首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 757 毫秒
1.
Bacillus subtilis is one of the most studied gram‐positive bacteria. In this work, YvgN and YtbE from B. subtilis, assigned as AKR5G1 and AKR5G2 of aldo‐keto reductase (AKR) superfamily. AKR catalyzes the NADPH‐dependent reduction of aldehyde or aldose substrates to alcohols. YvgN and YtbE were studied by crystallographic and enzymatic analyses. The apo structures of these proteins were determined by molecular replacement, and the structure of holoenzyme YvgN with NADPH was also solved, revealing the conformational changes upon cofactor binding. Our biochemical data suggest both YvgN and YtbE have preferential specificity for derivatives of benzaldehyde, such as nitryl or halogen group substitution at the 2 or 4 positions. These proteins also showed broad catalytic activity on many standard substrates of AKR, such as glyoxal, dihydroxyacetone, and DL‐glyceraldehyde, suggesting a possible role in bacterial detoxification.  相似文献   

2.
The 2.2 Å X-ray crystal structure of Candida tenuis xylose reductase (AKR2B5) bound with NADP+ reveals that Phe-114 contributes to the substrate binding pocket of the enzyme. In the related human aldose reductase (AKR1B1), this phenylalanine is replaced by a tryptophan. The side chain of Trp was previously implicated in forming a hydrogen bond with bound substrate or inhibitor. The apparent Michaelis constant of AKR2B5 for xylose (Km≈90 mM) is 60 times that of AKR1B1, perhaps because critical enzyme–substrate interactions of Trp are not available to Phe-114. We, therefore, prepared a Phe-114→Trp mutant (F114W) of AKR2B5, to mimic the aldose reductase relationship in xylose reductase. Detailed analysis of the kinetic consequences in purified F114W revealed that the Km values for xylose and xylitol at pH 7.0 and 25°C were increased 5.1- and 4.4-fold, respectively, in the mutant compared with the wild-type. Turnover numbers (kcat) of F114W for xylose reduction and xylitol oxidation were half those of the wild-type. Apparent dissociation constants of NADH (KiNADH=44 µM) and NAD+ (KiNAD+=177 µM) were increased 1.6- and 1.4-fold in comparison with values of KiNADH and KiNAD+ for the wild-type, respectively. Catalytic efficiencies (kcat/Km) for NADH-dependent reduction of different aldehydes were between 3.1- and 31.5-fold lower than the corresponding kcat/Km values of the wild-type. Therefore, replacement of Phe-114 with Trp weakens rather than strengthens apparent substrate binding by AKR2B5, suggesting that xylose reductase exploits residue 114 in a different manner from aldose reductase.  相似文献   

3.
Conjugated polyketone reductase (CPR‐C1) from Candida parapsilosis IFO 0708 is a member of the aldo–keto reductase (AKR) superfamily and reduces ketopantoyl lactone to d ‐pantoyl lactone in a NADPH‐dependent and stereospecific manner. We determined the crystal structure of CPR‐C1.NADPH complex at 2.20 Å resolution. CPR‐C1 adopted a triose‐phosphate isomerase (TIM) barrel fold at the core of the structure in which Thr25 and Lys26 of the GXGTX motif bind uniquely to the adenosine 2′‐phosphate group of NADPH. This finding provides a novel structural basis for NADPH binding of the AKR superfamily. Proteins 2013; 81:2059–2063. © 2013 Wiley Periodicals, Inc.  相似文献   

