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 46 毫秒
1.
Todd MJ  Hausinger RP 《Biochemistry》2000,39(18):5389-5396
Klebsiella aerogenes urease uses a dinuclear nickel active site to catalyze the hydrolysis of urea. Here, we describe the steady-state and pre-steady-state kinetics of urease inhibition by fluoride. Urease is slowly inhibited by fluoride in both the presence and absence of substrate. Steady-state rate studies yield parallel double-reciprocal plots; however, we show that fluoride interaction with urease is not compatible with classical uncompetitive inhibition. Rather, we propose that fluoride binds to an enzyme state (E) that is in equilibrium with resting enzyme (E) and produced during catalysis. Fluoride binding rates are directly proportional to inhibitor concentration. Substrate reduces both the rate of fluoride binding to urease and the rate of fluoride dissociation from the complex, consistent with urea binding to E and E.F in addition to E. Fluoride inhibition is pH-dependent due to a protonation event linked to fluoride dissociation. Fluoride binding is pH-independent, suggesting that fluoride anion, not HF, is the actual inhibitor. We assess the kinetic results in terms of the known protein crystal structure and evaluate possible molecular interpretations for the structure of the E state, the site of fluoride binding, and the factors associated with fluoride release. Finally, we note that the apparent uncompetitive inhibition by fluoride as reported for several other metalloenzymes may need to be reinterpreted in terms of fluoride interaction with the corresponding E states.  相似文献   

2.
Tetrachloro-o-benzoquinone (TCoBQ) and tetrachloro-p-benzoquinone (TCpBQ) were studied as inhibitors of jack bean urease in 20 mM phosphate buffer, pH 7.0, 1 mM EDTA, 25 degrees C. The mechanisms of inhibition were evaluated by analysis of the progress curves obtained with two procedures: the reaction initiated by addition of the enzyme and the reaction initiated by addition of the substrate after preincubation of the enzyme with the inhibitor. The obtained results were characteristic of slow-binding inhibition. The effects of different inhibitor concentrations on the initial and steady-state velocities obeyed the relationships of two-step enzyme-inhibitor interaction, qualified as mechanism B. It was found that TCoBQ and TCpBQ are strong urease inhibitors. TCpBQ is more effective than TCoBQ with the overall inhibition constant of K(i)* = 4.5 x 10(-7) mM. The respective inhibition constant of TCoBQ was equal to: K(i)* = 2.4 x 10(-6) mM. The protective experiment proved that the urease active site is involved in the tetrachlorobenzoquinone inhibition process. High effectiveness of thiol protectors against inhibition by TCoBQ and TCpBQ indicates the strategic role of the active site sulfhydryl group in the blocking process. The stability of the complexes: urease-TCoBQ and urease-TCpBQ was tested in two ways: by dilution or addition of dithiothreitol. No recovery of urease activity bound in the urease-inhibitor complexes proves that the complexes are stable and strong.  相似文献   

3.
In this study, we examined the anti-Helicobactor pylori effects of the main protoberberine-type alkaloids in Rhizoma Coptidis. Coptisine exerted varying antibacterial and bactericidal effects against three standard H. pylori strains and eleven clinical isolates, including four drug-resistant strains, with minimum inhibitory concentrations ranging from 25 to 50?μg/mL and minimal bactericidal concentrations ranging from 37.5 to 125?μg/mL. Coptisine’s anti-H. pylori effects derived from specific inhibition of urease in vivo. In vitro, coptisine inactivated urease in a concentration-dependent manner through slow-binding inhibition and involved binding to the urease active site sulfhydryl group. Coptisine inhibition of H. pylori urease (HPU) was mixed type, while inhibition of jack bean urease was non-competitive. Importantly, coptisine also inhibited HPU by binding to its nickel metallocentre. Besides, coptisine interfered with urease maturation by inhibiting activity of prototypical urease accessory protein UreG and formation of UreG dimers and by promoting dissociation of nickel from UreG dimers. These findings demonstrate that coptisine inhibits urease activity by targeting its active site and inhibiting its maturation, thereby effectively inhibiting H. pylori. Coptisine may thus be an effective anti-H. pylori agent.  相似文献   

