Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 42
Filter
Add more filters










Publication year range
1.
Clin Pharmacol Ther ; 91(3): 442-9, 2012 Mar.
Article in English | MEDLINE | ID: mdl-22048224

ABSTRACT

The allosteric effect of fluconazole (effector) on the formation of 1'-hydroxymidazolam (1'-OH-MDZ) and 4-hydroxymidazolam (4-OH-MDZ) from midazolam (MDZ), a substrate of CYP3A4/5--members of the cytochrome P450 superfamily of enzymes--was examined in healthy volunteers. Following pretreatment with fluconazole, the ratio of the areas under the curve (AUCs) for 4-OH-MDZ and MDZ (AUC(4-OH)/AUC(MDZ)) increased by 35-62%, whereas the ratio AUC(1'-OH)/AUC(MDZ) decreased by 5-37%; the ratio AUC(1'-OH)/AUC(4-OH) decreased by 46-58% after fluconazole administration and had no association with the CYP3A5 genotype. The in vitro formation of 1'-OH-MDZ was more susceptible to inhibition by fluconazole than that of 4-OH-MDZ. Fluconazole decreased the intrinsic formation-clearance ratio of 1'-OH-MDZ/4-OH-MDZ to an extent that was quantitatively comparable to in vivo observations. The elimination clearance of MDZ metabolites appeared unaffected by fluconazole. This study demonstrated that fluconazole alters formation of MDZ metabolites, both in vivo and in vitro, in a manner consistent with an allosteric interaction. The 1'-OH-MDZ/4-OH-MDZ ratio may serve as a biomarker of such interactions among MDZ, CYP3A4/5, and other putative effectors.


Subject(s)
Cytochrome P-450 CYP3A/metabolism , Fluconazole/pharmacology , Midazolam/pharmacokinetics , Allosteric Regulation , Area Under Curve , Biomarkers/metabolism , Cytochrome P-450 CYP3A/genetics , Drug Interactions , Humans , Midazolam/analogs & derivatives , Midazolam/metabolism , Midazolam/pharmacology
2.
Methods Enzymol ; 464: 211-31, 2009.
Article in English | MEDLINE | ID: mdl-19903557

ABSTRACT

Self-assembled phospholipid bilayer Nanodiscs have become an important and versatile tool among model membrane systems to functionally reconstitute membrane proteins. Nanodiscs consist of lipid domains encased within an engineered derivative of apolipoprotein A-1 scaffold proteins, which can be tailored to yield homogeneous preparations of disks with different diameters, and with epitope tags for exploitation in various purification strategies. A critical aspect of the self-assembly of target membranes into Nanodiscs lies in the optimization of the lipid:protein ratio. Here we describe strategies for performing this optimization and provide examples for reconstituting bacteriorhodopsin as a trimer, rhodopsin, and functionally active P-glycoprotein. Together, these demonstrate the versatility of Nanodisc technology for preparing monodisperse samples of membrane proteins of wide-ranging structure.


Subject(s)
Lipid Bilayers/chemistry , Membrane Proteins/chemistry , Models, Biological , Nanostructures/chemistry , Phospholipids/chemistry , ATP Binding Cassette Transporter, Subfamily B, Member 1/chemistry , Animals , Bacteriorhodopsins/chemistry , Crystallography, X-Ray , Mice , Phosphatidylcholines/chemistry
3.
Cell Mol Life Sci ; 62(11): 1221-33, 2005 Jun.
Article in English | MEDLINE | ID: mdl-15798895

ABSTRACT

The cytosolic glutathione S-transferases are a family of structurally homologous enzymes with multiple functions, including xenobiotic detoxification, clearance of oxidative stress products, and modulation of cell proliferation and apoptosis signaling pathways. This wide-ranging functional repertoire leads to several possible therapeutic uses for isoform-specific GST inhibitors. These inhibitors may be used, in principle, to modulate tumor cell drug resistance, as sensitizers to therapeutically directed oxidative stress, to enhance cell proliferation and to augment anti-malarial drugs. With increasing knowledge of GST structural and function, rational design strategies and mechanism-based inhibitors have been exploited successfully. However, design of isoform specificity remains a significant challenge in GST inhibitor development. Strategies for further inhibitor design and their possible limitations, along with potential therapeutic uses, are summarized.


