Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 76
Filter
1.
Biochemistry ; 63(10): 1347-1358, 2024 May 21.
Article in English | MEDLINE | ID: mdl-38691339

ABSTRACT

The physiological role of dihydroorotate dehydrogenase (DHOD) enzymes is to catalyze the oxidation of dihydroorotate to orotate in pyrimidine biosynthesis. DHOD enzymes are structurally diverse existing as both soluble and membrane-associated forms. The Family 1 enzymes are soluble and act either as conventional single subunit flavin-dependent dehydrogenases known as Class 1A (DHODA) or as unusual heterodimeric enzymes known as Class 1B (DHODB). DHODBs possess two active sites separated by ∼20 Å, each with a noncovalently bound flavin cofactor. NAD is thought to interact at the FAD containing site, and the pyrimidine substrate is known to bind at the FMN containing site. At the approximate center of the protein is a single Fe2S2 center that is assumed to act as a conduit, facilitating one-electron transfers between the flavins. We present anaerobic transient state analysis of a DHODB enzyme from Lactoccocus lactis. The data presented primarily report the exothermic reaction that reduces orotate to dihydroorotate. The reductive half reaction reveals rapid two-electron reduction that is followed by the accumulation of a four-electron reduced state when NADH is added in excess, suggesting that the initial two electrons acquired reside on the FMN cofactor. Concomitant with the first reduction is the accumulation of a long-wavelength absorption feature consistent with the blue form of a flavin semiquinone. Spectral deconvolution and fitting to a model that includes reversibility for the second electron transfer reveals equilibrium accumulation of a flavin bisemiquinone state that has features of both red and blue semiquinones. Single turnover reactions with limiting NADH and excess orotate reveal that the flavin bisemiquinone accumulates with reduction of the enzyme by NADH and decays with reduction of the pyrimidine substrate, establishing the bisemiquinone as a fractional state of the two-electron reduced intermediate observed.


Subject(s)
Dihydroorotate Dehydrogenase , Lactococcus lactis , Biocatalysis , Catalysis , Catalytic Domain , Flavin Mononucleotide/metabolism , Flavin Mononucleotide/chemistry , Flavin-Adenine Dinucleotide/metabolism , Flavin-Adenine Dinucleotide/chemistry , Kinetics , Lactococcus lactis/enzymology , Lactococcus lactis/metabolism , NAD/metabolism , NAD/chemistry , Oxidation-Reduction
2.
Arch Biochem Biophys ; 755: 109966, 2024 May.
Article in English | MEDLINE | ID: mdl-38537870

ABSTRACT

Flavin disulfide reductases (FDRs) are FAD-dependent enzymes that transmit electrons from NAD(P)H to reduce specific oxidant substrate disulfides. These enzymes have been studied extensively, most particularly the paradigm examples: glutathione reductase and thioredoxin reductase. The common, though not universal, traits of the family include a tyrosine- or phenylalanine-gated binding pocket for NAD(P) nicotinamides adjacent to the FAD isoalloxazine re-face, and a disulfide stacked against the si-face of the isoalloxazine whose dithiol form is activated for subsequent exchange reactions by a nearby histidine acting as a base. This arrangement promotes transduction of the reducing equivalents for disulfide exchange relay reactions. From an observational standpoint the proximal parallel stacking of three redox moieties induces up to three opportunities for unique charge transfer interactions (NAD(P)H FAD, NAD(P)+•FADH2, and FAD•thiolate). In transient state, the charge transfer transitions provide discrete signals to assign reaction sequences. This review summarizes the lineage of observations for the FDR enzymes that have been extensively studied. Where applicable and in order to chart a consistent interpretation of the record, only data derived from studies that used anaerobic methods are cited. These data reveal a recurring theme for catalysis that is elaborated with specific additional functionalities for each oxidant substrate.

3.
Arch Biochem Biophys ; : 110162, 2024 Sep 23.
Article in English | MEDLINE | ID: mdl-39322101

ABSTRACT

Thioredoxin/glutathione reductase from Schistosoma mansoni (SmTGR) is a multifunctional enzyme that catalyzes the reduction of glutathione (GSSG) and thioredoxin, as well as the deglutathionylation of peptide and non-peptide substrates. SmTGR structurally resembles known glutathione reductases (GR) and thioredoxin reductases (TrxR) but with an appended N-terminal domain that has a typical glutaredoxin (Grx) fold. Despite structural homology with known GRs, the site of glutathione reduction has frequently been reported as the Grx domain, based primarily on aerobic, steady-state kinetic measurements and x-ray crystallography. Here, we present an anaerobic characterization of a series of variant SmTGRs to establish the site of GSSG reduction as the cysteine pair most proximal to the FAD, Cys154/Cys159, equivalent to the site of GSSG reduction in GRs. Anaerobic steady-state analysis of U597C, U597S, U597C+C31S, and I592STOP SmTGR demonstrate that the Grx domain is not involved in the catalytic reduction of GSSG, as redox silencing of the C-terminus results in no modulation of the observed turnover number (∼0.025 s-1) and redox silencing of the Grx domain results in an increased observed turnover number (∼0.08 s-1). Transient-state single turnover analysis of these variants corroborates this, as the slowest rate observed titrates hyperbolically with GSSG concentration and approaches a limit that coincides with the respective steady-state turnover number for each variant. Numerical integration fitting of the transient state data can only account for the observed trends when competitive binding of the C-terminus is included, indicating that the partitioning of electrons to either substrate occurs at the Cys154/Cys159 disulfide rather than the previously proposed Cys596/Sec597 sulfide/selenide. Paradoxically, truncating the C-terminus at Ile592 results in a loss of GR activity, indicating a crucial non-redox role for the C-terminus.

