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 836 毫秒
1.
Fluorescence spectrum of camel lens zeta-crystallin, a major protein in the lens of camelids and histicomorph rodents, showed maximum emission at 315 nm. This emission maximum is blue shifted compared to most proteins, including alpha-crystallin, and appeared to be due to tryptophan in highly hydrophobic environment. Interaction of NADPH with zeta-crystallin quenched the protein fluorescence and enhanced the fluorescence of bound NADPH. Analysis of fluorescence quenching suggested high-affinity interaction between NADPH and zeta-crystallin with an apparent Km<0.45 microM. This value is at least an order of magnitude lower than that suggested by activity measurements. Analysis of NADPH fluorescence showed a biphasic curve representing fluorescence of free- and bound-NADPH. The intersection between free- and bound-NADPH closely paralleled the enzyme concentration, suggesting one mole of NADPH was bound per subunit of the enzyme. Phenanthrenequinone (PQ), the substrate of zeta-crystallin, also was able to quench the fluorescence of zeta-crystallin, albeit weaker than NADPH. Quantitative analysis suggested that zeta-crystallin had low affinity for PQ in the absence of NADPH, and PQ binding induced significant conformational changes in zeta-crystallin.  相似文献   

2.
Camel lens zeta-crystallin was reversibly inhibited to various degrees by aspirin (acetyl salicylic acid) and the aspirin-like analgesics: paracetamol (acetaminophen) and ibuprofen (2-(4-isobutyl phenyl)-propionic acid). Among these, aspirin was the most potent inhibitor, causing nearly complete inhibition in a dose-dependent, but time-independent manner. Analysis of inhibition kinetics revealed that aspirin was uncompetitive inhibitor (K(i) 0.64 mM) with respect to NADPH and non-competitive inhibitor (K(i) 1.6 mM) with respect to the substrate, 9,10-phenanthrenequinone (PQ). Multiple-inhibition analysis showed that aspirin and pyridoxal 5' phosphate (PAL-P), a lysine specific reagent, simultaneously bound to a critical lysine residue located towards the NADPH binding region. Consistent with this, NADPH was able to substantially protect zeta-crystallin against aspirin, whereas PQ did not provide any protection. The results suggested that an essential lysine residue was the locus of aspirin binding. The inhibition of zeta-crystallin by aspirin and aspirin-like analgesics was reversible thus eliminating acetylation as a mechanism for inhibition. Reversible binding of aspirin to this lysine may cause steric hindrance resulting in uncompetitive inhibition with respect to NADPH.  相似文献   

3.
Interaction of camel lens zeta-crystallin with aspirin was investigated by activity and fluorescence measurements. Aspirin minimally inhibited the oxidoreductase activity of the enzyme and weakly quenched its fluorescence. However, significant fluorescence quenching of zeta-crystallin coincided with the appearance of a fluorescence signal characteristic of salicylic acid thereby raising the possibility that salicylic acid might have been the moiety responsible for inhibition and fluorescence quenching. Direct fluorescence measurements showed that zeta-crystallin had a much higher affinity for salicylic acid than aspirin (K(i) of about 24 microM for salicylic acid versus 630 microM for aspirin). Salicylic acid was also far more effective in inhibiting zeta-crystallin than aspirin (K(i) values were 23 microM versus 820 microM, respectively). Inhibition kinetics suggested that salicylic acid interacted with zeta-crystallin via a binding site that was distinct from that of NADPH. Salicylic acid also interacted with and quenched the fluorescence of camel lens alpha-crystallin suggesting a general mode of interaction with lens proteins. Within the normal therapeutic concentrations of salicylic acid or aspirin, only crystallin-salicylic acid interactions might be significant. These results showed that camel lens zeta- and alpha-crystallin exhibited remarkable selectivity for salicylic acid over aspirin, and thus, could be considered as salicylate-binding proteins.  相似文献   

