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.
The steady-state kinetics of oxidation of the mitochondrial NADH: ubiquinone oxidoreductase (complex I, EC 1.6.99.3) by artificial electron acceptors--p-quinones and inorganic complexes has been investigated. A limiting stage in the NADH: ferricyanide reductase reaction is a reductive half-reaction. Ferricyanide interacts with negative-charged protein groups taking part in the NADH binding. The rate constants of the quinone reduction by complex I vary from 1.10(6) to 4.10(3) M-1s-1. The NADH, NAD+ and ADP-ribose inhibition data indicate that oxidizers in the rotenono-insensitive reaction interact with the redox centre near the NAD+/NADH binding site, most probably with FMN.  相似文献   

2.
The rotenone-insensitive reduction of quinones and aromatic nitrocompounds by mitochondrial NADH: ubiquinone reductase (complex I, EC 1.6.99.3) has been studied. It was found that these reactions proceed via a mixed one- and two-electron transfer. The logarithms of the bimolecular rate constants of oxidation (TN/Km) are proportional to the one-electron-reduction potentials of oxidizers. The reactivities of nitrocompounds are close to those of quinones. Unlike the reduction of ferricyanide, these reactions are not inhibited by NADH. However, they are inhibited by NAD+ and ADP-ribose, which also act as the mixed-type inhibitors for ferricyanide. TN/Km of quinones and nitrocompounds depend on the NAD+/NADH ratio, but not on NAD+ concentration. They are diminished by the limiting factors of 2.5-3.5 at NAD+/NADH greater than 200. It seems that rotenone-insensitive reduction of quinones and nitrocompounds takes place near the NAD+/NADH and ferricyanide binding site, and the inhibition is caused by induced conformational changes after the binding of NAD+ or ADP-ribose.  相似文献   

3.
《FEBS letters》1986,202(2):327-330
The sodium-transport respiratory chain NADH:quinone reductase of a marine bacterium, Vibrio alginolyti-cus, is composed of three protein subunits, α,β and γ. The β-subunit contains FAD as a prosthetic group and corresponds to NADH dehydrogenase, which catalyses the reduction of ubiquinone to ubisemiquinone. In addition to β, subunits α. and γ are essential for the quinone reductase, which catalyses the reduction of ubiquinone to ubiquinol. The α-subunit contains FMN and the reaction catalysed by subunit α is related to the coupling site of the sodium pump in the quinone reductase.  相似文献   

4.
The NADH:ubiquinone, but not the NADH:ferricyanide, reductase activity of mitochondrial complex I (NADH:ubiquinone oxidoreductase) is inhibited by incubation of the enzyme at pH 6.0 and 0 degree C with ethoxyformic anhydride (EFA), and the inhibition is partially reversed by subsequent incubation of EFA-treated complex I with hydroxylamine. These results and spectral changes of EFA-treated complex I in the u.v. region are consistent with modification of essential histidyl or tyrosyl residues between the primary NADH dehydrogenase and the site of ubiquinone reduction. Treatment of complex I with EFA in the presence of high concentrations of Seconal or Demerol did not protect against EFA inactivation, suggesting that the site of EFA modification may not be the same as the inhibiton sites of Seconal and Demerol. However, the presence of NADH during incubation of complex I with EFA greatly enhanced the inhibition rate, indicating that the reduced conformation of complex I is more susceptible to attack by EFA.  相似文献   

5.
Deamino-NADH/ubiquinone 1 oxidoreductase activity in membrane preparations from Escherichia coli GR19N is 20-50% of NADH/ubiquinone 1 oxidoreductase activity. In comparison, membranes from E. coli IY91, which contain amplified levels of NADH dehydrogenase, exhibit about 100-fold higher NADH/ubiquinone 1 reductase activity but about 20-fold less deamino-NADH/ubiquinone 1 reductase activity. Deamino-NADH/ubiquinone 1 reductase is more sensitive than NADH/ubiquinone 1 reductase activity to inhibition by 3-undecyl-2-hydroxyl-1,4-naphthoquinone, piericidin A, or myxothiazol. Furthermore, GR19N membranes exhibit two apparent Kms for NADH but only one for deamino-NADH. Inside-out membrane vesicles from E. coli GR19N generate a H+ electrochemical gradient (interior positive and acid) during electron transfer from deamino-NADH to ubiquinone 1 that is large and stable relative to that observed with NADH as substrate. Generation of the H+ electrochemical gradient in the presence of deamino-NADH is inhibited by 3-undecyl-2-hydroxy-1,4-naphthoquinone and is not observed in IY91 membrane vesicles or in vesicles from GR19N that are deficient in deamino-NADH/ubiquinone 1 reductase activity. The data provide a strong indication that the E. coli aerobic respiratory chain contains two species of NADH dehydrogenases: (i) an enzyme (NADH dh I) that reacts with deamino-NADH or NADH whose turnover leads to generation of a H+ electrochemical gradient at a site between the primary dehydrogenase and ubiquinone and (ii) an enzyme (NADH dh II) that reacts with NADH exclusively whose turnover does not lead to generation of a H+ electrochemical gradient between the primary dehydrogenase and ubiquinone 1.  相似文献   