4.
Biochemistry ; 62(9): 1497-1508, 2023 05 02.
Article in English | MEDLINE | ID: mdl-37071546

ABSTRACT

Thioredoxin/glutathione reductase from Schistosoma mansoni (SmTGR) catalyzes the reduction of both oxidized thioredoxin and glutathione with electrons from reduced nicotinamide adenine dinucleotide phosphate (NADPH). SmTGR is a drug target for the treatment of Schistosomiasis, an infection caused by Schistosoma platyhelminths residing in the blood vessels of the host. Schistosoma spp. are reliant on TGR enzymes as they lack catalase and so use reduced thioredoxin and glutathione to regenerate peroxiredoxins consumed in the detoxification of reactive oxygen species. SmTGR is a flavin adenine dinucleotide (FAD)-dependent enzyme, and we have used the flavin as a spectrophotometric reporter to observe the movement of electrons within the enzyme. The data show that NADPH fractionally reduces the active site flavin with an observed rate constant estimated in this study to be ∼3000 s-1. The flavin then reoxidizes by passing electrons at a similar rate to the proximal Cys159-Cys154 disulfide pair. The dissociation of NADP+ occurs with a rate of ∼180 s-1, which induces the deprotonation of Cys159, and this coincides with the accumulation of an intense FAD-thiolate charge transfer band. It is proposed that the electrons then pass to the Cys596-Cys597 disulfide pair of the associated subunit in the dimer with a net rate constant of ∼2 s-1. (Note: Cys597 is Sec597 in wild-type (WT) SmTGR.) From this position, the electrons can be passed to oxidized thioredoxin or further into the protein to reduce the Cys28-Cys31 disulfide pair of the originating subunit of the dimer. From the Cys28-Cys31 center, electrons can then pass to oxidized glutathione that has a binding site directly adjacent.


Subject(s)
Flavin-Adenine Dinucleotide , Schistosoma mansoni , Animals , Schistosoma mansoni/metabolism , Glutathione Reductase/metabolism , NADP/metabolism , Flavin-Adenine Dinucleotide/metabolism , Thioredoxin-Disulfide Reductase/metabolism , Glutathione/metabolism , Disulfides , Thioredoxins/metabolism , Oxidation-Reduction
5.
J Biol Chem ; 298(6): 101969, 2022 06.
Article in English | MEDLINE | ID: mdl-35460691

ABSTRACT

Hepatocellular carcinoma (HCC) is the most common primary cancer of the liver and occurs predominantly in patients with underlying chronic liver diseases. Over the past decade, human ornithine aminotransferase (hOAT), which is an enzyme that catalyzes the metabolic conversion of ornithine into an intermediate for proline or glutamate synthesis, has been found to be overexpressed in HCC cells. hOAT has since emerged as a promising target for novel anticancer therapies, especially for the ongoing rational design effort to discover mechanism-based inactivators (MBIs). Despite the significance of hOAT in human metabolism and its clinical potential as a drug target against HCC, there are significant knowledge deficits with regard to its catalytic mechanism and structural characteristics. Ongoing MBI design efforts require in-depth knowledge of the enzyme active site, in particular, pKa values of potential nucleophiles and residues necessary for the molecular recognition of ligands. Here, we conducted a study detailing the fundamental active-site properties of hOAT using stopped-flow spectrophotometry and X-ray crystallography. Our results quantitatively revealed the pH dependence of the multistep reaction mechanism and illuminated the roles of ornithine α-amino and δ-amino groups in substrate recognition and in facilitating catalytic turnover. These findings provided insights of the catalytic mechanism that could benefit the rational design of MBIs against hOAT. In addition, substrate recognition and turnover of several fragment-sized alternative substrates of hOATs, which could serve as structural templates for MBI design, were also elucidated.


Subject(s)
Ornithine-Oxo-Acid Transaminase/metabolism , Carcinoma, Hepatocellular , Crystallography, X-Ray , Humans , Hydrogen-Ion Concentration , Kinetics , Liver Neoplasms , Models, Molecular , Ornithine/chemistry , Ornithine-Oxo-Acid Transaminase/chemistry , Substrate Specificity
6.
Arch Biochem Biophys ; 748: 109772, 2023 10 15.
Article in English | MEDLINE | ID: mdl-37820757

