Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 44
Filter
1.
Hum Vaccin Immunother ; 15(3): 549-559, 2019.
Article in English | MEDLINE | ID: mdl-30689507

ABSTRACT

BACKGROUND: Two new formulations of an investigational 15-valent pneumococcal conjugate vaccine (PCV15-A and PCV15-B) were developed using 2 different protein-polysaccharide conjugation processes and evaluated in separate phase I/II studies (NCT02037984 [V114-004] and NCT02531373 [V114-005]) to assess optimal concentrations of pneumococcal polysaccharide (PnPs) and Aluminum Phosphate Adjuvant. METHODS: Various lots of PCV15-A and PCV15-B containing different concentrations of PnPs and/or adjuvant were compared to PCV13 in young adults and infants. Adults received single dose and infants received 4 doses at 2, 4, 6, and 12-15 months of age. Adverse events (AEs) were collected after each dose. Serotype-specific immunoglobulin G (IgG) concentrations and opsonophagocytic activity (OPA) were measured prior and 30 days postvaccination in adults, at 1 month postdose 3 (PD3), pre-dose4, and postdose 4 (PD4) in infants. RESULTS: Safety profiles were comparable across vaccination groups. At PD3, serotype-specific IgG GMCs were generally lower for either PCV15 formulation than PCV13 for most shared serotypes. PCV15 consistently elicited higher antibody responses to the 2 serotypes unique to the vaccine (22F and 33F) and serotype 3 for which PCV13 was shown to be ineffective. Except for serotypes 6A and 6B, no dose-response effect was observed with increasing concentrations of PnPs and/or adjuvant. CONCLUSION: PCV15 is safe and induces IgG and OPA responses to all 15 serotypes in the vaccine. No significant differences in antibody responses were observed with increases in PnPs and/or Aluminum Phosphate Adjuvant.


Subject(s)
Antibodies, Bacterial/blood , Immunogenicity, Vaccine , Pneumococcal Vaccines/immunology , Streptococcus pneumoniae/immunology , Adolescent , Adult , Dose-Response Relationship, Immunologic , Healthy Volunteers , Humans , Immunoglobulin G/blood , Infant , Middle Aged , Pneumococcal Infections/prevention & control , Pneumococcal Vaccines/administration & dosage , Pneumococcal Vaccines/chemistry , Vaccines, Conjugate/administration & dosage , Vaccines, Conjugate/chemistry , Vaccines, Conjugate/immunology , Young Adult
2.
Infect Immun ; 70(8): 4433-40, 2002 Aug.
Article in English | MEDLINE | ID: mdl-12117954

ABSTRACT

Staphylococcus aureus and Staphylococcus epidermidis often elaborate adherent biofilms, which contain the capsular polysaccharide-adhesin (PS/A) that mediates the initial cell adherence to biomaterials. Biofilm cells produce another antigen, termed polysaccharide intercellular adhesin (PIA), which is composed of a approximately 28 kDa soluble linear beta(1-6)-linked N-acetylglucosamine. We developed a new method to purify PS/A from S. aureus MN8m, a strain hyperproducing PS/A. Using multiple analytical techniques, we determined that the chemical structure of PS/A is also beta(1-6)-N-acetylglucosamine (PNAG). We were unable to find N-succinylglucosamine residues in any of our preparations in contrast to previously reported findings (D. McKenney, K. Pouliot, Y. Wang, V. Murthy, M. Ulrich, G. Doring, J. C. Lee, D. A Goldmann, and G. B. Pier, Science 284:1523-1527, 1999). PNAG was produced with a wide range of molecular masses that could be divided into three major fractions with average molecular masses of 460 kDa (PNAG-I), 100 kDa (PNAG-II), and 21 kDa (PNAG-III). The purified antigens were not soluble at neutral pH unless first dissolved in 5 M HCl and then neutralized with 5 M NaOH. PNAG-I was very immunogenic in rabbits, but the responses of individual animals were variable. Immunization of mice with various doses (100, 50, or 10 microg) of PNAG-I, -II, and -III demonstrated that only PNAG-I was able to elicit an immunoglobulin G (IgG) immune response with the highest titers obtained with 100-microg dose. When we purified a small fraction of PNAG with a molecular mass of approximately 780 kDa (PNAG-780) from PNAG-I, significantly higher IgG titers than those in mice immunized with the same doses of PNAG-I were obtained, suggesting the importance of the molecular mass of PNAG in the antibody response. These results further clarify the chemical structure of PS/A and help to differentiate it from PIA on the basis of immunogenicity, molecular size, and solubility.


Subject(s)
Acetylglucosamine/immunology , Adhesins, Bacterial/immunology , Polysaccharides, Bacterial/immunology , Staphylococcus aureus/immunology , Acetylglucosamine/chemistry , Acetylglucosamine/classification , Acetylglucosamine/isolation & purification , Adhesins, Bacterial/chemistry , Adhesins, Bacterial/classification , Adhesins, Bacterial/isolation & purification , Animals , Female , Immunochemistry , Mice , Nuclear Magnetic Resonance, Biomolecular/methods , Opsonin Proteins/immunology , Phagocytosis/immunology , Polysaccharides, Bacterial/chemistry , Polysaccharides, Bacterial/classification , Polysaccharides, Bacterial/isolation & purification , Rabbits , Staphylococcus aureus/chemistry
3.
Anal Biochem ; 279(2): 226-40, 2000 Mar 15.
Article in English | MEDLINE | ID: mdl-10706792

ABSTRACT

A method utilizing NMR spectroscopy has been developed to confirm the identity of bacterial polysaccharides used to formulate a polyvalent pneumococcal polysaccharide vaccine. The method is based on 600 MHz proton NMR spectra of individual serotype-specific polysaccharides. A portion of the anomeric region of each spectrum (5.89 to 4.64 ppm) is compared to spectra generated for designated reference samples for each polysaccharide of interest. The selected region offers a spectral window that is unique to a given polysaccharide and is sensitive to any structural alteration of the repeating units. The similarity of any two spectral profiles is evaluated using a correlation coefficient (rho), where rho >/= 0.95 between a sample and reference profile indicates a positive identification of the sample polysaccharide. This method has been shown to be extremely selective in its ability to discriminate between serotype-specific polysaccharides, some of which differ by no more than a single glycosidic linkage. Furthermore, the method is rapid and does not require extensive sample manipulations or pretreatments. The method was validated as a qualitative identity assay and will be incorporated into routine quality control testing of polysaccharide powders to be used in preparation of the polyvalent pneumococcal vaccine PNEUMOVAX 23. The specificity and reproducibility of the NMR-based identity assay is superior to the currently used colorimetric assays and can be readily adapted for use with other bacterial polysaccharide preparations as well.


