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 31 毫秒
1.
Fluorescein isothiocyanate (FITC) reactivity with the (Na+ + K+)-ATPase was studied at pH 6.5 and 9.0. Reaction with FITC is nearly complete in 30 min and is irreversible at both pH values. Differential inhibition of enzyme activity is observed at the two pH values as follows: at pH 6.5 the maximal inhibition reached is only 35–45% of the ATPase or p-nitrophenylphosphatase activities, whereas at pH 9.0 ATPase activity can be completely inhibited while maximal phosphatase inhibition is ca. 50%. At all concentrations of FITC tested, more FITC is incorporated into the enzyme at pH 9.0 than at 6.5. At both pH values NaCl increases the inhibition due to FITC while KCl protects against the inhibition. ATP protects the enzyme at both pH values with a K0.5 in the range of 8–20 μm. Enzyme that is partially inactivated at either pH shows no significant change in the K0.5 values for Na+ or K+ or in the Km app for ATP or p-nitrophenylphosphate for the remaining activity. The binding of 48VO4 is not changed by reaction with FITC at either pH, while [3H]ouabain binding is inhibited after reaction at pH 9.0 only in the presence of Mg+2 + Na+ + ATP. [3H]Ouabain binding in the presence of Mg+2 + inorganic phosphate is not inhibited by FITC reaction. Enzyme reacted at both pH values exhibits the expected fluorescein fluorescence (λex = 490, λem = 520) but only with enzyme reacted at pH 9.0 is fluorescence quenching by K+ or reversal by Na+ observed. These results suggest that different classes of amino groups react with FITC at the two pH values tested, and that these groups have distinct roles in the different activities of the enzyme.  相似文献   

2.
The egg jelly-induced acrosome reaction of sea urchin sperm requires the presence of Ca2+ and Na+ in seawater at its normal pH 8. Sperm suspended in seawater at pH 9 undergo the acrosome reaction in the absence of jelly. We have attempted to understand the role of external Na+ in this reaction. Sperm were suspended in Na+-free seawater and the percentage of acrosome reaction and the amount of Ca2+ uptake were determined as a function of external pH. High pH (9.0) in Na+-free medium without jelly triggered a high percentage (above 65%) of sperm acrosome reactions and a two to fourfold increase in Ca2+ uptake. Both the percentage of acrosome reactions and the amount of Ca2+ uptake were similar to those induced by either jelly or pH 9 in Na+-containing seawater. On the other hand, the absence of Na+ in seawater inhibits jelly from inducing Ca2+ uptake and acrosome reactions at pH 8.0 and even at pH 8.5. These results indicate that the Na+ requirement for the acrosome reaction induced by jelly is lost when triggering is by high pH. In contrast, Ca2+ was strictly required since sperm did not react in Ca2+-free seawater at pH 9. We also found that like the jelly-induced acrosome reaction the high-pH-induced acrosome reaction and Ca2+ uptake in complete and Na+-free seawater were inhibited by D600. This finding suggests that the same transport system for Ca2+ uptake associated with the acrosome reaction operates at both triggering conditions, i.e., jelly or pH 9. Although D600 is not now considered a specific blocker, its effect has suggested the involvement of Ca2+ channels in the acrosome reaction. This proposal is supported by our results with nisoldipine, a highly specific inhibitor of calcium channels. The drug inhibited both the sperm acrosome reaction and Ca2+ uptake induced by jelly or pH 9 in complete seawater.  相似文献   

3.
Stemler A 《Plant physiology》1980,65(6):1160-1165
High concentrations of both bicarbonate and formate inhibit photosynthetic O2 evolution at pH 8.0. At this pH, only 2.4% of the total dissolved carbon dioxide exists as CO2. At pH 7.3, where 11% of the total dissolved carbon dioxide exists as CO2, HCO3 no longer inhibits. While formate still inhibits O2 evolution at pH 7.3, its effect can be partially overcome if CO2 is also present. The rate of binding of added 14C-labeled inorganic carbon is nearly 10-fold more rapid when the internal pH of thylakoid membranes is at 6.0 than when it is at 7.8. These observations suggest that CO2, not HCO3, is initially bound to the photosystem II reaction center and that the location of the binding site is on the inside thylakoid surface. However, additional data presented here suggest that, after binding, CO2 is hydrated to HCO3 + H+ in a pH-dependent reaction. Two possible explanations of the “bicarbonate effect” are presented.  相似文献   