4.
Klebsiella aerogenes urease was purified 1,070-fold with a 25% yield by a simple procedure involving DEAE-Sepharose, phenyl-Sepharose, Mono Q, and Superose 6 chromatographies. The enzyme preparation was comprised of three polypeptides with estimated Mr = 72,000, 11,000, and 9,000 in a alpha 2 beta 4 gamma 4 quaternary structure. The three components remained associated during native gel electrophoresis, Mono Q chromatography, and Superose 6 chromatography despite the presence of thiols, glycols, detergents, and varied buffer conditions. The apparent compositional complexity of K. aerogenes urease contrasts with the simple well-characterized homohexameric structure for jack bean urease (Dixon, N. E., Hinds, J. A., Fihelly, A. K., Gazzola, C., Winzor, D. J., Blakeley, R. L., and Zerner, B. (1980) Can. J. Biochem. 58, 1323-1334); however, heteromeric subunit compositions were also observed for the enzymes from Proteus mirabilis, Sporosarcina ureae, and Selemonomas ruminantium. K. aerogenes urease exhibited a Km for urea of 2.8 +/- 0.6 mM and a Vmax of 2,800 +/- 200 mumol of urea min-1 mg-1 at 37 degrees C in 25 mM N-2-hydroxyethylpiperazineN'-2-ethanesulfonic acid, 5.0 mM EDTA buffer, pH 7.75. The enzyme activity was stable in 1% sodium dodecyl sulfate, 5% Triton X-100, 1 M KCl, and over a pH range from 5 to 10.5, with maximum activity observed at pH 7.75. Two active site groups were defined by their pKa values of 6.55 and 8.85. The amino acid composition of K. aerogenes urease more closely resembled that for the enzyme from Brevibacter ammoniagenes (Nakano, H., Takenishi, S., and Watanabe, Y. (1984) Agric. Biol. Chem. 48, 1495-1502) than those for plant ureases. Atomic absorption analysis was used to establish the presence of 2.1 +/- 0.3 mol of nickel per mol of 72,000-dalton subunit in K. aerogenes urease.  相似文献   

5.
The improvement in the characterization of slow-binding inhibitors achieved by performing experiments at elevated enzyme concentrations is presented. In particular, the characterization of slow-binding inhibitors conforming to a two-step mode of inhibition with a steady-state dissociation constant that is much lower than the initial dissociation constant with enzyme is discussed. For these systems, inhibition is rapid and low steady-state product concentrations are produced at saturating inhibitor concentrations. By working at elevated enzyme concentrations, improved signal-to-noise ratios are achieved and data may be collected at saturating inhibitor levels. Numerical simulations confirmed that improved parameter estimates are obtained and useful data to discern the mechanism of slow-binding inhibition are produced by working at elevated enzyme concentrations. The saturation kinetics that were unobservable in two previous studies of an enzyme inhibitor system were measured by performing experiments at an elevated enzyme concentration. These results indicate that consideration of the quality of the data acquired using a particular assay is an important factor when selecting the enzyme concentration at which to perform experiments used to characterize the class of enzyme inhibitors examined herein.  相似文献   