6.
Lauryl sulfate inhibits the Deltamu;(H)(+)-dependent reverse electron transfer reactions catalyzed by NADH:ubiquinone oxidoreductase (Complex I) in coupled bovine heart submitochondrial particles and in vesicles derived from Paracoccus denitrificans. The inhibitor affects neither NADH oxidase (coupled or uncoupled) nor NADH:ferricyanide reductase and succinate oxidase activities at the concentrations that selectively prevent the succinate-supported, rotenone-sensitive NAD(+) or ferricyanide reduction. Possible uncoupling effects of the inhibitor are ruled out: in contrast to oligomycin and gramicidin, which increases and decreases the rate of the reverse electron transfer, respectively, in parallel with their coupling and uncoupling effects, lauryl sulfate does not affect the respiratory control ratio. A mechanistic model for the unidirectional effect of lauryl sulfate on the Complex I catalyzed oxidoreduction is proposed.  相似文献   

7.
F. J. Ruzicka  F. L. Crane 《BBA》1971,226(2):221-233
1. Enzymatic reduction of 2,3,5,6-tetramethyl-1,4-benzoquinone (duroquinone) by NADH can be used in an assay procedure for the NADH dehydrogenase. The reduction of this quinone occurs in the region of the electron transport system between the primary dehydrogenase and the cytochrome system as defined by the almost complete loss of reductase activity following piericidin A treatment.

2. Duroquinone reduction can be distinguished from ubiquinone 2 reduction by the marked inhibition of the former following phospholipase C, poly- -lysine, or chloroquine diphosphate treatment. In addition, duroquinone reduction requires the presence of endogenous ubiquinone 10 specifically whereas ubiquinone 2 reduction does not require the presence of endogenous quinone. These observations are consistent with the nonequivalency of the reduction sites of duroquinone and ubiquinone 2.

3. Duroquinol can be utilized as an electron donor for the energy-linked reduction, of NAD+. Duroquinol reduction of NAD+ is dependent upon the presence of ATP, is inhibited by oligomycin, carbonyl cyanide p-trifluoro methoxyphenylhydrazone and piericidin A, and is not inhibited by antimycin A at levels which inhibit electron transport.

4. Duroquinone reduction as well as ubiquinone 2 reduction are inhibited almost completely by phospholipase A, p-chloromercuribenzoate, o-phenanthroline, and Triton X100 treatments.  相似文献   


8.
Maki Hayashi  Tsutomu Unemoto   《BBA》1984,767(3):470-478
The Na+-dependent respiratory chain NADH: quinone oxidoreductase of the marine bacterium, Vibrio alginolyticus, was extracted from membrane by a detergent, Liponox DCH, and was purified by chromatography on QAE-Sephadex and Bio-Gel HTP. The activity of NADH oxidation was separated into two fractions. The one fraction could react with several artificial electron acceptors including Q-1, but could not reduce ubiquinone and menaquinone such as Q-5 and menaquinone-4, which was called NADH dehydrogenase. The other fraction could reduce Q-5 and menaquinone-4 in addition to the NADH dehydrogenase activity, which was called quinone reductase. The purified NADH dehydrogenase consumed NADH in excess of the amount of Q-1 and the reduced Q-1 (quinol) was not produced at all due to an oxidation-reduction cycle of semiquinone radicals. The quinone reductase, however, consumed NADH with the quantitative formation of quinol on account of a dismutation reaction of semiquinone radicals. Identical to the membrane-bound NADH: quinone oxidoreductase, the quinone reductase specifically required Na+ for the activity and was inhibited by 2-heptyl-4-hydroxyquinoline N-oxide. The electron transfer in the quinone reductase was formulated in a form of quinone cycle and the dismutation reaction of semiquinone radicals was assigned to be coupled to the Na+ pump in the respiratory chain of this organism.  相似文献   