Subject(s)
Magnetic Resonance Spectroscopy/methods , Polysaccharides, Bacterial/analysis , Bacterial Vaccines/analysis , Bacterial Vaccines/chemistry , Carbohydrate Sequence , Deuterium Oxide , Evaluation Studies as Topic , Hydrogen/chemistry , Magnetic Resonance Spectroscopy/statistics & numerical data , Molecular Sequence Data , Pneumococcal Vaccines , Polysaccharides, Bacterial/chemistry , Reproducibility of Results , Serotyping , Streptococcus pneumoniae/chemistry , Streptococcus pneumoniae/classification , Temperature
5.
Biochemistry ; 38(38): 12343-57, 1999 Sep 21.
Article in English | MEDLINE | ID: mdl-10493802

ABSTRACT

Three arginine residues (Arg-11, Arg-39, Arg-61) are found at the active site of 4-oxalocrotonate tautomerase in the X-ray structure of the affinity-labeled enzyme [Taylor, A. B., Czerwinski, R. M., Johnson, R. M., Jr., Whitman, C. P., and Hackert, M. L. (1998) Biochemistry 37, 14692-14700]. The catalytic roles of these arginines were examined by mutagenesis, kinetic, and heteronuclear NMR studies. With a 1,6-dicarboxylate substrate (2-hydroxymuconate), the R61A mutation showed no kinetic effects, while the R11A mutation decreased k(cat) 88-fold and increased K(m) 8.6-fold, suggesting both binding and catalytic roles for Arg-11. With a 1-monocarboxylate substrate (2-hydroxy-2,4-pentadienoate), no kinetic effects of the R11A mutation were found, indicating that Arg-11 interacts with the 6-carboxylate of the substrate. The stereoselectivity of the R11A-catalyzed protonation at C-5 of the dicarboxylate substrate decreased, while the stereoselectivity of protonation at C-3 of the monocarboxylate substrate increased in comparison with wild-type 4-OT, indicating the importance of Arg-11 in properly orienting the dicarboxylate substrate by interacting with the charged 6-carboxylate group. With 2-hydroxymuconate, the R39A and R39Q mutations decreased k(cat) by 125- and 389-fold and increased K(m) by 1.5- and 2.6-fold, respectively, suggesting a largely catalytic role for Arg-39. The activity of the R11A/R39A double mutant was at least 10(4)-fold lower than that of the wild-type enzyme, indicating approximate additivity of the effects of the two arginine mutants on k(cat). For both R11A and R39Q, 2D (1)H-(15)N HSQC and 3D (1)H-(15)N NOESY-HSQC spectra showed chemical shift changes mainly near the mutated residues, indicating otherwise intact protein structures. The changes in the R39Q mutant were mainly in the beta-hairpin from residues 50 to 57 which covers the active site. HSQC titration of R11A with the substrate analogue cis, cis-muconate yielded a K(d) of 22 mM, 37-fold greater than the K(d) found with wild-type 4-OT (0.6 mM). With the R39Q mutant, cis, cis-muconate showed negative cooperativity in active site binding with two K(d) values, 3.5 and 29 mM. This observation together with the low K(m) of 2-hydroxymuconate (0.47 mM) suggests that only the tight binding sites function catalytically in the R39Q mutant. The (15)Nepsilon resonances of all six Arg residues of 4-OT were assigned, and the assignments of Arg-11, -39, and -61 were confirmed by mutagenesis. The binding of cis,cis-muconate to wild-type 4-OT upshifts Arg-11 Nepsilon (by 0.05 ppm) and downshifts Arg-39 Nepsilon (by 1.19 ppm), indicating differing electronic delocalizations in the guanidinium groups. A mechanism is proposed in which Arg-11 interacts with the 6-carboxylate of the substrate to facilitate both substrate binding and catalysis and Arg-39 interacts with the 1-carboxylate and the 2-keto group of the substrate to promote carbonyl polarization and catalysis, while Pro-1 transfers protons from C-3 to C-5. This mechanism, together with the effects of mutations of catalytic residues on k(cat), provides a quantitative explanation of the 10(7)-fold catalytic power of 4-OT. Despite its presence in the active site in the crystal structure of the affinity-labeled enzyme, Arg-61 does not play a significant role in either substrate binding or catalysis.


Subject(s)
Arginine/genetics , Isomerases/chemistry , Isomerases/genetics , Mutagenesis, Site-Directed , Alanine/genetics , Binding Sites/genetics , Catalysis , Glutamine/genetics , Isomerases/biosynthesis , Isomerases/metabolism , Kinetics , Models, Molecular , Nitrogen Isotopes , Nuclear Magnetic Resonance, Biomolecular , Pseudomonas putida/enzymology , Pseudomonas putida/genetics , Recombinant Proteins/biosynthesis , Recombinant Proteins/chemical synthesis , Recombinant Proteins/chemistry , Sorbic Acid/analogs & derivatives , Sorbic Acid/metabolism , Stereoisomerism , Titrimetry
6.
Article in English | MEDLINE | ID: mdl-10218109

ABSTRACT

The MutT enzyme prevents errors in DNA replication by hydrolyzing mutagenic nucleotide substrates such as 8-oxo-dGTP. It does so by catalyzing nucleophilic attack at the electron rich P beta of nucleoside triphosphates. Members of this small mechanistic class of enzymes require two divalent cations per active site for activity--one coordinated by the enzyme and the other by the enzyme-bound NTP--and show low catalytic powers of 10(7)- to 10(9)-fold. The first structure of an enzyme of this class, obtained by NMR methods in solution, shows MutT to be a compact globular protein with an alpha + beta-fold. The binding of the essential divalent cation activator Mg2+ and the substrate analog Mg(2+)-AMPCPP to the MutT enzyme to form the quaternary E-Mg(2+)-AMPCPP-Mg2+ complex does not alter the global fold of the enzyme but produces localized small conformational changes at or near the metal and substrate binding sites. The adenine-ribose moiety binds in a hydrophobic cleft near 3-strands of a mixed beta-sheet, with the 6-NH2 group of the purine ring approaching the -NH2 side chain of Asn-119. With a 6-keto group, GTP would interact more favorably with Asn-119, consistent with the substrate preference of MutT for guanine over adenine nucleotides. The enzyme-bound metal is coordinated by three conserved Glu residues (Glu-56, Glu-57, and Glu-98), the backbone carbonyl of a conserved Gly residue (Gly-38), and by two water ligands. The metal-triphosphate moiety of the metal-AMPCPP complex binds in the second coordination sphere of the enzyme-bound divalent cation. One of the water ligands of the enzyme-bound metal ion is well positioned to attack P beta with inversion and to be deprotonated or oriented by Glu-53, which in turn may be oriented by Arg-52. Lys-39 is positioned to interact electrostatically with the alpha-phosphoryl group and thereby to facilitate the departure of the NMP-leaving group. Quantitatively, the 10(9)-fold rate acceleration produced by the MutT enzyme may be ascribed to catalysis by approximation and polarization of the attacking water by the enzyme-bound metal ion (> or = 10(5)-fold), activation of the NMP leaving group by Lys-39 (10-fold), charge neutralization at P beta by the nucleotide-bound divalent cation (> or = 10-fold), and orientation and/or deprotonation of the attacking water by Glu-53 (> or = 10(2)-fold). This reaction mechanism, derived from the solution structure of the quaternary MutT complex, is both qualitatively and quantitatively consistent with the results of mutagenesis studies and may well be applicable to other enzymes that catalyze nucleophilic substitution at the electron-rich P beta of NTP substrates.


