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 15 毫秒
1.
J M Storey  K B Storey 《Cryobiology》1982,19(2):185-194
The kinetic properties of cytoplasmic glycerol-3-P dehydrogenase from the third instar larva of the gall fly, Eurosta solidaginis, were studied with emphasis on temperature effects on the enzyme and the regulation of enzyme activity during the synthesis of the cryoprotectant, glycerol. Isoelectrofocusing revealed one major and two minor forms of the enzyme with no alteration in the pI's or relative activities of the forms in larvae acclimated to 24 versus ?30 °C. Kinetic properties of the enzyme were also the same in larvae acclimated to high and low temperatures. Arrhenius plots were linear over a 30 to 0 °C range with an activation energy of 12,630 ± 185 cal/mol and a Q10 of 2.16. The Km for dihydroxyacetone-P was constant, at 50 μM, between 30 and 10 °C but increased by 75% at 0 °C; this increase may be a factor in the cessation of glycerol synthesis which occurs below 5 °C in this species. The Km(NADH), by contrast, was higher (5–6 μM) at 30 °C but decreased (3 μM) at lower temperatures. In the reverse direction, Km's were 340 μM for glycerol-3-P and 12 μM for NAD+. Effects of most inhibitors (of the forward reaction), glycerol-3-P (Ki = 2.4 mM), NAD+ (Ki = 0.2 mM), ATP, Mg·ATP, and Pi, were unaltered by assay temperature but ADP effects were potentiated by low temperature while citrate inhibition was greatest at high temperatures. Glycerol and sorbitol, which accumulate as cryoprotectants in E. solidaginis, had no significant effects on kinetic constants at any temperature but decreased the Vmax activity of the enzyme. Thermal inactivation studies showed an increased thermal stability of the larval enzyme compared to the homologous enzyme from rabbit muscle while added polyols stabilized enzyme activity, decreasing the rate of enzyme inactivation at 50 °C.  相似文献   

2.
Ribulosebisphosphate carboxylase can exist in two forms having different kinetic properties. The fraction of enzyme present in each of the two forms is determined by both the absolute and the relative amounts of substrates and other effector molecules in solution with the enzyme. High CO2 levels induce formation of an active CO2 form of the enzyme while high ribulosebisphosphate (RuBP) levels cause formation of a much less active form. The CO2 form is characterized by a high Km(RuBP), low Km(CO2), and relatively high V. The ribulosebisphosphate form has comparatively lower Km(RuBP) and V and higher Km(CO2). CO2 appears to bind before RuBP in the reaction sequence catalyzed by the CO2 form; this catalytic binding order is apparently reversed in the RuBP form. Steady state rates of enzyme reaction reflect the contributions of both these forms. A brief model, based on the cooperative effects of binding at a small number of catalytic or activator sites in the multimeric enzyme, is presented to account for the changes in enzyme activity with varying substrate and effector molecule concentrations.  相似文献   

3.
ω-Hydroxyfatty acid:NADP oxidoreductase, an enzyme involved in suberin biosynthesis, is induced by wounding potato tubers. Initial velocity and product inhibition studies with the purified enzyme suggested an ordered sequential mechanism, where NADPH is added first, followed by 16-oxohexadecanoate, and NADP is released after 16-hydroxyhexadecanoate. Substrate inhibition by NADPH was observed at concentrations higher than 0.2 mm. The inhibitory NADPH molecule competes with 16-oxohexadecanoate, indicating that it forms a dead-end complex with the E-NADPH form of the enzyme. The kinetics for the NADPH inhibition suggested that n > 1 in the rate equation v = V[NADPH](Km + [NADPH]+ [NADPH]n+1Ki); i.e., more than two NADPH molecules bind to enzyme. The Km for 16-oxohexadecanoate did not change from pH 7.5 to 9.0 but increased about 10-fold from pH 9.0 to 10.0, whereas the Km for NADPH and hexadecanal did not vary significantly in this pH range. Phenylglyoxal inactivated the enzyme; NADPH and AMP (which competes with NADPH; Ki = 1.1 mM) provided protection against such inactivation. Diethylpyrocarbonate also caused inactivation which was reversed by hydroxylamine; NADPH but not AMP protected the enzyme from this inhibition. Pyridoxal-5′-phosphate reversibly inactivated the enzyme and NaBH4 reduction of the pyridoxal phosphate-treated enzyme resulted in irreversible inhibition; a combination of NADPH and ω-oxo C16 acid provided protection against such inactivation. As the chain length of alkanals increased from C3 to C8, the Km for the substrate decreased drastically from 7000 to 90μm and a further increase in chain length from C8 to C20 resulted in only a small decrease in Km. The Km and V for 8-oxooctanoate and 10-oxodecanoate are compared with the values obtained for 16-oxohexadecanoate. Based on these results, it is proposed that arginine acts as the binding site for NADPH, a hydrophobic crevice with lysine at the bottom forms the binding site for 16-oxohexadecanoate and histidine participates in the reaction as the proton donor.  相似文献   