ABSTRACT

Dihydropyrimidine dehydrogenase (DPD) is an enzyme that uses an elaborate architecture to catalyze a simple net reaction: the reduction of the vinylic bond of uracil and thymine. Known DPDs have two active sites separated by approximately 60 Å. One active site has an FAD cofactor and binds NAD(P) and the other has an FMN cofactor and binds pyrimidines. The intervening distance is spanned by four Fe4S4 centers that act as an electron conduit. Recent advancements with porcine DPD have revealed unexpected chemical sequences where the enzyme undergoes reductive activation by transferring two electrons from NADPH to the FMN via the FAD such that the active form has the cofactor set FAD•4(Fe4S4)•FMNH2. Here we describe the first comprehensive kinetic investigation of a bacterial form of DPD. Using primarily transient state methods, DPD from E. coli (EcDPD) was shown to have a similar mechanism to that observed with the mammalian form in that EcDPD is observed to undergo reductive activation before pyrimidine reduction and displays half-of-sites activity. However, two distinct aspects of the EcDPD reaction relative to the mammalian enzyme were observed that relate to the effector roles for substrates: (i) the enzyme will rapidly take up electrons from NADH, reducing a flavin in the absence of pyrimidine substrate, and (ii) the activated form of the enzyme can become fully oxidized by transferring electrons to pyrimidine substrates in the absence of NADH.


Subject(s)
Escherichia coli , NAD , Swine , Animals , NAD/metabolism , Dihydrouracil Dehydrogenase (NADP)/chemistry , Kinetics , Uracil , Mammals/metabolism , Oxidation-Reduction , Flavin-Adenine Dinucleotide/chemistry
7.
Arch Biochem Biophys ; 736: 109517, 2023 03 01.
Article in English | MEDLINE | ID: mdl-36681231

ABSTRACT

Dihydropyrimidine dehydrogenase (DPD) is a flavin dependent enzyme that catalyzes the reduction of the 5,6-vinylic bond of pyrimidines uracil and thymine with electrons from NADPH. DPD has two active sites that are separated by ∼60 Å. At one site NADPH binds adjacent to an FAD cofactor and at the other pyrimidine binds proximal to an FMN. Four Fe4S4 centers span the distance between these active sites. It has recently been established that the enzyme undergoes reductive activation prior to reducing the pyrimidine. In this initial process NADPH is oxidized at the FAD site and electrons are transmitted to the FMN via the Fe4S4 centers to yield the active state with a cofactor set of FAD•4(Fe4S4)•FMNH2. The catalytic chemistry of DPD can be studied in transient-state by observation of either NADPH consumption or charge transfer absorption associated with complexation of NADPH adjacent to the FAD. Here we have utilized both sets of absorption transitions to find evidence for specific additional aspects of the DPD mechanism. Competition for binding with NADP+ indicates that the two charge transfer species observed in activation/single turnover reactions arise from NADPH populating the FAD site before and after reductive activation. An additional charge transfer species is observed to accumulate at longer times when high NADPH concentrations are mixed with the enzyme•pyrimidine complex and this data can be modelled based on asymmetry in the homodimer. It was also shown that, like pyrimidines, dihydropyrimidines induce rapid reductive activation indicating that the reduced pyrimidine formed in turnover can stimulate the reinstatement of the active state of the enzyme. Investigation of the reverse reaction revealed that dihydropyrimidines alone can reductively activate the enzyme, albeit inefficiently. In the presence of dihydropyrimidine and NADP+ DPD will form NADPH but apparently without measurable reductive activation. Pyrimidines that have 5-substituent halogens were utilized to probe both reductive activation and turnover. The linearity of the Hammett plot based on the rate of hydride transfer to the pyrimidine establishes that, at least to the radius of an iodo-group, the 5-substituent volume does not have influence on the observed kinetics of pyrimidine reduction.


Subject(s)
Dihydrouracil Dehydrogenase (NADP) , Pyrimidines , Animals , Oxidation-Reduction , Dihydrouracil Dehydrogenase (NADP)/chemistry , NADP/metabolism , Spectrophotometry , Pyrimidines/metabolism , Kinetics , Flavin-Adenine Dinucleotide/chemistry , Mammals/metabolism
8.
Molecules ; 28(3)2023 Jan 23.
Article in English | MEDLINE | ID: mdl-36770800

ABSTRACT

Ornithine aminotransferase (OAT) is overexpressed in hepatocellular carcinoma (HCC), and we previously showed that inactivation of OAT inhibits the growth of HCC. Recently, we found that (3S,4S)-3-amino-4-fluorocyclopentenecarboxylic acid (5) was a potent inactivator of γ-aminobutyric acid aminotransferase (GABA-AT), proceeding by an enamine mechanism. Here we describe our investigations into the activity and mechanism of 5 as an inactivator of human OAT. We have found that 5 exhibits 10-fold less inactivation efficiency (kinact/KI) against hOAT than GABA-AT. A comprehensive mechanistic study was carried out to understand its inactivation mechanism with hOAT. pKa and electrostatic potential calculations were performed to further support the notion that the α,ß-unsaturated alkene of 5 is critical for enhancing acidity and nucleophilicity of the corresponding intermediates and ultimately responsible for the improved inactivation efficiency of 5 over the corresponding saturated analogue (4). Intact protein mass spectrometry and the crystal structure complex with hOAT provide evidence to conclude that 5 mainly inactivates hOAT through noncovalent interactions, and that, unlike with GABA-AT, covalent binding with hOAT is a minor component of the total inhibition which is unique relative to other monofluoro-substituted derivatives. Furthermore, based on the results of transient-state measurements and free energy calculations, it is suggested that the α,ß-unsaturated carboxylate group of PLP-bound 5 may be directly involved in the inactivation cascade by forming an enolate intermediate. Overall, compound 5 exhibits unusual structural conversions which are catalyzed by specific residues within hOAT, ultimately leading to an enamine mechanism-based inactivation of hOAT through noncovalent interactions and covalent modification.


