The Enethiolate Anion Reaction Products of EpiD
pKa VALUE OF THE ENETHIOL SIDE CHAIN IS LOWER THAN THAT OF THE THIOL SIDE CHAIN OF PEPTIDES*

(Received for publication, November 18, 1996)

Thomas Kupke Dagger and Friedrich Götz

From the Mikrobielle Genetik, Universität Tübingen, Waldhäuserstrasse 70/8, 72076 Tübingen, Germany

ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
FOOTNOTES
Acknowledgments
REFERENCES


ABSTRACT

One of the steps involved in the biosynthesis of the lantibiotic epidermin is the oxidative decarboxylation reaction of peptides catalyzed by the flavoenzyme EpiD. EpiD catalyzes the formation of a (Z)-enethiol derivative from the C-terminal cysteine residue of the precursor peptide of epidermin and related peptides. The UV-visible spectra of the reaction products of EpiD are pH-dependent, indicating that the enethiol side chain is converted to an enethiolate anion. The pKa value of the enethiol group was determined to be 6.0 and is substantially lower than the pKa value of the thiol side chain of cysteine residues. The increased acid strength of the enethiol side chain compared with that of the thiol group is attributed to the resonance stabilization of the negative charge of the anion.


INTRODUCTION

Several new posttranslational modification reactions, such as dehydration of serine and threonine residues, thioether formation, and formation of D-alanine residues from L-serine residues, are involved in the biosynthesis of the ribosomally synthesized lantibiotics (reviewed in Refs. 1 and 2). The lantibiotic epidermin contains the alpha ,beta -unsaturated amino acid didehydroaminobutyric acid, S-[(Z)-2-aminovinyl]-D-cysteine, 3-methyl-lanthionine, and meso-lanthionine (3). The formation of epidermin from the precursor peptide EpiA (4) includes the oxidative decarboxylation of the C-terminal cysteine residue of EpiA to a (Z)-enethiol catalyzed by the FMN-containing enzyme EpiD (5, 6). Two reducing equivalents from the C-terminal cysteine residue are removed, a double bond is formed, and the coenzyme FMN is reduced to FMNH2. The decarboxylation occurs spontaneously or is also catalyzed by EpiD. The (Z)-enethiol derivative is the intermediate in the formation of the C-terminal S-[(Z)-2-aminovinyl]-D-cysteine residue of epidermin (6). The unusual enethiol structure has been confirmed by using mass spectrometry, tandem mass spectrometry, UV-visible spectroscopy, two-dimensional NMR spectroscopy, and conversion of the enethiol to a mixed disulfide with 5,5'-dithiobis(2-nitrobenzoic acid) (Ellman's reagent (7, 8)). EpiD has a low substrate specificity, and most of the peptides with the sequence [V/I/L/(M)/F/Y/W]-[A/S/V/T/C/(I/L)]-C at the C terminus are substrates of EpiD, as elucidated by analysis of the reaction of EpiD with single peptides and peptide libraries, respectively (7).

Here we report the further characterization of the reaction products of EpiD. We demonstrate that at physiological pH, the dominant form of the reaction product is the enethiolate anion and that the pKa value of the enethiol group is 6.0. We assume that the enethiolate anion is strongly nucleophilic and is the substrate for the Michael addition-like formation of S-[(Z)-2-aminovinyl]-D-cysteine. Furthermore, we discuss the possibility that enzymes exist in nature that catalyze the formation of enethiol groups from non-C-terminal cysteine residues.


EXPERIMENTAL PROCEDURES

Synthesis and Purification of SFNSYV-NH-CH=CH-SH and SFNSYT-NH-CH=CH-SH

SFNSYVC and SFNSYTC were synthesized as described previously (7). For generation of the enethiol derivative of SFNSYVC, approximately 25 mg of peptide were dissolved in 50 ml of 20 mM Tris/HCl (pH 8.0) containing 4 mM DTT1 and 1 mg of MBP-EpiD (5). For generation of the enethiol derivative of SFNSYTC, approximately 12 mg of peptide and 500 µg of MBP-EpiD were used. After 25 min of incubation at 37 °C, 500 µl of trifluoroacetic acid was added. The reaction mixture was then separated by reversed-phase chromatography using a ProRPC HR10/10 column (Pharmacia Biotech Inc.). Peptides were eluted with a linear gradient of 0-80% acetonitrile, 0.1% trifluoroacetic acid in 150 ml at a flow rate of 2 ml/min. Peptides with a high absorbance at 260 nm were collected and diluted with H2O, 0.1% trifluoroacetic acid. DTT was added to 3 mM, and the samples were subjected to a second reversed-phase chromatography using a µRPC C2/C18 SC 2.1/10 column (Pharmacia). Peptides were eluted with a linear gradient of 0-50% acetonitrile, 0.1% trifluoroacetic acid in 5.8 ml at a flow rate of 200 µl/min. The absorbance was measured simultaneously at 214, 260, and 280 nm. The modified peptide was identified by the high A260/A214 ratio and collected by the peak fractionation method using the SMARTTM system (Pharmacia, Freiburg, FRG). Samples were dried with a vacuum concentrator (yields: approximately 4.4 mg of SFNSYV-NH-CH=CH-SH and 1 mg of SFNSYT-NH-CH=CH-SH, respectively) and stored at -70 °C. The modified peptides were analyzed by electrospray mass spectrometry as described recently (7).