9.
The purified respiratory chain NADH dehydrogenase of Escherichia coli oxidizes NADH with either dichlorophenolindophenol (DCIP). ferricyanide, or menadione as electron acceptors, with values for NADH are similar with the three electron acceptors (approximately 50 muM). The purified enzyme contains no flavin and has an absolute requirement for FAD, with Km values around 4 muM. The pH optimum of the enzyme appears to be between 6.5 and 7; the optimum is difficult to establish because of nonenzymatic reduction of DCIP at the lower pH values. Potassium cyanide stimulates the DCIP reductase activity about 2-fold, but has no effect on ferricyanide reductase. The enzyme exhibits hyperbolic kinetics with respect to NADH concentration in both the ferricyanide and DCIP reductase assays, but cooperatively is seen in the menadione reductase reaction. NAD+ is an effective competitive inhibitor of the reaction (Ki congruent to 20 muM); in the presence of NAD+, the NADH saturation curve becomes cooperative, even in the DCIP reductase assay. Many adenine containing nucleotides are competitive inhibitors of the enzyme. The apparent Ki values for these nucleotides as inhibitors of the purified enzyme, the membrane-bound NADH dehydrogenase, and the NADH oxidase are equivalent. An examination of inhibitory effects of a series of adenine nucleotides suggests that the inhibitors act as analogues of NAD+, which is the true physiological inhibitor. The results suggest that the enzyme in situ is always partially inhibited by the levels of NAD- in the E coli cell, and thus behaves in a cooperative fashion to changes in the NAD+/NADH ratio. An antibody has been elicited against the purified NADH dehydrogenase. Immunodiffusion and crossed immunoelectrophoresis show that the antibody is directed principally against the NADH dehydrogenase, with some activity against minor contaminants in the purified preparation. The antibody inhibits NADH dehydrogenase activity 50% at saturating levels. When this antibody preparation is used to examine solubilized membrane preparations, two major immunoprecipitates are found. A parallel inhibition of the membrane-bound NADH dehydrogenase and NADH oxidase activities is seen, supporting the hypothesis that the purified enzyme is indeed a component of the respiratory chain-dependent NADH oxidase pathway.  相似文献   

10.
《BBA》1987,890(1):47-54
The sodium-transport respiratory chain NADH: quinone reductase of a marine bacterium, Vibrio alginolyticus, was purified by high-performance liquid chromatography. The purified quinone reductase, which catalyses the reduction of ubiquinone to ubiquinol, was composed of three subunits, α, β and γ, with apparent molecular weights of 52 000, 46 000 and 32 000, respectively. The subunit β contained one molecule of FAD per molecule and catalysed the reduction of ubiquinone to ubisemiquinone. The subunit α contained FMN as a prosthetic group. The quinone reductase was reconstituted from α and βγ, but not from α and β, and the maximum activity was obtained at the equimolar amounts of FAD(β) and FMN(α). The molecular weight of quinone reductase complex was estimated to be 254 000, which corresponded to a dimer of αβγ complex or α2β2γ2. The subunit γ increased the affinity of β for ubiquinone-1. The reaction catalysed by FMN-containing α-subunit was essential for the generation of membrane potential in proteoliposomes and the coupling site of sodium pump in the quinone reductase was localised to this reaction step.  相似文献   

11.
The reduction of duroquinone (DQ) and 2,3-dimethoxy-5-methyl-6-decyl-1,4-benzoquinone (DB) by NADH and ethanol was investigated in intact yeast mitochondria with good respiratory control ratios. In these mitochondria, exogenous NADH is oxidized by the NADH dehydrogenase localized on the outer surface of the inner membrane, whereas the NADH produced by ethanol oxidation in the mitochondrial matrix is oxidized by the NADH dehydrogenase localized on the inner surface of the inner membrane. The reduction of DQ by ethanol was inhibited 86% by myxothiazol; however, the reduction of DQ by NADH was inhibited 18% by myxothiazol, suggesting that protein-protein interactions between the internal (but not the external) NADH: ubiquinone oxidoreductase and ubiquinol:cytochrome c oxidoreductase (the cytochrome bc1 complex) are involved in the reduction of DQ by NADH. The reduction of DQ and DB by NADH and ethanol was also investigated in mutants of yeast lacking cytochrome b, the iron-sulfur protein, and ubiquinone. The reduction of both quinone analogues by exogenous NADH was reduced to levels that were 10 to 20% of those observed in wild-type mitochondria; however, the rate of their reduction by ethanol in the mutants was equal to or greater than that observed in the wild-type mitochondria. Furthermore, the reduction of DQ in the cytochrome b and iron-sulfur protein lacking mitochondria was myxothiazol sensitive, suggesting that neither of these proteins is an essential binding site for myxothiazol. The mitochondria from the three mutants also contained significant amounts of antimycin- and myxothiazol-insensitive NADH:cytochrome c reductase activity, but had no detectable succinate:cytochrome c reductase activity. These results suggest that the mutants lacking a functional cytochrome bc1 complex have adapted to oxidize NADH.  相似文献   