4.
Interaction of camel lens zeta-crystallin, an NADPH:quinone oxidoreductase, with several quinone derivatives was examined by fluorescence spectroscopy and activity measurements. Fluorescence of zeta-crystallin was quenched to different levels by the different quinones:juglone (5-OH, 1,4 naphthoquinone), 1,4 naphthoquinone (1,4-NQ), and 1,2 naphthoquinone (1,2-NQ) considerably quenched the fluorescence of zeta-crystallin, where as the commonly used substrate, 9,10-phenanthrenequinone (PQ) did not induce significant quenching. Activity measurements showed only PQ served as a substrate for camel lens zeta-crystallin, while juglone, 1,4-NQ, and 1,2-NQ were inhibitors. Thus quinones that interacted with zeta-crystallin directly inhibited the enzyme, whereas the substrate had very low affinity for the enzyme in the absence of NADPH. Another substrate, dichlorophenol indophenol (DCIP), conformed to the same pattern; DCIP did not quench the fluorescence of the enzyme significantly, but served as a substrate. This pattern is consistent with an ordered mechanism of catalysis with quinone being the second substrate. All three naphthoquinones were uncompetitive inhibitors with respect to NADPH and noncompetitive with respect to PQ. These kinetics are similar to those exhibited by cysteine- and/or lysine-modifying agents. Juglone, 1,4-NQ, and 1,2-NQ interacted with and quenched the fluorescence of camel lens alpha-crystallin, but to lesser extent than that of zeta-crystallin.  相似文献   

5.
Fluorescence spectrum of camel lens ζ-crystallin, a major protein in the lens of camelids and histicomorph rodents, showed maximum emission at 315 nm. This emission maximum is blue shifted compared to most proteins, including α-crystallin, and appeared to be due to tryptophan in highly hydrophobic environment. Interaction of NADPH with ζ-crystallin quenched the protein fluorescence and enhanced the fluorescence of bound NADPH. Analysis of fluorescence quenching suggested high-affinity interaction between NADPH and ζ-crystallin with an apparent Km<0.45 μM. This value is at least an order of magnitude lower than that suggested by activity measurements. Analysis of NADPH fluorescence showed a biphasic curve representing fluorescence of free- and bound-NADPH. The intersection between free- and bound-NADPH closely paralleled the enzyme concentration, suggesting one mole of NADPH was bound per subunit of the enzyme. Phenanthrenequinone (PQ), the substrate of ζ-crystallin, also was able to quench the fluorescence of ζ-crystallin, albeit weaker than NADPH. Quantitative analysis suggested that ζ-crystallin had low affinity for PQ in the absence of NADPH, and PQ binding induced significant conformational changes in ζ-crystallin.  相似文献   

6.
The binding interactions between dimeric glutathione transferase from Schistosoma japonicum (Sj26GST) and bromosulfophthalein (BS) or 8-anilino-1-naphthalene sulfonate (ANS) were characterised by fluorescence spectroscopy and isothermal titration calorimetry (ITC). Both ligands inhibit the enzymatic activity of Sj26GST in a non-competitive form. A stoichiometry of 1 molecule of ligand per mole of dimeric enzyme was obtained for the binding of these ligands. The affinity of BS is higher (K(d)=3.2 microM) than that for ANS (K(d)=195 microM). The thermodynamic parameters obtained by calorimetric titrations are pH-independent in the range of 5.5 to 7.5. The interaction process is enthalpically driven at all the studied temperatures. This enthalpic contribution is larger for the ANS anion than for BS. The strongly favourable enthalpic contribution for the binding of ANS to Sj26GST is compensated by a negative entropy change, due to enthalpy-entropy compensation. DeltaG degrees remains almost invariant over the temperature range studied. The free energy change for the binding of BS to Sj26GST is also favoured by entropic contributions at temperatures below 32 degrees C, thus indicating a strong hydrophobic interaction. Heat capacity change obtained for BS (DeltaC(p) degrees =(-580.3+/-54.2) cal x K(-1) mol(-1)) is twofold larger (in absolute value) than for ANS (DeltaC(p) degrees =(-294.8+/-15.8) cal x K(-1) mol(-1)). Taking together the thermodynamic parameters obtained for these inhibitors, it can be argued that the possible hydrophobic interactions in the binding of these inhibitors to L-site must be accompanied by other interactions whose contribution is enthalpic. Therefore, the non-substrate binding site (designed as ligandin) on Sj26GST may not be fully hydrophobic.  相似文献   

