; Watson, J .T., Proceedings of the 40th Annual Conference on Mass Spectrometry and Allied Topics, Washington, DC, 1992. BS. Wagner, Ph.D. Dissertation, Michigan State University, 1992. Friedrich, K.; Henning, H.-G., Chem. Ber., 1959, 92, 2756-2760. Still, W.C.; Kahn, M.; Mitra, A., J. Org. Chem., 1978, 43, 2923-2925. Keough, P.T., Grayson, M., J. Org. Chem., 1964, 29, 631 -635. Schweizer, E.E.; Bach, R.D., J. Org. Chem., 1964, 29, 1746-1751. Chistol, H.; Cristau, H.J.; Soleiman, Tetrahedron Lett., 1975, 16, 1381-1384. Dubois, H.J.; Lin, C.-C.L.; Beisler, J .A., J. Med. Chem., 1978, 21(3), 303-306. 13. 14. EEFSEE .52 l 04 Sanyal, U.; Chatterjee, R.S.; Das, S.K., Chakraborti, Neoplasma, 1984, 31(2), 149-155. Dean, J .A., Ed., Lange's Handbook of Chemistry, 13th Ed., 1985, Me- Graw-Hill, Inc., NY, Table 5.8, 5-55. Swan, J .M.; Wright, S.H.B., Aust. J. Chem., 1971, 24, 777-783. Kunz, H., Chem. Ber., 1976, 109, 2670-2683. Kunz, H., Angew. Chem. Int. Ed. Engl., 1978, 17, 67-68. Kunz, H.; Lerchen, H.-G., Angew. Chem. Int. Ed. Engl., 1984, 23(10), 808-809. Kunz, H.; Dauth, H., Leibigs Ann. Chem., 1983, 3, 337-359. Kauth, H.; Kunz, H., Leibigs Ann. Chem., 1983, 3, 360-366. Kunz, H., Angew. Chem. Int. Ed. Engl., 1987, 26, 294-308. Sheehan, J .C.; Ledis, S.L., J. Am. Chem. Soc., 1973, 95, 875. Aebersold, R.; Pipes, G.D., Wettenhall, R.E.H.; Nika, H., Hood, L.E., Anal. Biochem., 1990, 187, 56-65. Jensen, N.J.; Tomer, K.B.; Gross, M.L., Anal. Chem., 1985, 57, 2018- 2021. McLafl'erty, F.W., Ed. Tandem Mass Spectrometry, John Wiley and Sons, New York, N.Y., 1983. Jennings, K.B.; Dolnikowski, G.G., Methods in Enzymology, 1990, 193, 37-61. Chapter III DEVELOPMENT OF ON-LINE ETHYL-TRIPHENYLPHOSPHONIUM DERIVATIZATION A. Introduction Analyte derivatization is a necessary evil for many chromatographic separations to increase the detectability of targeted solutes. Advances in the design of faster and more selective reactions, synthesis of new derivatizing reagents, and development of novel derivatization methods for almost every type of compound have occurred in the last several years. Postcolumn derivatization offers inherent advantages over precolumn methods in terms of precision and ease of automation; a single reaction pathway is not necessary. However, these schemes require more care in design to minimize extracolumn broadening, and they require reasonably fast reaction kinetics [1]. Derivatization can be performed by homogeneous solution reactions or heterogeneously with a solid phase reagent. Solid phase reactors employing enzymes or chemical reagents are becoming increasingly p0pular, especially in combination with flow injection analysis because the disadvantages associated with homogeneous reactions can be overcome with use of a solid phase reagent [2]. Krull’s group has been especially active in the area of solid phase derivatizations. His group has reported the use of polymeric agents for weak nucleophiles, including primary and secondary alcohols [3], amines, amino alcohols, and amino acids [4]. They also devised a mixed bed reactor that contains several different reagents to simultaneously prepare derivatives of several functional groups on a single 105 106 analyte [5]. A mixture of silica gel and a polymeric anhydride containing the o-acetylsalicyl group as the labeling moiety is useful for the simultaneous collection and derivatization of airborne primary aliphatic amines with a solid phase reactor [6]. Two different polymeric reagents to derivatize fatty acids have also been described [7 ,8]. By placing a solid phase reactor on-line with a high performance liquid chromatography (HPLC) column and a mass spectrometer, the possibility of injecting picomole amounts of an enzymatic digest into this apparatus and deducing a protein's entire amino acid sequence just hours later is foreseeable. This is the ultimate goal for developing a system that will combine a reactor for ethyl-TPP derivatization, HPLC separation, and FAB-CAD-MS/MS. Further development would include an enzyme reactor that could perform enzymatic digestion of proteins on-line [9]. Vestal