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 0 毫秒
1.
Flavin-containing reductases are involved in a wide variety of physiological reactions such as photosynthesis, nitric oxide synthesis, and detoxification of foreign compounds, including therapeutic drugs. Ferredoxin-NADP(H)-reductase (FNR) is the prototypical enzyme of this family. The fold of this protein is highly conserved and occurs as one domain of several multidomain enzymes such as the members of the diflavin reductase family. The enzymes of this family have emerged as fusion of a FNR and a flavodoxin. Although the active sites of these enzymes are very similar, different enzymes function in opposite directions, that is, some reduce oxidized nicotinamide adenine dinucleotide phosphate (NADP(+)) and some oxidize reduced nicotinamide adenine dinucleotide phosphate (NADPH). In this work, we analyze the protonation behavior of titratable residues of these enzymes through electrostatic calculations. We find that a highly conserved carboxylic acid in the active site shows a different titration behavior in different flavin reductases. This residue is deprotonated in flavin reductases present in plastids, but protonated in bacterial counterparts and in diflavin reductases. The protonation state of the carboxylic acid may also influence substrate binding. The physiological substrate for plastidic enzymes is NADP(+), but it is NADPH for the other mentioned reductases. In this article, we discuss the relevance of the environment of this residue for its protonation and its importance in catalysis. Our results allow to reinterpret and explain experimental data.  相似文献   

2.
3.
The quenching of coenzyme fluorescence in glycogen phosphorylase b is reinvestigated. Data with anionic quenchers show deviations from the original Stern-Volmer kinetics. A kinetic analysis based on measured lifetime data indicates a collisional quenching process, which is, however, not diffusion-controlled. It is proposed, that the quenching takes place primarily by enzyme-bound quencher species. The observed inhibition of the enzyme reaction by I- and IO-3 is consistent with this hypothesis. The inhibition pattern and spectral investigation refer to a true competition with the substrate, glucose-1-phosphate. So, this dynamic quenching can be regarded as an indicator of rapid conformational fluctuations which bring the two important active-site groups in contact. Effect of ligand binding on the quenching of coenzyme fluorescence should also be revaluated according to these results.  相似文献   

4.
The active site of myo-inositol dehydrogenase (IDH, EC 1.1.1.18) from Bacillus subtilis recognizes a variety of mono- and disaccharides, as well as 1l-4-O-substituted inositol derivatives. It catalyzes the NAD+-dependent oxidation of the axial alcohol of these substrates with comparable kinetic constants. We have found that 4-O-p-toluenesulfonyl-myo-inositol does not act as a substrate for IDH, in contrast to structurally similar compounds such as those bearing substituted benzyl substituents in the same position. X-ray crystallographic analysis of 4-O-p-toluenesulfonyl-myo-inositol and 4-O-(2-naphthyl)methyl-myo-inositol, which is a substrate for IDH, shows a distinct difference in the preferred conformation of the aryl substituent. Conformational analysis of known substrates of IDH suggests that this conformational difference may account for the difference in reactivity of 4-O-p-toluenesulfonyl-myo-inositol in the presence of IDH. A sequence alignment of IDH with the homologous glucose-fructose oxidoreductase allowed the construction of an homology model of inositol dehydrogenase, to which NADH and 4-O-benzyl-scyllo-inosose were docked and the active site energy minimized. The active site model is consistent with all experimental results and suggests that a conserved tyrosine-glycine-tyrosine motif forms the hydrophobic pocket adjoining the site of inositol recognition. Y233F and Y235F retain activity, while Y233R and Y235R do not. A histidine-aspartate pair, H176 and D172, are proposed to act as a dyad in which H176 is the active site acid/base. The enzyme is inactivated by diethyl pyrocarbonate, and the mutants H176A and D172N show a marked loss of activity. Kinetic isotope effect experiments with D172N indicate that chemistry is rate-determining for this mutant.  相似文献   