6.
Tetrachloro-o-benzoquinone (TCoBQ) and tetrachloro-p-benzoquinone (TCpBQ) were studied as inhibitors of jack bean urease in 20 mM phosphate buffer, pH 7.0, 1 mM EDTA, 25°C. The mechanisms of inhibition were evaluated by analysis of the progress curves obtained with two procedures: the reaction initiated by addition of the enzyme and the reaction initiated by addition of the substrate after preincubation of the enzyme with the inhibitor. The obtained results were characteristic of slow-binding inhibition. The effects of different inhibitor concentrations on the initial and steady-state velocities obeyed the relationships of two-step enzyme-inhibitor interaction, qualified as mechanism B. It was found that TCoBQ and TCpBQ are strong urease inhibitors. TCpBQ is more effective than TCoBQ with the overall inhibition constant of Ki* = 4.5 × 10? 7 mM. The respective inhibition constant of TCoBQ was equal to: Ki* = 2.4 × 10? 6 mM. The protective experiment proved that the urease active site is involved in the tetrachlorobenzoquinone inhibition process. High effectiveness of thiol protectors against inhibition by TCoBQ and TCpBQ indicates the strategic role of the active site sulfhydryl group in the blocking process. The stability of the complexes: urease-TCoBQ and urease-TCpBQ was tested in two ways: by dilution or addition of dithiothreitol. No recovery of urease activity bound in the urease-inhibitor complexes proves that the complexes are stable and strong.  相似文献   

7.
During reaction with [14C]iodoacetamide at pH 6.3, radioactivity was incorporated primarily into a single Klebsiella aerogenes urease peptide concomitant with activity loss. This peptide was protected from modification at pH 6.3 by inclusion of phosphate, a competitive inhibitor of urease, which also protected the enzyme from inactivation. At pH 8.5, several peptides were alkylated; however, modification of one peptide, identical to that modified at pH 6.3, paralleled activity loss. The N-terminal amino acid sequence and composition of the peptide containing the essential thiol was determined. Previous enzyme inactivation studies of K. aerogenes urease could not distinguish whether one or two essential thiols were present per active site (Todd, M. J., and Hausinger, R. P. (1991) J. Biol. Chem. 266, 10260-10267); we conclude that there is a single essential thiol present and identify this residue as Cys319 in the large subunit of the heteropolymeric enzyme.  相似文献   

8.
The urease accessory protein encoded by ureE from Klebsiella aerogenes is proposed to deliver Ni(II) to the urease apoprotein during enzyme activation. Native UreE possesses a histidine-rich region at its carboxyl terminus that binds several equivalents of Ni(2+); however, a truncated form of this protein (H144*UreE) binds only 2 Ni(2+) per dimer and is functionally active (Brayman, T. G., and Hausinger, R. P. (1996) J. Bacteriol. 178, 5410-5416). The urease activation kinetics were studied in vivo by monitoring the development of urease activity upon adding Ni(2+) to spectinomycin-treated Escherichia coli cells that expressed the complete K. aerogenes urease gene cluster with altered forms of ureE. Site-specific alterations of H144*UreE decrease the rate of in vivo urease activation, with the most dramatic changes observed for the H96A, H110A, D111A, and H112A substitutions. Notably, urease activity in cells producing H96A/H144*UreE was lower than cells containing a ureE deletion. Prior studies had shown that H110A and H112A variants each bound a single Ni(2+) per dimer with elevated K(d) values compared with control H144*UreE, whereas the H96A and D111A variants bound 2 Ni(2+) per dimer with unperturbed K(d) values (Colpas, G. J., Brayman, T. G., Ming, L.-J., and Hausinger, R. P. (1999) Biochemistry 38, 4078-4088). To understand why cells containing the latter two proteins showed reduced rates of urease activation, we characterized their metal binding/dissociation kinetics and compared the results to those obtained for H144*UreE. The truncated protein was shown to sequentially bind two Ni(2+) with k(1) approximately 18 and k(2) approximately 100 M(-1) s(-1), and with dissociation rates k(-1) approximately 3 x 10(-3) and k(-2) approximately 10(-4) s(-1). Similar apparent rates of binding and dissociation were noted for the two mutant proteins, suggesting that altered H144*UreE interactions with Ni(2+) do not account for the changes in cellular urease activation. These conclusions are further supported by in vitro experiments demonstrating that addition of H144*UreE to urease apoprotein activation mixtures inhibited the rate and extent of urease formation. Our results highlight the importance of other urease accessory proteins in assisting UreE-dependent urease maturation.  相似文献   