et al. reported the HPLC-thermospray MS detection of peptides resulting from digestion in a reactor containing immobilized enzymes [1 0-1 3]. Voyksner et al. [14] reported a 1 0 to 40-fold signal enhancement for peptides obtained by hydrolysis in a enzyme reactor over unhydrolyzed peptide. The increased sensitivity allowed detection and qunatitation of hydrolysis products down to the 800 fmol level. On-line coupling of HPLC and mass spectrometry (LC-MS) has already been realized and has made a major contribution to analytical methodology in protein chemistry [1 5]. One LC-MS configuration utilizes a sample introduction probe that provides a continuous flow of solvent that may contain sample into the mass spectrometer ion source and onto a frit where FAB ionization takes place. The mobile phase flow rate is in the range of 3-10 uL’min, generally. For peptide analysis with LC-MS, 1-5% glycerol is added to the carrier solution to maintain stable operating 107 conditions. There are two advantages for using continuous flow CF-FAB- MS in preference to FAB with a direct probe: i) lower limits of detection are possible (the low picomole level for peptides) and ii) there is a decreased ion suppression effect. Typically, packed, capillary-bore, fused-silica columns are employed with LC-MS because they offer good separation and increased sensitivity at low flow rates (3-10 [LL/min). although they are fragile and dificult to work with. Complex mixtures of tryptic peptides routinely are analyzed with packed capillary columns coupled on-line with FAB-MS at picomole levels [16]. Packed capillary columns allow all the injected peptide to be introduced into the mass spectrometer and consequently are advantageous for separating complex mixtures of peptides derived from the enzymatic digestion of proteins. Enzymatic digestion in combination with LC-MS quickly verifies the primary structure of recombinant proteins by a technique called peptide mapping. When tandem MS is utilized in combination with HPLC, identification of posttranslational modifications of peptides and single amino acid substitutions is possible. Improved sensitivity, minimized suppression of hydrophilic peptides, and efficient sample introduction are just some of the advantages of coupling HPLC online with FAB-MS. B. Advantages of an on-line derivatization reactor In general, the advantages already realized with the use of solid phase reagents for sample derivatization in combination with chromatography include: i) fast, mild, and efficient derivatization, ii) improved reagent stability, iii) increased selectivity and production of fewer side products, iv) less contamination caused by an excess of derivatization reagent, v) higher reaction capacities due to high 108 concentration of derivatizing reagent, vi) ease of regenerating the solid phase reagent, and vii) the option of performing on-line derivatization pre- or post-column. In some cases, it is possible to filter and reuse the derivatizing reagents. Krull and co-investigators have developed several reactors for on-line derivatization of amines employing a derivatizing agent covalently attached to a polymeric support [1 7-1 9]. Covalent linkage of the derivatizing agent to a polymer provides the best stability for the reagent and causes the polymer-bound derivatizing reagent to be compatible with a larger range of solvents. Derivatization yields near 90% for primary amines and 70% for secondary amines were realized with these reactors. Detection limits in the low parts-per-billion range were achieved in conjunction with improved sample separation, achieving rapid, accurate, and precise quantitation of amines in real-world sample matrices. The proposed setup for on-line, ethyl-TPP derivatization in combina- tion with LC-MS has several advantages beyond those common to solid phase derivatization. Because only low-picomole amounts of peptides are available for analysis in many cases, the most important advantage of combining peptide derivatization, separation, and mass analysis on-line is the reduced sample losses resulting from elimination of the sample transfers necessary if these procedures were carried out manually. The proposed setup should provide high derivatization yields, low detection limits, increased sensitivity, fewer interferences, and faster quantitation than is available by manual performance of these procedures. Further, because these procedures will be automated, they can be performed quickly and with decreased labor input. 