4.
4-trans-(N,N-dimethylamino)cinnamaldehyde (DACA) is a chromophoric and fluorogenic substrate of aldehyde dehydrogenase. Fluorescence of DACA is enhanced by binding to aldehyde dehydrogenase in the absence of catalysis both in the presence and absence of the coenzyme analogue 5′AMP. DACA binds to aldehyde dehydrogenase with a dissociation constant of 1–3 μM and stoichiometry of 2 mol mol−1 enzyme. Incorporation of DACA during catalysis was also investigated and found to be 2 mol DACA mol−1 enzyme. Effect of pH on the stoichiometry of DACA incorporation during catalysis has shown that DACA incorporation remained constant at 2 mol DACA mol−1 enzyme, despite a 74-fold velocity enhancement between pH 5.0 and 9.0. Increase of pH increased decomposition of enzyme–acyl intermediate without affecting the rate-limiting step of the reaction. At pH 7.0 the pH stimulated velocity enhancement was 10-fold over that at pH 5.0; further velocity enhancement (11.5-fold that of pH 7.0) was achieved by 150 μM Mg2+ ions. The velocity at pH 7.0 with Mg2+ exceeded that of pH 9.0, and that at maximal pH stimulation at pH 9.5. It was observed that level of intermediate decreased to about 1 mol mol−1 enzyme, indicating that Mg2+ ions increased the rate of decomposition of the enzyme–acyl intermediate and shifted the rate-limiting step of the reaction to another step in the reaction sequence.  相似文献   

5.
1. The NADPH-oxidizing activity of a 100 000 × g particulate fraction of the postnuclear supernatant obtained from guinea-pig phagocytosing polymorphonuclear leucocytes has been assayed by simultaneous determination of oxygen consumption, NADPH oxidation and O?2 generation at pH 5.5 and 7.0 and with 0.15 mM and 1 mM NADPH.2. The measurements of oxygen consumption and NADPH oxidation gave comparable results. The stoichiometry between the oxygen consumed and the NADPH oxidized was 1 : 1.3. A markedly lower enzymatic activity was observed, under all the experimental conditions used, when the O?2 generation assay was employed as compared to the assays of oxygen uptake and NADPH oxidation.4. The explanation of this difference came from the analysis of the effect of superoxide dismutase and of cytochrome c which removes O?2 formed during the oxidation of NADPH.5. Both superoxide dismutase and cytochrome c inhibited the NADPH-oxidizing reaction at pH 5.5. The inhibition was higher with 1 mM NADPH than with 0.15 mM NADPH.6. Both superoxide dismutase and cytochrome c inhibited the NADPH-oxidizing reaction at pH 7.0 with 1 mM NADPH but less than at pH 5.5 with 1 mM NADPH.7. The effect of superoxide dismutase at pH 7.0 with 0.15 mM NADPH was negligible.8. In all instances the inhibitory effect of cytochrome c was greater than that of superoxide dismutase.9. It was concluded that the NADPH-oxidizing reaction studied here is made up of three components: an enzymatic univalent reduction of O2; an enzymatic, apparently non-univalent, O2 reduction and a non-enzymatic chain reaction.10. These three components are variably and independently affected by the experimental conditions used. For example, the chain reaction is freely operative at pH 5.5 with 1 mM NADPH but is almost absent at pH 7.0 with 0.15 mM NADPH, whereas the univalent reduction of O2 is optimal at pH 7.0 with 1 mM NADPH.  相似文献   

6.
An octanol/aqueous two-phase process for the enzymatic production of (R)-phenylacetylcarbinol (PAC) has been investigated further with regard to optimal pH control and replacement of 2.5?M MOPS buffer by a low cost solute. The specific rate of PAC production in the 2.5?M MOPS system controlled at pH?7 was 0.60?mg?U?1?h?1 (reaction completed at 34?h), a 1.6 times improvement over the same 2.5?M MOPS system without pH control (0.39?mg?U?1?h?1 at 49?h). An improved stability of PDC was evident at the end of biotransformation for the pH-controlled system with 84% residual carboligase activity, while 23% of enzyme activity remained in the absence of pH control. Lowering the MOPS concentration to 20?mM resulted in a lower benzaldehyde concentration in the aqueous phase with a major increase in the formation of by-product acetoin and three times decreased PAC production (0.21?mg?U?1?h?1). Biotransformation with 20?mM MOPS and 2.5?M DPG as inexpensive replacement of high MOPS concentrations provided similar aqueous phase benzaldehyde concentrations compared to 2.5?M MOPS and resulted in a comparable PAC concentration (92.1?g?L?1 in the total reaction volume in 47?h) with modest formation of acetoin.  相似文献   