9.
The Klebsiella aerogenes ureE gene product was previously shown to facilitate assembly of the urease metallocenter (Lee, M.H., et al., 1992, J. Bacteriol. 174, 4324-4330). UreE protein has now been purified and characterized. Although it behaves as a soluble protein, UreE is predicted to possess an amphipathic beta-strand and exhibits unusually tight binding to phenyl-Sepharose resin. Immunogold electron microscopic studies confirm that UreE is a cytoplasmic protein. Each dimeric UreE molecule (M(r) = 35,000) binds 6.05 + 0.25 nickel ions (Kd of 9.6 +/- 1.3 microM) with high specificity according to equilibrium dialysis measurements. The nickel site in UreE was probed by X-ray absorption and variable-temperature magnetic circular dichroism spectroscopies. The data are most consistent with the presence of Ni(II) in pseudo-octahedral geometry with 3-5 histidyl imidazole ligands. The remaining ligands are nitrogen or oxygen donors. UreE apoprotein has been crystallized and analyzed by X-ray diffraction methods. Addition of nickel ion to apoprotein crystals leads to the development of fractures, consistent with a conformational change upon binding nickel ion. We hypothesize that UreE binds intracellular nickel ion and functions as a nickel donor during metallocenter assembly into the urease apoprotein.  相似文献   

10.
The nature of the interaction between Escherichia coli cytidine deaminase and the phosphapyrimidine nucleoside 1 has been studied kinetically and spectrophotometrically. Compound 1 was designed as a transition-state analog, and is a potent, slow-binding inhibitor of cytidine deaminase (Ashley, G. W., and Bartlett, P. A. (1982) Biochem. Biophys. Res. Commun. 108, 1467-1474). We present evidence that the binding of 1 is reversible, with no covalent linkage between the enzyme and 1. At pH 6, the rate of recovery of enzyme activity from dissociation of the E X I complex is strongly dependent on the concentration of E X I, indicating that the inhibitor dissociates reversibly. UV difference spectroscopy reveals that the chromophore of 1 is unaltered on binding to the enzyme, thus eliminating the possibility of reversible, covalent modification of the enzyme. For the binding of the active beta-anomers of 1 to cytidine deaminase, the following kinetic parameters were determined at pH 6: kon = 8300 M-1 S-1, koff = 7.8 X 10(-6) S-1, Ki = 0.9 nM. We were also able to observe and characterize time-dependent inhibition of E. coli cytidine deaminase by tetrahydrouridine, 3. This interaction involves involves initial formation of a loose complex (KD = 1.2 microM), followed by isomerization in a slow step to give a more tightly bound complex (Ki = 0.24 microM) with forward and reverse rate constants kf = 3.81 min-1 and kr = 0.95 min-1, respectively.  相似文献   

11.
S-Adenosylmethionine (AdoMet) synthetase catalyzes the only known route of biosynthesis of the primary in vivo alkylating agent. Inhibitors of this enzyme could provide useful modifiers of biological methylation and polyamine biosynthetic processes. The AdoMet synthetase catalyzed reaction converts ATP and L-methionine to AdoMet, PP(i), and P(i), with formation of tripolyphosphate as a tightly bound intermediate. This work describes a nonhydrolyzable analogue of the tripolyphosphate (PPP(i)) reaction intermediate, diimidotriphosphate (O(3)P-NH-PO(2)-NH-PO(3)(5)(-)), as a potent inhibitor. In the presence of AdoMet, PNPNP is a slow-binding inhibitor with an overall inhibition constant (K(i)) of 2 nM and a dissociation rate of 0.6 h(-)(1). In contrast, in the absence of AdoMet PNPNP is a classical competitive inhibitor with a K(i) of 0.5 microM, a slightly higher affinity than PPP(i) itself (K(i) = 3 microM). The imido analogue of the product pyrophosphate, imidodiphosphate (O(3)P-NH-PO(3)(4)(-)) also displays slow onset inhibition only in the presence of AdoMet, with a K(i) of 0.8 microM, compared to K(i) of 250 microM for PP(i). Circular dichroism spectra of the unliganded enzyme and various complexes are indistinguishable indicating that the protein secondary structure is not greatly altered upon complex formation, suggesting local rearrangements at the active site during the slow binding process. A model based on ionization of the bridging -NH- moiety is presented which could account for the potent inhibition by PNP and PNPNP.  相似文献   