UV-visible Spectroscopy

UV-visible spectra were obtained with a Beckman DU 7500 spectrophotometer. Spectral data were recorded in the range of 200-600 nm at 21 °C at various pH values in quartz cuvettes (1-cm light path; the reference cuvette contained the used buffers and on occasion 4 mM DTT). For determination of the pKa, 0.6-ml aliquots of a stock solution of the oxidatively decarboxylated peptide SFNSYV-NH-CH=CH-SH were mixed with 0.15 ml of 50 mM sodium phosphate (pKa = 2.15, 7.20, and 12.38), 50 mM sodium citrate buffer (pKa = 3.13, 4.76, and 6.40) (final buffer concentration was 10 mM each for phosphate and citrate). The pH of the buffers was adjusted by adding HCl. pH values of the 10 mM phosphate/citrate buffers and of the buffered peptide solutions were determined using a Knick pH meter (761 Calimatic).


RESULTS

pH Dependence of UV-visible Spectra of Oxidatively Decarboxylated Peptides

The reaction products of the flavoenzyme EpiD are oxidatively decarboxylated peptides with a C-terminal enethiol side chain (6-8). Recently, we demonstrated that the reaction products of EpiD have an absorption maximum at 260 nm when measured in 0.1% trifluoroacetic acid, 30% acetonitrile, 70% H2O (6, 7). In order to determine the acid strength of the -SH group of enethiol-containing peptides, we investigated whether the UV-visible spectra of SFNSYV-NH-CH=CH-SH and SFNSYT-NH-CH=CH-SH are pH-dependent. It is well established that the thiolate anion of cysteine can be detected by its absorbance at 240 nm (9, 10) and that the pKa value of the -SH group of cysteine-containing peptides is in the range of 8.5-9.5.

The UV-visible spectrum of the peptide SFNSYV-NH-CH=CH-SH purified by reversed-phase chromatography was determined in the pH range 2.5-8.0. To ensure that spectral changes were not due to oxidation of the peptide to the dimer, the UV spectra were recorded in the presence of 4 mM DTT. Interestingly, the UV spectrum was pH-dependent (Fig. 1) with an absorption maximum of 259 nm under acidic conditions (pH <=  4.2) and of 283 nm at pH 7.0. Due to ionization of the tyrosine residue, the absorbance increased significantly below 260 nm at pH 8.0; however, there was no further shift of the absorption maximum to wavelengths greater than 283 nm (data not shown). Similar pH-dependent spectral changes were also observed for the peptide SFNSYT-NH-CH=CH-SH (data not shown but see Fig. 2).


Fig. 1. UV-visible spectra of SFNSYV-NH-CH=CH-SH at various pH values. A, 0.145 mM SFNSYV-NH-CH=CH-SH was placed in the sample cuvettes in 10 mM phosphate/citrate buffer at the indicated pH values and containing 4 mM DTT. The proposed ionization of the enethiol group to the enethiolate anion is shown. The absorption maxima of the enethiol form (259 nm) and of the enethiolate form (283 nm) are indicated; the isosbestic point is 271 nm. Similar spectra were obtained in the absence of DTT (data not shown). B, the absorbance values from A at 283 nm (filled circles) are shown as a function of the pH value. The dotted line represents the fit using the Henderson-Hasselbalch equation with pKa = 5.97. If the absorbance values at 320 nm (very low absorbance of the present aromatic amino acids Phe and Tyr) were used for curve fitting the obtained pKa was 6.02.
[View Larger Version of this Image (19K GIF file)]