109 The second important advantage of an apparatus to perform peptide derivatization with the ethyl-TPP moiety on-line is the increased amount of information generated with analysis by FAB-CAD-MS/MS of the derivatized peptide beyond that obtained for the underivatized peptide. Covalently linking the ethyl-TPP group to a peptide increases its hydrophobicity and places a positive charge at a fixed position on the peptide. Therefore, ethyl-TPP derivatization enhances the efficiency of ionization for hydrophilic peptides and generates a series of ions from the terminus at which the peptide has been modified. Fixing a positive charge at the N-terminus or C-terminus facilitates the production of d" or w" ions, respectively, which helps distinguish amino acid residues in the peptide; these ions are especially helpful in difl'erentiating the isomeric amino acids leucine and isoleucine. C. Proposed setup The proposed apparatus for on-line derivatization of peptides consists of three major components: a reactor, a packed microbore LC column, and a mass spectrometer capable of LC-MS (Figure 3.1 ). A microsyringe pump will supply the solvent at low flow rates (3-10 uL/min). The sample of peptide fragments obtained from an enzymatic digest will be introduced to the derivatization reactor through an injector loop. Once inside the reactor, the peptides will be tagged with the ethyl-TPP moiety. The derivatized peptides then will pass through a reverse phase, packed microcolumn for chromatographic separation. If the derivatization reactor is pre-column, the derivatized peptides will be more homogeneous in hydrophobicity and chromatographic properties, probably requiring a shallower solvent gradient for separation. Following separation, the presence of derivatized peptides will be indicated by an on-line UV-Vis 110 Reactor LC Column X0919." Figure 3.1. Proposed on-line, ethyl-TPP derivatization setup. The most effective design probably will perform solid phase, precolumn derivatization prior to LC separation and analysis by FAB-MS. 1 l l detector before they pass into the mass spectrometer. A mass spectrometer adapted for sample analysis by LC-MS (a JEOL HX-llO modified with a frit-FAB interface) will determine the mass of the derivatized peptides when operated in the FAB ionization mode. Product ions resulting from CAD-MS/MS will aid in determining the amino acid sequence of the derivatized peptides. A polymer-bound reagent containing the ethyl-TPP group will be the active component of the reactor causing peptide derivatization. Functionalized polymers have found widespread use in organic synthesis and related fields [20]. They are employed as reagent carriers, ion exchange resins, and matrices for immobilized enzymes and cells [21]. Styrene resins containing more than fifty percent vinyl benzene are employed for HPLC applications because they can be safely subjected to pressures above 1000 psi [22]. Polymer resins containing immobilized reagent are beneficial for use in derivatization procedures because they can be recycled repeatedly and conveniently, can be modified to contain a large excess of reagent to drive reactions to completion, are compatible with a wide range of solvents, and do not leach. Krull and coworkers have reported the development of several reactors containing polymer-bound magenta for on-line derivatizations; these immobilized reagents were stable during more than fifty injection periods before regeneration was necessary [17-19, 23]. D. Chemical methodology The overall proposed procedure for preparing the immobilized reagent for on-column derivatization of peptides is represented in Scheme 3.1 . This procedure utilizes a pre-made polymer support covalently linked to o-nitrophenol. The o-nitrophenol group will be reacted with the “Kunz” 112 .uaowaou .uasM. on... 853 consensus». on a? 93:8 25. 