12.
C E Nakamura  R H Abeles 《Biochemistry》1985,24(6):1364-1376
The sodium salts of compactin (1) and trans-6-[2-(2,4- dichloro-6-hydroxyphenyl)ethyl]-3,4,5,6-tetrahydro-4-hydroxy-2H-pyran- 2-one (3) are inhibitors of yeast beta-hydroxy-beta-methylglutaryl coenzyme A (HMG-CoA) reductase. The dissociation constants are 0.24 X 10(-9) and 0.28 X 10(-9) M, respectively. Similar values have been reported for HMG-CoA reductase from mammalian sources [Endo, A., Kuroda, M., & Tanzawa, K. (1976) FEBS Lett. 72, 323; Alberts, A. W., et al. (1980) Proc. Natl. Acad. Sci. U.S.A. 77, 3957]. The structures of these compounds marginally resemble that of any substrates of HMG-CoA reductase. We, therefore, investigated the basis for the strong interaction between HMG-CoA reductase and these inhibitors. HMG-CoA and coenzyme A (CoASH), but not reduced nicotinamide adenine dinucleotide phosphate (NADPH), prevent binding of compactin to the enzyme. HMG-CoA, but not CoASH or NADPH, prevents binding of 3 to the enzyme. We also investigated the inhibitory activity of molecules that resemble structural components of compactin. Compactin consists of a moiety resembling 3,5-dihydroxyvaleric acid that is attached to a decalin structure. The sodium salt of DL-3,5-dihydroxyvaleric acid inhibits HMG-CoA reductase competitively with respect to HMG-CoA and noncompetitively with respect to NADPH. The dissociation constant for DL-3,5-dihydroxyvaleric acid, derived from protection against inactivation of enzyme by iodoacetic acid, is (2.1 +/- 0.9) X 10(-2) M. Two decalin derivatives (structurally identical with or closely related to the decalin moiety of compactin) showed no detectable inhibition. If the lack of inhibition is due to their limited solubility, the dissociation constant of these decalin derivatives may be conservatively estimated to be greater than or equal to 0.5 mM. Simultaneous addition of decalin derivatives and DL-3,5-dihydroxyvaleric acid does not lead to enhanced inhibition. The sodium salt of (E)-6-[2-(2-methoxy-1-naphthalenyl)ethenyl]-3,4,5,6- tetrahydro-4-hydroxy-2H-pyran-2-one (6) inhibits HMG-CoA reductase competitively with respect to HMG-CoA and noncompetitively with respect to NADPH. The inhibition constant (vs. HMG-CoA) is 0.8 microM. CoASH does not prevent binding of 6 to enzyme. Compound 6, therefore, behaves analogously to compound 3. We propose that these inhibitors occupy two sites on the enzyme: one site is the hydroxymethylglutaryl binding domain of the enzyme active site and the other site is a hydrophobic pocket located adjacent to the active site.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

13.
Prolyl endopeptidase is a serine proteinase that specifically cleaves peptides on the carboxy side of proline residues. Wilk & Orlowski [(1983) J. Neurochem. 41, 69-75] have shown that benzyloxycarbonyl-prolyl-prolinal (Z-prolyl-prolinal) is a potent inhibitor of prolyl endopeptidase. We show that Z-prolyl-prolinal is a slow-binding inhibitor of mouse brain prolyl endopeptidase with Ki 0.35 +/- 0.05 nM. Kinetic analysis indicates that the mechanism is a simple, but slow, reversible equilibrium between free and bound enzyme (E + I in equilibrium EI) with rate constants for association (kon) and dissociation (koff) of 1.6 X 10(5) M-1.s-1 and approx. 4 X 10(-5) s-1 respectively. Slow-binding inhibition is dependent on the presence of the aldehyde group since the alcohol (Z-prolyl-prolinol) is a rapid and 50,000-fold poorer inhibitor (Ki 19 microM). Prolyl endopeptidase from human brain is also inhibited by Z-prolyl-prolinal with kinetics similar to those of the mouse brain enzyme.  相似文献   