Subject(s)
Glutathione Transferase , Glutathione Transferase/antagonists & inhibitors , Malaria/enzymology , Neoplasms/enzymology , Prodrugs , Biology , Chemical Phenomena , Chemistry , Drug Design , Glutathione Transferase/chemistry , Glutathione Transferase/physiology , Humans , Malaria/drug therapy , Molecular Conformation , Neoplasms/drug therapy , Prodrugs/pharmacology , Prodrugs/therapeutic use , Structure-Activity Relationship
4.
J Phys Chem B ; 109(41): 19484-9, 2005 Oct 20.
Article in English | MEDLINE | ID: mdl-16853517

ABSTRACT

The photophysics of hypericin have been studied in its complex with two different isoforms, A1-1 and P1-1, of the protein glutathione S-transferase (GST). One molecule of hypericin binds to each of the two GST subunits. Comparisons are made with our previous results for the hypericin/human serum albumin complex (Photochem. Photobiol. 1999, 69, 633-645). Hypericin binds with high affinity to the GSTs: 0.65 microM for the A1-1 isoform and 0.51 microM for the P1-1 isoform (Biochemistry 2004, 43, 12761-12769). The photophysics and activity of hypericin are strongly modulated by the binding protein. Intramolecular hydrogen-atom transfer is suppressed in both cases. Most importantly, while there is significant singlet oxygen generation from hypericin bound to GST A1-1, binding to GST P1-1 suppresses singlet oxygen generation to almost negligible levels. The data are rationalized in terms of a simple model in which the hypericin photophysics depends entirely upon the decay of the triplet state by two competing processes, quenching by oxygen to yield singlet oxygen and ionization, the latter of these two are proposed to be modulated by A1-1 and P1-1.


Subject(s)
Glutathione Transferase/chemistry , Perylene/analogs & derivatives , Anthracenes , Chemical Phenomena , Chemistry, Physical , Dimethyl Sulfoxide , Kinetics , Light , Oxygen/chemistry , Perylene/chemistry , Photochemistry , Photolysis , Protons , Serum Albumin/chemistry , Spectrophotometry, Ultraviolet
5.
Biochemistry ; 40(35): 10614-24, 2001 Sep 04.
Article in English | MEDLINE | ID: mdl-11524005

ABSTRACT

Most cytosolic glutathione S-transferases (GSTs) exploit a hydrogen bond between an active site Tyr and the bound glutathione (GSH) cofactor to lower the pK(a) of the GSH and generate the nucleophilic thiolate anion, GS(-). In human (hGSTA1-1) and rat (rGSTA1-1) homologues, the active site Tyr-9 has a low pK(a) of 8.1-8.3, for which the functional significance is unknown. Crystal structures of GSTA1-1 suggest that weakly polar interactions between the electropositive ring edge of Phe-10 and the pi-cloud of Tyr-9, in the apoenzyme, could stabilize the tyrosinate anion and also modulate the pK(a) of GSH. Upon binding a product GSH conjugate, Phe-10 moves away from Tyr-9, allowing the highly dynamic C-terminus to "close" over the active site. To explore the role of Phe-10 in modulating the Tyr-9 pK(a) and in ligand binding, rGSTA1-1 mutants F10Y, F10L, and F10A were characterized. The pK(a)s of Tyr-9 in the apoenzymes were 8.2 +/- 0.2, 8.7 +/- 0.2, and 9.3 +/- 0.1, respectively, for F10Y, F10L, and F10A, compared to 8.3 +/- 0.2 for the "wild type". The experimentally determined pK(a)s qualitatively paralleled the energies required to remove a proton predicted by ab initio calculations using model compounds constrained to the coordinates of rGSTA1-1. The pK(a) of GSH in the binary complex was significantly less affected by these substitutions. In contrast, F220I and F220Y C-terminal mutations caused the pK(a) of Tyr-9 to decrease modestly. For the binary complex with S-hexyl-GSH, which induces the "closed" conformation, Tyr-9 retains a low pK(a) and the Phe-10 substitutions have significant effects. Presumably, Phe-10 plays a critical structural role in stabilizing the closed conformation. The mutations F10L and F10A also slowed the rate of GSH conjugate binding by 10-20-fold, as measured by stopped-flow fluorescence. The effects of Phe-10 substitution were large for both steps of the biphasic binding reaction, suggesting the importance of aromatic interactions throughout the reaction coordinate. A unified view of the C-terminal dynamics of GSTA1-1 is discussed, which emphasizes the coupling between Tyr-9 ionization, active site solvation, and C-terminal dynamics.


Subject(s)
Glutathione Transferase/metabolism , Tyrosine/metabolism , Animals , Escherichia coli , Glutathione/metabolism , Glutathione Transferase/genetics , Kinetics , Ligands , Mutagenesis, Site-Directed , Mutation , Phenylalanine/metabolism , Protein Binding , Protein Conformation , Rats , Recombinant Fusion Proteins/genetics , Recombinant Fusion Proteins/metabolism , Spectrometry, Fluorescence , Spectrophotometry, Ultraviolet
6.
J Am Chem Soc ; 123(19): 4408-13, 2001 May 16.
Article in English | MEDLINE | ID: mdl-11457225