4.
In this study, we characterized rat and mouse aldo-keto reductases (AKR1C16 and AKR1C13, respectively) with 92% sequence identity. The recombinant enzymes oxidized non-steroidal alcohols using NAD+ as the preferred coenzyme, and showed low 3α/17β/20α-hydroxysteroid dehydrogenase (HSD) activities. The substrate specificity differs from that of rat NAD+-dependent 3α-HSD (AKR1C17) that shares 95% sequence identity with AKR1C16. To elucidate the residues determining the substrate specificity of the enzymes, we performed site-directed mutagenesis of Tyr24, Asp128 and Phe129 of AKR1C16 with the corresponding residues (Ser, Tyr and Leu, respectively) of AKR1C17. The double mutation (Asp128/Tyr-Phe129/Leu) had few effects on the substrate specificity, while the Tyr24/Ser mutant showed only 3α-HSD activity, and the triple mutation of the three residues produced an enzyme that had almost the same properties as AKR1C17. The importance of the residue 24 for substrate recognition was verified by the mutagenesis of Ser24/Tyr of AKR1C17 which resulted in a decrease in 3α-HSD activity and appearance of 17β- and 20α-HSD activities. AKR1C16 is also 92% identical with rat NAD+-dependent 17β-HSD (AKR1C24), which possesses Tyr24. The replacement of Asp128, Phe129 and Ser137 of AKR1C16 with the corresponding residues (Glu, Ser and Phe, respectively) of AKR1C24 increased the catalytic efficiency for 17β- and 20α-hydroxysteroids.  相似文献   

5.
A new member of the aldo–keto reductase (AKR) superfamily with 3-dehydroecdysone reductase activity was found in the silkworm Bombyx mori upon induction by the insecticide diazinon. The amino acid sequence showed that this enzyme belongs to the AKR2 family, and the protein was assigned the systematic name AKR2E4. In this study, recombinant AKR2E4 was expressed, purified to near homogeneity, and kinetically characterized. Additionally, its ternary structure in complex with NADP+ and citrate was refined at 1.3 Å resolution to elucidate substrate binding and catalysis. The enzyme is a 33-kDa monomer and reduces dicarbonyl compounds such as isatin and 17α-hydroxy progesterone using NADPH as a cosubstrate. No NADH-dependent activity was detected. Robust activity toward the substrate inhibitor 3-dehydroecdysone was observed, which suggests that this enzyme plays a role in regulation of the important molting hormone ecdysone. This structure constitutes the first insect AKR structure determined. Bound NADPH is located at the center of the TIM- or (β/α)8-barrel, and residues involved in catalysis are conserved.  相似文献   

6.
Aldo-keto reductases (AKRs) constitute a large protein superfamily of mainly NAD(P)-dependent oxidoreductases involved in carbonyl metabolism. Catalysis is promoted by a conserved tetrad of active site residues (Tyr, Lys, Asp and His). Recent results of structure-function relationship studies for xylose reductase (AKR2B5) require an update of the proposed catalytic mechanism. Electrostatic stabilization by the epsilon-NH3+ group of Lys is a key source of catalytic power of xylose reductase. A molecular-level analysis of the substrate binding pocket of xylose reductase provides a case of how a very broadly specific AKR achieves the requisite selectivity for its physiological substrate and could serve as the basis for the design of novel reductases with improved specificities for biocatalytic applications.  相似文献   

7.
CYP107W1 from Streptomyces avermitilis is a cytochrome P450 enzyme involved in the biosynthesis of macrolide oligomycin A. A previous study reported that CYP107W1 regioselectively hydroxylated C12 of oligomycin C to produce oligomycin A, and the crystal structure of ligand free CYP107W1 was determined. Here, we analyzed the structural properties of the CYP107W1-oligomycin A complex and characterized the functional role of the Trp178 residue in CYP107W1. The crystal structure of the CYP107W1 complex with oligomycin A was determined at a resolution of 2.6 Å. Oligomycin A is bound in the substrate access channel on the upper side of the prosthetic heme mainly by hydrophobic interactions. In particular, the Trp178 residue in the active site intercalates into the large macrolide ring, thereby guiding the substrate into the correct binding orientation for a productive P450 reaction. A Trp178 to Gly mutation resulted in the distortion of binding titration spectra with oligomycin A, whereas binding spectra with azoles were not affected. The Gly178 mutant’s catalytic turnover number for the 12-hydroxylation reaction of oligomycin C was highly reduced. These results indicate that Trp178, located in the open pocket of the active site, may be a critical residue for the productive binding conformation of large macrolide substrates.  相似文献   