14.
J T Slama  A M Simmons 《Biochemistry》1991,30(9):2527-2534
Two new photoactive analogues of oxidized nicotinamide adenine dinucleotide (NAD+) which are resistant to cleavage by NAD glycohydrolase were synthesized and characterized. The beta-D-ribonucleotide ring of the nicotinamide riboside moiety of NAD+ was replaced with a 2,3-dihydroxycyclopentane ring forming a carbocyclic dinucleotide analogue. Photoreactivity was achieved by the incorporation of an azido group at the 8-position of the adenosyl ring. The previously published synthesis of carbocyclic pyridine dinucleotide analogues [Slama, J. T., & Simmons, A. M. (1988) Biochemistry 27, 183] was modified by resolving the carbocyclic 1-aminoribose analogues and producing optically pure (+)-(1S)- or (-)-(1R)-4 beta-amino-2 alpha,3 alpha-dihydroxy-1 beta-cyclopentanemethanol. Each of these was converted to the corresponding carbocyclic nicotinamide 5'-nucleotide analogue and coupled with 8-azidoadenosine 5'-monophosphate. Two photoactive and isomeric NAD+ analogues were thus prepared. 8-Azidoadenosyl carba-NAD is the analogue in which D-dihydroxycyclopentane is substituted for the D-ribose of the nicotinamide nucleoside moiety. 8-Azido-adenosyl pseudocarba-NAD contains the L-carbocycle in place of the D-ribotide ring. 8-Azidoadenosyl carba-NAD was shown to inhibit the NAD glycohydrolase from Bungarus fasciatus venom competitively with an inhibitor dissociation constant of 187 microM. 8-Azidoadenosyl pseudocarba-NAD was shown to inhibit the same enzyme competitively with a Ki of 73 microM. The superior NADase inhibitor, 8-azidoadenosyl pseudocarba-NAD, was characterized kinetically and shown to fulfill the criteria required of a specific active site directed photoaffinity probe. Irradiation of mixtures of the photoprobe and NAD glycohydrolase with short-wave ultraviolet light resulted in the rapid and irreversible loss of enzyme activity.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Urease activity in microbiologically-induced calcite precipitation.   总被引:1,自引:0,他引:1  
The role of microbial urease in calcite precipitation was studied utilizing a recombinant Escherichia coli HB101 containing a plasmid, pBU11, that encodes Bacillus pasteurii urease. The calcite precipitation by E. coli HB101 (pBU11) was significant although its precipitation level was not as high as that by B. pasteurii. Addition of low concentrations (5-100 microM) of nickel, the cofactor of urease, to the medium further enhanced calcite precipitation by E. coli (pBU11). Calcite precipitation induced by both B. pasteurii and E. coli (pBU11) was inhibited in the presence of a urease inhibitor, acetohydroxamic acid (AHA). These observations on the recombinant urease have confirmed that urease activity is essential for microbiologically-induced calcite precipitation. Partially purified B. pasteurii urease was immobilized in polyurethane (PU) foam to compare the efficacy of calcite precipitation between the free and immobilized enzymes. The immobilized urease showed higher K(m) and lower V(max) values, which were reflected by a slower overall calcite precipitation. However, scanning electron micrographs (SEM) identified that the calcite precipitation occurred throughout the matrices of polyurethane. Furthermore, PU-immobilized urease retained higher enzymatic activities at high temperatures and in the presence of a high concentration of pronase, indicating that immobilization protects the enzyme activity from environmental changes.  相似文献   