12.
Barker CD  Reda T  Hirst J 《Biochemistry》2007,46(11):3454-3464
Complex I (NADH:ubiquinone oxidoreductase) from bovine heart mitochondria contains 45 different subunits and nine redox cofactors. NADH is oxidized by a noncovalently bound flavin mononucleotide (FMN), then seven iron-sulfur clusters transfer the two electrons to quinone, and four protons are pumped across the inner mitochondrial membrane. Here, we use protein film voltammetry to investigate the mechanisms of NADH oxidation and NAD+ reduction in the simplest catalytically active subcomplex of complex I, the flavoprotein (Fp) subcomplex. The Fp subcomplex was prepared using chromatography and contained the 51 and 24 kDa subunits, the FMN, one [4Fe-4S] cluster, and one [2Fe-2S] cluster. The reduction potential of the FMN in the enzyme's active site is lower than that of free FMN (thus, the oxidized state of the FMN is most strongly bound) and close to the reduction potential of NAD+. Consequently, the catalytic transformation is reversible. Electrocatalytic NADH oxidation by subcomplex Fp can be explained by a model comprising substrate mass transport, the Michaelis-Menten equation, and interfacial electron transfer kinetics. The difference between the "catalytic" potential and the FMN potential suggests that the flavin is reoxidized before NAD+ is released or that intramolecular electron transfer from the flavin to the [4Fe-4S] cluster influences the catalytic rate. NAD+ reduction displays a marked activity maximum, below which the catalytic rate decreases sharply as the driving force increases. Two possible models reproduce the observed catalytic waveshapes: one describing an effect from reducing the proximal [2Fe-2S] cluster and the other the enhanced catalytic ability of the semiflavin state.  相似文献   

13.
Low concentrations of NAD+ inhibit the NADH: acceptor reductase reactions catalyzed by soluble NADH dehydrogenase from bovine heart mitochondria. The degree of incomplete inhibition of the enzyme depends on the nature and concentration of artificial electron acceptors and is manifested only at low concentrations of the latter. Marked inhibition was demonstrated for the 2.6-dichlorophenolindophenol-, ferricyanide- and O2-reductase reactions, being weakly pronounced during the measurement of the NADH: cytochrome c reductase activity. The inhibition of the above reactions by oxidized NAD+ isn't competitive towards NADH. A kinetic scheme is proposed, which postulates NADH: acceptor reductase reactions occurrence via two mechanisms, namely, a ping-pong mechanism and oxidation of the product-enzyme complex by the acceptor. It was shown that low concentrations of NAD+ also inhibit the NADH oxidase reaction catalyzed by complex I.  相似文献   

14.
(1) The steady-state kinetics of the NADH dehydrogenase activity of Type-II (low molecular weight) NADH dehydrogenase with the acceptors ferricyanide, cytochrome c and 2,6-dichloroindophenol are consistent with the simultaneous operation of an ordered and a ping-pong mechanism. Thus, depending on the acceptor concentration, the reduced enzyme is preferentially oxidized before or after NAD+ disociates from it. (2) The acceptors are able to oxidize the reduced enzyme and its NAD+ complex equally well. In contrast to the kinetics of the Type-I (high molecular weight) enzyme, double substrate inhibition is not found, implying that the site of oxidation of the reduced enzyme by acceptors and the NADH-binding site are remote. (3) With the indophenol, in the concentration range measured, the ordered mechanism is mainly operative. At infinite NADH and acceptor concentrations the rate constant of the reduction of enzyme by bound NADH is measured. (4) With ferricyanide and cytochrome c, in the concentration range measured, erroneous conclusions may be drawn from extrapolations owing to the fact that extrapolated lines in double-reciprocal plots of turnover number against acceptor concentration, at different NADH concentrations, intersect in the third quadrant. A method is described that allows the extrapolation of these data to zero acceptor concentrations. (5) The relation between activity and NADH concentration is sigmoidal (h = 2.0) with ferricyanide or cytochrome c as acceptor, but hyperbolic with 2,6-dichloroindophenol. The latter is also an inhibitor, competitive with respect to NADH. It is concluded that this two-electron acceptor, like ubiquinone, acts as an allosteric effector. (6) Type II is isolated from Type I without gross changes in tertiary structure, as judged by the unaltered rate constants of dissociation of NADH (k-1) and NAD+ (k4) and association of NADH (k1). (7) Type II differs from Type I in two respects, (a) The accessibility of the acceptors is greater by at least two orders of magnitude (k3). (b) The redox potential of the prosthetic group FMN is 120 mV less, as judged by a drop in the value of k2 by four orders of magnitude. It is suggested that one or more of the iron-sulphur proteins present in Type-I but lacking in Type-II dehydrogenase functions as an effector, regulating the redox potential of the FMN.  相似文献   