8.
A procedure for the synthesis of a11C‐labeled oligopeptide containing [1‐11C]1,2,3,4‐tetrahydro‐β‐carboline‐3‐carboxylic acid ([1‐11C]Tpi) from the corresponding Trp?HCl‐containing peptides has been developed involving a Pictet‐Spengler reaction with [11C]formaldehyde. The synthesis of [1‐11C]Tpi from Trp and [11C]formaldehyde was examined as a model reaction with the aim of developing a facile and effective method for the labeling of peptides with carbon‐11. The Pictet‐Spengler reaction of Trp and [11C]formaldehyde in acidic media (TsOH or HCl) afforded the desired [1‐11C]Tpi in a moderate radiochemical yield. Herein, the application of a Pictet‐Spengler reaction to an aqueous solution of Trp?HCl gave the desired product with a radiochemical yield of 45.2%. The RGD peptide cyclo[Arg‐Gly‐Asp‐D‐Tyr‐Lys] was then selected as a substrate for the labeling reaction with [11C]formaldehyde. The radiolabeling of a Trp?HCl‐containing RGD peptide using the Pictet‐Spengler reaction was successful. Furthermore, the remote‐controlled synthesis of a [1‐11C]Tpi‐containing RGD peptide was attempted by using an automatic production system to generate [11C]CH3I. The radiochemical yield of the [1‐11C]Tpi‐containing RGD at the end of synthesis (EOS) was 5.9 ± 1.9% (n = 4), for a total synthesis time of about 35 min. The specific activity was 85.7 ± 9.4 GBq/µmol at the EOS. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
Xu H  West AH  Cook PF 《Biochemistry》2007,46(25):7625-7636
A survey of NADH, alpha-Kg, and lysine analogues has been undertaken in an attempt to define the substrate specificity of saccharopine dehydrogenase and to identify functional groups on all substrates and dinucleotides important for substrate binding. A number of NAD analogues, including NADP, 3-acetylpyridine adenine dinucleotide (3-APAD), 3-pyridinealdehyde adenine dinucleotide (3-PAAD), and thionicotinamide adenine dinucleotide (thio-NAD), can serve as a substrate in the oxidative deamination reaction, as can a number of alpha-keto analogues, including glyoxylate, pyruvate, alpha-ketobutyrate, alpha-ketovalerate, alpha-ketomalonate, and alpha-ketoadipate. Inhibition studies using nucleotide analogues suggest that the majority of the binding energy of the dinucleotides comes from the AMP portion and that distinctly different conformations are generated upon binding of the oxidized and reduced dinucleotides. Addition of the 2'-phosphate as in NADPH causes poor binding of subsequent substrates but has little effect on coenzyme binding and catalysis. In addition, the 10-fold decrease in affinity of 3-APAD in comparison to NAD suggests that the nicotinamide ring binding pocket is hydrophilic. Extensive inhibition studies using aliphatic and aromatic keto acid analogues have been carried out to gain insight into the keto acid binding pocket. Data suggest that a side chain with three carbons (from the alpha-keto group up to and including the side chain carboxylate) is optimal. In addition, the distance between the C1-C2 unit and the C5 carboxylate of the alpha-keto acid is also important for binding; the alpha-oxo group contributes a factor of 10 to affinity. The keto acid binding pocket is relatively large and flexible and can accommodate the bulky aromatic ring of a pyridine dicarboxylic acid and a negative charge at the C3 but not the C4 position. However, the amino acid binding site is hydrophobic, and the optimal length of the hydrophobic portion of the amino acid carbon side chain is three or four carbons. In addition, the amino acid binding pocket can accommodate a branch at the gamma-carbon, but not at the beta-carbon.  相似文献   