ABSTRACT

The oxidized disulfide form of the ubiquitous tripeptide glutathione (gamma-glu-cys-gly) (GSSG) is shown to produce transparent, thermoreversible gels in aqueous solutions of dimethyl sulfoxide, dimethylformamide, and methanol, at GSSG concentrations as low as 1.5 mM. The gels bind Congo Red and exhibit dramatic green birefringence when observed between crossed polarizers, characteristic of amyloid structures. By transmission electron microscopy, the gels appear to consist of a network of fibrous structures about 75 nm in diameter. Several structurally related peptides, including the glutathione isomer glu-cys-gly and the aspartyl analogue of glutathione (beta-asp-cys-gly), failed to produce gels under similar conditions. These results suggest that the interactions which produce gelation are highly specific and that the unusual peptide geometry introduced by gamma-glu-cys linkage is critical to the gelation behavior. (1)H NMR indicates solvent-dependent perturbation of the gamma-glutamyl alpha- and beta-protons and circular dichroism reveals a shift in the geometry of the disulfide bond under conditions producing gelation. We propose that in appropriate organic solvents, GSSG self-assembles into an extended network of beta-sheetlike structures capable of immobilizing bulk solvent. While obviously speculative, it is interesting to consider possible physiological consequences of glutathione self-recognition in such processes as abnormal protein aggregation and the thiol-disulfide exchange which is believed to participate in protein folding.


Subject(s)
Glutathione/chemistry , Circular Dichroism , Coloring Agents , Congo Red , Gels , Mass Spectrometry , Microscopy, Electron , Microscopy, Polarization , Nuclear Magnetic Resonance, Biomolecular , Oxidation-Reduction , Protein Structure, Secondary , Solvents
7.
Biochemistry ; 40(12): 3536-43, 2001 Mar 27.
Article in English | MEDLINE | ID: mdl-11297419

ABSTRACT

Binding of a hydrophobic glutathione product conjugate to rGST A1-1 proceeds via a two-step mechanism, including rapid ligand docking, followed by a slow isomerization to the final [GST.ligand] complex, which involves the localization of the flexible C-terminal helix. These kinetically resolved steps have been observed previously by stopped-flow fluorescence with the wild-type rGST A1-1, which contains a native Trp-21 approximately 20 A from the ligand binding site at the intrasubunit domain-domain interface. To confirm this binding mechanism, as well as elucidate the effects of truncation of the C-terminus, we have further characterized the binding and dissociation of the glutathione-ethacrynic acid product conjugate (GS-EA) to wild-type, F222W:W21F, and Delta209-222 rGST A1-1 and wild-type hGST A1-1. Although modest kinetic differences were observed between the hGST A1-1 and rGST A1-1, stopped-flow binding studies with GS-EA verified that the two-step mechanism of ligand binding is not unique to the GST A1-1 isoform from rat. An F222W:W21F rGST A1-1 double mutant provides a direct fluorescence probe of changes in the environment of the C-terminal residue. The observation of two relaxation times during ligand binding and dissociation to F222W:W21F suggests that the C-terminus has an intermediate conformation following ligand docking, which is distinct from its conformation in the apoenzyme or localized helical state. For the wild-type, Delta209-222, and F222W:W21F proteins, variable-temperature stopped-flow experiments were performed and activation parameters calculated for the individual steps of the binding reaction. Activation parameters for the binding reaction coordinate illustrate that the C-terminus provides a significant entropic contribution to ligand binding, which is completely realized within the initial docking step of the binding mechanism. In contrast, the slow isomerization step is enthalpically driven. The partitioning of entropic and enthalpic components of binding energy was confirmed by isothermal titration calorimetry with wild-type and Delta209-222 rGST A1-1.


Subject(s)
Entropy , Glutathione Transferase/metabolism , Peptide Fragments/metabolism , Animals , Calorimetry , Glutathione Transferase/genetics , Humans , Isoenzymes/genetics , Isoenzymes/metabolism , Kinetics , Ligands , Mutagenesis, Site-Directed , Peptide Fragments/genetics , Phenylalanine/genetics , Protein Binding/genetics , Rats , Temperature , Thermodynamics , Tryptophan/genetics
9.
Proteins ; 42(2): 192-200, 2001 Feb 01.
Article in English | MEDLINE | ID: mdl-11119643

ABSTRACT

Twelve C-terminal residues of human glutathione S-transferase A1-1 form a helix in the presence of glutathione-conjugate, or substrate alone, and partly cover the active site. According to X-ray structures, the helix is disordered in the absence of glutathione, but it is not known if it is helical and delocalized, or in a random-coil conformation. Mutation to a tyrosine of residue 220 within this helix was previously shown to affect the pK(a) of Tyr-9 at the active site, in the apo form of the enzyme, and it was proposed that an on-face hydrogen bond between Tyr-220 and Tyr-9 provided a means for affecting this pK(a). In the current study, X-ray structures of the W21F and of the C-terminal mutation, W21F/F220Y, with glutathione sulfonate bound, show that the C-terminal helix is disordered (or delocalized) in the W21F crystal but is visible and ordered in a novel location, a crystal packing crevice, in one of three monomers in the W21F/F220Y crystal, and the proposed hydrogen bond is not formed. Fluorescence spectroscopy studies using an engineered F222W mutant show that the C-terminus remains delocalized in the absence of glutathione or when only the glutathione binding site is occupied, but is ordered and localized in the presence of substrate or conjugate, consistent with these and previous crystallographic studies. Proteins 2001;42:192-200.