Subject(s)
Bacterial Proteins/chemistry , Bacterial Proteins/metabolism , Escherichia coli Proteins , Escherichia coli/enzymology , Phosphoric Monoester Hydrolases/chemistry , Phosphoric Monoester Hydrolases/metabolism , Binding Sites , Catalysis , Cations, Divalent/pharmacology , Protein Binding , Protein Conformation , Pyrophosphatases , Solutions
8.
J Med Chem ; 41(26): 5287-96, 1998 Dec 17.
Article in English | MEDLINE | ID: mdl-9857096

ABSTRACT

Induction of phase 2 enzymes (e.g., glutathione transferases, NAD(P)H:quinone reductase, glucuronosyltransferases, epoxide hydrolase) is a major strategy for reducing the susceptibility of animal cells to neoplasia and other forms of electrophile toxicity. In a search for new chemoprotective enzyme inducers, a structure-activity analysis was carried out on two types of naturally occurring and synthetic substituted phenylpropenoids: (a) Ar-CH=CH-CO-R, where R is OH, OCH3, CH3, or Ar, including cinnamic, coumaric, ferulic, and sinapic acid derivatives, their ketone analogues, and chalcones; and (b) bis(benzylidene)cycloalkanones, Ar-CH=C(CH2)n(CO)C=CH-Ar, where n = 5, 6, or 7. The potencies of these compounds in inducing NAD(P)H:quinone reductase activity in murine hepatoma cells paralleled their Michael reaction acceptor activity (Talalay, P.; De Long, M. J.; Prochaska, H. J. Proc. Natl. Acad. Sci. U.S.A. 85, 1988, 8261-8265). Unexpectedly, the bis(benzylidene)cycloalkanones also powerfully quenched the lucigenin-derived chemiluminescence evoked by superoxide radicals. Introduction of o-hydroxyl groups on the aromatic rings of these phenylpropenoids dramatically enhanced their potencies not only as inducers for quinone reductase but also as quenchers of superoxide. These potentiating o-hydroxyl groups are hydrogen-bonded, as shown by moderate downfield shift of their proton NMR resonances and their sensitivities to the solvent environment. The finding that the potencies of a series of bis(benzylidene)cycloalkanones in inducing quinone reductase appear to be correlated with their ability to quench superoxide radicals suggests that the regulation of phase 2 enzymes may involve both Michael reaction reactivity and radical quenching mechanisms.


Subject(s)
Anticarcinogenic Agents/pharmacology , Chalcone/pharmacology , Cinnamates/pharmacology , Coumarins/pharmacology , Enzyme Induction , Free Radical Scavengers/pharmacology , NAD(P)H Dehydrogenase (Quinone)/metabolism , Animals , Anticarcinogenic Agents/chemistry , Chalcone/analogs & derivatives , Chalcone/chemistry , Cinnamates/chemistry , Coumarins/chemistry , Free Radical Scavengers/chemistry , Liver Neoplasms, Experimental/enzymology , Liver Neoplasms, Experimental/pathology , Luminescent Measurements , Magnetic Resonance Spectroscopy , Mice , Molecular Conformation , Structure-Activity Relationship , Superoxides/metabolism , Tumor Cells, Cultured
9.
Biochemistry ; 37(42): 14701-12, 1998 Oct 20.
Article in English | MEDLINE | ID: mdl-9778345

ABSTRACT

The solution structure of the ketosteroid isomerase homodimer complexed with the product analogue 19-nortestosterone hemisuccinate (19-NTHS) was solved by heteronuclear multidimensional NMR methods using 1647 distance restraints, 77 dihedral angle (phi) restraints, and 67 hydrogen bond restraints per monomer. The refined secondary structure of each subunit consists of three alpha-helices, eight beta-strands, four turns, and two beta-bulges. The beta-strands form a mixed beta-sheet. One of the five proline residues, Pro-39, is cis and begins a nonclassical turn. A self-consistent ensemble of 15 tertiary/quaternary structures of the enzyme dimer-steroid complex, with no distance violations greater than 0.35 A, was generated by simulated annealing and energy minimization with the program X-PLOR. The mean pairwise RMSD of the secondary structural elements was 0.63 A for the average subunit and 1.25 A for the dimer. Within each subunit, the three alpha-helices are packed onto the concave surface of the beta-sheet with a groove between them into which the steroid binds at a site defined by 14 intermolecular distances. In the productive complex, Tyr-14, from alpha-helix 1, approaches both Asp-99 and the 3-keto group of 19-NTHS while, from beta-strand 1, the carboxylate of Asp-38 approaches the beta-face of the steroid near C4 and C6, between which it transfers a proton during catalysis. Thus the solution structure of the isomerase-steroid complex can accommodate the catalytic diad mechanism in which Asp-99 donates a hydrogen bond to Tyr-14 which in turn is hydrogen bonded to the 3-oxygen of the steroid. While direct hydrogen bonding of Asp-99 to the steroid oxygen is less likely, it cannot be excluded. All other interactions of the steroid with the enzyme are hydrophobic. The dimer interface, which is between the convex surfaces of the beta-sheets, is defined by 28 intersubunit NOEs between hydrophobic residues in the 13C-filtered NOESY-HSQC spectrum of a 13C/12C-heterolabeled dimer. Both hydrophobic and polar interactions occur at the dimer interface which contains no space that would permit additional steroid binding. Comparison of the complexed enzyme with the solution structure of the free enzyme [Wu et al. (1997) Science 276, 415-418] reveals that the three helices change position in the steroid complex, becoming more closely packed onto the concave surface of the beta-sheet, thus bringing Tyr-14 closer to Asp-99 and the substrate. Comparison of the enzyme-steroid complex in solution with the free enzyme in the crystalline state reveals similar differences between the positions of the helices.