4.
UDP-Galactose 4′-epimerase was purified ca 800-fold through a multi-step procedure which included affinity chromatography using NAD+ -Agarose. Three forms of the enzyme were separated by gel-filtration but only the major form was purified. The pH optimum of the enzyme was 9.5. Exogenous NAD+ was not required for enzymic activity but its removal caused inactivation. The enzyme was unstable below pH 7.0 but stable at pH 8.0 in the presence of glycerol and at ?20° for two months. The equilibrium constant for the enzyme-catalysed reaction was 3.2 ± 0.15. The Km for UDP-galactose and UDP-glucose were 0.12 mM and 0.25 mM, respectively. The inhibition by NADH was competitive, with a Ki of 5 μM. The MW of the enzyme was 78 000; the two minor forms showed the values of 158 000 and 39 000, respectively.  相似文献   

5.
The cyclic nucleotide phosphodiesterases in crude homogenate, soluble material, and particulate preparations of adult Drosophila melanogaster flies, hydrolyze cyclic AMP with nonlinear kinetics. Cyclic GMP is hydrolyzed by the phosphodiesterases in crude homogenate and soluble material with linear kinetics. Physical separation techniques of gel filtration, velocity sedimentation, and ion-exchange chromatography reveal that Drosophila soluble fraction contains two major forms of cyclic nucleotide phosphodiesterase. Form I hydrolyzes both cyclic AMP and cyclic GMP. Inhibition experiments suggest that the hydrolysis of both cyclic nucleotides by Form I occurs at a single active site. The Km's for hydrolysis of both substrates are about 4 μm. This form has a molecular weight of about 168,000 as estimated by gel nitration. Form II cyclic nucleotide phosphodiesterase is specific for cyclic AMP as substrate. Gel filtration indicates that this form has a molecular weight of about 68,000. The Km for cyclic AMP is about 2 μm.  相似文献   

6.
Light inactivation of glucose-6-phosphate dehydrogenase within the pea (Pisum sativum L.) leaf chloroplast has a narrow pH optimum between 7.2 and 7.4 and is NADP-sensitive. The pH optimum for dark activation is slightly lower. Inactivation apparently results in a simple decrease in maximal velocity of the chloroplastic and cytoplasmic forms of the enzyme with no concomitant change in pH optimum or Km (glucose 6-phosphate).  相似文献   

7.
Antibody binding of cartilage-specific proteoglycans   总被引:4,自引:0,他引:4  
The spectroscopically observable acid dissociation constant of aspartate aminotransferase (EC 2.6.1.1) varies to different degrees upon the addition of different monovalent anions. These interactions may be described by the minimal scheme
where XEH and XE represent anion complexes with the acidic (EH) and basic (E) forms of the enzyme, respectively. Both graphical and computer procedures were used to determine the three equilibrium constants which describe such a system. The analysis was based upon the effects of salt concentration (X) upon the apparent pKa of the enzyme determined spectrophotometrically. The affinities for anions of the basic enyzme are less than those of the acidic form of the enzyme so that the apparent pKa rises with anion concentration to a limiting value; pK4 of the enzyme anion complex. That different anion-enzyme complexes have different pK4's reflects the fact that the interaction specificities of the basic and acidic enzymes differ. Cacodylate did not appear either to cause significant effects on the chromophoric pK's or to compete with the binding of halide or carboxylate anions which cause a perturbation of the pKa.  相似文献   