Subject(s)
Glutathione Transferase/chemistry , Animals , Binding Sites , Crystallography, X-Ray , Glutathione Transferase/genetics , Isoenzymes , Models, Molecular , Mutation , Peptide Fragments/chemistry , Protein Conformation , Rats , Spectrometry, Fluorescence
10.
J Biol Chem ; 275(23): 17447-51, 2000 Jun 09.
Article in English | MEDLINE | ID: mdl-10751412

ABSTRACT

The glutathione S-transferase enzymes (GSTs) have a tyrosine or serine residue at their active site that hydrogen bonds to and stabilizes the thiolate anion of glutathione, GS(-). The importance of this hydrogen bond is obvious, in light of the enhanced nucleophilicity of GS(-) versus the protonated thiol. Several A-class GSTs contain a C-terminal segment that undergoes a ligand-dependent local folding reaction. Here, we demonstrate the effects of the Y9F substitution on binding affinity for glutathione conjugates and on rates of the order-disorder transition of the C terminus in rat GST A1-1. The equilibrium binding affinity of the glutathione conjugate, GS-NBD (NBD-Cl, 7-chloro-4-nitrobenzo-2-oxa-1, 3-diazole), was decreased from 4.09 microm to 0.641 microm upon substitution of Tyr-9 with Phe. This result was supported by isothermal titration calorimetry, with K(d) values of 1.51 microm and 0.391 microm for wild type and Y9F, respectively. The increase in binding affinity for the mutant is associated with dramatic decreases in rates for the C-terminal order-disorder transition, based on a stopped-flow kinetic analysis. The same effects were observed, qualitatively, for a second GSH conjugate, GS-ethacrynic acid. Apparently, the phenolic hydroxyl group of Tyr-9 is critical for orchestrating C-terminal dynamics and efficient product release, in addition to its role in lowering the pK(a) of GSH.


Subject(s)
Glutathione Transferase/chemistry , Glutathione Transferase/metabolism , Tyrosine , 4-Chloro-7-nitrobenzofurazan , Amino Acid Substitution , Animals , Binding Sites , Calorimetry , Catalytic Domain , Glutathione/metabolism , Humans , Hydrogen Bonding , Isoenzymes/chemistry , Isoenzymes/metabolism , Kinetics , Models, Molecular , Phenylalanine , Protein Conformation , Rats , Serine , Thermodynamics
11.
Drug Metab Dispos ; 28(3): 360-6, 2000 Mar.
Article in English | MEDLINE | ID: mdl-10681383

ABSTRACT

Testosterone, terfenadine, midazolam, and nifedipine, four commonly used substrates for human cytochrome P-450 3A4 (CYP3A4), were studied in pairs in human liver microsomes and in microsomes from cells containing recombinant human CYP3A4 and P-450 reductase, to investigate in vitro substrate-substrate interaction with CYP3A4. The interaction patterns between compounds with CYP3A4 were found to be substrate-dependent. Mutual inhibition, partial inhibition, and activation were observed in the testosterone-terfenadine, testosterone-midazolam, or terfenadine-midazolam interactions. However, the most unusual result was the interaction between testosterone and nifedipine. Although nifedipine inhibited testosterone 6beta-hydroxylation in a concentration-dependent manner, testosterone did not inhibit nifedipine oxidation. Furthermore, the effect of testosterone and 7,8-benzoflavone on midazolam 1'-hydroxylation and 4-hydroxylation demonstrated different regiospecificities. These results may be explained by a model in which multiple substrates or ligands can bind concurrently to the active site of a single CYP3A4 molecule. However, the contribution of separate allosteric sites and conformational heterogeneity to the atypical kinetics of CYP3A4 can not be ruled out in this model.


Subject(s)
Cytochrome P-450 Enzyme System/metabolism , Microsomes, Liver/metabolism , Microsomes/metabolism , Mixed Function Oxygenases/metabolism , Cytochrome P-450 CYP3A , Drug Interactions , Humans , Hydroxylation/drug effects , Kinetics , Microsomes/drug effects , Microsomes, Liver/drug effects , Midazolam/metabolism , Midazolam/pharmacology , Nifedipine/metabolism , Nifedipine/pharmacology , Recombinant Proteins/metabolism , Substrate Specificity , Terfenadine/metabolism , Terfenadine/pharmacology , Testosterone/metabolism , Testosterone/pharmacology
12.
J Biol Chem ; 274(39): 27963-8, 1999 Sep 24.
Article in English | MEDLINE | ID: mdl-10488145