633% ca... mfinflwzuov accuse." 333 330350 on... 4:3 338 333m .95 Safioqcnmoufid-e.a§umx on... Show 3 “oaongefidé panacea fit» Beach a accuse.“ gun—am. dofleuflwzuov name—.33 .393 v38 .Sm c8058 .8533 33825 A.» sacs—om @ AHHHV E Fm “Sig + ©I+mlumolamoloI_o_l:¢_aln_.lolvom _ __ Noz Q 0 mm 0 m.2... a mlefizlwlolvom ___ no A: ©Iflvalnmolamo I o IoLo fluu CC + mlamoiumoIOIolo RM a ©I+ __ oawuonmmom Q o No Q 6 a ©l+mlumolamoIOIol6 + no : e . Q 113 reagent [2-(triphenylphosphonio)ethylchloroformate chloride] in a nucleophilic acyl substitution reaction that eliminates HCl and immobilizes the “Kunz” reagent. The immobilized o-nitrophenolate of the “Kunz” reagent and its polymeric support then will be packed into a column. When the polymeric reagent comes in contact with peptides dissolved in the mobile phase containing peptides is pumped through the column, the N- terminal amine group of peptides will act as a nucleophile in a nucleophilic substitution reaction to produce peptide derivatized with the ethyl-TPP group. The starting reagent, immobilized o-nitrophenol, will be the other product of this reaction. The column reactor can be regenerated quickly by simply passing a solution containing the “Kunz” reagent through the reactor bed. E. Separation of a peptide mixture by LC-MS 1. Experimean The analysis of a mixture containing approximately 1 00 pmol of four peptides (GLA, YGGFL, VGVAPG, and VQAADYING) was performed by LC-MS with a JEOL HX-110 modified to house a frit-FAB interface. The magnet was scanned from m/z 50 to m/z 1000 every 10 seconds. Two buffers were used to introduce the sample into the mass spectrometer. Bufi'er'A contained 98.4% H20, 1.5% glycerol, and 0.1% TFA; buffer B contained 98.4% ACN, 1.5% glycerol, and 0.1% TFA. A gradient of 040% B in 10 minutes, 40 to 100% B from 10 to 60 minutes, and 100% B from 60 to 70 minutes achieved chromatographic separation of the peptide mixture with a fused-silica microbore column [0.32 x 300 mm, C13, 5 pm, capillary outlet; LC Packings (U.S.A.), Inc.; San Francisco, CA]. The solvent flow rate rate was 10 uL/min as delivered by a syringe pump (Brownlee Labs; Santa Clara, CA). The mobile phase was monitored post-column with a 113 reagent [2-(tripheny1phosphonio)ethylchloroformate chloride] in a nucleophilic acyl substitution reaction that eliminates HCl and immobilizes the “Kunz” reagent. The immobilized o-nitrophenolate of the “Kunz” reagent and its polymeric support then will be packed into a column. When the polymeric reagent comes in contact with peptides dissolved in the mobile phase containing peptides is pumped through the column, the N- terminal amine group of peptides will act as a nucleophile in a nucleophilic substitution reaction to produce peptide derivatized with the ethyl-TPP group. The starting reagent, immobilized o-nitrophenol, will be the other product of this reaction. The column reactor can be regenerated quickly by simply passing a solution containing the “Kunz” reagent through the reactor bed. E. Separation of a peptide mixture by LC-MS 1. Experimental The analysis of a mixture containing approximately 100 pmol of four peptides (GLA, YGGFL, VGVAPG, and VQAADYING) was performed by LC-MS with a JEOL HX-110 modified to house a frit-FAB interface. The magnet was scanned from m/z 50 to m/z 1000 every 10 seconds. Two buffers were used to introduce the sample into the mass spectrometer. Bufl'erA contained 98.4% H20, 1.5% glycerol, and 0.1% TFA; buffer B contained 98.4% ACN, 1.5% glycerol, and 0.1% TFA. A gradient of 0-40% B in 10 minutes, 40 to 100% B from 10 to 60 minutes, and 100% B from 60 to 70 minutes achieved chromatographic separation of the peptide mixture with a fused-silica microbore column [0.32 x 300 mm, C13, 5 pm, capillary outlet; LC Packings (U.S.A.), Inc.; San Francisco, CA]. The solvent flow rate rate was 10 uL/min as delivered by a syringe pump (Brownlee Labs; Santa Clara, CA). The mobile phase was monitored post-column with a 1 14 Kratos Spectroflow 783 UV-Vis detector (Kratos; Foster City, CA) before entering the mass spectrometer. 