Subject(s)
Nandrolone/analogs & derivatives , Steroid Isomerases/chemistry , Binding Sites , Crystallography, X-Ray , Dimerization , Macromolecular Substances , Models, Molecular , Molecular Sequence Data , Nandrolone/chemistry , Nuclear Magnetic Resonance, Biomolecular , Protein Binding , Protein Conformation , Protein Structure, Secondary , Protein Structure, Tertiary , Protons , Solutions , Stereoisomerism
10.
Infect Immun ; 65(12): 5035-41, 1997 Dec.
Article in English | MEDLINE | ID: mdl-9393793

ABSTRACT

Lectin-mediated interactions between oral viridans group streptococci and actinomyces may play an important role in microbial colonization of the tooth surface. The presence of two host-like motifs, either GalNAc beta1-->3Gal (Gn) or Gal beta1-->3GalNAc (G), in the cell wall polysaccharides of five streptococcal strains accounts for the lactose-sensitive coaggregations of these bacteria with Actinomyces naeslundii. Three streptococcal strains which have Gn-containing polysaccharides also participate in GalNAc-sensitive coaggregations with strains of Streptococcus gordonii and S. sanguis. Each Gn- or G-containing polysaccharide is composed of a distinct phosphodiester-linked hexa- or heptasaccharide repeating unit. The occurrence of these polysaccharides on 19 additional viridans group streptococcal strains that participate in lactose-sensitive coaggregations with actinomyces was examined. Negatively charged polysaccharides that reacted with Bauhinia purpurea agglutinin, a Gal and GalNAc binding plant lectin, were isolated from 17 strains by anion exchange column chromatography of mutanolysin-cell wall digests. Results from nuclear magnetic resonance and immunodiffusion identified each of 16 polysaccharides as a known Gn- or G-containing structural type and one polysaccharide as a new but closely related Gn-containing type. Unlike the reactions of lectins, the cross-reactions of most rabbit antisera with these polysaccharides were correlated with structural features other than the host-like motifs. Gn-containing polysaccharides occurred primarily on the strains of S. sanguis and S. oralis while G-containing polysaccharides were more common among the strains of S. gordonii and S. mitis examined. The findings strongly support the hypothesis that lectin-mediated recognition of these streptococci by other oral bacteria depends on a family of antigenically diverse Gn- and G-containing cell wall polysaccharides, the occurrence of which may differ between streptococcal species.


Subject(s)
Cell Wall/metabolism , Polysaccharides, Bacterial/metabolism , Receptors, Cell Surface/metabolism , Streptococcaceae/metabolism , Actinomyces/metabolism , Animals , Carbohydrate Sequence , Lectins , Molecular Sequence Data , Polysaccharides, Bacterial/chemistry , Polysaccharides, Bacterial/immunology , Rabbits , Receptors, Cell Surface/immunology
11.
Biochemistry ; 36(48): 14616-26, 1997 Dec 02.
Article in English | MEDLINE | ID: mdl-9398180

ABSTRACT

The solution secondary structure of the highly active Y55F/Y88F "Tyr-14-only" mutant of delta 5-3-ketosteroid isomerase complexed with 19-nortestosterone hemisuccinate has been shown to consist of three helices, a six-stranded mixed beta-sheet, and five turns. The steroid binds near the general acid, Tyr-14, on helix 1, near the general base, Asp-38, on the first strand of the beta-sheet, and on the hydrophobic face of the beta-sheet [Zhao, Q., Abeygunawardana, C., & Mildvan, A. S. (1997) Biochemistry 36, 3458-3472]. On this hydrophobic face, Asp-99 is the only polar residue. Free isomerase shows a deshielded exchangeable proton resonance at 13.1 ppm assigned to the N epsilon H of neutral His-100. Its fractionation factor (phi = 0.79) and slow exchange with solvent suggest it to be buried or involved in an H-bond. The binding of dihydroequilenin or estradiol to isomerase induces the appearance of two additional deshielded proton resonances, one at 18.2 ppm assigned to the gamma-carboxyl proton of Asp-99, and the other, at 11.6 ppm, assigned to the zeta-OH proton of Tyr-14. While mutation of Asp-99 to Ala results in the disappearance of only the resonance near 18 ppm [Wu, R. W., Ebrahemian, S., Zwrotny, M. E., Thornberg, L. D., Perez-Alverado, G. C., Brothers, P., Pollack, R. M., & Summers, M. F. (1997) Science 276, 415-418], both of these resonances disappear in mutants lacking Tyr-14, suggesting an H-bonded catalytic diad, Asp-99-COOH--Tyr14-OH--O-steroid enolate. The catalytic diad is further supported by NOEs from the beta 1 and beta 2 protons of Asp-99 to the epsilon protons of Tyr-14, and from the zeta-OH proton of Tyr-14 to the gamma-carboxyl proton of Asp-99, indicating close proximity of these two residues, and by other data from the literature. A strong, low-barrier H-bond between Asp-99 and Tyr-14 is indicated by the 6.2 ppm deshielding, low fractionation factor (phi = 0.34) and slow exchange of the resonance at 18.2 ppm. A normal H-bond between Tyr-14 and the steroid is indicated by the 1.8 ppm deshielding, fractionation factor of 0.97 and the slow exchange of the resonance at 11.6 ppm. It is suggested that the 10(4.7)-fold contribution of Tyr-14 to catalysis is made possible by strong H-bonding from Asp-99 in the catalytic diad which strengthens general acid catalysis by Tyr-14. It is also noted that highly deshielded proton resonance on enzymes between 15 and 20 ppm, assigned to low-barrier H-bonds, generally involve carboxyl groups.