5.
6.
The decahaem homodimeric cytochrome c nitrite reductase (NrfA) is expressed within the periplasm of a wide range of Gamma-, Delta- and Epsilon-proteobacteria and is responsible for the six-electron reduction of nitrite to ammonia. This allows nitrite to be used as a terminal electron acceptor, facilitating anaerobic respiration while allowing nitrogen to remain in a biologically available form. NrfA has also been reported to reduce nitric oxide (a reaction intermediate) and sulfite to ammonia and sulfide respectively, suggesting a potential secondary role as a detoxification enzyme. The protein sequences and crystal structures of NrfA from different bacteria and the closely related octahaem nitrite reductase from Thioalkalivibrio nitratireducens (TvNir) reveal that these enzymes are homologous. The NrfA proteins contain five covalently attached haem groups, four of which are bis-histidine-co-ordinated, with the proximal histidine being provided by the highly conserved CXXCH motif. These haems are responsible for intraprotein electron transfer. The remaining haem is the site for nitrite reduction, which is ligated by a novel lysine residue provided by a CXXCK haem-binding motif. The TvNir nitrite reductase has five haems that are structurally similar to those of NrfA and three extra bis-histidine-coordinated haems that precede the NrfA conserved region. The present review compares the protein sequences and structures of NrfA and TvNir and discusses the subtle differences related to active-site architecture and Ca2+ binding that may have an impact on substrate reduction.  相似文献   

7.
Kinetic isotope effects in reactions involving hydride transfer and their temperature dependence are powerful tools to explore dynamics of enzyme catalytic sites. In plant-type ferredoxin-NADP+ reductases the FAD cofactor exchanges a hydride with the NADP(H) coenzyme. Rates for these processes are considerably faster for the plastidic members (FNR) of the family than for those belonging to the bacterial class (FPR). Hydride transfer (HT) and deuteride transfer (DT) rates for the NADP+ coenzyme reduction of four plant-type FNRs (two representatives of the plastidic type FNRs and the other two from the bacterial class), and their temperature dependences are here examined applying a full tunnelling model with coupled environmental fluctuations. Parameters for the two plastidic FNRs confirm a tunnelling reaction with active dynamics contributions, but isotope effects on Arrhenius factors indicate a larger contribution for donor–acceptor distance (DAD) dynamics in the Pisum sativum FNR reaction than in the Anabaena FNR reaction. On the other hand, parameters for bacterial FPRs are consistent with passive environmental reorganisation movements dominating the HT coordinate and no contribution of DAD sampling or gating fluctuations. This indicates that active sites of FPRs are more organised and rigid than those of FNRs. These differences must be due to adaptation of the active sites and catalytic mechanisms to fulfil their particular metabolic roles, establishing a compromise between protein flexibility and functional optimisation. Analysis of site-directed mutants in plastidic enzymes additionally indicates the requirement of a minimal optimal architecture in the catalytic complex to provide a favourable gating contribution.  相似文献   

8.
Active site titrations provide essential information for studies of enzymes, e.g., functional purity and half-site reactivity. The use of a hitherto overlooked class of hydrolase substrates, the oxazolones, for an easy synthesis of reagents possessing desirable properties, e.g., specificity, solubility, and optical parameters, is described. The operational normality may be determined in aqueous media or organic solvents, e.g., dimethyl sulfoxide, which stabilize the enzymes for extended periods at room temperature. Absorbance or fluorescence measurements may be used for detection. The lower limit permitting analyses in this series was found to be 60 pmol of enzyme.  相似文献   

9.
The pathway of sterol biosynthesis is highly conserved in all eucaryotic cells. We demonstrated structural and functional conservation of the rate-limiting enzyme of the mammalian pathway, 3-hydroxy-3-methyl-glutaryl coenzyme A reductase (HMG-CoA reductase), between the yeast Saccharomyces cerevisiae and humans. The amino acid sequence of the two yeast HMG-CoA reductase isozymes was deduced from DNA sequence analysis of the HMG1 and HMG2 genes. Extensive sequence similarity existed between the region of the mammalian enzyme encoding the active site and the corresponding region of the two yeast isozymes. Moreover, each of the yeast isozymes, like the mammalian enzyme, contained seven potential membrane-spanning domains in the NH2-terminal region of the protein. Expression of cDNA clones encoding either hamster or human HMG-CoA reductase rescued the viability of hmg1 hmg2 yeast cells lacking this enzyme. Thus, mammalian HMG-CoA reductase can provide sufficient catalytic function to replace both yeast isozymes in vivo. The availability of yeast cells whose growth depends on human HMG-CoA reductase may provide a microbial screen to identify new drugs that can modulate cholesterol biosynthesis.  相似文献   