2. Results and discussion The UV-Vis trace obtained with injection of the four-component peptide mixture is shown in Figure 3.2. In figure 3.3 a) is a total ion current (TIC) plot obtained when the peptides entered the mass spectrometer. Data acquisition was begun when the fastest-running peptide was detected by the UV-Vis detector. Figure 3.3 b) contains the reconstructed mass chromatograms for the individual protonated peptides. The order of elution was VGVAPG (mlz 499), GLA (m/z 260), VQAADYING (mlz 950), and finally YGGFL (mlz 556). The background-subtracted mass spectra of the peptides are contained in Figure 3.4. The peptides all display good signal-to-noise ratios, with protonated GLA represented by the base peak in Figure 3.4 b). Although the UV-Vis trace shows the chromatographic separation of the reaction mixture was not baseline resolved, Figure 3.4 demonstrates that individual peptides can be detected by the mass spectrometer. The performance of the JEOL HX-110 modified with a fiit-FAB interface demonstrates that it is adaptable for analysis of peptide mixtures by LC-MS. F. Reaction in homogeneous solution To investigate the feasibility of the proposed reaction sequence (Scheme 3.1), the chemical reactions were performed in homogeneous solution with non-mobilized reagents according to Scheme 3.2. This allows the feasibility of the reaction to be investigated quickly and conveniently without actually fabricating the column to be used for on-column derivatization. 115 ;- - e g. .. A 4 L g- _ g fiM 4' __ atsgéégazéigfiuén Time(min) Figure 3.2. UV-Vis trace of the four-component peptide mixture containing 100 pmol of VGVAPG, GLA, YGGFL, and VQAADYING following separation with a microbore column, prior to analysis by fiit-FAB- MS. 116 RetentionTime(min) -9' r-§.- .19. -.12. . -15. (a) N ..s 8 O 49 M V gnu-ugoogr—q m<-¢+p~raw wnwnpanaH ocwnn-‘o , A 1 1 l 1 l» l I I if I 1 on O a ? 55 0A-- - - ~‘ ---- - - m/z950 h 132 m/z 556 A x2.7 m/z499 JL 11.8 f - j 1 - ' - - . j“ - f - . - ‘ - - M260 100 200 300 400 Scannumber Figure 3.3. Reconstructed total ion current (TIC) plot (a) obtained when the peptide mixture entered the mass spectrometer. Scanning was commenced when the first-running peptide was detected by the UV-Vis detector positioned prior to the mass spectrometer. Figure 3.3 (b) contains the reconstructed mass chromatograms of the individual components of the peptide mixture. 117 (a) 100 a‘ Sea 162 . n . 1 V-G-V-A-P-G "12%” I 80‘ t 601 p . 134 ’ 40‘ 1 3 _ T %"fl'flflflflgt—i o<~.”’~°w ffi T' Y 2“ 256 1 al. 7 100 ll kg“ Bindiil 14H. 1 1 . (b) : Scan 185 (£33,. : G-L-A 80‘ . 40‘ p- 92 20" 171 . Killlfljn if ' J L .444?9_. A r a - a? J l Scans182-194 i V-Q-A-A-D-Y-I-N-G . wow-"nuanur-n o<-on~ow ..__9__-—" ‘— H P 100 fir “‘ 10?— w (c) on -?-- a: -?-- r 2): .5 O 20 1 7 A” :°-: «139%»- 80 “8... ' U 1 V V j v u ' Vflfllflflf’flh‘ O<"'f’D—‘¢Dw H F" N 8---?---.-- I; (d) Scan 231 . Y-G-G-F-L . _ wen-nonuv—n saw-«p—ow ml: 600 80° . ...... Figure 3.4. The background-subtracted mass spectra of individual peptides (a-d) eluted during LC-MS. The scan rate was one scan every 10 seconds. 118 a Q @013 . m—c—o—cn,—cn,—p+@ (1) m/z 369 a <3 @o-c-o-cnz—cnz—(Crg + HCl (II) m/z472 fl 3‘ Ho(-—c—c|:—N)n—H 1 En; o H 0 Q No2 II I II + Ho(—c—<':—1~II)n—c-o—CHZ—CH2—P-© + @011 “n“ (III) Scheme 3.2. Homogenous-phase chemistry simulating the on-line reaction sequence. N onmobilized o-nitrophenol reacts with the ‘Kunz’ reagent (1) to produce the “Kunz”-o-nitrophenolate (II). This product (II), in turn, derivatizes the peptide to yield the “Kunz” modified peptide (III). 