16.
M Tagaya  K Yamano  T Fukui 《Biochemistry》1989,28(11):4670-4675
Pyridoxal kinase from pig liver has been purified 10,000-fold to apparent homogeneity. The enzyme is a dimer of subunits of Mr 32,000. The enzyme is strongly inhibited by the product pyridoxal 5'-phosphate. Liver pyridoxamine phosphate oxidase, another enzyme involved in the biosynthesis of pyridoxal 5'-phosphate, is also strongly inhibited by this compound [Wada, H., & Snell, E. E. (1961) J. Biol. Chem. 236, 2089-2095]. Thus, the biosynthesis of pyridoxal 5'-phosphate in the liver might be regulated by the product inhibition of both pyridoxamine phosphate oxidase and pyridoxal kinase. Kinetic studies revealed that the catalytic reaction of liver pyridoxal kinase follows an ordered mechanism in which pyridoxal and ATP bind to the enzyme and ADP and pyridoxal 5'-phosphate are released from the enzyme, in this order. Adenosine tetraphosphopyridoxal was found to be a slow-binding inhibitor of pyridoxal kinase. Pre-steady-state kinetics of the inhibition revealed that the inhibitor and the enzyme form an initial weak complex prior to the formation of a tighter and slowly reversing complex. The overall inhibition constant was 2.4 microM. ATP markedly protects the enzyme against time-dependent inhibition by the inhibitor, whereas another substrate pyridoxal affords no protection. By contrast, adenosine triphosphopyridoxal is not a slow-binding inhibitor of this enzyme.  相似文献   

17.
The effects of tricyclic antidepressants drugs (TCA) amitriptyline, imipramine and nortriptyline, on purified Electrophorus electricus (L.) acetylcholinesterase (AChE; acetylcholine hydrolase, EC 3.1.1.7) were studied using kinetic methods and specific fluorescent probe propidium. The antidepressants inhibited AChE activity by a non-competitive mechanism. Inhibition constants range from 200 to 400 microM. Dimethylated amitriptyline and imipramine were more potent inhibitors than the monomethylated nortriptyline. Fluorescence measurements using bis-quaternary ligand propidium were used to monitor ligand-binding properties of these cationic antidepressants to the AChE peripheral anionic site (PAS). This ligand exhibited an eight-fold fluorescence enhancement upon binding to the peripheral anionic site of AChE from E. electricus (L.) with K(D)=7 x 10(-7)M. It was observed that TCA drugs displaced propidium from the enzyme. On the basis of the displacement experiments antidepressant dissociation constants were determined. Similar values for the inhibition constants suggest that these drugs have similar affinity to the peripheral anionic site. The results also indicate that the catalytic active center of AChE does not participate in the interaction of enzyme with tricyclic antidepressants. These studies suggest that the binding site for tricyclic antidepressants is located at the peripheral anionic site of E. electricus (L.) acetylcholinesterase.  相似文献   

18.
In view of the complexity of the role of the active site flap cysteine in the urease catalysis, in this work we studied how the presence of typical active-site binding inhibitors of urease, phenylphosphorodiamidate (PPD), acetohydroxamic acid (AHA), boric acid and fluoride, affects the reactivity of enzyme thiol groups, the active site flap thiol in particular. For that the inhibitor-urease complexes were prepared with excess inhibitors and had their thiol groups titrated with DTNB. The effects observed were analyzed in terms of the structures of the inhibitor-urease complexes reported in the literature. We found that the effectiveness in preventing the active site cysteine from the modification by disulfides, varied among the inhibitors studied, even though they all bind to the active site. The variations were accounted for by different extents of geometrical distortion in the active site that the inhibitors introduced upon binding, leaving the flap either open in AHA-, boric acid- and fluoride-inhibited urease, like in the native enzyme or closed in PPD-inhibited urease. Among the inhibitors, only PPD was found to be able to thoroughly protect the flap cysteines from the further reaction with disulfides, this apparently resulting from the closed conformation of the flap. Accordingly, in practical terms PPD may be regarded as the most suitable inhibitor for active-site protection experiments in inhibition studies of urease.  相似文献   