10.
Completed eukaryotic genomes were screened for medium-chain dehydrogenases/reductases (MDR). In the human genome, 23 MDR forms were found, a number that probably will increase, because the genome is not yet fully interpreted. Partial sequences already indicate that at least three further members exist. Within the MDR superfamily, at least eight families were distinguished. Three families are formed by dimeric alcohol dehydrogenases (ADH; originally detected in animals/plants), cinnamyl alcohol dehydrogenases (originally detected in plants) and tetrameric alcohol dehydrogenases (originally detected in yeast). Three further families are centred around forms initially detected as mitochondrial respiratory function proteins, acetyl-CoA reductases of fatty acid synthases, and leukotriene B4 dehydrogenases. The two remaining families with polyol dehydrogenases (originally detected as sorbitol dehydrogenase) and quinone reductases (originally detected as zeta-crystallin) are also distinct but with variable sequences. The most abundant families in the human genome are the dimeric ADH forms and the quinone oxidoreductases. The eukaryotic patterns are different from those of Escherichia coli. The different families were further evaluated by molecular modelling of their active sites as to geometry, hydrophobicity and volume of substrate-binding pockets. Finally, sequence patterns were derived that are diagnostic for the different families and can be used in genome annotations.  相似文献   

11.
The content of coenzyme A-SH (CoASH) and acetyl-CoA of suspensions of rat heart mitochondria was stabilized by the addition of DL-carnitine and acetyl-DL-carnitine, in the presence of the respiratory inhibitor rotenone. The mitochondrial content of NAD+ and NADH was similarly stabilized by the addition of acetoacetate and DL-3-hydroxybutyrate, and the content of ADP and ATP was imposed by the addition of these nucleotides to the mitochondrial suspension, in the presence of uncoupling agent and oligomycin, to inhibit ATPase. Under these conditions, mitochondrial CoASH/acetyl-CoA, NAD+/ NADH, and ADP/ATP ratios could be varied independently, and the effect on the interconversion of active and inactive pyruvate dehydrogenase could be studied. Decreases in both CoASH/acetyl-CoA and NAD+/NADH ratios were shown to be inhibitory to the steady state activity of pyruvate dehydrogenase, and this effect is described at three different ADP/ATP ratios and different concentrations of added MgCl2. A new steady state level of activity was achieved within 10 min of a change in either CoASH/acetyl-CoA or NAD+/NADH ratio; the rate of inactivation was much higher than the rate of reactivation under these conditions. Effects of CoASH/acetyl-CoA and NAD+/NADH may be additive but are still quantitatively lesser than the changes in activity of pyruvate dehydrogenase induced by changes in ADP/ATP ratio. The variation in activity of pyruvate dehydrogenase with ADP/ATP ratio is described in the absence of changes in the other two ratios, conditions which were not met in earlier studies which employed the oxidation of different substrates to generate changes in all three ratios.  相似文献   

12.
J Turk  G R Marshall 《Biochemistry》1975,14(12):2631-2635
Although optical resolution of alpha-methylphenylalanine (alpha-Me-Phe) has been achieved by the action of carboxypeptidase A on the N-trifluoroacetyl derivative of the amino acid (TFA-alpha-Me-Phe), it is improbable that an alpha-methyl substrate could bind in the same orientation as glycyl-L-tyrosine, due to steric interaction of the alpha-methyl group with an atom in the imidazole ring of zinc ligand His-196. The kinetic parameters for TFA-alpha-Me-Phe and for an ester substrate bearing an alpha-methyl group (beta-hippuryl-alpha-methylphenyllactic acid, HMPL) have been determined and compared to those for the appropriate nonmethylated control substrates. Both TFA-alpha-Me-Phe and HMPL appear to be bound nearly as well as are their respective controls, and HMPL is hydrolyzed nearly as rapidly as its control. TFA-alpha-Me-Phe, however, is hydrolyzed only about one-fiftieth as rapidly as is the nonmethylated substrate. These findings are consistent with the possibilities that: (1) the proposed substrate-induced conformational shift of Tyr-248 is hindered when the methylated substrates are bound; (2) the orientation in which alpha-methyl substrates are bound precludes optimal positioning of Tyr-248 and the scissile bond even after the rotation of Tyr-248 has occurred; (3) amides and esters are bound in different orientations, and in the amide orientation an alpha-methyl group is so directed as to interfere with the approach of Glu-270 to the scissile bond.  相似文献   