8.
The activity of 3-hydroxy-3-methylglutaryl-coenzyme A reductase (hydroxymethylglutaryl-CoA reductase) was considerably inhibited during incubation with ATP+Mg2+. The inactivated enzyme was reactivated on further incubation with partially purified cytosolic phosphoprotein phosphatase. The inactivation was associated with a decrease in the apparent Km of the reductase for hydroxymethylglutaryl-CoA, and this was reversed on reactivation. The slight increase in activity observed during incubation of microsomal fraction without ATP was not associated with a change in apparent Km and, unlike the effect of the phosphatase, was not inhibited by NaF. Liver microsomal fraction from rats given cholesterol exhibited a low activity of hydroxymethylglutaryl-CoA reductase with a low apparent Km for hydroxymethylglutaryl-CoA. Mícrosomal fraction from rats fed cholestyramine exhibited a high activity with a high Km. To discover whether these changes had resulted from phosphorylation and dephosphorylation of the reductase, microsomal fraction from rats fed the supplemented diets and the standard diet were inactivated with ATP and reactivated with phosphoprotein phosphatase. Inactivation reduced the maximal activity of the reductase in each microsomal preparation and also reduced the apparent Km for hydroxymethylglutaryl-CoA. There was no difference between the preparations in the degree of inactivation produced by ATP. Treatment with phosphatase restored both the maximal activity and the apparent Km of each preparation, but never significantly increased the activity above that observed with untreated microsomal fraction. It is concluded that hydroxymethylglutaryl-CoA reductase in microsomal fraction prepared by standard procedures is almost entirely in the dephosphorylated form, and that the difference in kinetic properties in untreated microsomal fraction from rats fed the three diets cannot be explained by differences in the degree of phosphorylation of the enzyme.  相似文献   

9.
Both uncomplexed subunits of the anthranilate synthetase-phosphoribosyltransferase enzyme complex from Salmonella typhimurium have an absolute requirement for divalent metal ions which can be satisfied by Mg2+, Mn2+, or Co2+. The metal ion kinetics for uncomplexed anthranilate synthetase give biphasic double-reciprocal plots and higher apparent Km values than those for anthranilate synthetase in the enzyme complex. In contrast, the apparent Km values for phosphoribosyltransferase are the same whether the enzyme is uncomplexed or complexed with anthranilate synthetase. This suggests that the metal ion sites on anthranilate synthetase, but not those on phosphoribosyltransferase, are altered upon formation of the enzyme complex. These results and the results of studies reported by others, suggest that complex formation between anthranilate synthetase and phosphoribosyltransferase leads to marked alterations at the active site of the former, but not the latter enzyme. Uncomplexed anthranilate synthetase can be stoichiometrically labeled with Co(III) under conditions which lead to inactivation of 75% of its activity. A comparison of the effects of anthranilate and tryptophan on phosphoribosyltransferase activity in the uncomplexed and complexed forms shows that anthranilate, but not tryptophan, inhibits the uncomplexed enzyme. The complexed phosphoribosyltransferase shows substrate inhibition by anthranilate binding to the phosphoribosyltransferase subunits. In contrast, in a tryptophan-hypersensitive variant complex, anthranilate inhibits phosphoribosyltransferase activity by acting on the anthranilate synthetase subunits. The data are interpreted to mean that there are two classes of binding sites for anthranilate, one on each type of subunit, which may participate in the regulation of anthranilate synthetase and phosphoribosyltransferase under different conditions.  相似文献   

10.
Malate saturation isotherms for the NAD+ malic enzyme determined at widely differing, but saturating, concentrations (8, 80, 160 mm) of magnesium show the same response to malate concentration only when velocity is plotted against the concentration of free malate2?. This identification of the ionized malic acid as the true substrate for this enzyme, together with the observation that the complex of Mg with malate has no influence on the reaction rate even at very high concentrations, indicates that the metal ion activator of the enzyme must also bind in the ionized form. A kinetic analysis shows that, with respect to magnesium and malate, the malic enzyme catalyzes a rapid equilibrium reaction of the intersecting type. Either Mg2+ or malate2? may bind first but the fact that the Km's for both Mg2+ and malate2? are smaller than the respective Ki's suggests that, when either metal ion or malate is present on the enzyme, the other is bound more tightly than when it binds to the free enzyme. This demonstration of the nature of the true substrates for this enzyme has implications for studies of the NAD+ malic enzyme in which conditions influencing the amount of free magnesium and malate, e.g., changes in pH, addition of weak acid effectors etc., are involved.  相似文献   