19.
1. Anthranilate synthase and phosphoribosyltransferase from Aerobacter aerogenes purify simultaneously and sediment together on sucrose gradients, showing that they occur as an enzyme aggregate. Both activities of the intact aggregate are subject to inhibition by tryptophan. 2. By using appropriate auxotrophic mutants it was shown that an intact active enzyme aggregate is formed when the components come from separate mutant strains. An intact active aggregate can also be formed when one component is from Escherichia coli and the other from A. aerogenes. 3. Phosphoribosyltransferase of A. aerogenes is active when not in an aggregate with anthranilate synthase, but is not subject to tryptophan inhibition, indicating that the inhibitor site is on the anthranilate synthase component. 4. Anthranilate synthase can be active and sensitive to tryptophan inhibition when complexed with an inactive phosphoribosyltransferase. 5. Kinetic studies on the anthranilate synthase activity show that tryptophan is a competitive inhibitor with respect to chorismate and a non-competitive inhibitor with respect to either glutamine or NH(4) (+) ions. This is consistent with a sequential mechanism of the ordered type in which chorismate is the first reactant.  相似文献   

20.
Defining the structural and catalytic properties of the ion transport site(s) of enzyme-phosphorylating ATPases is of key importance in understanding the mechanism of ion transport by these enzymes. In the case of the H+, K(+)-ATPase, SCH 28080 (3-(cyanomethyl)-2-methyl-8-(phenylmethoxy)imidazo[1,2a]-pyridine) has been shown to act as a high affinity, extracytosolic, K(+)-competitive inhibitor of Mg2+, K(+)-ATPase activity (Wallmark, B., Briving, C., Fryklund, J., Munson, K., Jackson, R., Mendlein, J., Rabon, E., and Sachs, G. (1987) J. Biol. Chem. 262, 2077-2084). To define the nature of the SCH 28080-binding site in relation to the catalytic cycle of the enzyme, we have investigated the effects of this potential K+ transport site probe on the steady-state and partial reactions of the H+, K(+)-ATPase. In the absence of K+, SCH 28080 inhibits Mg2(+)-ATPase activity with high affinity (apparent Ki = 30 nM). Inhibition is due to K(+)-like prevention of phosphoenzyme formation. SCH 28080 has no effect on Mg2(+)-catalyzed dephosphorylation. SCH 28080, at concentrations less than 0.5 microM, increases the apparent Km for K+ for Mg2+, K(+)-ATPase activity with little effect on the maximum velocity. At higher concentrations of SCH 28080, reversal of inhibition by higher K+ concentrations is not complete, due to inhibition of ATPase activity by high K+. In contrast, SCH 28080 inhibits K(+)-stimulated dephosphorylation by competitively displacing K+ from phosphoenzyme with an extracytosolic conformation of the monovalent cation site (E2P) at low concentrations of SCH 28080 and K+. At higher concentrations, 10 microM SCH 28080 and 50 mM K+, a slowly dephosphorylating complex with both SCH 28080 and K+ bound to E2P may form which represents a small fraction of the total E2P (15-25%). Preincubation of SCH 28080 with E2P completely blocks K(+)-stimulated dephosphorylation, and K+ is unable to reverse this preincubation effect, indicating that the SCH 28080 dissociation rate is at least as slow as K(+)-independent dephosphorylation of E2P. These findings indicate that SCH 28080 inhibits K(+)-stimulated ATPase activity by competing with K+ for binding to E2P and blocking K(+)-stimulated dephosphorylation. In the absence of K+, SCH 28080 has a higher apparent affinity for E2P, but it permits K(+)-independent dephosphorylation. Since the dissociation rate of SCH 28080 from the enzyme is slow, phosphoenzyme formation is prevented by SCH 28080 remaining bound to the extracytosolic conformation of the monovalent cation site, thereby reducing the steady-state level of phosphoenzyme.  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