7.
《Free radical research》2013,47(9):1150-1156
Abstract

Oxidation of tyrosine moieties by radicals involved in lipid peroxidation is of current interest; while a rate constant has been reported for reaction of lipid peroxyl radicals with a tyrosine model, little is known about the reaction between tyrosine and alkoxyl radicals (also intermediates in the lipid peroxidation chain reaction). In this study, the reaction between a model alkoxyl radical, the tert-butoxyl radical and tyrosine was followed using steady-state and pulse radiolysis. Acetone, a product of the β-fragmentation of the tert-butoxyl radical, was measured; the yield was reduced by the presence of tyrosine in a concentration- and pH-dependent manner. From these data, a rate constant for the reaction between tert-butoxyl and tyrosine was estimated as 6?±?1 × 107 M?1 s?1 at pH 10. Tyrosine phenoxyl radicals were also monitored directly by kinetic spectrophotometry following generation of tert-butoxyl radicals by pulse radiolysis of solutions containing tyrosine. From the yield of tyrosyl radicals (measured before they decayed) as a function of tyrosine concentration, a rate constant for the reaction between tert-butoxyl and tyrosine was estimated as 7?±?3 × 107 M?1 s?1 at pH 10 (the reaction was not observable at pH 7). We conclude that reaction involves oxidation of tyrosine phenolate rather than undissociated phenol; since the pKa of phenolic hydroxyl dissociation in tyrosine is ~ 10.3, this infers a much lower rate constant, about 3 × 105 M?1 s?1, for the reaction between this alkoxyl radical and tyrosine at pH 7.4.  相似文献   

8.
The effect of an extracellular electron acceptor, ferricyanide, on the light-induced polar leaf pH changes of the submerged angiosperm Elodea canadensis in light and in darkness was determined. The rate of transmembrane ferricyanide reduction was stimulated by increased light intensity and was inhibited by inorganic carbon, indicating that changes in the redox state of the chloroplast were reflected at the plasma membrane. The addition of ferricyanide inhibited the light-induced polar leaf pH reaction. This effect could be balanced by increasing the light intensity. In the dark, the acidification induced by ferricyanide was not influenced by diethylstilbestrol at concentrations that completely inhibited the polar leaf pH changes. This indicates that the ferricyanide-induced H+ extrusion and the H+ transport during the polar reaction were mediated by different mechanisms.  相似文献   

9.
Summary The distribution of Mg+ +-ATPase in osteoclasts along the endosteal surface of the chick tibia was investigated by neutral and alkaline pH cytochemical methods at the electron-microscopic level. Reaction product was observed in mitochondria, cytoplasmic vesicles, and ruffled-border membrane. Levamisole, ouabain, and vanadate did not affect the enzymatic activity. Para-chloromercuribenzoic acid (PCMB) prevented staining of mitochondria, ruffled border, and most cytoplasmic vesicles. Tri-n-butyltin decreased the amount of reaction product in cytoplasmic vesicles and ruffled-border membrane, but did not inhibit reaction product formation within mitochondria. Duramycin, which is a potent inhibitor for proton-pump ATPase, blocked reaction-product formation along the ruffled-border membrane, in mitochondria, and in cytoplasmic vesicles at alkaline pH, but not at neutral pH. It is concluded that the alkaline pH method for Mg+ +-ATPase appears to demonstrate sites of proton-pump ATPase activity.  相似文献   