10.
Human aldo-keto reductase 1D1 (AKR1D1) and AKR1C enzymes are essential for bile acid biosynthesis and steroid hormone metabolism. AKR1D1 catalyzes the 5β-reduction of Δ(4)-3-ketosteroids, whereas AKR1C enzymes are hydroxysteroid dehydrogenases (HSDs). These enzymes share high sequence identity and catalyze 4-pro-(R)-hydride transfer from NADPH to an electrophilic carbon but differ in that one residue in the conserved AKR catalytic tetrad, His(120) (AKR1D1 numbering), is substituted by a glutamate in AKR1D1. We find that the AKR1D1 E120H mutant abolishes 5β-reductase activity and introduces HSD activity. However, the E120H mutant unexpectedly favors dihydrosteroids with the 5α-configuration and, unlike most of the AKR1C enzymes, shows a dominant stereochemical preference to act as a 3β-HSD as opposed to a 3α-HSD. The catalytic efficiency achieved for 3β-HSD activity is higher than that observed for any AKR to date. High resolution crystal structures of the E120H mutant in complex with epiandrosterone, 5β-dihydrotestosterone, and Δ(4)-androstene-3,17-dione elucidated the structural basis for this functional change. The glutamate-histidine substitution prevents a 3-ketosteroid from penetrating the active site so that hydride transfer is directed toward the C3 carbonyl group rather than the Δ(4)-double bond and confers 3β-HSD activity on the 5β-reductase. Structures indicate that stereospecificity of HSD activity is achieved because the steroid flips over to present its α-face to the A-face of NADPH. This is in contrast to the AKR1C enzymes, which can invert stereochemistry when the steroid swings across the binding pocket. These studies show how a single point mutation in AKR1D1 can introduce HSD activity with unexpected configurational and stereochemical preference.  相似文献   

11.
Cofactor specificity in the aldo‐keto reductase (AKR) superfamily has been well studied, and several groups have reported the rational alteration of cofactor specificity in these enzymes. Although most efforts have focused on mesostable AKRs, several putative AKRs have recently been identified from hyperthermophiles. The few that have been characterized exhibit a strong preference for NAD(H) as a cofactor, in contrast to the NADP(H) preference of the mesophilic AKRs. Using the design rules elucidated from mesostable AKRs, we introduced two site‐directed mutations in the cofactor binding pocket to investigate cofactor specificity in a thermostable AKR, AdhD, which is an alcohol dehydrogenase from Pyrococcus furiosus. The resulting double mutant exhibited significantly improved activity and broadened cofactor specificity as compared to the wild‐type. Results of previous pre‐steady‐state kinetic experiments suggest that the high affinity of the mesostable AKRs for NADP(H) stems from a conformational change upon cofactor binding which is mediated by interactions between a canonical arginine and the 2′‐phosphate of the cofactor. Pre‐steady‐state kinetics with AdhD and the new mutants show a rich conformational behavior that is independent of the canonical arginine or the 2′‐phosphate. Additionally, experiments with the highly active double mutant using NADPH as a cofactor demonstrate an unprecedented transient behavior where the binding mechanism appears to be dependent on cofactor concentration. These results suggest that the structural features involved in cofactor specificity in the AKRs are conserved within the superfamily, but the dynamic interactions of the enzyme with cofactors are unexpectedly complex. Biotechnol. Bioeng. 2010;107: 763–774. © 2010 Wiley Periodicals, Inc.  相似文献   