13.
14.
Monoclonal antibodies (mAbs) against antipterin immunoglobulin and dihydropteridine reductase (DHPR) and also polyclonal antibodies against human dihydrofolate reductase (DHFR) were obtained. The anti-idiotypic mAbs and anti-DHPR mAbs bind specifically to human DHFR, Escherichia coli DHFR, soybean seedling DHFR, and human DHPR in solid-phase immunoassays. Further, the mAbs bind to the native but not to the denatured forms of DHFRs. The monoclonal antibodies also inhibit the enzymatic activity of human DHFR but not that of human DHPR. Competitive solid-phase immunoassays show stoichiometric inhibition by methotrexate and partial inhibition by NADPH of mAb binding to human DHFR. Cyanogen bromide fragments derived from human DHFR (residues 15-52 and 53-111), containing several active site residues, bind partially to some of the monoclonal antibodies. Accordingly, polyclonal antibodies to peptide 53-111 of human DHFR cross-react to some extent with human DHPR. Data from competitive immunoassays in which the binding of the various mAbs was tested singly and in combination with other mAbs suggest that these antibodies bind to a common region on human DHFR. The results also indicate that the mAbs display some heterogeneity with respect to specific epitopes. These data suggest that despite the absence of significant amino acid sequence homologies among the various DHFRs and DHPR, they have a fundamentally similar topography at the site of binding of the pterin moiety that is recognized by the anti-idiotypic mAbs generated by pterin. In the relatively simple structure of the pterin ring system there are different substituent groups at positions C4 and C6 in methotrexate, 7,8-dihydrofolate, and 7,8-dihydrobiopterin, suggesting that these antibodies are specific for regions on various proteins that interact with the remainder of the pterin moiety. These mAbs and similar mAbs specified by substituent groups on pterin may thus be used as specific probes or inhibitors of various folate-dependent enzymes and transport proteins. They should also provide insights into some of the general features of antibody recognition of protein antigens.  相似文献   

15.
The effect of two reagents on glutamin (asparagin) ase from Pseudomonas aurantiaca-548 has been studied. 2,3-butanedione which modified arginine residues was ineffective for the inactivation of the enzyme. The enzyme was completely inactivated in the presence of N-ethyl-5-phenylisoxazolium-3'-sulfonate (Woodward's reagent K). The effects of pH, reagent concentration, competitive inhibitors and their analogues on the rate or degree of enzyme inactivation were studied. The experimental results suggest that the carboxyl groups localized at the active site of glutamin (asparagin) ase are probably essential for the substrate binding.  相似文献   

16.
17.
D C Phelps  Y Hatefi 《Biochemistry》1985,24(14):3503-3507
Membrane-bound and purified mitochondrial energy-linked nicotinamide nucleotide transhydrogenase (TH) was inhibited by incubation with 5'-[p-(fluorosulfonyl)benzoyl]adenosine (FSBA), which is an analogue of TH substrates and their competitive inhibitors, namely, 5'-, 2'-, or 3'-AMP. NAD(H) and analogues, NADP, 5'-AMP, 5'-ADP, and 2'-AMP/3'-AMP mixed isomers protected TH against inhibition by FSBA, but NADPH accelerated the inhibition rate. In the absence of protective ligands or in the presence of NADP, FSBA appeared to modify the NAD(H) binding site of TH, because, unlike unmodified TH, the enzyme modified by FSBA under these conditions did not bind to an NAD-affinity column (NAD-agarose). However, when the NAD(H) binding site of TH was protected in the presence of 5'-AMP or NAD, then FSBA modification resulted in an inhibited enzyme that did bind to NAD-agarose, suggesting FSBA modification of the NADP(H) binding site or an essential residue outside the active site. [3H]FSBA was covalently bound to TH, and complete inhibition corresponded to the binding of about 0.5 mol of [3H]FSBA/mol of TH. Since purified TH is known to be dimeric in the isolated state, this binding stoichiometry suggests half-of-the-sites reactivity. A similar binding stoichiometry was found earlier for complete inhibition of TH by [14C]DCCD [Phelps, D.C., & Hatefi, Y. (1984) Biochemistry 23, 4475-4480]. The active site directed labeling of TH by radioactive FSBA should allow isolation of appropriate peptides for sequence analysis of the NAD(H) and possibly the NADP(H) binding domains.  相似文献   