10.
11.
When D-glucosaminate dehydratase (GADH) was incubated with D-glucosaminate (GlcNA) in veronal buffer (VB; 0.01 M, pH 8.0), GlcNA was converted stoichiometrically to glyceraldehyde, pyruvate, and ammonia (aldolase reaction A). This reaction occurred in addition to the dehydratase reaction (conversion of GlcNA to 2-keto-3-deoxy-o-gluconate and ammonia: α-elimination reaction, B). The ratio of the activities (A:B) was about 1:4. However, in potassium phosphate buffer (KPB; 0.04 M, pH 8.0), the aldolase reaction was inhibited to 3–4% of that in VB, and also inhibited by various derivatives of glycerol, in particular, glycerol-3-phosphate (glycerol-3-P) and glyceraldehyde-3-phosphate (glyceraldehyde-3-P) in VB. The native enzyme was inhibited by incubation with 0.1 M EDTA, and the activity was restored by incubation of the EDTA-treated enzyme with (Mn2+ + pyridoxal 5′-phosphate (PLP)). When the EDTA-treated enzyme was incubated with (Mn2+ + PLP + glycerol-3-P), the activity of reaction B increased to 131% but that of reaction A decreased to 21%. These results suggested that Mn2+, PLP, and the phosphate group of glycerol-3-P are involved in formation of the active enzyme. In the case of the aldolase reaction, Mn2+ ion, which might be essential for the reaction, is chelated by the phosphate group of glycerol-3-P with resultant inhibition of the aldolase reaction.  相似文献   

12.
Glutamate dehydrogenase, GDH (l-glutamate: NAD+ oxidoreductase (deaminating) EC 1.4.1.2) was purified from the plant fraction of lupin nodules and the purity of the preparation established by gel electrophoresis and electrofocusing. The purified enzyme existed as 4 charge isozymes with a MW of 270000. The subunit MW, as determined by dodecyl sulphate electrophoresis, was 45 000. On the basis of the results of the MW determinations a hexameric structure is proposed for lupin-nodule GDH. The pH optima for the enzyme were pH 8.2 for the amination reaction and pH 8.8 for the deamination reaction. GDH from lupin nodules showed a marked preference for NADH over NADPH in the amination reaction and used only NAD+ for the deamination reaction. Pyridoxal-5′-P and EDTA inhibited activity. The enzyme displayed Michaelis-Menten kinetics with respect to all substrates except NAD+. When NAD+ was the varied substrate, there was a deviation from Michaelis-Menten behaviour towards higher activity at high concentrations of NAD+.  相似文献   

13.
Summary Hydroxyl radicals (OH') can be formed in aqueous solution by direct reaction of hydrogen peroxide (H2O2) with ferrous salt (Fenton reaction). OH' damage to deoxyribose, measured as formation of thiobarbituric acid-reactive material, was evaluated at different pHs to study the mechanism of action of classical OH' scavengers. OH' scavenger effect on Fe2+ oxidation was also evaluated in the same experimental conditions. In the absence of OH' scavengers, OH' damage to deoxyribose is higher at acidic compared to neutral and moderately basic pH. At acidic pH deoxiribose is per se able to inhibit Fe2+ oxidation by H202. Most of OH' scavengers tested inhibit deoxyribose damage and Fe2+ oxidation in a similar manner: both inhibitions are most relevant at acidic pH and decrease by increasing the pH. These results are not due to OH' scavenger inhibition of Fenton reaction. The influence of pH on the parameters studied appears to be due to the competition of deoxyribose and OH' scavengers for iron. These results suggest the prominent role of iron binding in the degradation of deoxyribose and in the OH' scavenging ability of different compounds. Results obtained with triethylenetetramine, a iron chelator with a low rate constant with OH', confirm that both deoxyribose and the OH' scavengers interact with iron bringing about a site specific Fenton reaction; that the OH' formed at these sites oxidize these molecules to their radical forms which in turn reduce the Fe3– produced by Fenton reaction. The results presented indicate that most of classical OH' scavengers exert their effect predominantly by preventing the site specific reaction between Fe2+ and H202 on the deoxyribose molecule.  相似文献   

14.
Several factors involved in induction of the acrosomal reaction in sperm of the sea urchin, Arbacia punctulata, have been investigated quantitatively using a simple substrate film technique to monitor extension of the acrosomal process by electron microscopy. Verification of typical acrosomal process formation has been accomplished using thin sections. Sperm were found to undergo the acrosomal reaction in artificial sea water in the absence of egg jelly coat at pH values above 9.6. In the presence of egg jelly a high percentage of sperm react at pH 8.6. At this pH, the fraction of sperm that undergo the acrosomal reaction is directly proportional to the concentration of egg jelly. The Ca2+ ionophore A23187 induces the acrosomal reaction in the absence of egg jelly at pH 8.6. The proportion of sperm that react is dependent on the concentration of ionophore and on the concentration of Ca2+ in the medium. Pretreatment of sperm with low levels of La3+ ion, which is known to be a Ca2+ ion antagonist, results in inhibition of egg jelly induction of the acrosomal reaction. These findings suggest that there are marked similarities between the acrosomal reaction in sea urchin sperm and membrane fusion dependent secretory processes in other cell types.  相似文献   