Subject(s)
Steroid Isomerases/chemistry , Aspartic Acid/chemistry , Binding Sites , Hydrogen Bonding , Models, Chemical , Mutation , Nuclear Magnetic Resonance, Biomolecular , Protein Structure, Secondary , Recombinant Proteins/chemistry , Steroid Isomerases/genetics , Tyrosine/chemistry
12.
Biochemistry ; 36(48): 14661-75, 1997 Dec 02.
Article in English | MEDLINE | ID: mdl-9398185

ABSTRACT

Triosephosphate isomerase (TIM) catalyzes the reversible interconversion of dihydroxyacetone phosphate (DHAP) and glyceraldehyde-3-phosphate (GAP), with Glu-165 removing the pro-R proton from C1 of DHAP and neutral His-95 polarizing the carbonyl group of the substrate. TIM and its complexes with the reactive intermediate analogs, phosphoglycolic acid (PGA) and phosphoglycolohydroxamic acid (PGH), were studied by 1H NMR at 600 MHz and at low temperature (-4.8 degrees C). His-95 shows an N epsilon H resonance at 13.1 ppm which shifts to 13.3 ppm in the TIM-PGA complex and to 13.5 ppm in the TIM-PGH complex. In the TIM-PGH complex, His-95 N epsilon H shows a slow, pH-independent exchange rate with water (kex = 80 s-1 at 30 degrees C, Eact = 19 kcal/mol), which is 44-fold slower than that of an exposed histidine suggesting partial shielding from bulk solvent, and a fractionation factor phi = 0.71 +/- 0.02 consistent with its donation of a normal hydrogen bond. The formation of the TIM-PGH complex results in the appearance of several deshielded proton resonances, including one at 14.9 ppm and one at 10.9 ppm which overlaps with another resonance. The resonance at 14.9 ppm is absent and the resonance at 10.9 ppm is much weaker in the TIM complex of PGA, which lacks the hydroxamic acid (-NHOH) moiety. 15N-labeled PGH was synthesized and the NH proton of free [15N]PGH shows a single 1H-15N HMQC cross peak with delta (1H) = 10.3 ppm and delta (15N) = 168 ppm which shifts to delta (1H) = 10.9 ppm and delta (15N) = 174 ppm in the TIM complex of [15N]PGH. The 15N-1H coupling in the complex indicates covalent N-H bonding, and the deshielded delta (15N) indicates a significant contribution of the imidate resonance form of PGH. The 14.9 ppm resonance is assigned to the NOH proton of bound PGH. This resonance shows a pH-independent exchange rate with water (kex = 3900 s-1 at 30 degrees C, Eact = 8.9 kcal/mol) which may reflect the dissociation of the TIM-PGH complex, and meets the criteria for a low-barrier hydrogen bond on the basis of the significant downfield shift of 6.2 ppm from the NOH proton of the model compound acetohydroxamic acid, and a very low fractionation factor phi = 0.38 +/- 0.06. In the X-ray structure of the TIM-PGH complex [Davenport, R. C., Bash, P. A., Seaton, B. A., Karplus, M., Petsko, G. A., and Ringe, D. (1991) Biochemistry 30, 5821], the NOH proton of bound PGH is hydrogen bonded to Glu-165. A low-barrier hydrogen bond from PGH NOH to Glu-165 suggests a dual role for Glu-165 in catalysis of proton transfer not only between the C1 and C2 carbons but also between the O1 and O2 oxygens in the interconversion of DHAP and GAP in wild type TIM. Such a mechanism, together with the measured exchange rate of the His-95 N epsilon H proton with solvent protons can accommodate the classical measurements of tritium incorporation from DHAP into GAP.


Subject(s)
Triose-Phosphate Isomerase/metabolism , Computer Simulation , Histidine/chemistry , Histidine/metabolism , Hydrogen Bonding , Hydroxamic Acids/chemistry , Hydroxamic Acids/metabolism , Models, Chemical , Nitrogen Isotopes , Nuclear Magnetic Resonance, Biomolecular , Titrimetry , Triose-Phosphate Isomerase/chemistry , Yeasts/enzymology
13.
Biochemistry ; 36(12): 3458-72, 1997 Mar 25.
Article in English | MEDLINE | ID: mdl-9131995

ABSTRACT

Backbone and side chain resonances of steroid-bound delta5-3-ketosteroid isomerase (EC 5.3.3.1), a homodimeric enzyme with 125 residues per monomer, have been assigned by heteronuclear NMR methods with the 15N- and 13C-labeled enzyme. The secondary structure in solution of steroid-bound isomerase, based on interproton NOE's and differences in chemical shifts of backbone H alpha, C alpha, C beta, and CO resonances from random coil values, consists of two alpha-helices (residues 5-21, 48-60), one 3(10) helix (residues 23-30), seven beta-strands (residues 34-38, 44-47, 62-67, 71-73, 78-87, 92-104, and 111-116), and five turns (residues 39-42, 74-77, 88-91, 105-108, and 119-122). Thus isomerase consists of 30% helix, 38% beta-sheet, and 16% turns. The remaining 20 residues (16%) are assumed to form coils. With the exception of a parallel interaction between beta-strands 1 and 7, all beta-strand interactions are antiparallel, forming both a beta-hairpin (beta1, beta2) and a four-stranded beta-sheet in which the first strand is interrupted (beta3-beta4, beta5, beta6, beta7). 1H-15N HSQC titrations of the free enzyme with the substrate analog 19-nortestosterone hemisuccinate revealed steroid-induced changes in backbone 15N and NH chemical shifts throughout the enzyme, with maximal effects on helix I (Val-15), beta-strand 1 of the beta-hairpin (Asp-38), the loop between helix 3 and beta-strand 3 (Leu-61), beta-strand 3 (Ala-64), beta-strand 5 (Phe-82, Ser-85, Glu-87), beta-strand 6 (Ile-98), and beta-strand 7 (Ala-114, Phe-116) of the beta-sheet, thus indicating the secondary structural components involved in steroid binding. These effects include regions near the catalytic residues Tyr-14 and Asp-38 which function as the general acid and base, respectively, in the ketosteroid isomerase reaction. Intermolecular NOE's between 19-nortestosterone hemisuccinate and isomerase indicate that the steroid binds near alpha-helices 1 and 3, which form one wall of the active site, and one end of the four-stranded beta-sheet which forms the other wall. Consistent with these observations, doxyldihydrotestosterone, a steroid that is spin-labeled at its solvent-exposed end [Kuliopulos, A., Westbrook, E. M., Talalay, P., & Mildvan, A. S. (1987) Biochemistry 26, 3927-3937], induced the selective attenuation in the 1H-15N HSQC spectra of cross peaks of residues at the end of helix 3 (Ser-58, Leu-59, Lys-60, Leu-61), beta-strand 5 (Val-84, Ser-85), and beta-strand 6 (Val-95), due to the proximity of the nitroxide radical to the backbone 15N and NH nuclei of these residues, thus confirming the location of the D ring of the bound steroid and defining the mouth of the active site.