7.
Native beta-lactoglobulin (Blg) binds 1 mole of palmitic acid per mole of protein with a dissociation constant of 0.6 microM for the primary fatty acid binding site. Chemical modification of Cys 121, which lies at the external putative hydrophobic binding site of Blg, does not affect retinol or 4,4'-bis 1-(phenylamino)-8-naphthalenesulfonate (bis-ANS) binding to the protein, indicating that the incorporated appendages do not perturb the internal hydrophobic site within the beta-barrel of Blg (i.e., the retinoid site is unaffected). On the other hand, methylation of Cys 121, reduces the affinity of Blg for palmitic acid by 10-fold as monitored by intrinsic fluorescence. Modification of the Cys 121 with methylmethanethiosulfonate or a thiol-specific spin label appears to either further weaken or totally eliminate fatty acid binding, respectively, due to steric hindrance. Furthermore, this binding pattern has been independently verified using a spin labeled fatty acid analog and monitoring ESR as well as by bis-ANS fluorescence when bound to the protein. These results suggest that fatty acids bind at the "external site" of beta-lactoglobulin, between the sole alpha-helix and the beta-barrel. In addition, structural stability studies of native and chemically modified Blg appear to confirm this observation as well.  相似文献   

8.
The interaction between n-octyl-beta-D-glucopyranoside (octyl glucoside) and bovine liver glutamate dehydrogenase (GDH) was studied using techniques including equilibrium dialysis, UV-spectrophotometry, circular dichroism (CD), fluorescence energy transfer and extrinsic spectrofluorometry in 50 mM sodium phosphate buffer solution (pH 7.6). The equilibrium dialysis experiment showed a higher binding of octyl glucoside to GDH that induces up to 80% enzyme inhibition in 20 mM octyl glucoside solution. The CD study indicated that GDH retains its secondary structure in the presence of octyl glucoside, but loses a degree of its tertiary structure by acquiring a more extended tertiary structure. Measurement of the binding of a hydrophobic fluorescent probe, 1-anilino-naphthalene-8-sulfonate (ANS), to GDH revealed that the binding of ANS to GDH is increased in the presence of octyl glucoside, a finding that may be interpreted in terms of the increment of surface hydrophobic patch(es) of GDH because of its binding to octyl glucoside. Fluorescence energy transfer studies also showed more binding of the reduced coenzyme (NADH) to GDH and the Lineweaver-Burk plots (with respect to NADH) indicate the existence of substrate inhibition in the presence of octyl glucoside. These observations are aimed at explaining the formation of the molten globule-like structure of GDH, which is induced by a non-ionic detergent such as octyl glucoside.  相似文献   

9.
Chloroplast coupling factor 1 (CF1) contains a high-affinity binding site for 8-anilino-1-napthalene sulphonate (ANS,Kd = 5-6 microM). The binding of ANS to the enzyme is associated with a fluorescence enhancement and a blue-shift in the emission spectrum. ANS only slightly inhibits ATP hydrolysis by CF1. Adenine nucleotides and inorganic phosphate induce a fast ANS fluorescence quenching of about 50% which is due to a decrease in the affinity of the enzyme for ANS (Kd increases from 6 microM to 22 microM) and in the fluorescence quantum yield of the bound probe (by 33%) but not in the number of ANS sites (n = 1). Conversely, Mg and Ca ions induce a fluorescence enhancement of bound ANS. Inactivation of the enzyme enhances ANS fluorescence, eliminates the response to adenine nucleotides and inorganic phosphate but increases the response to divalent metals. The affinity of latent CF1 for ADP (Kd = 12 microM) is considerably higher than for ATP (Kd = 95 microM) in buffer containing EDTA. The Kd for inorganic phosphate is 140 microM. Mg increases the apparent affinity for ATP (Kd = 28 microM) but not for ADP or Pi. Binding of ATP to the tight-sites does not inhibit the ADP or Pi-induced fluorescence quenching but decreases the affinity for ADP (Kd = 34 microM) and for inorganic phosphate (Kd = 320 microM). These results suggest that the ADP and phosphate binding sites are different but not independent from the tight sites. Activation of a Mg-specific ATPase in CF1 by octyl glucoside decreases the affinity for ADP and inorganic phosphate by about threefold but increases the affinity for ATP. ATPase activation of CF1 also increases the Ki for ADP inhibition of ATP hydrolysis. ATPase activation also influences the ANS responses to Ca and Mg. Ca-ATPase activation increases the fluorescence enhancement and the apparent affinity for Ca whereas Mg-ATPase activation specifically increases the Mg-induced fluorescence enhancement. The fluorescence of CF1-bound ANS is enhanced by Dio-9 and quenched by phloridzin, quercetin, Nbf-Cl and FITC. Nbf-Cl and FITC completely inhibit the ADP-induced fluorescence quenching whereas Dio-9 inhibits the Mg-induced fluorescence enhancement. ANS does not relieve the quercetin or phloridzin inhibition of ATP hydrolysis indicating that these inhibitors do not compete with ANS for a common binding site. ANS may be used, therefore, as a sensitive probe to detect conformational changes in CF1 in response to activation or inactivation and to binding of substrates and of inhibitors.  相似文献   