15.
The reductant of ferricytochrome c2 in Rhodopseudomonas sphaeroides is a component, Z, which has an equilibrium oxidation-reduction reaction involving two electrons and two protons with a midpoint potential of 155 mV at pH 7. Under energy coupled conditions, the reduction of ferricytochrome c2 by ZH2 is obligatorily coupled to an apparently electrogenic reaction which is monitored by a red shift of the endogeneous carotenoids. Both ferricytochrome c2 reduction and the associated carotenoid bandshift are similarly affected by the concentrations of ZH2 and ferricytochrome c2, pH, temperature the inhibitors diphenylamine and antimycin, and the presence of ubiquinone. The second-order rate constant for ferricytochrome c2 reduction at pH 7.0 and at 24°C was 2 · 109 M?1 · s?1, but this varied with pH, being 5.1 · 108 M?1 · s?1 at pH 5.2 and 4.3 · 109 M?1 · s?1 at pH 9.3. At pH 7 the reaction had an activation energy of 10.3 kcal/mol.  相似文献   

16.
The kinetics of O·-2 reaction with semi-oxidized tryptophan radicals in lysozyme, Trp·(Lyz) have been investigated at various pHs and conformational states by pulse radiolysis. The Trp·(Lyz) radicals were formed by Br·-2 oxidation of the 3–4 exposed Trp residues in the protein. At pH lower than 6.2, the apparent bimolecular rate is about 2 × 108M-1s-1; but drops to 8 × 107M-1s-1 or less above pH 6.3 and in CTAC micelles. Similarly, the apparent bimolecular rate constant for the intermolecular Trp·(Lyz) + Trp·(Lyz) recombination reaction is about (4-7 × 106M-1s-1) at/or below pH 6.2 then drops to 1.3-1.6 × 106M-1s-1 at higher pH or in micelles. This behavior suggests important conformational and/or microenvironmental rearrangement with pH, leading to less accessible semioxidized Trp· residues upon Br·-2 reaction. The kinetics of Trp·(Lyz) with ascorbate, a reducing species rather larger than O·-2 have been measured for comparison. The well-established long range intramolecular electron transfer from Tyr residues to Trp radicals-leading to the repair of the semi-oxidized Trp·(Lyz) and formation of the tyrosyl phenoxyl radical is inhibited by the Trp·(Lyz)+O·-2 reaction, as is most of the Trp·(Lyz)+Trp·(Lyz) reaction. However, the kinetic behavior of Trp·(Lyz) suggests that not all oxidized Trp residues are involved in the intermolecular recombination or reaction with O·-2. As the kinetics are found to be quite pH sensitive, this study demonstrates the effect of the protein conformation on O·-2 reactivity. To our knowledge, this is the first report on the kinetics of a protein-O·-2 reaction not involving the detection of change in the redox state of a prosthetic group to probe the reactivity of the superoxide anion.  相似文献   

17.
Flavonoids, plant polyphenols, ubiquitous components of human diet, are excellent antioxidants. Hypochlorous acid (HOCl), produced by activated neutrophils, is highly reactive chlorinating and oxidizing species. It has been reported earlier that flavonoids are chlorinated by HOCl. Here we show that flavonoids from flavonol subclass are also oxidized by HOCl, but only if the latter is in a large molar excess (≥?10). The kinetics of this reaction was studied by stopped-flow spectrophotometry, at different pH. We found that flavonols were oxidized by HOCl with the rate constants of the order of 104–105 M?1 s?1 at pH 7.5. Antioxidant activity of HOCl-modified flavonoids was measured by 2,2′-azinobis-(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) method. Slightly higher antioxidant activity, compared to parent compounds, was observed for flavonols after their reaction with equimolar or moderate excess of HOCl whereas flavonols treated with high molar excess of HOCl exhibited decrease in antioxidant activity. The mechanism of flavonoid reaction with HOCl at physiological pH is proposed, and biological consequences of this reaction are discussed.  相似文献   