Subject(s)
Enzyme Inhibitors/chemistry , Nandrolone/analogs & derivatives , Steroid Isomerases/chemistry , Amino Acid Sequence , Binding Sites , Electron Spin Resonance Spectroscopy , Enzyme Inhibitors/metabolism , Macromolecular Substances , Magnetic Resonance Spectroscopy , Magnetics , Models, Molecular , Molecular Sequence Data , Mutagenesis , Nandrolone/metabolism , Protein Structure, Secondary , Solutions , Steroid Isomerases/metabolism
14.
Biochemistry ; 36(6): 1199-211, 1997 Feb 11.
Article in English | MEDLINE | ID: mdl-9063868

ABSTRACT

The MutT enzyme (129 residues) catalyzes the hydrolysis of nucleoside triphosphates (NTP) by substitution at the rarely attacked beta-P, to yield NMP and pyrophosphate. It requires two divalent cations, forming an active E-M2+-NTP-M2+ complex. The solution structure of the free enzyme consists of a five-stranded mixed beta-sheet connected by loop I-alpha-helix I-loop II, by two tight turns, and by loop III and terminated by loop IV-alpha-helix II [Abeygunawardana, C., et al. (1995) Biochemistry 34, 14997-15005]. Assignments of backbone 15N and NH resonances and side chain 15N and NH2 resonances of the quaternary complex were made by 1H-15N HSQC titrations of the free enzyme with MgCl2 followed by equimolar AMPCPP/MgCl2. H(alpha) assignments were made by 1H-15N 3D TOCSY HSQC, and 1H-13C CT-HSQC spectra and backbone and side chain 1H and 13C assignments were made by 3D HCCH TOCSY experiments. Ligands donated by the protein to the enzyme-bound divalent cation, identified by paramagnetic effects of Co2+ and Mn2+ on CO(C)H spectra, are the carboxylate groups of Glu-56, -57, and -98 and the amide carbonyl of Gly-38. The solution structure of the complex was computed with XPLOR using a total of 2168 NOE and 83 phi restraints for the protein, 11 intramolecular NOEs for bound Mg2+ AMPCPP, 22 intermolecular NOEs between MutT and AMPCPP, and distances from the enzyme-bound Co2+ to the three phosphorus atoms of Co3+(NH3)4AMPCPP from paramagnetic effects of Co2+ on their T1 values. The fold of the MutT enzyme in the complex is very similar to that of the free enzyme, with minor changes in the metal and substrate binding sites. The adenine ring binds in a hydrophobic cleft, interacting with Leu-4 and Ile-6 on beta-strand A and with Ile-80 on beta-strand D. The 6-NH2 group of adenine approaches the side chain NH2 of Asn-119. This unfavorable interaction is consistent with the stronger binding by MutT of guanine nucleotides, which have a 6-keto group. The ribose binds with its hydroxyl groups oriented toward the solvent and its hydrophobic face interacting with Leu-4, Ile-6, and the gamma-CH2 of Lys-39 of loop I. The metal-triphosphate moiety appears to bind in the second coordination sphere of the enzyme-bound divalent cation. One of two intervening water ligands is well positioned to attack P(beta) with inversion and to donate a hydrogen bond to the conserved residue, Glu-53, which may deprotonate or orient the attacking water ligand. Lys-39 which is positioned to interact electrostatically with the alpha-phosphoryl group may facilitate the departure of the leaving NMP. On the basis of the structure of the quaternary complex, a mechanism of the MutT reaction is proposed which is qualitatively and quantitatively consistent with kinetic and mutagenesis studies. It is suggested that similar mechanisms may be operative for other enzymes that catalyze substitution at P(beta) of NTP substrates.


Subject(s)
Adenosine Triphosphate/analogs & derivatives , Bacterial Proteins/chemistry , Escherichia coli Proteins , Magnesium/metabolism , Phosphoric Monoester Hydrolases/chemistry , Adenosine Triphosphate/metabolism , Bacterial Proteins/metabolism , Models, Molecular , Molecular Sequence Data , Phosphoric Monoester Hydrolases/metabolism , Protein Conformation , Pyrophosphatases , Solutions
15.
Biochemistry ; 35(50): 16036-47, 1996 Dec 17.
Article in English | MEDLINE | ID: mdl-8973173

ABSTRACT

The solution secondary structure of 4-oxalocrotonate tautomerase (4-OT), a 41 kDa homohexamer with 62 residues per subunit, consists of an alpha-helix, two beta-strands, a beta-hairpin, two loops, two turns, and a C-terminal coil [Stivers et al. (1996) Protein Sci. 5, 729-741]. The general base, proline-1, as well as the two loops and the beta-hairpin have been shown to comprise the active site [Stivers et al. (1996) Biochemistry 35, 814-823]. The backbone dynamics of both the free enzyme and its complex with a substrate analog have been studied by 1H-detected 15N relaxation rates and NOE determinations at 500 and 600 MHz. Analysis of the data using the model-free formalism showed that the nanosecond to picosecond motion of 53 of the 60 backbone 15N-H vectors was highly restricted with a mean order parameter mean value of S2 = 0.87 +/- 0.03. The lowest backbone mobility (S2 > 0.90) is found in the beta 1-strand, loop 2, and turn 2. Greater backbone mobility is found in the active site (0.5 < or = S2 < or = 0.83) and at C-terminal residues 58-62 (0.03 < or = S2 < or = 0.70). A tau m value for the free hexamer of 13.7 ns at 42 degrees C was determined, consistent with a compact globular molecule of 41 kDa. Saturation of 4-OT with the analog of the dienolic intermediate and linear competitive inhibitor cis, cis-muconate (4) (KD = 0.59 mM) increased the backbone S2 of seven residues and decreased the backbone S2 of another eight residues, both at the active site and at the antiparallel beta 1-beta 1 interface. The S2 values of the other 44 detectable NH vectors were not altered by the binding of 4. The increases in S2, resulting from the "freezing" of the backbone NH vectors of seven residues upon the binding of 4, correspond to an unfavorable entropic contribution to delta Gbinding of 3.2 +/- 1.1 kcal/mol. This freezing is partially compensated for by the mobilization of the other eight residues, since the decreases in S2 for these residues correspond to an entropic contribution to binding of -1.9 +/- 0.1 kcal/mol. These entropy changes, resulting solely from alterations in high-frequency motion, are significant compared to the overall delta Gbinding = -4.6 kcal/mol for 4. Other effects of the binding of 4 include (1) changes in 15N and NH chemical shifts localized to the active site and (2) increases in the exchange contributions (R(ex)) to 1/T2 of backbone 15N resonances at the active site and at the subunit interface, reflecting microsecond to millisecond motions which may play a role in substrate binding (k(on) > or = 4 x 10(6) M-1 s-1) and/or catalysis (kcat = 10(3) s-1).