11.
The data on the pH dependence of the Km for Mg-ATP and the Vm of the ATPase of pig heart mitochondrial F1 indicate the presence of two groups of different pK's which modify the enzyme activity. The first pK at pH 9.6 ± 0.2 may be related to the possible presence of arginine and/or tyrosine residues in the ATPase site; the second pK at pH 7.2 ± 0.2 could be due to the presence of a histidine residue in the ATPase site or to the involvement of amino groups in the ATPase site. The inhibition induced by photooxidation in the presence of Rose Bengal is not pH dependent in the pH range corresponding to the pK of histidine. The inhibition induced by diethylpyrocarbonate cannot be reversed by hydroxylamine and the characteristics of this inhibition rather correspond to the reaction of the inhibitor with amino groups. Pyridoxal phosphate also inhibits the ATPase activity of F1 by reaction with amino groups. The presence of ATP or phosphate partially protects against the inhibition induced by diethylpyrocarbonate or pyridoxal phosphate, which indicates that amino groups may be directly or indirectly involved in the binding of nucleotide and phosphate to F1. Glutaraldehyde also inhibits the enzyme by reacting with amino groups and inducing a crosslinking of the subunits. The disappearance of subunit C is well correlated with the decrease of ATPase activity, indicating that subunit C is essential in the ATPase activity.  相似文献   

12.
A detailed study of the pH dependence of the Michaelis-Menten constants (V and Km) of aryl sulfatase A (EC 3.1.6.1) from rabbit liver indicates that at least two functional groups (pK's ~4.3 and ~7 in the enzyme-substrate complex) participate in the enzymic degradation of substrate. Aryl sulfatase A is inactivated by diethyl pyrocarbonate (ethoxyformic anhydride). The enzyme that has been modified with this reagent can in turn be reactivated by treatment with hydroxylamine. The pH dependence of inactivation reveals a reactive group having a pK of 6.5–7.0. The results indicate that at least one histidine plays an important catalytic role in rabbit liver aryl sulfatase A, consistent with the results of earlier workers who employed diazotized sulfanilic acid. Phosphate ion, a competitive inhibitor, partially protects the enzyme from inactivation by diethyl pyrocarbonate whereas sulfate ion, also a competitive inhibitor, increases the rate of inactivation by diethyl pyrocarbonate. This result is of particular significance in view of the anomalous kinetics of aryl sulfatase A. The kinetic effects of even small amounts of sulfate ion impurities in many commercial sulfate ester substrate preparations is also discussed.  相似文献   

13.
The maximal velocity, V, for isocitrate cleavage by isocitrate lysase from Pseudomonas indigofera was dependent on two dissociable groups (pKa's of 6.9 and 8.6). The pH dependence of the pKi for succinate, a product of isocitrate cleavage, implied that a dissociable group (pKa of 6.0) on the enzyme functions in binding succinate. The pKi's for maleate and itaconate (succinate analogs) were similarly pH dependent. The pKi for oxalate, an analog of glyoxylate which is also a product of isocitrate cleavage, was pH independent. In contrast the pKi's of the four-carbon dicarboxylic acid inhibitors, fumarate and meso-tartrate, both of which affect the glyoxylate site, were dependent on a dissociable group on the enzyme-inhibitor complex. Comparison of the pH dependence of the pKm for isocitrate and the pKi for succinate (and succinate analogs) indicated that the binding of isocitrate was dependent on an acidic dissociable group on the enzyme (pKa of 5.8). The pH dependence of the pKi for homoisocitrate was similar. In addition the Ki for succinate and Km for isocitrate were dependent upon Mg2+ concentration. Inhibition by phosphoenolpyruvate, which binds to the succinate site and may regulate isocitrate lyase from P. indigofera, was twice as pH dependent as that for succinate. Two dissociable groups, one on the enzyme (pKa of 5.8) and one on phosphoenolpyruvate (pKa of 6.35), contributed to the pH dependence observed with phosphoenolpyruvate.  相似文献   