Subject(s)
Carcinoma, Hepatocellular , Liver Neoplasms , Humans , Amino Acids/pharmacology , Enzyme Inhibitors/pharmacology , Ornithine-Oxo-Acid Transaminase/chemistry , Ornithine-Oxo-Acid Transaminase/metabolism , gamma-Aminobutyric Acid , Carboxylic Acids/pharmacology , Carboxylic Acids/chemistry , Ornithine
9.
J Am Chem Soc ; 144(28): 12769-12780, 2022 07 20.
Article in English | MEDLINE | ID: mdl-35802469

ABSTRACT

RibB (3,4-dihydroxy-2-butanone 4-phosphate synthase) is a magnesium-dependent enzyme that excises the C4 of d-ribulose-5-phosphate (d-Ru5P) as formate. RibB generates the four-carbon substrate for lumazine synthase that is incorporated into the xylene moiety of lumazine and ultimately the riboflavin isoalloxazine. The reaction was first identified by Bacher and co-workers in the 1990s, and their chemical mechanism hypothesis became canonical despite minimal direct evidence. X-ray crystal structures of RibB typically show two metal ions when solved in the presence of non-native metals and/or liganding non-substrate analogues, and the consensus hypothetical mechanism has incorporated this cofactor set. We have used a variety of biochemical approaches to further characterize the chemistry catalyzed by RibB from Vibrio cholera (VcRibB). We show that full activity is achieved at metal ion concentrations equal to the enzyme concentration. This was confirmed by electron paramagnetic resonance of the enzyme reconstituted with manganese and crystal structures liganded with Mn2+ and a variety of sugar phosphates. Two transient species prior to the formation of products were identified using acid quench of single turnover reactions in combination with NMR for singly and fully 13C-labeled d-Ru5P. These data indicate that dehydration of C1 forms the first transient species, which undergoes rearrangement by a 1,2 migration, fusing C5 to C3 and generating a hydrated C4 that is poised for elimination as formate. Structures determined from time-dependent Mn2+ soaks of VcRibB-d-Ru5P crystals show accumulation in crystallo of the same intermediates. Collectively, these data reveal for the first time crucial transient chemical states in the mechanism of RibB.


Subject(s)
Intramolecular Transferases , Riboflavin , Butanones , Formates , Intramolecular Transferases/chemistry , Phosphates , Riboflavin/biosynthesis , Riboflavin/chemistry , Riboflavin Synthase/chemistry
10.
J Am Chem Soc ; 144(12): 5629-5642, 2022 03 30.
Article in English | MEDLINE | ID: mdl-35293728

ABSTRACT

Human ornithine aminotransferase (hOAT) is a pyridoxal 5'-phosphate (PLP)-dependent enzyme that contains a similar active site to that of γ-aminobutyric acid aminotransferase (GABA-AT). Recently, pharmacological inhibition of hOAT was recognized as a potential therapeutic approach for hepatocellular carcinoma. In this work, we first studied the inactivation mechanisms of hOAT by two well-known GABA-AT inactivators (CPP-115 and OV329). Inspired by the inactivation mechanistic difference between these two aminotransferases, a series of analogues were designed and synthesized, leading to the discovery of analogue 10b as a highly selective and potent hOAT inhibitor. Intact protein mass spectrometry, protein crystallography, and dialysis experiments indicated that 10b was converted to an irreversible tight-binding adduct (34) in the active site of hOAT, as was the unsaturated analogue (11). The comparison of kinetic studies between 10b and 11 suggested that the active intermediate (17b) was only generated in hOAT and not in GABA-AT. Molecular docking studies and pKa computational calculations highlighted the importance of chirality and the endocyclic double bond for inhibitory activity. The turnover mechanism of 10b was supported by mass spectrometric analysis of dissociable products and fluoride ion release experiments. Notably, the stopped-flow experiments were highly consistent with the proposed mechanism, suggesting a relatively slow hydrolysis rate for hOAT. The novel second-deprotonation mechanism of 10b contributes to its high potency and significantly enhanced selectivity for hOAT inhibition.


Subject(s)
4-Aminobutyrate Transaminase , Liver Neoplasms , Carboxylic Acids , Enzyme Inhibitors/chemistry , Humans , Kinetics , Molecular Docking Simulation , Ornithine-Oxo-Acid Transaminase , Phenylacetates , Pyridoxal Phosphate/chemistry , gamma-Aminobutyric Acid
11.
Biochemistry ; 60(22): 1764-1775, 2021 06 08.
Article in English | MEDLINE | ID: mdl-34032117