Subject(s)
Enzyme Inhibitors/metabolism , Isomerases/antagonists & inhibitors , Isomerases/chemistry , Protein Structure, Secondary , Sorbic Acid/analogs & derivatives , Binding Sites , Entropy , Enzyme Inhibitors/chemistry , Enzyme Inhibitors/pharmacology , Kinetics , Magnetic Resonance Spectroscopy , Models, Chemical , Models, Structural , Nitrogen Isotopes , Proline , Sorbic Acid/chemistry , Sorbic Acid/metabolism , Sorbic Acid/pharmacology , Stereoisomerism
16.
Anal Biochem ; 239(2): 168-79, 1996 Aug 01.
Article in English | MEDLINE | ID: mdl-8811902

ABSTRACT

Much effort has been devoted to developing methods for the efficient isolation and identification of glucosinolates. Existing methods for separation involve ion exchange, GLC, and HPLC (mostly after chemical modification by enzymatic sulfate removal and/or silylation). We demonstrate a simple and direct strategy for analyzing the glucosinolate content of plant extracts, made possible by a new combination of widely available techniques: (a) reverse-phase paired-ion chromatography (PIC) of plant extracts, (b) hydrolysis of glucosinolates by myrosinase and quantitation of resulting isothiocyanates by cyclocondensation with 1, 2-benzenedithiol; (c) a novel method for replacing the PIC counterions by ammonium ions, permitting direct bioassay, mass, and 1H NMR spectrometry; (d) mass spectrometric analysis of ammonium salts by negative-ion fast atom bombardment (FAB) to determine m/z of the [M - H]- ion, and by chemical ionization (CI) in ammonia to obtain accurate masses of characteristic fragment ions, principally [R-CN:NH4]+, [R-CH=NOH:H]+ and [R-CH=NOH:NH4]+; and (e) high-resolution 1H NMR spectroscopy of intact glucosinolates. FAB and CI mass spectra, as well as high-resolution 1H NMR spectra were obtained for a variety of glucosinolate standards. The results provide guidance for the isolation and characterization of unknown glucosinolates from plants. These combined procedures were applied to a sample of broccoli (cultivar SAGA), in order to resolve and identify its major glucosinolates: 4-methylsulfinylbutyl glucosinolate (glucoraphanin) and 4-methylthiobutyl glucosinolate (glucoerucin). Thus, this analytical strategy provides a powerful technique for identifying and quantitating glucosinolates in plant extracts without resorting to derivatization.


Subject(s)
Glucosinolates/isolation & purification , Chromatography, High Pressure Liquid/methods , Chromatography, Ion Exchange/methods , Glucosinolates/chemistry , Magnetic Resonance Spectroscopy/methods , Plant Extracts/chemistry , Spectrometry, Mass, Fast Atom Bombardment/methods
17.
Proc Natl Acad Sci U S A ; 93(16): 8220-4, 1996 Aug 06.
Article in English | MEDLINE | ID: mdl-8710850

ABSTRACT

Delta 5-3-Ketosteroid isomerase (EC 5.3.3.1) promotes an allylic rearrangement involving intramolecular proton transfer via a dienolic intermediate. This enzyme enhances the catalytic rate by a factor of 10(10). Two residues, Tyr-14, the general acid that polarizes the steroid 3-carbonyl group and facilitates enolization, and Asp-38 the general base that abstracts and transfers the 4 beta-proton to the 6 beta-position, contribute 10(4.7) and 10(5.6) to the rate increase, respectively. A major mechanistic enigma is the huge disparity between the pKa values of the catalytic groups and their targets. Upon binding of an analog of the dienolate intermediate to isomerase, proton NMR detects a highly deshielded resonance at 18.15 ppm in proximity to aromatic protons, and with a 3-fold preference for protium over deuterium (fractionation factor, phi = 0.34), consistent with formation of a short, strong (low-barrier) hydrogen bond to Tyr-14. The strength of this hydrogen bond is estimated to be at least 7.1 kcal/mol. This bond is relatively inaccessible to bulk solvent and is pH insensitive. Low-barrier hydrogen bonding of Tyr-14 to the intermediate, in conjunction with the previously demonstrated tunneling contribution to the proton transfer by Asp-38, provide a plausible and quantitative explanation for the high catalytic power of this isomerase.


Subject(s)
Steroid Isomerases/chemistry , Bacterial Proteins/chemistry , Bacterial Proteins/metabolism , Catalysis , Hydrogen Bonding , Magnetic Resonance Spectroscopy , Mutagenesis, Site-Directed , Pseudomonas/enzymology , Recombinant Proteins , Steroid Isomerases/metabolism , Structure-Activity Relationship
18.
Biochemistry ; 35(21): 6715-26, 1996 May 28.
Article in English | MEDLINE | ID: mdl-8639622