ABSTRACT

Escherichia coli glutamine synthetase (GS) is a dodecameric assembly of identical subunits arranged as two back-to-back hexagonal rings. In the presence of divalent metal ions, the dodecamers "stack" along their six-fold axis of symmetry to yield elongated tubes. This self-assembly process provides a useful model for probing metal-dependent protein-protein interactions. However, no direct spectroscopic or structural data have confirmed the identity of the ligands to the shared metal ions in "stacked" GS. Here, 9-GHz Cu(2+) EPR studies have been used to probe the ligand structure and stoichiometry of the metal binding sites. The wild type protein, with N-terminal sequence (His-4)-X(3)-(Met-8)-X(3)-(His-12), exhibits a classic Cu(2+)-nitrogen spectrum, with g = 2.06 G, g = 2.24 G, and A = 19.3 x 10(-3) cm(-1). No superhyperfine structure is observed. The H4C mutant affords a spectrum that is the combination of two spectra at all stages of saturation. One of the overlapping spectra is nearly identical to the spectrum of wild type, and is due to His ligation. The second spectrum observed yields g = 2.28 and A = 17.1 x 10(-3) cm(-1). The linewidth and tensor values of the second component have been assigned to Cu(2+)-S ligation. In contrast, the H12C mutant exhibits an EPR spectrum at low Cu(2+) occupancy that is very similar to the second set of spectral features observed for H4C, and which is assigned to Cu(2+)-S ligation. No Cu(2+)-His ligation is apparent until the Cu(2+)/N-terminal helices ratio is >1.0. At saturation, the g = 2.00-2.06 region of the spectrum is essentially a mirror image of the spectrum obtained with H4C, and is due to overlapping Cu(2+)-N and Cu(2+)-S EPR spectra. The M8L and M8C mutants were also studied, in order to probe the role of position 8 in the N-terminal helix. Spectral parameters of these mutants are nearly identical to each other and to the wild type spectrum at saturating Cu(2+), suggesting that Met-8 does not act as a direct metal ligand. Together, the results provide the first direct evidence for a binuclear metal ion site between each N-terminal helix pair at the GS-GS interface, with both His-4 and His-12 providing metal ligands.


Subject(s)
Copper/metabolism , Escherichia coli/enzymology , Glutamate-Ammonia Ligase/chemistry , Glutamate-Ammonia Ligase/ultrastructure , Binding Sites , Electron Spin Resonance Spectroscopy , Glutamate-Ammonia Ligase/metabolism , Kinetics , Macromolecular Substances , Models, Molecular , Mutagenesis, Site-Directed , Protein Structure, Secondary , Recombinant Proteins/chemistry , Recombinant Proteins/metabolism , Recombinant Proteins/ultrastructure
13.
Arch Biochem Biophys ; 370(1): 59-65, 1999 Oct 01.
Article in English | MEDLINE | ID: mdl-10496977

ABSTRACT

(R)-(+)-Menthofuran is the proximate toxic metabolite of pulegone, the major constituent of the pennyroyal oil, that contributes significantly to the hepatotoxicity resulting from ingestion of this folklore abortifacient pennyroyal oil. Recently, menthofuran was shown to be metabolized by cytochrome P450 to form (R)-2-hydroxymenthofuran. In this paper it is demonstrated that glutathione S-transferase (GST) catalyzes the tautomerization of 2-hydroxymenthofuran to mintlactone and isomintlactone, apparently without the formation of stable glutathione (GSH) conjugates. The reaction strictly required GSH; S-methyl GSH, which binds to the active site and leaves the active site Tyr-9 partly ionized, did not support GST-catalyzed isomerization. It was also determined that the tautomerization reaction requires the active site tyrosine, Tyr-9. The rat GSTA1-1 mutant (Y9F), with the active site tyrosine replaced with phenylalanine, demonstrated no catalytic activity. Rat cytosolic GST A1-1, in the presence of GSH, tautomerized 2-hydroxymenthofuran with apparent K(M) and V(max) values of 110 microM and 190 nmol/min/nmol GST, respectively. However, the site-directed mutant (F220Y), in which Tyr-9 and GSH in the binary complex [GST. GSH] have lower pK(a)s, exhibited K(M) and V(max) values of 97 microM and 280 nmol/min/nmol GST, respectively. Similarly, human liver cytosol catalyzed the tautomerization of 2-hydroxymenthofuran in a GST-dependent reaction. The mechanism most consistent with the data is a general-base catalyzed isomerization with GS(-) serving to deprotonate the substrate to initiate the reaction.