ABSTRACT

The native function of dihydropyrimidine dehydrogenase (DPD) is to reduce the 5,6-vinylic bond of pyrimidines uracil and thymine with electrons obtained from NADPH. NADPH and pyrimidines bind at separate active sites separated by ∼60 Šthat are bridged by four Fe4S4 centers. We have shown that DPD undergoes reductive activation, taking up two electrons from NADPH [Beaupre, B. A., et al. (2020) Biochemistry 59, 2419-2431]. pH studies indicate that the rate of turnover is not controlled by the protonation state of the general acid, cysteine 671. The activation of the C671 variants is delineated into two phases particularly at low pH values. Spectral deconvolution of the delineated reductive activation reaction reveals that the initial phase results in the accumulation of charge transfer absorption added to the binding difference spectrum for NADPH. The second phase results in reduction of one of the two flavins. X-ray crystal structure analysis of the C671S variant soaked with NADPH and the slow substrate, thymine, in a low-oxygen atmosphere resolved the presumed activated form of the enzyme that has the FMN cofactor reduced. These data reveal that charge transfer arises from the proximity of the NADPH and FAD bases and that the ensuing flavin is a result of rapid transfer of electrons to the FMN without accumulation of reduced forms of the FAD or Fe4S4 centers. These data suggest that the slow rate of turnover of DPD is governed by the movement of a mobile structural feature that carries the C671 residue.


Subject(s)
Dihydrouracil Dehydrogenase (NADP)/chemistry , Dihydrouracil Dehydrogenase (NADP)/metabolism , Hydrogen/metabolism , Animals , Catalytic Domain , Dihydrouracil Dehydrogenase (NADP)/physiology , Flavin-Adenine Dinucleotide/chemistry , Flavins/chemistry , Hydrogen/chemistry , Kinetics , NADP/metabolism , NADPH-Ferrihemoprotein Reductase/metabolism , Oxidation-Reduction , Protein Structure, Tertiary , Pyrimidines/chemistry , Pyrimidines/metabolism , Spectrophotometry/methods , Swine
12.
Biochemistry ; 60(14): 1120-1132, 2021 04 13.
Article in English | MEDLINE | ID: mdl-33755421

ABSTRACT

Dihydropyrimidine dehydrogenase (DPD) is a complex enzyme that reduces the 5,6-vinylic bond of pyrimidines, uracil, and thymine. 5-Fluorouracil (5FU) is also a substrate for DPD and a common chemotherapeutic agent used to treat numerous cancers. The reduction of 5FU to 5-fluoro-5,6-dihydrouracil negates its toxicity and efficacy. Patients with high DPD activity levels typically have poor outcomes when treated with 5FU. DPD is thus a central mitigating factor in the treatment of a variety of cancers. 5-Ethynyluracil (5EU) covalently inactivates DPD by cross-linking with the active-site general acid cysteine in the pyrimidine binding site. This reaction is dependent on the simultaneous binding of 5EU and nicotinamide adenine dinucleotide phosphate (NADPH). This ternary complex induces DPD to become activated by taking up two electrons from the NADPH. The covalent inactivation of DPD by 5EU occurs concomitantly with this reductive activation with a rate constant of ∼0.2 s-1. This kinact value is correlated with the rate of reduction of one of the two flavin cofactors and the localization of a mobile loop in the pyrimidine active site that places the cysteine that serves as the general acid in catalysis proximal to the 5EU ethynyl group. Efficient cross-linking is reliant on enzyme activation, but this process appears to also have a conformational aspect in that nonreductive NADPH analogues can also induce a partial inactivation. Cross-linking then renders DPD inactive by severing the proton-coupled electron transfer mechanism that transmits electrons 56 Šacross the protein.


Subject(s)
Dihydrouracil Dehydrogenase (NADP)/metabolism , Uracil/analogs & derivatives , Animals , Catalytic Domain , Dihydrouracil Dehydrogenase (NADP)/chemistry , Drug Synergism , Protein Binding , Swine , Uracil/metabolism , Uracil/pharmacology
13.
Biochemistry ; 60(40): 3027-3039, 2021 10 12.
Article in English | MEDLINE | ID: mdl-34569786

ABSTRACT

Guanosine triphosphate (GTP) cyclohydrolase II (RibA) is one of three enzymes that hydrolytically cleave the C8-N9 bond of the GTP guanine. RibA also catalyzes a subsequent hydrolytic attack at the base liberating formate and in addition cleaves the α-ß phosphodiester bond of the triphosphate to form pyrophosphate (PPi). These hydrolytic reactions are promoted by tandem active-site metal ions, zinc and magnesium, that respectively function at the GTP guanine and triphosphate moieties. The RibA reaction is part of riboflavin biosynthesis and forms 2,5-diamino-6-ß-pyrimidinone 5'-phosphate, an exocyclic pyrimidine nucleotide that ultimately forms the pyrimidine ring of the isoalloxazine of riboflavin. The stoichiometry of the RibA reaction was defined in the study that first identified this activity in Escherichia coli (Foor, F., Brown, G. M. J. Biol. Chem., 1975, 250, 9, 3545-3551) and has not been quantitatively evaluated in subsequent works. Using primarily transient state approaches we examined the interaction of RibA from E. coli with the GTP, inosine triphosphate, and PPi. Our data indicate that PPi is a slow substrate for RibA that is cleaved to form two phosphate ions (Pi). A combination of real-time enzymatically coupled Pi reporter assays and end-point 31P NMR revealed that Pi is formed at a catalytically relevant rate in the native reaction of RibA with GTP, redefining the reaction stoichiometry. Furthermore, our data indicate that both PPi and GTP stimulate conformational changes prior to hydrolytic chemistry, and we conclude that the cleavage of PPi bound as a substrate or an intermediate state results in conformational relaxation.