Fig. 2. Analysis of the reaction products of SFNSYTC by reversed-phase chromatography using acidic and neutral buffer systems. The reaction of MBP-EpiD with SFNSYTC was investigated by separation of the reaction mixture on a Pharmacia µRPC C2/C18 SC 2.1/10 column under acidic (A1, no enzyme; A2, with MBP-EpiD) (buffer A, 0.1% trifluoroacetic acid, H2O; buffer B, 0.1% trifluoroacetic acid, acetonitrile) and neutral buffer conditions (B1, no enzyme; B2, with MBP-EpiD) (buffer A, 25 mM of TEAA/H2O; buffer B, 25 mM of TEAA/20% H2O, 80% acetonitrile). The assay (total volume of 2 ml) was carried out in 20 mM Tris/HCl (pH 8.0) containing 3.5 mM DTT, 100 µg of the peptide, and approximately 10 µg of MBP-EpiD. The reaction was stopped after 30 min of incubation at 37 °C by adding 20 µl of trifluoroacetic acid. Aliquots of the reaction mixture were applied to reversed-phase chromatography using acid and neutral buffer conditions, respectively. After washing the reversed-phase column with 2 ml of buffer A, the peptides were eluted with a linear gradient of 0-50% buffer B in 5.8 ml at a flow rate of 200 µl/min. The elution was followed by absorbance at 214 nm (upper thin line), 260 nm (thick line), and 280 nm (lower thin line), and the peak fractions were analyzed by electrospray mass spectrometry (determined m/z values are indicated in the panels; substrate peptide, 821 m/z; oxidatively decarboxylated peptide, 775 m/z).
[View Larger Version of this Image (28K GIF file)]


The pH dependence of the UV spectrum of the oxidatively decarboxylated peptide indicates the ionization of the enethiol form to the enethiolate anion form (Fig. 1, A and B). From the UV spectra, the apparent pKa value of the enethiol group was determined to be pH 6.0 (Fig. 1B). The molar extinction coefficient (epsilon ) of SFNSYV-NH-CH=CH-SH at 259 nm and of SFNSYV-NH-CH=CH-S- at 283 nm was determined to be at least epsilon  approx  6,800 and epsilon  approx  7,200 M-1 cm-1, respectively. However, this value may even be higher due to error in weighing out the small amounts of available peptide and due to the presence of sodium chloride and water in the peptide sample. Assuming that the peptides SFNSYVC and SFNSYV-NH-CH=CH-SH have the same epsilon 214 value, epsilon 259 of SFNSYV-NH-CH=CH-SH and epsilon 283 of SFNSYV-NH-CH=CH-S- were calculated to be approximately 10,000 M-1 cm-1.

Reversed-phase Chromatography of Enethiol-containing Peptides under Acidic and Neutral Conditions

In order to confirm the pH dependence of the UV spectrum of enethiol-containing peptides, the reaction mixture containing the substrate peptide SFNSYTC (821 m/z), the reaction product, and MBP-EpiD was analyzed by reversed-phase chromatography under acidic conditions (0.1% trifluoroacetic acid, approximately pH 2.0) and under neutral conditions (25 mM TEAA, pH 7.0, Fig. 2). The reaction product (775 m/z) was identified by high resolution electrospray mass spectrometry as a monomer (data not shown). Degradation products (741 and 759 m/z) of the reaction product were also present. The A260/A280 ratio of the reaction product was 2.3 in 0.1% trifluoroacetic acid/acetonitrile/H2O and 0.6 in 25 mM TEAA/acetonitrile/H2O (compare Fig. 2, A2 and B2). Comparable results were obtained for the reaction products of SFNSYVC and SFNSYCC (data not shown).


DISCUSSION

The determined pKa value of 6.0 for the enethiol group of SFNSYV-NH-CH=CH-SH is significantly (3 pH units) lower than the reported pKa value for the thiol group of cysteine residues in small peptides. Peptides with a C-terminal cysteine residue in which the negative charge of the carboxyl group is removed (cysteine amides, a negative charge in the vicinity of the thiol group decreases the acidity of the thiol group (11)), as is also the case for the investigated oxidatively decarboxylated peptides SFNSYV-NH-CH=CH-SH and SFNSYT-NH-CH=CH-SH, are reported to have a pKa value in the range of pH 8.8 (12). The increased acid strength of the enethiol group compared with the acid strength of the thiol group is explained by a resonance effect in which the negative charge on the sulfur anion is delocalized in the enethiolate. Delocalization of the negative charge is the reason for stabilization of enolate ions so that the acidity of enols is increased compared with alcohols (13).