10.
The technique of analytical affinity chromatography was extended to characterize binding of ions and hydrophobic probes to proteins. Using the immobilized protein mode of chromatography, alpha-lactalbumin and kappa-casein were covalently attached to 200-nm-pore-diameter controlled-pore glass beads and accommodated for high-performance liquid chromatography. The existence of a high affinity binding site (Kdiss = 0.16 microM) (site I) for calcium ion in alpha-lactalbumin was confirmed by chromatography of [45Ca2+]. In addition, chromatography of the hydrophobic probes, 1-(phenylamino)-8-naphthalene-sulfonate (ANS)2 and 4,4'-bis[1-(phenylamino)-8-naphthalenesulfonate (bis-ANS) indicated that Ca2+ bound to a second site (presumably the zinc site or site II) with weaker affinity. Dissociation constants obtained for apo-alpha-lactalbumin were about 80 microM for ANS and 4.7 microM for bis-ANS in the absence of sodium ion. Addition of Ca2+ initially caused a reduction in surface hydrophobicity (lowered affinity for the probe dyes) followed by an increase at higher Ca2+ concentrations (greater than 0.5 mM), suggesting that occupancy of site II restores an apo-like conformation to the protein. Moreover, the effect of Zn2+ was similar to that observed in the higher Ca2+ concentration range, whereas Na+ apparently bound to site I. A calcium binding site of moderate affinity also exists in kappa-casein (Kdiss = 15.6 microM). A cluster of negative charges, probably including the orthophosphate group, most likely comprise this binding site. By preventing self-association, analytical affinity chromatography permits microscale characterization of ligand equilibria in proteins that are unaffected by protein-protein interactions.  相似文献   

11.
1-anilino-8-naphtalenesulfonate (ANS) is a hydrophobic dipole previously used to demonstrate that the proton for potassium exchange by the gastric HK-ATPase is electroneutral. In this paper, we demonstrate that ANS binds to gastric membranes and probes conformational changes of the HK-ATPase independently of any active H for K exchange. Conformational changes require the presence of potassium-valinomycin and are not triggered by sodium. Potassium effect is enhanced by ATP, in the presence and in the absence of magnesium and, by ADP, in the presence of magnesium. Labeling of the pig HK-ATPase K518 by fluorescein-5-isothiocyanate inhibits the enzyme activity and knocks out the ATP effect on ANS fluorescence. Scherring 28080 and the monoclonal antibody 95-111, two competitive inhibitors of K-activated ATPase dephosphorylation, do not modify K-effect on ANS fluorescence but inhibit ATP effects. This supports that ANS does not probe K-site between the H1–H2 loop. Treatment of gastric membranes with trypsin does not inhibit the ANS response to potassium but does inhibit the response to ATP. This suggests that the ATP site inducing the ANS response is cytoplasmic and the potassium site is intramembranous. Titration reveals that one mole of ANS interacts with one mole of ATPase. We suggest that ANS probes a hydrophobic potassium site of gastric ATPase and that addition of ATP and ADP-Mg embed that site. Received: 16 July 1997/Revised: 10 June 1998  相似文献   