Subject(s)
Escherichia coli Proteins/chemistry , Escherichia coli/enzymology , GTP Cyclohydrolase/chemistry , Biocatalysis , Diphosphates/metabolism , Escherichia coli Proteins/metabolism , GTP Cyclohydrolase/metabolism , Guanosine Triphosphate/metabolism , Inosine Triphosphate/metabolism , Kinetics , Protein Binding , Pyrophosphatases/chemistry , Pyrophosphatases/metabolism
14.
J Am Chem Soc ; 143(23): 8689-8703, 2021 06 16.
Article in English | MEDLINE | ID: mdl-34097381

ABSTRACT

The inhibition of human ornithine δ-aminotransferase (hOAT) is a potential therapeutic approach to treat hepatocellular carcinoma. In this work, (S)-3-amino-4,4-difluorocyclopent-1-enecarboxylic acid (SS-1-148, 6) was identified as a potent mechanism-based inactivator of hOAT while showing excellent selectivity over other related aminotransferases (e.g., GABA-AT). An integrated mechanistic study was performed to investigate the turnover and inactivation mechanisms of 6. A monofluorinated ketone (M10) was identified as the primary metabolite of 6 in hOAT. By soaking hOAT holoenzyme crystals with 6, a precursor to M10 was successfully captured. This gem-diamine intermediate, covalently bound to Lys292, observed for the first time in hOAT/ligand crystals, validates the turnover mechanism proposed for 6. Co-crystallization yielded hOAT in complex with 6 and revealed a novel noncovalent inactivation mechanism in hOAT. Native protein mass spectrometry was utilized for the first time in a study of an aminotransferase inactivator to validate the noncovalent interactions between the ligand and the enzyme; a covalently bonded complex was also identified as a minor form observed in the denaturing intact protein mass spectrum. Spectral and stopped-flow kinetic experiments supported a lysine-assisted E2 fluoride ion elimination, which has never been observed experimentally in other studies of related aminotransferase inactivators. This elimination generated the second external aldimine directly from the initial external aldimine, rather than the typical E1cB elimination mechanism, forming a quinonoid transient state between the two external aldimines. The use of native protein mass spectrometry, X-ray crystallography employing both soaking and co-crystallization methods, and stopped-flow kinetics allowed for the detailed elucidation of unusual turnover and inactivation pathways.


Subject(s)
Ornithine-Oxo-Acid Transaminase/metabolism , Humans , Molecular Structure , Ornithine-Oxo-Acid Transaminase/chemistry
15.
Arch Biochem Biophys ; 714: 109066, 2021 12 15.
Article in English | MEDLINE | ID: mdl-34717904

ABSTRACT

Dihydropyrimidine dehydrogenase (DPD) catalyzes the two-electron reduction of pyrimidine bases uracil and thymine as the first step in pyrimidine catabolism. The enzyme achieves this simple chemistry using a complex cofactor set including two flavins and four Fe4S4 centers. The flavins, FAD and FMN, interact with respective NADPH and pyrimidine substrates and the iron-sulfur centers form an electron transfer wire that links the two active sites that are separated by 56 Å. DPD accepts the common antineoplastic agent 5-fluorouracil as a substrate and so undermines the establishment of efficacious toxicity. Though studied for multiple decades, a precise description of the behavior of the enzyme had remained elusive. It was recently shown that the active form of DPD has the cofactor set of FAD-4(Fe4S4)-FMNH2. This two-electron reduced state is consistent with fewer mechanistic possibilities and data suggests that the instigating and rate determining step in the catalytic cycle is reduction of the pyrimidine substrate that is followed by relatively rapid oxidation of NADPH at the FAD that, via the electron conduit of the 4(Fe4S4) centers, reinstates the FMNH2 cofactor for subsequent catalytic turnover.


Subject(s)
Dihydrouracil Dehydrogenase (NADP)/metabolism , Animals , Catalysis , Dihydrouracil Dehydrogenase (NADP)/chemistry , Flavin Mononucleotide/metabolism , Flavin-Adenine Dinucleotide/metabolism , Hydroquinones/metabolism , Mammals , Oxidation-Reduction , Protein Conformation , Protein Domains , Substrate Specificity
16.
Biochemistry ; 59(26): 2419-2431, 2020 07 07.
Article in English | MEDLINE | ID: mdl-32516529