For several enzymes, such as the protein thiol-disulfide oxidoreductase DsbA from Escherichia coli, human glutathione transferase, and cysteine peptidases, it has been shown that the catalytic active cysteine thiol groups have a pKa value near 4.0 (14-16). This low pKa value is partially explained by stabilization of the thiolate anion by interaction with histidine or lysine side chains in the tertiary structure of the proteins. It is very unlikely that the peptides SFNSYV-NH-CH=CH-SH and SFNSYT-NH-CH=CH-SH have a defined three-dimensional structure and that this structure is responsible for stabilization of the enethiolate anion.

There is a strong correlation between the low pKa value of the thiol group of Cys30 of DsbA and the redox potential of DsbA (17). The two cysteine residues of the active site Cys30-Pro31-His32-Cys33 form a disulfide bridge, and disulfide-bonded DsbA is a potent oxidant that generates protein disulfide bonds (18, 19). Because of the correlation between the pKa value of the thiol group and the redox potential, we are interested in determining the redox potential of peptides containing both a cysteine residue and a C-terminal enethiol group (e.g. oxidatively decarboxylated precursor peptide EpiA), and we are interested in investigating the reaction between protein thiol-disulfide oxidoreductases and such substrates.

As known for thiols (20), we assume that enethiols are far more reactive as the enethiolate anions. Therefore, reasonable chemical intuition suggests that the intermediate in the Michael addition-like formation of the S-[(Z)-2-aminovinyl]-D-cysteine residue of epidermin is the enethiolate and not the enethiol form of posttranslationally modified precursor peptide EpiA. Furthermore, we postulate that in the case of lanthionine formation the thiolate anions of the involved cysteine residues are stabilized by interaction with positive charges within the precursor peptide or by interaction with the enzyme responsible for the thioether formation (EpiC and/or EpiB). Glutathione transferase catalyzes a similar reaction, the nucleophilic addition of glutathione to xenobiotic compounds containing an electrophilic center. It has been discussed that the pKa of the thiol group of glutathione is shifted from 9.0 in aqueous solution to 6.6 in the active site of glutathione transferase so that the predominant species in the active site is the thiolate anion of glutathione (21).

A fascinating question is whether enzymes exist in nature that catalyze the formation of enethiols/enethiolates from cysteine residues that are not located at the C terminus. Principally, such a reaction can be involved in the formation of the thiazole backbone modification of the gyrase inhibitor microcin B17, whose structure has been elucidated recently (22, 23) (Fig. 3). Enzymes involved in microcin B17 modification have not been characterized, and there is no experimental evidence for the existence of the proposed enethiol-forming activity in microcin B17 modification. By converting the thiol side chain of cysteine residues in peptides and proteins to enethiol/enethiolates the (pH dependence of the) activity and stability of proteins may be influenced to a large extent. However, nothing is known about the stability of these enethiol side chains. Peptides with a C-terminal enethiolate group are unstable, but they are stabilized in the presence of zinc ions (6-8). In addition to the flavoprotein EpiD, another enzyme catalyzing an alpha ,beta -dehydrogenation reaction of peptides has already been characterized; heme-containing L-tryptophan 2',3'-oxidase from Chromobacterium violaceum is responsible for the synthesis of alpha ,beta -dehydrotryptophan derivatives (24, 25).


Fig. 3. Alternative model for the thiazole backbone modification of microcin B17. The recently postulated mechanism for the posttranslational modification of -Gly-Cys-Gly- segments of the microcin precursor peptide to the thiazole backbone modification of microcin B17 involved the attack of the thiol group of cysteine on the peptide carbonyl of the preceding glycine residue, the elimination of water, and the oxidation (removal of H2 (22, 23)). Here we describe that, in principle, the oxidation can be the first reaction and that an enethiol is formed as the first intermediate. Due to the increased acidity of the enethiol, the reactive nucleophile is the enethiolate anion. A similar reaction is the equilibrium between thiamine (vitamin B1) and its enethiolate form (26, 27).
[View Larger Version of this Image (12K GIF file)]



FOOTNOTES

*   This work was supported by grants from the Deutsche Forschungsgemeinschaft (SFB 323). The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be hereby marked "advertisement" in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.
Dagger    To whom correspondence should be addressed. Fax: 49 7071 295937.
1    The abbreviations used are: DTT, dithiothreitol; MBP, maltose-binding protein; TEAA, triethylammonium acetate.

Acknowledgments

We thank Christoph Kempter for assisting in electrospray mass spectrometry and Karen A. Brune for critically reading the manuscript.