12.
zeta-Crystallin is a major protein in the lens of certain mammals. In guinea pigs it comprises 10% of the total lens protein, and it has been shown that a mutation in the zeta-crystallin gene is associated with autosomal dominant congenital cataract. As with several other lens crystallins of limited phylogenetic distribution, zeta-crystallin has been characterized as an "enzyme/crystallin" based on its ability to reduce catalytically the electron acceptor 2,6-dichlorophenolindophenol. We report here that certain naturally occurring quinones are good substrates for the enzymatic activity of zeta-crystallin. Among the various quinones tested, the orthoquinones 1,2-naphthoquinone and 9,10-phenanthrenequinone were the best substrates whereas menadione, ubiquinone, 9,10-anthraquinone, vitamins K1 and K2 were inactive as substrates. This quinone reductase activity was NADPH specific and exhibited typical Michaelis-Menten kinetics. Activity was sensitive to heat and sulfhydryl reagents but was very stable on freezing. Dicumarol (Ki = 1.3 x 10(-5) M) and nitrofurantoin (Ki = 1.4 x 10(-5) M) inhibited the activity competitively with respect to the electron acceptor, quinone. NADPH protected the enzyme against inactivation caused by heat, N-ethylmaleimide, or H2O2. Electron paramagnetic resonance spectroscopy of the reaction products showed formation of a semiquinone radical. The enzyme activity was associated with O2 consumption, generation of O2- and H2O2, and reduction of ferricytochrome c. These properties indicate that the enzyme acts through a one-electron transfer process. The substrate specificity, reaction characteristics, and physicochemical properties of zeta-crystallin demonstrate that it is an active NADPH:quinone oxidoreductase distinct from quinone reductases described previously.  相似文献   

13.
The mechanism of inhibition of Ca2+-transport activity of rabbit sarcoplasmic reticulum Ca 2+-ATPase (SERCA) by anisodamine (a drug isolated from a medicinal herb Hyoscyamuns niger L) was investigated by using ANS (1-anilino-8-naphthalenesulfonate) fluorescence probe, intrinsic fluorescence quenching and Ca 2+-transport activity assays. The number of ANS binding sites for apo Ca2+-ATPase was determined as 8, using a multiple-identical binding site model. Both anisodamine and Ca2+ at millimolar level enhanced the ANS binding fluorescence intensities. Only anisodamine increased the number of ANS molecules bound by SERCA from 8 to 14. The dissociation constants of ANS to the enzyme without any ligand, with 30 mM anisodamine and with 15 mM Ca 2 were found to be 53.0 microM, 85.0 microM and 50.1 microM, respectively. Both anisodamine and Ca2+ enhanced the ANS binding fluorescenc with apparent dissociation constants of 7.6 mM and 2.3 mM, respectively, at a constant concentration of the enzyme. Binding of anisodamine significantly decreased the binding capacity of Ca2+ with the dissociation constant of 9.5 mM, but binding of Ca2+ had no obvious effect on binding of anisodamine. Intrinsic fluorescence quenching and Ca2+-transport activity assays gave the dissociation constants of anisodamine to SERCA as 9.7 and 5.4 mM, respectively, which were consistent with those obtained from ANS-binding fluorescence changes during titration of SERCA with anisodamine and anisodamine + 15 mM Ca2+, respectively. The results suggest that anisodamine regulates Ca2+-transport activity of the enzyme, by stabilizing the trans-membrane domain in an expanded, inactive conformation, at least at its annular ring region.  相似文献   