ABSTRACT

Dihydropyrimidine dehydrogenase (DPD) catalyzes the initial step in the catabolism of the pyrimidines uracil and thymine. Crystal structures have revealed an elaborate subunit architecture consisting of two flavin cofactors, apparently linked by four Fe4S4 centers. Analysis of the DPD reaction(s) equilibrium position under anaerobic conditions revealed a reaction that favors dihydropyrimidine formation. Single-turnover analysis shows biphasic kinetics. The serine variant of the candidate general acid, cysteine 671, provided enhanced kinetic resolution for these phases. In the first event, one subunit of the DPD dimer takes up two electrons from NADPH in a reductive activation. Spectrophotometric deconvolution suggests that these electrons reside on one of the two flavins. The fact that oxidation of the enzyme by dioxygen can be suppressed by the addition of pyrimidine is consistent with these electrons residing on the FMN. The second phase involves further oxidation of NADPH and concomitant reduction of the pyrimidine substrate. During this phase no net reduction of DPD cofactors is observed, indicating that the entire cofactor set acts as a wire, transmitting electrons from NADPH to the pyrimidine rapidly. This indicates that the availability of the proton from the C671 general acid controls the transmittance of electrons from NADPH to the pyrimidine. Acid quench and high-performance liquid chromatography product analysis of single-turnover reactions with limiting NADPH confirmed a 2:1 NADPH:pyrimidine stoichiometry for the enzyme, accounting for successive activation and pyrimidine reduction. These data support an alternating subunit model in which one protomer is activated and turns over before the other subunit can be activated and enter catalysis.


Subject(s)
Dihydrouracil Dehydrogenase (NADP)/chemistry , Flavin Mononucleotide/chemistry , NADP/chemistry , Pyrimidines/chemistry , Sus scrofa , Animals , Dihydrouracil Dehydrogenase (NADP)/genetics , Enzyme Activation
17.
J Biol Chem ; 294(47): 17988-18001, 2019 11 22.
Article in English | MEDLINE | ID: mdl-31615895

ABSTRACT

Pseudopaline and staphylopine are opine metallophores biosynthesized by Pseudomonas aeruginosa and Staphylococcus aureus, respectively. The final step in opine metallophore biosynthesis is the condensation of the product of a nicotianamine (NA) synthase reaction (i.e. l-HisNA for pseudopaline and d-HisNA for staphylopine) with an α-keto acid (α-ketoglutarate for pseudopaline and pyruvate for staphylopine), which is performed by an opine dehydrogenase. We hypothesized that the opine dehydrogenase reaction would be reversible only for the opine metallophore product with (R)-stereochemistry at carbon C2 of the α-keto acid (prochiral prior to catalysis). A kinetic analysis using stopped-flow spectrometry with (R)- or (S)-staphylopine and kinetic and structural analysis with (R)- and (S)-pseudopaline confirmed catalysis in the reverse direction for only (R)-staphylopine and (R)-pseudopaline, verifying the stereochemistry of these two opine metallophores. Structural analysis at 1.57-1.85 Å resolution captured the hydrolysis of (R)-pseudopaline and allowed identification of a binding pocket for the l-histidine moiety of pseudopaline formed through a repositioning of Phe-340 and Tyr-289 during the catalytic cycle. Transient-state kinetic analysis revealed an ordered release of NADP+ followed by staphylopine, with staphylopine release being the rate-limiting step in catalysis. Knowledge of the stereochemistry for opine metallophores has implications for future studies involving kinetic analysis, as well as opine metallophore transport, metal coordination, and the generation of chiral amines for pharmaceutical development.


Subject(s)
Biocatalysis , Imidazoles/metabolism , Oligopeptides/metabolism , Oxidoreductases/metabolism , Pseudomonas aeruginosa/enzymology , Staphylococcus aureus/enzymology , Fluorescence , Hydrolysis , Imidazoles/chemistry , Ions , Ketoglutaric Acids/metabolism , Kinetics , Metals/pharmacology , Models, Molecular , Oligopeptides/chemistry , Stereoisomerism
18.
Appl Environ Microbiol ; 86(22)2020 10 28.
Article in English | MEDLINE | ID: mdl-32887719

ABSTRACT

Soil bacteria can detoxify Cr(VI) ions by reduction. Within the last 2 decades, numerous reports of chromate reductase enzymes have been published. These reports describe catalytic reduction of chromate ions by specific enzymes. These enzymes each have sequence similarity to known redox-active flavoproteins. We investigated the enzyme NfoR from Staphylococcus aureus, which was reported to be upregulated in chromate-rich soils and to have chromate reductase activity (H. Han, Z. Ling, T. Zhou, R. Xu, et al., Sci Rep 7:15481, 2017, https://doi.org/10.1038/s41598-017-15588-y). We show that NfoR has structural similarity to known flavin mononucleotide (FMN) reductases and reduces FMN as a substrate. NfoR binds FMN with a dissociation constant of 0.4 µM. The enzyme then binds NADPH with a dissociation constant of 140 µM and reduces the flavin at a rate of 1,350 s-1 Turnover of the enzyme is apparently limited by the rate of product release that occurs, with a net rate constant of 0.45 s-1 The rate of product release limits the rate of observed chromate reduction, so the net rate of chromate reduction by NfoR is orders of magnitude lower than when this process occurs in solution. We propose that NfoR is an FMN reductase and that the criterion required to define chromate reduction as enzymatic has not been met. That NfoR expression is increased in the presence of chromate suggests that the survival adaption was to increase the net rate of chromate reduction by facile, adventitious redox processes.IMPORTANCE Chromate is a toxic by-product of multiple industrial processes. Chromate reduction is an important biological activity that ameliorates Cr(VI) toxicity. Numerous researchers have identified chromate reductase activity by observing chromate reduction. However, all identified chromate reductase enzymes have flavin as a cofactor or use a flavin as a substrate. We show here that NfoR, an enzyme claimed to be a chromate reductase, is in fact an FMN reductase. In addition, we show that reduction of a flavin is a viable way to transfer electrons to chromate but that it is unlikely to be the native function of enzymes. We propose that upregulation of a redox-active flavoprotein is a viable means to detoxify chromate that relies on adventitious reduction that is not catalyzed.