12.
A three-dimensional model of the neuropeptide Y (NPY) - rat Y1 (rY1) receptor complex and of the NPY 13-36 - rY1 receptor complex was constructed by molecular modeling based on the electron density projection map of rhodopsin and on site-directed mutagenesis studies of neuropeptide receptors. In order to further guide the modeling, the nucleotide sequences encoding Trp287, Cys295 and His297 in the third extracellular loop of the rY1 receptor, were altered by site-directed mutagenesis experiments. Single-point mutated receptors were expressed in COS-7 cells, and tested for their ability to bind radio labelled NPY (3H-NPY). Mutations of Trp287 and His297 completely abolished binding of 3H-NPY. The Cys295Ser mutation only slightly decreased the binding of 3H-NPY, suggesting that the involvement of Cys295 in a disulphide bond is not essential for maintaining the correct three-dimensional structure of the binding site for NPY. Molecular dynamics simulations of NPY-rY1 receptor interactions suggested that Asp199, Asp103 and Asp286 in the receptor interact, respectively, with Lys4, Arg33 and Arg35 of NPY. The simulations also suggested that His297 acts as a hydrogen acceptor from Arg35 in NPY, and that Tyr1 of NPY interacts with a binding pocket on the receptor formed by Asn115, Asp286, Trp287 and His297. Tyr36 in NPY interacted both with Thr41 and Tyr99 via hydrogen bonds, and also with Asn296, His297 and Phe301. The present study suggests that amino acid residues at the extracellular end of the transmembrane helices and in the extracellular loops are strongly involved in binding to NPY and NPY13-36.Electronic Supplementary Material available.  相似文献   

13.
Streptococcus pneumoniae Sp1610, a Class‐I fold S‐adenosylmethionine (AdoMet)‐dependent methyltransferase, is a member of the COG2384 family in the Clusters of Orthologous Groups database, which catalyzes the methylation of N1‐adenosine at position 22 of bacterial tRNA. We determined the crystal structure of Sp1610 in the ligand‐free and the AdoMet‐bound forms at resolutions of 2.0 and 3.0 Å, respectively. The protein is organized into two structural domains: the N‐terminal catalytic domain with a Class I AdoMet‐dependent methyltransferase fold, and the C‐terminal substrate recognition domain with a novel fold of four α‐helices. Observations of the electrostatic potential surface revealed that the concave surface located near the AdoMet binding pocket was predominantly positively charged, and thus this was predicted to be an RNA binding area. Based on the results of sequence alignment and structural analysis, the putative catalytic residues responsible for substrate recognition are also proposed.  相似文献   

14.
The structure of aldehyde reductase (ALR1) in ternary complex with the coenzyme NADPH and 3,5-dichlorosalicylic acid (DCL), a potent inhibitor of human 20α-hydroxysteroid dehydrogenase (AKR1C1), was determined at a resolution of 2.41 Å. The inhibitor formed a network of hydrogen bonds with the active site residues Trp22, Tyr50, His113, Trp114 and Arg312. Molecular modelling calculations together with inhibitory activity measurements indicated that DCL was a less potent inhibitor of ALR1 (256-fold) when compared to AKR1C1. In AKR1C1, the inhibitor formed a 10-fold stronger binding interaction with the catalytic residue (Tyr55), non-conserved hydrogen bonding interaction with His222, and additional van der Waals contacts with the non-conserved C-terminal residues Leu306, Leu308 and Phe311 that contribute to the inhibitor’s selectivity advantage for AKR1C1 over ALR1.  相似文献   