Subject(s)
Benzofurans/pharmacokinetics , Glutathione Transferase/metabolism , Lactones , Monoterpenes , Amino Acid Substitution , Animals , Benzofurans/chemistry , Catalysis , Cyclohexane Monoterpenes , Cytosol/enzymology , Humans , Isoenzymes/metabolism , Isomerism , Kinetics , Liver/enzymology , Menthol/analogs & derivatives , Menthol/chemistry , Molecular Structure , Mutagenesis, Site-Directed , Rats , Recombinant Proteins/metabolism
14.
Pac Symp Biocomput ; : 554-65, 1999.
Article in English | MEDLINE | ID: mdl-10380227

ABSTRACT

On the basis of available x-ray structures, A-class glutathione S-transferases (GSTs) contain at their C-termini a short alpha-helix that provides a 'lid' over the active site in the presence of the reaction products, glutathione-conjugates. However, in the ligand-free enzyme this helix is disordered and crystallographically invisible. An aromatic cluster including Phe-10, Phe-220, and the catalytic Tyr-9 within the C-terminal strand control the order of this helix. Here, preliminary x-ray crystallographic analyses of the wild type and F220Y rGSTA1-1 in the presence of GSH are described. Also, a transition state analysis is presented for ligand-dependent formation of the helix, based on variable temperature stopped-flow fluorescence. Together, the results suggest that the ligand-dependent ordering of the C-terminal strand occurs with a transition state that is highly desolvated, but with few intramolecular hydrogen bonds or electrostatic interactions. However, substitutions at Phe-220 modulate the activation parameters through interactions with the side chain of Tyr-9.


Subject(s)
Glutathione Transferase/chemistry , Protein Structure, Secondary , Computer Graphics , Crystallography, X-Ray , Escherichia coli , Glutathione/chemistry , Glutathione/metabolism , Glutathione Transferase/metabolism , Humans , Isoenzymes/chemistry , Isoenzymes/metabolism , Macromolecular Substances , Models, Molecular , Mutagenesis, Site-Directed , Protein Denaturation , Recombinant Proteins/chemistry
15.
Biochemistry ; 38(21): 6971-80, 1999 May 25.
Article in English | MEDLINE | ID: mdl-10346919

ABSTRACT

Structural studies have suggested that the glutathione S-transferase (GST) A1-1 isozyme contains a dynamic C-terminus which undergoes a ligand-dependent disorder-order transition and sequesters substrates within the active site. Here, the contribution of the C-terminus to the kinetics and thermodynamics of ligand binding and dissociation has been determined. Steady-state turnover rates of the wild type (WT) and a C-terminal truncated (Delta209-222) rGST A1-1 with ethacrynic acid (EA) were measured in the presence of variable concentrations of viscogen. The results indicate that a physical step involving segmental protein motion is at least partially rate limiting at temperatures between 10 and 40 degrees C for WT. Dissociation rates of the glutathione-ethacrynic acid product conjugate (GS-EA), determined by stopped-flow fluorescence, correspond to the steady-state turnover rates. In contrast, the chemical step governs the turnover reaction by Delta209-222, suggesting that the slow rate of product release for WT is controlled by the dynamics of the C-terminal coil-helix transition. In addition, the association reaction of WT rGST A1-1 with GS-EA established that the binding was biphasic and included ligand docking followed by slow isomerization of the enzyme-ligand complex. In contrast, binding of GS-EA to Delta209-222 was a monophasic, bimolecular reaction. These results indicate that the binding of GS-EA to WT rGST A1-1 proceeds via an induced fit mechanism, with a slow conformational step that corresponds to the coil-helix transition. However, the biphasic dissociation kinetics for the wild type, and the recovered kinetic parameters, suggest that a significant fraction of the [GST.GS-EA] complex ( approximately 15%) retains a persistent disordered state at equilibrium.


Subject(s)
Glutathione Transferase/chemistry , Glutathione Transferase/metabolism , Binding Sites/genetics , Glutathione Transferase/genetics , Humans , Isoenzymes/chemistry , Isoenzymes/genetics , Isoenzymes/metabolism , Kinetics , Ligands , Peptide Fragments/chemistry , Peptide Fragments/genetics , Peptide Fragments/metabolism , Protein Denaturation/genetics , Protein Structure, Secondary , Recombinant Proteins/chemistry , Recombinant Proteins/genetics , Recombinant Proteins/metabolism , Spectrophotometry, Ultraviolet , Substrate Specificity/genetics , Viscosity
16.
Biochemistry ; 38(16): 5065-75, 1999 Apr 20.
Article in English | MEDLINE | ID: mdl-10213609