Subject(s)
Bacterial Proteins/genetics , FMN Reductase/genetics , Gene Expression Regulation, Bacterial , Oxidoreductases/genetics , Staphylococcus aureus/genetics , Bacterial Proteins/metabolism , FMN Reductase/metabolism , Oxidoreductases/metabolism , Staphylococcus aureus/enzymology
19.
Arch Biochem Biophys ; 690: 108474, 2020 09 15.
Article in English | MEDLINE | ID: mdl-32687799

ABSTRACT

Kynurenine 3-monoxygenase (KMO) catalyzes the conversion of l-kynurenine (L-Kyn) to 3-hydroxykynurenine (3-OHKyn) in the pathway for tryptophan catabolism. We have investigated the effects of pH and deuterium substitution on the oxidative half-reaction of KMO from P. fluorescens (PfKMO). The three phases observed during the oxidative half reaction are formation of the hydroperoxyflavin, hydroxylation and product release. The measured rate constants for these phases proved largely unchanging with pH, suggesting that the KMO active site is insulated from exchange with solvent during catalysis. A solvent inventory study indicated that a solvent isotope effect of 2-3 is observed for the hydroxylation phase and that two or more protons are in flight during this step. An inverse isotope effect of 0.84 ± 0.01 on the rate constant for the hydroxylation step with ring perdeutero-L-Kyn as a substrate indicates a shift from sp2 to sp3 hybridization in the transition state leading to the formation of a non-aromatic intermediate. The pH dependence of transient state data collected for the substrate analog meta-nitrobenzoylalanine indicate that groups proximal to the hydroperoxyflavin are titrated in the range pH 5-8.5 and can be described by a pKa of 8.8. That higher pH values do not slow the rate of hydroxylation precludes that the pKa measured pertains to the proton of the hydroperoxflavin. Together, these observations indicate that the C4a-hydroperoxyflavin has a pKa ≫ 8.5, that a non-aromatic species is the immediate product of hydroxylation and that at least two solvent derived protons are in-flight during oxygen insertion to the substrate aromatic ring. A unifying mechanistic proposal for these observations is proposed.


Subject(s)
Hydrogen/chemistry , Kynurenine 3-Monooxygenase/chemistry , Kynurenine 3-Monooxygenase/metabolism , Kynurenine/chemistry , Pseudomonas fluorescens/chemistry , Catalysis , Catalytic Domain , Deuterium/chemistry , Dinitrocresols/metabolism , Flavins/chemistry , Hydrogen-Ion Concentration , Hydroxylation , Kinetics , Mixed Function Oxygenases/metabolism , Oxidation-Reduction , Oxidative Stress , Oxygen/chemistry , Protons , Solvents/chemistry
20.
Protein Expr Purif ; 171: 105610, 2020 07.
Article in English | MEDLINE | ID: mdl-32088324

ABSTRACT

Dihydropyrimidine dehydrogenase (DPD) catalyzes the reduction of uracil and thymine bases with electrons derived from NADPH. The mammalian DPD enzyme is a functional homodimer and has an elaborate cofactor arrangement. Two flavin cofactors (FAD and FMN) reside in two active site cavities that are separated by around 60 Å. The flavins are apparently bridged by four Fe4S4 clusters, two of which are provided by the partner protomer of the dimer. The study of DPD has been hampered by modest yield from both native sources and from heterologous expression in E. coli. In addition, minimal active enzyme is obtained when the DPD gene is fused to an N-terminal 6His-tag. This limitation has dictated the use of traditional purification methods that are made more challenging by apparent over-expression of truncated and/or non-active forms of DPD. Here we detail methods of expression and purification that result in a ~4-fold improvement in the yield of active porcine DPD when expressed in E. coli BL21 DE3 cells via the pET plasmid expression system. The addition of ferrous ions and sulfate during induction provide a small increase in purified active enzyme. However, the addition of FAD and FMN during cell lysis results in a substantial increase in activity that also reduces the relative proportion of non-active, high molecular weight protein contaminants. We also describe methods that permit correlation of the flavin content with the amount of active enzyme and thus permit simple, rapid quantitation and evaluation of purified DPD sample.


Subject(s)
Dihydrouracil Dehydrogenase (NADP) , Gene Expression , Recombinant Fusion Proteins , Animals , Dihydrouracil Dehydrogenase (NADP)/biosynthesis , Dihydrouracil Dehydrogenase (NADP)/chemistry , Dihydrouracil Dehydrogenase (NADP)/genetics , Dihydrouracil Dehydrogenase (NADP)/isolation & purification , Escherichia coli/enzymology , Escherichia coli/genetics , Recombinant Fusion Proteins/biosynthesis , Recombinant Fusion Proteins/chemistry , Recombinant Fusion Proteins/genetics , Recombinant Fusion Proteins/isolation & purification , Swine/genetics
SELECTION OF CITATIONS
SEARCH DETAIL