15.
Zhu X  Lapthorn AJ  Ellis EM 《Biochemistry》2006,45(6):1562-1570
The aldo-keto reductases make up a superfamily of enzymes which can reduce a variety of aldehydes and ketones to their corresponding alcohols. Within each family are distinct preferences for certain substrates, presumably reflecting their role within the cell. The original member of the AKR7A subfamily was purified from liver as an aflatoxin dialdehyde reductase AKR7A1. However, recent additions to the family have revealed that even closely related enzymes have clear substrate preferences with AKR7A2, AKR7A4, and AKR7A5 showing much higher affinities for succinic semialdehyde (SSA) than does AKR7A1. To investigate the structural basis of this specificity, the crystal structure of mouse AKR7A5 has been determined to better than 2.5 A resolution. The structure is of the ternary complex of the enzyme with NADP+ and tartrate as an inhibitor. This structure has the same overall fold as the previously determined structure of AKR7A1; however, there are a number of differences in loops around the active site that contribute to observed differences in the substrate specificity between the AKR7A enzymes. Several differences are the result of bulky hydrophobic residues found in AKR7A5, namely, Met44, Trp77, and Trp224, which significantly restrict the size and modify the architecture of the substrate-binding pocket, producing a tighter or less flexible binding site for SSA than in AKR7A1. Site-directed mutagenesis was used to introduce Met44, Trp77, and Trp224 individually into AKR7A1, to test if they improved the affinity of the enzyme for SSA. Each mutation showed improved affinity for SSA, with Trp77Met having the largest effect. This confirms the role of these amino acids as substrate determinants for SSA.  相似文献   

16.
Two hormone-sensitive lipase (HSL) family esterases (RmEstA and RmEstB) from the thermophilic fungus Rhizomucor miehei, exhibiting distinct substrate specificity, have been recently reported to show great potential in industrial applications. In this study, the crystal structures of RmEstA and RmEstB were determined at 2.15 Å and 2.43 Å resolutions, respectively. The structures of RmEstA and RmEstB showed two distinctive domains, a catalytic domain and a cap domain, with the classical α/β-hydrolase fold. Catalytic triads consisting of residues Ser161, Asp262, and His292 in RmEstA, and Ser164, Asp261, and His291 in RmEstB were found in the respective canonical positions. Structural comparison of RmEstA and RmEstB revealed that their distinct substrate specificity might be attributed to their different substrate-binding pockets. The aromatic amino acids Phe222 and Trp92, located in the center of the substrate-binding pocket of RmEstB, blocked this pocket, thus narrowing its catalytic range for substrates (C2–C8). Two mutants (F222A and W92F in RmEstB) showing higher catalytic activity toward long-chain substrates further confirmed the hypothesized interference. This is the first report of HSL family esterase structures from filamentous fungi.jlr The information on structure-function relationships could open important avenues of exploration for further industrial applications of esterases.  相似文献   

17.
The human Monoamine oxidase (hMAO) metabolizes several biogenic amine neurotransmitters and is involved in different neurological disorders. Extensive MD simulation studies of dopamine-docked hMAO B structures have revealed the stabilization of amino-terminal of the substrate by a direct and water-mediated interaction of catalytic tyrosines, Gln206, and Leu171 residues. The catechol ring of the substrate is stabilized by Leu171(C–H)?π(Dop)?(H–C) Ile199 interaction. Several conserved water molecules are observed to play a role in the recognition of substrate to the enzyme, where W1 and W2 associate in dopamine– FAD interaction, reversible dynamics of W3 and W4 influenced the coupling of Tyr435 to Trp432 and FAD, and W5 and W8 stabilized the catalytic Tyr188/398 residues. The W6, W7, and W8 water centers are involved in the recognition of catalytic residues and FAD with the N+- site of dopamine through hydrogen bonding interaction. The recognition of substrate to gating residues is made through W9, W10, and W11 water centers. Beside the interplay of water molecules, the catalytic aromatic cage has also been stabilized by π?water, π?C–H, and π?π interactions. The topology of conserved water molecular sites along with the hydration dynamics of catalytic residues, FAD, and dopamine has added a new feature on the substrate binding chemistry in hMAO B which may be useful for substrate analog inhibitor design.  相似文献   