18.
During the ascidian sperm reaction the single large cylindrical mitochondrion which lies next to the nucleus in the head swells, becomes spherical, and migrates along the tail to be lost when it reaches the end. This sequence is initiated by eggs, egg water, high pH, low Na+, or the ionophore X537A. Accompanying the sperm reaction induced by low Na+ are H+ efflux and Ca2+ influx in a ratio of near 100:1 as determined by 45Ca2+ and atomic absorption analysis. Simultaneous pH and Ca2+ electrode measurements suggest that the movement of H+ begins 10–13 sec before the movement of Ca2+. Ca2+ uptake can be inhibited by verapamil without affecting H+ efflux or the sperm reaction. Acid release and Ca2+ uptake are proportional to the initial pH of the medium when the reaction is triggered by high pH. Acid release initiated by low Na+ is proportional to Ca2+ concentrations above 2 mM. H+ and Ca2+ movements differ in magnitude, kinetics, and inhibition by verapamil, thus suggesting that H+ is probably not exchanged for Ca2+. Instead we propose that loss of H+ triggers the uptake of Ca2+, which initiates the sperm reaction.  相似文献   

19.
It has been known for some time that pokeweed antiviral protein acts by enzymatically inhibiting protein synthesis on eucaryotic ribosome systems. The site of this action is known to be the ribosome itself. In this paper we show that the pokeweed antiviral protein reaction against ribosomes is a strong function of salt concentrations, where 160 mM K+ and 3 mM Mg2+ retards the reaction, while 20 mM K+ and 2 mM Mg2+ allows maximum reaction rate. It is also shown, however, that an unidentified protein in the postribosomal supernatant solution, together with ATP, allows the ribosome to be attacked even in the presence of high salt. Kinetic analysis of the antiviral protein reaction has been carried out under both sets of conditions, and reveals that the turnover number for the enzyme is about 300–400 mol/mol per min. in each case. The Km for ribosomes is 1 μM in the presence of low salt and 0.2 μM at higher salt in the presence of postribosomal supernatant factors plus ATP. The antiviral protein reaction is also shown to be pH dependent and is controlled by a residue with pKa value of approx. 7.0, apparently a histidine. Stoichiometric reaction of the enzyme with iodoacetamide results in a significant loss of antiribosomal activity.  相似文献   

20.
An inhibitor of phoapholipase A has been isolated from Bothrops neuwiedii venom after gel filtrations through Sephadex G-50 (pH 4.5), Sephadex G-25 (pH 7.6), Sephadex G-15 (pH 4.0), and chromatography on SE-Sephadex C-25 (pH 4.2–4.5). When subjected to paper electrophoresis, the inhibitor migrates as a simple compound with isoelectric point near pH 6.8. Aminoacid composition, sensitivity toward proteases, and the absorption spectrum fit in well with a polypeptide structure lacking tyrosine and tryptophan. In the absence of EDTA, an inactive, anionic derivative appears in inhibitor preparations; the reaction can be reversed by 2-mercaptoethanol. Direct interaction of enzyme and inhibitor is proved by the inhibition of enzyme activity and the chromatography of enzyme-inhibitor mixtures. Titration of inhibitor with venom phospholipases A (isoenzymes P-1 and P-2) yields sigmoid-shaped concentration-inhibition curves, with P-1 far more sensitive than P-2. The enzyme-inhibitor interaction depends on pH since it is tight at pH 4.5 but does not occur at pH 7.5. Presence of thiol groups in inhibitor is consistent with (a) characteristic spectral changes after reaction of inhibitor with PMB 4 and NEM; (b) the inhibitor inhibition by PMB, NEM, iodoacetate, and Hg2+, and (c) the reversal of PMB inhibition with reduced glutathione. Since phospholipase A is insensitive towards Hg2+, addition of Hg2+ to enzyme-inhibitor mixtures (or crude venom samples) causes an apparent enzyme activation (deinhibition). When substrate (egg-yolk lipoprotein) is added to enzyme-inhibitor mixtures, the reaction kinetics show an initial “lag-period” which is proportional to the inhibitor concentration. The “lag-period” does not occur in the absence of inhibitor or in the presence of Hg2+, that inactivates the inhibitor.  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