18.
The cytochrome P450 catalyzes hydroxylation of many substrates in the presence of O(2) and specific electron transport system. The ternary complex S-Fe(+)O(2) with substrate and O(2) bound to their respective sites on the reduced enzyme is an important intermediate in the formation of the hydroxylating species. Then the active site may be considered as having two sub-sites geared for entirely different types of functionally relevant interactions. The two sites are the substrate binding site, the specific protein residues (Site I), and the L(6) position of the iron (Site II) to which O(2) binds upon reduction. In the ferric enzyme, when substrate binds to Site I, the low spin six-coordinated P450 is converted to the readily reducible high spin five coordinated state. Certain amines and OH compounds, such as products of P450-catalyzed reactions, can bind to Site II resulting in six coordinated inhibited complexes. Then the substrate and product interactions with the two sub-sites can regulate the functional state of the enzyme during catalysis. Product interactions have received very little attention. CYP101 is the only P450 in which X-ray and spectroscopic data on all three structures, the substrate-free, camphor-bound and the 5-exo-OHcamphor-bound are available. The substrate-free CYP101 is low spin and six-coordinated with a water molecule ligated at the L(6) position of the iron. The substrate camphor binds to Site I, and releases the L(6) water despite its inability to bind to this site, indicating that Site I binding can inhibit Site II ligation. The product 5-exo-OHcamphor in addition to binding to Site I, binds to Site II through its -OH group forming Fe-O bond, resulting in the low spin six-coordinated complex. New temperature-jump relaxation kinetic data indicating that Site II ligation inhibits Site I binding are presented. It appears that the Site I and Site II function as interacting sub-sites. The inhibitory allosteric interactions between the two sub-sites are also reflected in the data on binding of the substrate camphor (S) in the presence of the product 5-exo-OH camphor (P) to CYP101 (E). The data are in accordance with the two-site model involving the ternary complex ESP. The affinity of the substrate to the product-bound enzyme as well as the affinity of the product to the substrate-bound enzyme decreased with increase in product concentration, which is consistent with mixed inhibition indicative of inhibitory allosteric interactions between the two sub-sites. Implications of these observations for coupling/uncoupling mechanisms are discussed in the light of the published findings consistent with the two-site behavior of the P450 active site. In addition, kinetic data indicating that the transient high spin intermediate may have to be taken into account for understanding how some P450s have been able to express appreciable hydroxylation activities in the absence of substrate-induced low to high spin transition, observable by the traditional static spectroscopy, are presented.  相似文献   

19.
Mapping the active site of meprin-A with peptide substrates and inhibitors   总被引:2,自引:0,他引:2  
R L Wolz  R B Harris  J S Bond 《Biochemistry》1991,30(34):8488-8493
The extended substrate-binding site of meprin-A, a tetrameric metalloendopeptidase from brush border membranes of mouse kidney proximal tubules, was mapped with a series of peptide substrates. Previous studies led to the development of the chromogenic substrate Phe5(4-nitro)bradykinin for meprin-A. With this substrate, several biologically active peptides were screened as alternate substrate inhibitors, and, of these, bradykinin (RPPGFSPFR) was found to be the best substrate with a single cleavage site (Phe5-Ser6). Three types of bradykinin analogues were used for a systematic investigation of substrate specificity: (1) nonchromogenic bradykinin analogues with substitutions in the P3 to P3' subsites were used as alternative substrate inhibitors of nitrobradykinin hydrolysis, (2) analogues of nitrobradykinin with variations in the P1' position were tested as substrates, and (3) intramolecularly quenched fluorogenic bradykinin analogues with substitutions in the P1 to P3 sites were tested as substrates. A wide variety of substitutions in P1' had little effect on KM (174-339 microM) but markedly affected kcat (51.5 s-1 = A greater than S greater than R greater than F greater than K greater than T greater than E = 0). Substitutions in P1 had a greater effect on KM (366 microM-2.46 mM) and also strongly affected kcat (98.5 s-1 = A greater than F much greater than L greater than E greater than K = 2.4 s-1). The variety of allowed cleavages indicates that meprin-A does not have strict requirements for residues adjacent to the cleavage site. Substitutions farther from the scissle bond also affected binding and hydrolysis, demonstrating that multiple subsite interactions are involved in meprin-A action.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Acyl coenzyme A (acyl-CoA) ligase (acyl-CoA synthetase [ACoAS]) from Pseudomonas putida U was purified to homogeneity (252-fold) after this bacterium was grown in a chemically defined medium containing octanoic acid as the sole carbon source. The enzyme, which has a mass of 67 kDa, showed maximal activity at 40 degrees C in 10 mM K2PO4H-NaPO4H2 buffer (pH 7.0) containing 20% (wt/vol) glycerol. Under these conditions, ACoAS showed hyperbolic behavior against acetate, CoA, and ATP; the Kms calculated for these substrates were 4.0, 0.7, and 5.2 mM, respectively. Acyl-CoA ligase recognizes several aliphatic molecules (acetic, propionic, butyric, valeric, hexanoic, heptanoic, and octanoic acids) as substrates, as well as some aromatic compounds (phenylacetic and phenoxyacetic acids). The broad substrate specificity of ACoAS from P. putida was confirmed by coupling it with acyl-CoA:6-aminopenicillanic acid acyltransferase from Penicillium chrysogenum to study the formation of several penicillins.  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