ABSTRACT

In the accompanying paper [Storch et al. (1999) Biochemistry 38, 5054-5064] equilibrium denaturation studies and molecular dynamics (MD) simulations were used to investigate localized dynamics on the surface of cytochrome b5 (cyt b5) that result in the formation of a cleft. In those studies, an S18C:R47C disulfide mutant was engineered to inhibit cleft mobility. Temperature- and urea-induced denaturation studies revealed significant differences in Trp 22 fluorescence between the wild-type and mutant proteins. On the basis of the results, it was proposed that wild type populates a conformational ensemble that is unavailable to the disulfide mutant and is mediated by cleft mobility. As a result, the solvent accessibility of Trp 22 is decreased in S18C:R47C, suggesting that the local environment of this residue is less mobile due to the constraining effects of the disulfide on cleft dynamics. To further probe the structural effects on the local environment of Trp 22 caused by inhibition of cleft formation, we report here the results of steady-state and time-resolved fluorescence quenching, differential phase/modulation fluorescence anisotropy, and 1H NMR studies. In Trp fluorescence experiments, the Stern-Volmer quenching constant increases in wild type versus the oxidized disulfide mutant with increasing temperature. At 50 degrees C, KSV is nearly 1.5-fold greater in wild type compared to the oxidized disulfide mutant. In the reduced disulfide mutant, KSV was the same as wild type. The bimolecular collisional quenching constant, kq, for acrylamide quenching of Trp 22 increases 2.7-fold for wild type and only 1.8-fold for S18C:R47C, upon increasing the temperature from 25 to 50 degrees C. The time-resolved anisotropy decay at 25 degrees C was fit to a double-exponential decay for both the wild type and S18C:R47C. Both proteins exhibited a minor contribution from a low-amplitude fast decay, consistent with local motion of Trp 22. This component was more prevalent in the wild type, and the fractional contribution increased significantly upon raising the temperature. The fast rotational component of the S18C:R47C mutant was less sensitive to increasing temperature. A comparison of the 1H NMR monitored temperature titration of the delta-methyl protons of Ile 76 for wild type and oxidized disulfide mutant, S18C:R47C, showed a significantly smaller downfield shift for the mutant protein, suggesting that Trp 22 in the mutant protein experiences comparatively decreased cleft dynamics in core 2 at higher temperatures. Furthermore, comparison of the delta'-methyl protons of Leu 25 in the two proteins revealed a difference in the ratio of the equilibrium heme conformers of 1.2:1 for S18C:R47C versus 1.5:1 for wild type at 40 degrees C. The difference in equilibrium heme orientations between wild type and S18C:R47C suggests that the disulfide bond affects heme binding within core 1, possibly through damped cleft fluctuations. Taken together, the NMR and fluorescence studies support the proposal that an engineered disulfide bond inhibits the formation of a dynamic cleft on the surface of cyt b5.


Subject(s)
Cytochromes b5/chemistry , Disulfides/chemistry , Tryptophan/chemistry , Amino Acid Substitution/genetics , Animals , Arginine/chemistry , Arginine/genetics , Cytochromes b5/genetics , Fluorescence Polarization , Isoleucine/chemistry , Isoleucine/genetics , Models, Molecular , Nuclear Magnetic Resonance, Biomolecular , Protein Engineering , Rats , Serine/chemistry , Serine/genetics , Spectrometry, Fluorescence , Temperature , Thermodynamics
17.
Biochemistry ; 38(16): 5054-64, 1999 Apr 20.
Article in English | MEDLINE | ID: mdl-10213608

ABSTRACT

A previous molecular dynamics (MD) simulation of cytochrome b5 (cyt b5) at 25 degrees C displayed localized dynamics on the surface of the protein giving rise to the periodic formation of a cleft that provides access to the heme through a protected hydrophobic channel [Storch and Daggett (1995) Biochemistry 34, 9682]. Here we describe the production and testing of mutants designed to prevent the cleft from opening using a combination of experimental and theoretical techniques. Two mutants have been designed to close the surface cleft: S18D to introduce a salt bridge and S18C:R47C to incorporate a disulfide bond. The putative cleft forms between two separate cores of the protein: one is structural in nature and can be monitored through the fluorescence of Trp 22, and the other binds the heme prosthetic group and can be tracked via heme absorbance. An increase in motion localized to the cleft region was observed for each protein, except for the disulfide-containing variant, in MD simulations at 50 degrees C compared to simulations at 25 degrees C. For the disulfide-containing variant, the cleft remained closed. Both urea and temperature denaturation curves were nearly identical for wild-type and mutant proteins when heme absorbance was monitored. In contrast, fluorescence studies revealed oxidized S18C:R47C to be considerably more stable based on the midpoints of the denaturation transitions, Tm and U1/2. Moreover, the fluorescence changes for each protein were complete at approximately 50 degrees C and a urea concentration of approximately 3.9 M, significantly below the temperature and urea concentration (62 degrees C, 5 M urea) required to observe heme release. In addition, solvent accessibility based on acrylamide quenching of Trp 22 was lower in the S18C:R47C mutant, particularly at 50 degrees C, before heme release [presented in the accompanying paper (58)]. The results suggest that a constraining disulfide bond can be designed to inhibit dynamic cleft formation on the surface of cyt b5. Located near the heme, the native dynamics of the cleft may be functionally important for protein-protein recognition and/or complex stabilization.