14.
In this paper, cyclic adenosine-3′:5′-monophosphate-dependent protein kinase from yeast-like cells of Mucor rouxii is characterized. A scheme of partial purification is described together with Km for ATP (15 μm), histone (0.2 mg/ml), half-maximal activation constant for cyclic AMP (30 nm), and dissociation constant for the binding of cyclic AMP (40 nm). This enzyme is similar to type II protein kinases in two main aspects: the elution position in DEAE-cellulose chromatography and the readiness of its reassociation. But it has a singular characteristic: it does not dissociate completely with cyclic AMP alone (even at concentrations as high as 0.3 mm) unless histone or NaCl is present. NaCl displays several roles: helps dissociation, prevents inactivation of the catalytic subunit, inhibits enzyme activity, and does not prevent reassociation as occurs with type II protein kinases. Once the holoenzyme is dissociated, cyclic AMP is essential to maintain the enzyme in the dissociated state.  相似文献   

15.
The tumour growth inhibitor L-2-amino-4-methoxy-trans-3-butenoic acid (Ro07-7957) inhibits serine hydroxymethyltransferase in cytosolic extracts of Walker carcinoma non-competitively with respect to L-serine with an apparent inhibition constant similar to the Km-value for L-serine. The kinetics of inactivation suggest that it reacts as an irreversible substrate analogue. Incubation of Walker cells with Ro07-7957 causes an increase in serine hydroxymethyltransferase activity which is most pronounced at concentration ≤LD50. This increase in enzyme activity does not occur in the presence of cycloheximide. These results suggest that inhibition of serine hydroxymethyltransferase in intact cells is accompanied by an increase in enzyme biosynthesis and that the growth inhibitory property of Ro07-7957 does not involve interference with the conversion of serine to glycine.  相似文献   

16.
Inactivation of thrombin (T) by the serpins heparin cofactor II (HCII) and antithrombin (AT) is accelerated by a heparin template between the serpin and thrombin exosite II. Unlike AT, HCII also uses an allosteric interaction of its NH2-terminal segment with exosite I. Sucrose octasulfate (SOS) accelerated thrombin inactivation by HCII but not AT by 2000-fold. SOS bound to two sites on thrombin, with dissociation constants (KD) of 10 ± 4 μm and 400 ± 300 μm that were not kinetically resolvable, as evidenced by single hyperbolic SOS concentration dependences of the inactivation rate (kobs). SOS bound HCII with KD 1.45 ± 0.30 mm, and this binding was tightened in the T·SOS·HCII complex, characterized by Kcomplex of ∼0.20 μm. Inactivation data were incompatible with a model solely depending on HCII·SOS but fit an equilibrium linkage model employing T·SOS binding in the pathway to higher order complex formation. Hirudin-(54–65)(SO3) caused a hyperbolic decrease of the inactivation rates, suggesting partial competitive binding of hirudin-(54–65)(SO3) and HCII to exosite I. Meizothrombin(des-fragment 1), binding SOS with KD = 1600 ± 300 μm, and thrombin were inactivated at comparable rates, and an exosite II aptamer had no effect on the inactivation, suggesting limited exosite II involvement. SOS accelerated inactivation of meizothrombin 1000-fold, reflecting the contribution of direct exosite I interaction with HCII. Thrombin generation in plasma was suppressed by SOS, both in HCII-dependent and -independent processes. The ex vivo HCII-dependent process may utilize the proposed model and suggests a potential for oversulfated disaccharides in controlling HCII-regulated thrombin generation.  相似文献   

17.
The kinetics of tyrosine hydroxylase from the desalted high-speed supernatants of rat striatal homogenates were examined at pH 7.0 using different concentrations of the natural cofactor, (6R)-l-erythrotetrahydrobiopterin, ranging from 4 μM to 1.5 mM. All analyses were performed using two different buffering solutions and their appropriate reducing systems for maintaining cofactor in the reduced state. In the presence of phosphate buffer the results show that tyrosine hydroxylase exists in two kinetically different forms with apparent Km values for the cofactor of 16 μM (low Km) and 2.3 mM (high Km). Similar results were obtained using MOPS buffer. A comparative analysis of the appropriate Vmax values indicates that tyrosine hydroxylase as obtained by our standard preparation procedures is predominately (95%) in the high Km form. When the striatal supernatant was exposed to phosphorylating conditions and subsequently analyzed it appeared that the enzyme now existed totally in the low Km form with very little change in the overall Vmax. A comparison of the results using the two different buffering systems, phosphate and MOPS, revealed that following phosphorylation a large percentage of enzyme was maintained in the phosphorylated state only when using phosphate buffer. In light of the present results, we can for the first time suggest a functional significance not only for the two apparently different kinetic forms of the enzyme but also for a supporting role for phosphorylation. In vivo dopamine synthesis may be accomplished to a significant extent by the phosphorylated form of the enzyme while the non-phosphorylated form may constitute a relatively inactive reservoir which can be recruited for increased dopamine synthesis by phosphorylation.  相似文献   