ABSTRACT

The role of the conserved residue Glu-57 in the mechanism of the MutT enzyme from Escherichia coli was investigated by mutagenesis and heteronuclear NMR methods. The enzymatic activity of the E57Q mutant is at least 10(5)-fold lower than that of the wild type enzyme. The solution structure of the E57Q mutant, based on comparisons of 1H-15N NOESY HSQC spectra and 1H-15N HSQC spectra to those of the wild type enzyme, differs in a region near Glu-57. The dissociation constants (KD) of the E-Mg2+ and E-Mn2+ complexes increased 3.3- and 3.6-fold, respectively, in the E57Q mutant, while the KD of E-dGTP is unaltered from that of the wild type enzyme. The enhanced paramagnetic effect of enzyme-bound Mn2+ on 1/T1 of water protons is halved in the E57Q mutant indicating an altered metal-binding site. 1H-15N HSQC titrations of E57Q with MnCl2 show selective attenuation of the side chain NH signals of Gln-57 and the backbone NH signals of Gly-37, Gly-38, Lys-39, Glu-53, Glu-56, Gln-57, and Glu-98, indicating proximity of bound Nm2+ to these residues. The same resonances of the wild type and the E57Q mutant enzymes are attenuated by Mn2+, but significantly smaller paramagnetic effects (relative to the largest effect on Lys-39) are found on Gly-37, Gly-38, Val-58, and Glu-98 of the mutant, indicating an altered position of the bound divalent cation. Thus Glu-57 is probably a ligand to the enzyme-bound metal, and the profound loss of catalytic activity in the E57Q mutant results from structural and electronic changes at the site of the enzyme-bound divalent cation. 1H-15N HSQC titrations of the wild type enzyme with MgCl2 show changes in chemical shifts of 15N and NH resonances in regions closely overlapping those induced by the E57Q mutation itself, suggesting that the loss of the negative charge at Glu-57, either by mutation or by neutralization with Mg2+, induces a similar effect. In the E57Q mutant, the slow exchange of the side chain NH2 protons of Gln-57 and NOE's from the NH2 protons of Gln-57 to the beta and gamma protons of Glu-98 suggests hydrogen bonding of Gln-57 to Glu-98 in the free enzyme. 1H-15N HSQC titrations of both the wild type and mutant enzymes with dGTP show changes in 15N and NH chemicals shifts of residues in a cleft formed by beta-strands A, C, and D on one side and loop I, the end of loop IV, and the beginning of helix II on the other side, suggesting this cleft to be the nucleotide binding site. These changes in chemical shift were smaller or absent in titrations of the wild type or mutant enzymes with AMPCPP or Mg2+-AMPCPP, in accord with the strong preference of the MutT enzyme for guanine over adenine nucleotide substrates.


Subject(s)
Bacterial Proteins/chemistry , Bacterial Proteins/metabolism , Escherichia coli Proteins , Escherichia coli/enzymology , Glutamic Acid , Phosphoric Monoester Hydrolases/chemistry , Phosphoric Monoester Hydrolases/metabolism , Protein Structure, Secondary , Amino Acid Sequence , Base Sequence , Conserved Sequence , DNA Primers , Kinetics , Magnetic Resonance Spectroscopy , Models, Structural , Molecular Sequence Data , Mutagenesis, Site-Directed , Point Mutation , Polymerase Chain Reaction , Pyrophosphatases
19.
Protein Sci ; 5(4): 729-41, 1996 Apr.
Article in English | MEDLINE | ID: mdl-8845763

ABSTRACT

4-Oxalocrotonate tautomerase (4-OT), a homohexamer consisting of 62 residues per subunit, catalyzes the isomerization of unsaturated alpha-keto acids using Pro-1 as a general base (Stivers et al., 1996a, 1996b). We report the backbone and side-chain 1H, 15N, and 13C NMR assignments and the solution secondary structure for 4-OT using 2D and 3D homonuclear and heteronuclear NMR methods. The subunit secondary structure consists of an alpha-helix (residues 13-30), two beta-strands (beta 1, residues 2-8; beta 2, residues 39-45), a beta-hairpin (residues 50-57), two loops (I, residues 9-12; II, 34-38), and two turns (I, residues 30-33; II, 47-50). The remaining residues form coils. The beta 1 strand is parallel to the beta 2 strand of the same subunit on the basis of cross stand NH(i)-NH(j) NOEs in a 2D 15N-edited 1H-NOESY spectrum of hexameric 4-OT containing two 15N-labeled subunits/hexamer. The beta 1 strand is also antiparallel to another beta 1 strand from an adjacent subunit forming a subunit interface. Because only three such pairwise interactions are possible, the hexamer is a trimer of dimers. The diffusion constant, determined by dynamic light scattering, and the rotational correlation time (14.5 ns) estimated from 15N T1/T2 measurements, are consistent with the hexameric molecular weight of 41 kDa. Residue Phe-50 is in the active site on the basis of transferred NOEs to the bound partial substrate 2-oxo-1,6-hexanedioate. Modification of the general base, Pro-1, with the active site-directed irreversible inhibitor, 3-bromopyruvate, significantly alters the amide 15N and NH chemical shifts of residues in the beta-hairpin and in loop II, providing evidence that these regions change conformation when the active site is occupied.


Subject(s)
Isomerases/chemistry , Amino Acid Sequence , Binding Sites , Crystallography, X-Ray , Magnetic Resonance Spectroscopy , Molecular Sequence Data , Protein Structure, Secondary , Protein Structure, Tertiary , Solutions
20.
Biochemistry ; 35(5): 1525-32, 1996 Feb 06.
Article in English | MEDLINE | ID: mdl-8634283

ABSTRACT

Side chain and backbone dynamics of the catalytic residue, Tyr-14, in free and steroid-bound delta 5-3-ketosteroid isomerase (EC 5.3.3.1, homodimer, M(r) = 26.8 kDa) have been examined by measurements of longitudinal and transverse 13C relaxation rates and nuclear Overhauser effects at both 500 and 600 MHz (proton frequencies). The data, analyzed using the model-free formalism, yielded an optimized correlation time for overall molecular rotation (tau m) of 17.9 ns, in agreement with the result (18 ns) of fluorescence anisotropy decay measurements [Wu, P., Li, Y.-K., Talalay, P., & Brand, L. (1994) Biochemistry 33, 7415-7422] and Stokes' law calculation (20 ns). The order parameter of the side chain C epsilon of Tyr-14 (S2 = 0.74), which is a measure of the restriction of its high-frequency (nanosecond to picosecond) motion, was significantly lower than that of the backbone C alpha (S2 = 0.82), indicating greater restriction of backbone motion. Upon binding of the steroid ligand, 19-nortestosterone hemisuccinate, a product analog and substrate of the reverse isomerase reaction, S2 of the side chain C epsilon increased from 0.74 to 0.86, while that of the backbone C alpha did not change significantly. Thus, in the steroid complex, the amplitude of high-frequency side chain motion of Tyr-14 became more restricted than that of its backbone which could lower the entropy barrier to catalysis. Lower-frequency (millisecond to microsecond) motion of Tyr-14 at rates comparable to kcat were detected by exchange contributions to transverse relaxation of both C epsilon and C alpha. Steroid binding produced no change in this low-frequency motion of the side chain of Tyr-14, which could contribute to substrate binding and product release, but decreased the exchange contribution to transverse relaxation of the backbone.


Subject(s)
Steroid Isomerases/chemistry , Androstenedione/chemistry , Androstenedione/metabolism , Binding Sites , Carbon Isotopes , Computer Simulation , Magnetic Resonance Spectroscopy , Models, Molecular , Monte Carlo Method , Motion , Steroid Isomerases/metabolism , Tyrosine/chemistry
SELECTION OF CITATIONS
SEARCH DETAIL
...