Subject(s)
Cytochromes b5/chemistry , Disulfides/chemistry , Salts/chemistry , Animals , Cytochromes b5/genetics , Cytochromes b5/metabolism , Heme/chemistry , Heme/metabolism , Hot Temperature , Models, Molecular , Mutagenesis, Site-Directed , Oxidation-Reduction , Protein Binding/genetics , Protein Denaturation , Protein Engineering , Rats , Spectrometry, Fluorescence , Thermodynamics , Tryptophan/chemistry , Urea/chemistry
18.
Chem Biol ; 5(12): 689-97, 1998 Dec.
Article in English | MEDLINE | ID: mdl-9862795

ABSTRACT

BACKGROUND: Technologies that improve control of protein orientation on surfaces or in solution, through designed molecular recognition, will expand the range of proteins that are useful for biosensors, molecular devices and biomaterials. A limitation of some proteins is their biologically imposed symmetry, which results in indistinguishable recognition surfaces. Here, we have explored methods for modifying the symmetry of an oligomeric protein that exhibits useful self-assembly properties. RESULTS: Escherichia coli glutamine synthetase (GS) contains 24 solvent-exposed histidines on two symmetry-related surfaces. These histidines drive a metal-dependent self-assembly of GS tubes. Immobilization of GS on the affinity resin Ni2+-NTA followed by on-column modification with diethyl pyrocarbonate affords asymmetrically modified GS that self-assembles only to the extent of 'short' dimeric GS tubes, as demonstrated by electron microscopy, dynamic light scattering and atomic force microscopy. The utility of Ni2+-NTA as a chemical mask was also demonstrated for asymmetric modification of engineered cysteines adjacent to the natural histidines. CONCLUSIONS: Current genetic methods do not provide distinguishable recognition elements on symmetry-related surfaces of biologically assembled proteins. Ni2+-NTA serves as a mask to control chemical modification in vitro of residues within symmetry-related pairs, on proteins containing functional His-tags. This strategy may be extended to modification of a wide range of amino acids with a myriad of reagents.


Subject(s)
Glutamate-Ammonia Ligase/chemistry , Molecular Probes , Nickel/metabolism , Nitrilotriacetic Acid/analogs & derivatives , Organometallic Compounds , Protein Conformation , Biocompatible Materials/chemical synthesis , Biosensing Techniques , Chromatography, Affinity/methods , Dimerization , Escherichia coli/enzymology , Glutamate-Ammonia Ligase/ultrastructure , Microscopy, Electron , Models, Molecular
19.
Biochemistry ; 37(42): 14948-57, 1998 Oct 20.
Article in English | MEDLINE | ID: mdl-9778372

ABSTRACT

rGSTA1-1 has been shown to catalyze the hydrolysis of the thiol ester glutathionyl ethacrynate (E-SG). In contrast, neither the retro-Michael addition with the substrate EA-SG, to yield GSH and ethacrynic acid (EA), nor the conjugation reaction between GSH and EA to yield the thiol ester E-SG was catalyzed to any measurable extent under similar conditions. The steady state kcat and KM for hydrolysis of E-SG by wild type rGSTA1-1 were 0.11 +/- 0.009 min-1 and 15.7 +/- 1.6 mM, respectively. The site-directed mutant, Y9F, in which the catalytic Tyr-9 is substituted with Phe, was completely inactive in this reaction. To uncover a mechanistic signature that would distinguish between direct hydrolysis and covalent catalysis involving acylation of Tyr-9, solvent isotope exchange and mass spectrometry experiments were performed. No 18O incorporation into the starting thiol ester was detected with initial velocity solvent isotope exchange experiments. However, covalent adducts corresponding to acylated protein also were not observed by electrospray ionization mass spectrometry, even with an assay that minimized the experimental dead time and which allowed for detection of N-acetyltyrosine acylated with EA in a chemical model system. The kon and koff rate constants for association and dissociation of E-SG were determined, by stopped flow fluorescence, to be 5 x 10(5) s-1 M-1 and 6.7 s-1, respectively. Together with the isotope partitioning results, these rate constants were used to construct partial free energy profiles for the GST-catalyzed hydrolysis of E-SG, assuming that Tyr-9 acts as a general acid-base catalyst. The "one-way flux" of the thiol esterase reaction results directly from the thermodynamic stability of the products after rate-limiting attack of the thiol ester by H2O or Tyr-9, and is sufficient to drive the hydrolysis to completion, in contrast to GST-catalyzed breakdown of other GSH conjugates.


Subject(s)
Glutathione Transferase/chemistry , Sulfhydryl Compounds/chemistry , Animals , Catalysis , Chromatography, High Pressure Liquid , Esters , Ethacrynic Acid/chemistry , Humans , Hydrogen-Ion Concentration , Hydrolysis , Kinetics , Mass Spectrometry , Rats , Spectrometry, Fluorescence
SELECTION OF CITATIONS
SEARCH DETAIL
...