1 1 9 1. Formation of the “Kunz" reagent o-nitrophenolate a) Experimental A five-fold molar excess of o-nitrophenol (Aldrich Chem. Co.; Milwaukee, WI) and 0.1 uL of 2,4,6-trimethylpyridine (TMP) to act as a weak base were dissolved in 20 pL acetonitrile. One milligram of “Kunz” reagent (Fluka Chemika-BioChemika; Buchs, Switzerland) was added to the solution. The reaction mixture was vortexed for l 0 sec and 1 51L of the solution was placed on the FAB probe tip together with a 1:1 (v/v) glycerol/thioglycerol matrix. The mass spectrum was recorded after 30 sec of FAB bombardment with the conversion dynode voltage at -3.8 kV. A study of the kinetics for the reaction of the “Kunz” reagent with o- nitrophenol and stability of the product was performed by vortexing the reaction mixture for 5, 10, 20, 45, and 90 sec and measuring the ratio of peak intensity of the product (m/z 472) to that of (2-hydroxyethyl)TPP (m/z 307), which is a side product and also a product of degradation of the o- nitrophenolate of the “Kunz” reagent with H20. The mass spectra for this investigation were acquired 30 sec after FAB was commenced. All trials were performed in triplicate. b) Results and discussion Because the “Kunz” reagent and the reaction products contain positively charged triphenylphosphonium groups, the reaction was conveniently monitored by FAB-MS. The mass spectrum of the reaction mixture resulting from vortexing a solution containing a five-fold molar excess of o-nitrophenol and 1 milligram of “Kunz” reagent in the presence of trimethylpyridine (TMP) for 10 see is shown in Figure 3.5 a). The base peak (m/z 472) represents the desired product of this reaction (Compound 11, Scheme 3.2), which forms quickly and in high yield. The peaks at m/z 120 H O P (a) on O 14%; . l O) O A ‘31.. 122 M 0 & L A a l A a a 92 1 08 wnwnuonaH o HaN-MfldO-C-OH + co, o=c\ Kit *8; Scheme 3.3. Suggested reaction pathway for decarboxylation of the peptide derivatized with the “Kunz” reagent to form an ethyl-TPP derivatized product. 1 27 structure than the original product and is identical to peptides modified with the ethyl-TPP group by vinyl-TPP. The fragmentation pattern of the decarboxylated product of the AAA methyl ester produced by this reaction sequence (Scheme 3.2) is identical to the vinyl-TPP derivatized AAA methyl ester as the FAB-CAD-B/E spectra demonstrate (Figure 3.8 a) and b)). The yields of derivatization with the o-nitmphenolate of the “Kunz” reagent in the presence of TEA for 32 min and derivatization with vinyl-TPP in a 1 :1 pyridine/acetonitrile solvent mixture for ~8 hrs at 37°C are quite similar as Figures 3.9 a) and b) demonstrate. The reaction kinetics of the simulated on-line derivatization reaction under different conditions are shown in Figure 3.10. The extent of the reaction was determined by measuring the ratio of derivatized, decarboxylated AAA methyl ester peak intensity (mlz 534) to underivatized AAA methyl ester (mlz 246 represents the protonated molecule of AAA methyl ester). As expected, faster reaction kinetics were observed at higher temperatures (37°C vs. 25°C) and in the presence of TEA as a base catalyst. This study indicates the optimal derivatization yield would be produced with the use of a base catalyst (such as TEA) and heating of the reaction mixture. The practicality of this reaction sequence for higher molecular weight peptides was investigated by performing the reaction with the hexapeptide VGVAPG to simulate on-line, ethyl-TPP derivatization. Although this peptide contains an unprotected carboxyl group at the C- terminus, the reaction proceeded in high yield as Figure 3.11 a) demonstrates. The FAB-CAD-B/E spectrum of the “Kunz” derivatized, decarboxylated product of VGVAPG (Figure 3.11 b)) is similar to that of the peptide derivatized at the N-terminus with the ethyl-TPP group by vinyl- TPP, indicating the products are identical (compare Figure 2.21). 128 20- R « . (M)+ 2!? f (a) a 1 t15‘ ' 1 . 262 V e O 110 Q :1 -©-P+ CH, CH, A A A (ll—OCH, 289 . : a n . 8 5- I . t al Y 1E5 l ”I (M-0020H3)+ 100 200 m/z 300 400 500 1" : (b) 277 a 1 t15- i . 262 V 1 e 1 110 n . 289 t . O . n t 9 5- 1 1 ., y . 1E5 l 1' (M-COzCI-Isr' '160‘ ' ' '260' " 360' ' ' '460' ' ' 560 m/z Figure 3.8. FAB-CAD-B/E spectrum (a) of 1 nmol of AAA methyl ester derivatized with vinyl-TPP and (b) FAB-CAD-B/E spectrum of 1 nmol of AAA methyl ester derivatized with the “Kunz”-o-nitrophenolate.s 129 93 ...: O O (a) .‘é. . 3. - 8.. wewusoeuH o