REFERENCES

  1. Sahl, H.-G., Jack, R. W., and Bierbaum, G. (1995) Eur. J. Biochem. 230, 827-853 [Abstract]
  2. Kupke, T., and Götz, F. (1996) Antonie Leeuwenhoek 69, 139-150 [Medline] [Order article via Infotrieve]
  3. Allgaier, H., Jung, G., Werner, R. G., Schneider, U., and Zähner, H. (1985) Angew. Chem. Int. Ed. Engl. 24, 1051-1053
  4. Kupke, T., Stevanovic, S., Ottenwälder, B., Metzger, J. W., Jung, G., and Götz, F. (1993) FEMS Microbiol. Lett. 112, 43-48 [Medline] [Order article via Infotrieve]
  5. Kupke, T., Stevanovic, S., Sahl, H.-G., and Götz, F. (1992) J. Bacteriol. 174, 5354-5361 [Abstract]
  6. Kupke, T., Kempter, C., Gnau, V., Jung, G., and Götz, F. (1994) J. Biol. Chem. 269, 5653-5659 [Abstract/Free Full Text]
  7. Kupke, T., Kempter, C., Jung, G., and Götz, F. (1995) J. Biol. Chem. 270, 11282-11289 [Abstract/Free Full Text]
  8. Kempter, C., Kupke, T., Kaiser, D., Metzger, J. W., and Jung, G. (1996) Angew. Chem. Int. Ed. Engl. 35, 2104-2107
  9. Benesch, R. E., and Benesch, R. (1955) J. Am. Chem. Soc. 77, 5877-5881
  10. Noda, L. H., Kuby, S. A., and Lardy, H. A. (1953) J. Am. Chem. Soc. 75, 913-917
  11. Crampton, M. R. (1974) in The Chemistry of the Thiol Group (Patai, S., ed) Part 1, pp. 379-415, John Wiley & Sons, Inc., New York
  12. Kortemme, T., and Creighton, T. E. (1995) J. Mol. Biol. 253, 799-812 [CrossRef][Medline] [Order article via Infotrieve]
  13. Keeffe, J. R., and Kresge, A. J. (1990) in The Chemistry of Enols (Rappoport, Z., ed), pp. 399-480, John Wiley & Sons, Inc., New York
  14. Lo Bello, M., Parker, M. W., Desideri, A., Polticelli, F., Falconi, M., Del Boccio, G., Pennelli, A., Federici, G., and Ricci, G. (1993) J. Biol. Chem. 268, 19033-19038 [Abstract/Free Full Text]
  15. Nelson, J. W., and Creighton, T. E. (1994) Biochemistry 33, 5974-5983 [Medline] [Order article via Infotrieve]
  16. Roberts, D. D., Lewis, S. D., Ballou, D. P., Olson, S. T., and Shafer, J. A. (1986) Biochemistry 25, 5595-5601 [Medline] [Order article via Infotrieve]
  17. Grauschopf, U., Winther, R., Korber, P., Zander, T., Dallinger, P., and Bardwell, J. C. A. (1995) Cell 83, 947-955 [Medline] [Order article via Infotrieve]
  18. Zapun, A., Bardwell, J. C., and Creighton, T. E. (1993) Biochemistry 32, 5083-5092 [Medline] [Order article via Infotrieve]
  19. Bardwell, J. C. (1994) Mol. Microbiol. 14, 199-205 [Medline] [Order article via Infotrieve]
  20. Lundblad, R. L., and Noyes, C. M. (1984) Chemical Reagents for Protein Modification, CRC Press, Inc., Boca Raton, FL
  21. Graminski, G. F., Kubo, Y., and Armstrong, R. N. (1989) Biochemistry 28, 3562-3568 [Medline] [Order article via Infotrieve]
  22. Bayer, A., Freund, S., Nicholson, G., and Jung, G. (1993) Angew. Chem. Int. Ed. Engl. 32, 1336-1339
  23. Bayer, A., Freund, S., and Jung, G. (1995) Eur. J. Biochem. 234, 414-426 [Abstract]
  24. Genet, R., Denoyelle, C., and Ménez, A. (1994) J. Biol. Chem. 269, 18177-18184 [Abstract/Free Full Text]
  25. Genet, R., Bénetti, P.-H., Hammadi, A., and Ménez, A. (1995) J. Biol. Chem. 270, 23540-23545 [Abstract/Free Full Text]
  26. Tee, O. S., Spiropoulos, G. D., McDonald, R. S., Geldart, V. D., and Moore, D. (1986) J. Org. Chem. 51, 2150-2151
  27. Maier, G. D., and Metzler, D. E. (1957) J. Am. Chem. Soc. 79, 4386-4391

©1997 by The American Society for Biochemistry and Molecular Biology, Inc.