14.
Purified recombinant MurA (enolpyruvyl-UDP-GlcNAc synthase) overexpressed in Escherichia coli had significant amounts of UDP-MurNAc (UDP-N-acetylmuramic acid) bound after purification. UDP-MurNAc is the product of MurB, the next enzyme in peptidoglycan biosynthesis. About 25% of MurA was complexed with UDP-MurNAc after five steps during purification that should have removed it. UDP-MurNAc isolated from MurA was identified by mass spectrometry, NMR analysis, and comparison with authentic UDP-MurNAc. Subsequent investigation showed that UDP-MurNAc bound to MurA tightly, with K(d,UDP)(-)(MurNAc) = 0.94 +/- 0.04 microM, as determined by fluorescence titrations using ANS (8-anilino-1-naphthalenesulfonate) as an exogenous fluorophore. UDP-MurNAc binding was competitive with ANS and phosphate, the second product of MurA, and it inhibited MurA. The inhibition patterns were somewhat ambiguous, likely being competitive with the substrate PEP (phosphoenolpyruvate) and either competitive or noncompetitive with respect to the substrate UDP-GlcNAc (UDP-N-acetylglucosamine). These results indicate a possible role for UDP-MurNAc in regulating the biosynthesis of nucleotide precursors of peptidoglycan through feedback inhibition. Previous studies indicated that UDP-MurNAc binding to MurA was not tight enough to be physiologically relevant; however, this was likely an artifact of the assay conditions.  相似文献   

15.
The effects of the quinone analog dibromothymoquinone on electron transfer in isolated mung bean mitochondria are described. Both the main, cyanide-sensitive and the alternate, cyanide-insensitive pathways are inhibited by dibromothymoquinone but in markedly different fashions. Half-maximal inhibition appeared at 40 microM and 20 microM dibromothymoquinone for the cyanide-sensitive and alternate pathways, respectively. With succinate as the electron donor, dibromothymoquinone inhibited the alternate pathway at a single site; showing a mixed, non-competitive type inhibition. On the succinate, cyanide-sensitive pathway dibromothymoquinone showed two sites of inhibition and neither coincides with the site of inhibition associated with the alternate pathway. With malate as the electron donor, two sites of inhibition by dibromothymoquinone were observed regardless of the pathway measured. Dibromothymoquinone also inhibited the rate of valinomycin-induced swelling of isolated mung bean mitochondria. Steady-state kinetics showed the inhibition to be non-competitive with respect to valinomycin. Additionally dibromothymoquinone was observed to increase the fluorescence polarization associated with the hydrophobic probe 1,6-diphenylhexatriene. The results indicated that dibromothymoquinone decreased the fluidity of the inner mitochondrial membrane and suggested that the inhibition of mitochondrial electron transfer by dibromothymoquinone may be associated with this decrease in membrane fluidity. The relationship of the multisite nature of the inhibition of electron transfer by dibromothymoquinone and the possible role of mobile electron carriers such as ubiquinone on the main and alternate respiratory pathways of higher plants is discussed.  相似文献   

16.
Murataliev MB  Feyereisen R 《Biochemistry》2000,39(41):12699-12707
NADP(H) binding is essential for fast electron transfer through the flavoprotein domain of the fusion protein P450BM3. Here we characterize the interaction of NADP(H) with the oxidized and partially reduced enzyme and the effect of this interaction on the redox properties of flavin cofactors and electron transfer. Measurements by three different approaches demonstrated a relatively low affinity of oxidized P450BM3 for NADP(+), with a K(d) of about 10 microM. NADPH binding is also relatively weak (K(d) approximately 10 microM), but the affinity increases manyfold upon hydride ion transfer so that the active 2-electron reduced enzyme binds NADP(+) with a K(d) in the submicromolar range. NADP(H) binding induces conformational changes of the protein as demonstrated by tryptophan fluorescence quenching. Fluorescence quenching indicated preferential binding of NADPH by oxidized P450BM3, while no catalytically competent binding with reduced P450BM3 could be detected. The hydride ion transfer step, as well as the interflavin electron transfer steps, is readily reversible, as demonstrated by a hydride ion exchange (transhydrogenase) reaction between NADPH and NADP(+) or their analogues. Experiments with FMN-free mutants demonstrated that FAD is the only flavin cofactor required for the transhydrogenase activity. The equilibrium constants of each electron transfer step of the flavoprotein domain during catalytic turnover have been calculated. The values obtained differ from those calculated from equilibrium redox potentials by as much as 2 orders of magnitude. The differences result from the enzyme's interaction with NADP(H).  相似文献   