18.
We have previously shown that orotate phosphoribosyltransferase (OPRT) and orotidine 5′-monophosphate decarboxylase (OMPDC) in human malaria parasite Plasmodium falciparum form an enzyme complex, containing two subunits each of OPRT and OMPDC. To enable further characterization, we expressed and purified P. falciparum OPRT-OMPDC enzyme complex in Escherichia coli. The OPRT and OMPDC activities of the enzyme complex co-eluted in the chromatographic columns used during purification. Kinetic parameters (Km, kcat and kcat/Km) of the enzyme complex were 5- to 125-folds higher compared to the monofunctional enzyme. Interestingly, pyrophosphate was a potent inhibitor to the enzyme complex, but had a slightly inhibitory effect for the monofunctional enzyme. The enzyme complex resisted thermal inactivation at higher temperature than the monofunctional OPRT and OMPDC. The result suggests that the OPRT-OMPDC enzyme complex might have kinetic benefits and thermal stability significantly different from the monofunctional enzyme.  相似文献   

19.
D.K. Srivastava  L.E. Anderson 《BBA》1983,724(3):359-369
Glucose-6-phosphate dehydrogenase (D-glucose-6-phosphate: NADP oxidoreductase, EC 1.1.1.49) has been purified to electrophoretic homogeneity from pea chloroplasts. The enzyme, which has a Stokes radius of 52 Å, is a tetramer made up of four 56000 Da monomers. The pH optimum is around 8.2. The enzyme is absolutely specific for NADP. The apparent Km(NADP) is 2.4 ± 0.1 μM. NADPH inhibition of the enzyme is competitive with respect to NADP (mean Ki, 18 ± 5 μM) and is mixed (Kp >Km, Vmax >Vp) with respect to glucose 6-phosphate (mean crossover point, 0.5 ± 0.1 mM). The apparent Km(glucose 6-phosphate) is 0.37 ± 0.01 mM. The purified enzyme is inactivated in the light in the presence of dilute stroma and washed thylakoids, and by dithiothreitol. Enzyme which has been partially inactivated by treatment with dithiothreitol can be further inactivated in the light in the presence of dilute stroma and washed thylakoids and reactivated in the dark, but only to the extent of the reverse of light inactivation. Dithiothreitol-inactivated enzyme is not reactivated further by addition of crude stroma or oxidized thioredoxin. Dithiothreitol-dependent inactivation of the enzyme follows pseudo-first-order kinetics and shows rate saturation. The enzyme which has been partially inactivated by treatment with dithiothreitol does not differ from the untreated control with respect to thermal and tryptic inactivation. However, enzyme which has been partially light inactivated shows different thermal and tryptic inactivation patterns as compared to the dark control. These observations suggest that the changes in the enzyme brought about by light modulation are not necessarily identical with those brought about by dithiothreitol inactivation.  相似文献   

20.
We compared the inhibitory and catalytic effects of various monoamines on forms A and B of monoamine oxidase (MAO) on mitochondrial preparations from rat brain in mixed substrate experiments. MAO activity was determined by a radioisotopic assay. MAO showed lower Km values for tryptamine and β-phenylethylamine than for tyramine and serotonin. The Km values of the untreated preparation for tyramine, tryptamine, and β-phenylethylamine obtained were the same as those of the form B enzyme and the Km value for serotonin was the same as that of the form A enzyme. Tyramine and tryptamine were competitive inhibitors of serotonin oxidation and β-phenylethylamine did not bind with form A enzyme or inhibit the oxidation of serotonin, while tyramine and tryptamine were competitive inhibitors of β-phenylethylamine oxidation. Although serotonin was not oxidized by form B enzyme, serotonin was a competitive inhibitor of β-phenylethylamine oxidation. It is suggested that rat brain mitochondrial MAO is characterized by two kinds of binding sites.  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