18.
Methionine γ‐lyse (MGL) catalyzes the α, γ‐elimination of l ‐methionine and its derivatives as well as the α, β‐elimination of l ‐cysteine and its derivatives to produce α‐keto acids, volatile thiols, and ammonia. The reaction mechanism of MGL has been characterized by enzymological studies using several site‐directed mutants. The Pseudomonas putida MGL C116H mutant showed drastically reduced degradation activity toward methionine while retaining activity toward homocysteine. To understand the underlying mechanism and to discern the subtle differences between these substrates, we analyzed the crystal structures of the reaction intermediates. The complex formed between the C116H mutant and methionine demonstrated that a loop structure (Ala51–Asn64) in the adjacent subunit of the catalytic dimer cannot approach the cofactor pyridoxal 5′‐phosphate (PLP) because His116 disrupts the interaction of Asp241 with Lys240, and the liberated side chain of Lys240 causes steric hindrance with this loop. Conversely, in the complex formed between C116H mutant and homocysteine, the thiol moiety of the substrate conjugated with PLP offsets the imidazole ring of His116 via a water molecule, disrupting the interaction of His116 and Asp241 and restoring the interaction of Asp241 with Lys240. These structural data suggest that the Cys116 to His mutation renders the enzyme inactive toward the original substrate, but activity is restored when the substrate is homocysteine due to substrate‐assisted catalysis.  相似文献   

19.
Chen M  Drury JE  Penning TM 《Steroids》2011,76(5):484-490
Human steroid 5β-reductase (aldo-keto reductase 1D1) catalyzes the stereospecific NADPH-dependent reduction of the C4-C5 double bond of Δ4-ketosteroids to yield an A/B cis-ring junction. This cis-configuration is crucial for bile acid biosynthesis and plays important roles in steroid metabolism. The biochemical properties of the enzyme have not been thoroughly studied and conflicting data have been reported, partially due to the lack of highly homogeneous protein. In the present study, we systematically determined the substrate specificity of homogeneous human recombinant AKR1D1 using C18, C19, C21, and C27 Δ4-ketosteroids and assessed the pH-rate dependence of the enzyme. Our results show that AKR1D1 proficiently reduced all the steroids tested at physiological pH, indicating AKR1D1 is the only enzyme necessary for all the 5β-steroid metabolites present in humans. Substrate inhibition was observed with C18 to C21 steroids provided that the C11 position was unsubstituted. This structure activity relationship can be explained by the existence of a small alternative substrate binding pocket revealed by the AKR1D1 crystal structure. Non-steroidal anti-inflammatory drugs which are potent inhibitors of the related AKR1C enzymes do not inhibit AKR1D1. By contrast chenodeoxycholate and ursodeoxycholate were found to be potent non-competitive inhibitors suggesting that bile-acids may regulate their own synthesis at the level of AKR1D1 inhibition.  相似文献   

20.
N10‐formyltetrahydrofolate synthetase (FTHFS) is a folate enzyme that catalyzes the formylation of tetrahydrofolate (THF) in an ATP dependent manner. Structures of FTHFS from the thermophilic homoacetogen, Moorella thermoacetica, complexed with (1) a catalytic intermediate—formylphosphate (XPO) and product—ADP; (2) with an inhibitory substrate analog–folate; (3) with XPO and an inhibitory THF analog, ZD9331, were used to analyze the enzyme mechanism. Nucleophilic attack of the formate ion on the gamma phosphate of ATP leads to the formation of XPO and the first product ADP. A channel that leads to the putative formate binding pocket allows for the binding of ATP and formate in random order. Formate binding is due to interactions with the gamma‐phosphate moiety of ATP and additionally to two hydrogen bonds from the backbone nitrogen of Ala276 and the side chain of Arg97. Upon ADP dissociation, XPO reorients and moves to the position previously occupied by the beta‐phosphate of ATP. Conformational changes that occur due to the XPO presence apparently allow for the recruitment of the third substrate, THF, with its pterin moiety positioned between Phe384 and Trp412. This position overlaps with that of the bound nucleoside, which is consistent with a catalytic mechanism hypothesis that FTHFS works via a sequential ping‐pong mechanism. More specifically, a random bi uni uni bi ping‐pong ter ter mechanism is proposed. Additionally, the native structure originally reported at a 2.5 Å resolution was redetermined at a 2.2 Å resolution.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号