17.
1-Anilinonaphthalene-8-sulfonic acid (ANS) noncompetitively inhibited enzyme activity of glutathione S-transferase P for both glutathione and 1-chloro-2,4-dinitrobenzene (Ki = 30 microM). Dissociation constant for ANS.GST-P complex calculated from the binding study was 15 microM. From the similar values of the inhibition constant and the dissociation constant, it was concluded that specific ANS binding caused the loss of enzyme activity. In the protein structural analysis by circular dichroism, the secondary structures remarkably changed by ANS binding in accordance with the decrease of enzymatic activities. The conformational change of the protein and the decrease in enzymatic activity were reversed by dissociation of ANS. This fact strongly suggested that the enzymatic activity was regulated by a nonsubstrate hydrophobic ligand.  相似文献   

18.
The steady-state kinetics of human erythrocyte glucose-6-phosphate dehydrogenase (D-glucose-6-phosphate: NADP+ 1-oxidoreductase, EC 1.1.1.49) dimers were studied by initial rate measurement. These experiments gave intersecting double-reciprocal plots suggesting a ternary complex mechanism with a Km for NADP and glucose 6-phosphate of 11 microM and 43 microM, respectively. These studies were combined with rate measurements in the presence of one product (NADPH), dead-end inhibitors, as well as alternative substrates. The inhibition by NADPH was found to be competitive with respect to both substrates. Alternate substrates experiments gave linear double-reciprocal plots over a wide range of substrate concentrations. The results suggest that the dimeric enzyme follows either a random or a Theorell-Chance mechanism.  相似文献   

19.
An analysis of the kinetic mechanism of the microsomal NADPH-linked progesterone 5 alpha-reductase obtained from female rat anterior pituitaries was performed. Initial velocity, product inhibition and dead-end inhibition studies indicate that the kinetic mechanism for the progesterone 5 alpha-reductase is equilibrium ordered sequential. Analysis of the initial velocity data resulted in intersecting double reciprocal plots suggesting a sequential mechanism [apparent Km(progesterone) = 88.2 +/- 8.2 nM; apparent Kia(NADPH) = 7.7 +/- 1.1 microM]. Furthermore, the plot of 1/v vs 1/progesterone intersected on the ordinate which is indicative of an equilibrium ordered mechanism. Additional support for ordered substrate binding was provided by the product inhibition studies with NADPH versus NADP and progesterone versus NADP. NADP is a competitive inhibitor versus NADPH (apparent Kis = 7.8 +/- 1.0 microM) and a noncompetitive inhibitor versus progesterone (apparent Kis = 9.85 +/- 2.1 microM and apparent Kii = 63.2 +/- 12.5 microM). These inhibition patterns suggest that NADPH binds prior to progesterone. In sum, these kinetic studies indicate that NADPH binds to the microsomal enzyme in rapid equilibrium and preferentially precedes the binding of progesterone.  相似文献   

20.
S S Wong  P A Frey 《Biochemistry》1978,17(17):3551-3556
Escherichia coli UDP-galactose 4-epimerase in its native form (epimerase.NAD) binds 8-anilino-1-naphthalenesulfonate (ANS) at one tight binding site per dimer with a dissociation constant of 25.9 +/- 2.1 micrometer at pH 8.5 and 27 degrees C. This appears to be the substrate binding site, as indicated by the fact that ANS is a kinetically competitive reversible inhibitor with a Ki of 27.5 micrometer and by the fact that ANS competes with UMP for binding to the enzyme. Upon binding at this site the fluorescence quantum yield of ANS is enhanced 185-fold, and its emission spectrum is blue shifted from a lambdamax of 515 to 470.nm, which suggests that the binding site is shielded from water and probably hydrophobic. Competitive binding experiments with nucleosides and nucleotides indicate that nucleotide binding at this site involves coupled hydrophobic and electrostatic interactions. The reduced form of the enzyme (epimerase.NADH) has no detectable binding affinity for ANS. The marked difference in the affinities of the native and reduced enzymes for ANS is interpreted to be a manifestation of a conformational difference between these enzyme forms.  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