15.
Scheide D  Huber R  Friedrich T 《FEBS letters》2002,512(1-3):80-84
The proton-pumping NADH:ubiquinone oxidoreductase, also called complex I, is the first energy-transducing complex of many respiratory chains. Homologues of complex I are present in the three domains of life. Here, we report the properties of complex I in membranes of the hyperthermophilic bacterium Aquifex aeolicus. The complex reacted with NADH but not with NADPH and F(420)H(2) as electron donors. Short-chain analogues of ubiquinone like decyl-ubiquinone and ubiquinone-2 were suitable electron acceptors. The affinities towards NADH and ubiquinone-2 were comparable to the ones obtained with the Escherichia coli complex I. The reaction was inhibited by piericidin A at the same concentration as in E. coli. The complex showed an unusual pH optimum at pH 9 and a maximal rate at 80 degrees C. We found no evidence for the presence of an alternative, single subunit NADH dehydrogenase in A. aeolicus membranes. The NADH:ferricyanide reductase activity of detergent extracts of A. aeolicus membranes sedimented as a protein with a molecular mass of approximately 550 kDa. From the data we concluded that A. aeolicus contains a NADH:ubiquinone oxidoreductase resembling complex I of mesophilic bacteria.  相似文献   

16.
The relationship between the NADH:lipoamide reductase and NADH:quinone reductase reactions of pig heart lipoamide dehydrogenase (EC 1.6.4.3) was investigated. At pH 7.0 the catalytic constant of the quinone reductase reaction (kcat.) is 70 s-1 and the rate constant of the active-centre reduction by NADH (kcat./Km) is 9.2 x 10(5) M-1.s-1. These constants are almost an order lower than those for the lipoamide reductase reaction. The maximal quinone reductase activity is observed at pH 6.0-5.5. The use of [4(S)-2H]NADH as substrate decreases kcat./Km for the lipoamide reductase reaction and both kcat. and kcat./Km for the quinone reductase reaction. The kcat./Km values for quinones in this case are decreased 1.85-3.0-fold. NAD+ is a more effective inhibitor in the quinone reductase reaction than in the lipoamide reductase reaction. The pattern of inhibition reflects the shift of the reaction equilibrium. Various forms of the four-electron-reduced enzyme are believed to reduce quinones. Simple and 'hybrid ping-pong' mechanisms of this reaction are discussed. The logarithms of kcat./Km for quinones are hyperbolically dependent on their single-electron reduction potentials (E1(7]. A three-step mechanism for a mixed one-electron and two-electron reduction of quinones by lipoamide dehydrogenase is proposed.  相似文献   

17.
Citrate is fermented by Klebsiella pneumoniae to 2 acetate, 0.5 formate and 1.2 CO2. The formation of less than 1 formate and greater than 1 CO2 per citrate can be accounted for by the oxidation of formate to CO2 in order to provide reducing equivalents for the assimilation of citrate into cell carbon. A membrane-bound electron transport chain is apparently involved in NADH synthesis by these cells. The electrons from formate oxidation to CO2 are used to reduce ubiquinone to ubiquinol by membrane-bound formate dehydrogenase and ubiquinol further delivers its electrons to NAD+, if this endergonic reaction is powered by delta mu Na+. The endogenous NADH level of K. pneumoniae cells thus increased in the presence of formate in response to a delta pNa+ greater than -100 mV. NADH formation was completely abolished in the presence of oxygen or after addition of hydroxyquinoline-N-oxide, a specific inhibitor of the Na(+)-translocating NADH:ubiquinone oxidoreductase. The increase of endogenous NADH was dependent on the delta pNa+ applied to the cells. Inverted membrane vesicles of K. pneumoniae catalysed the reduction of NAD+ to NADH with formate as electron donor after application of delta mu Na+ of about 120 mV consisting of delta pNa+ of 60 mV and delta psi of the same magnitude. Neither the delta pNa+ nor the delta psi of this size alone was sufficient to drive the endergonic reaction. Strictly anaerobic conditions were required for NADH formation and hydroxyquinoline-N-oxide completely inactivated the reaction.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
The short-term incubation of HeLa cells in the presence of diferric transferrin or ferricyanide, which are reduced externally by the transplasma membrane reductase, produces a stoichiometric decrease in NADH and increase in NAD+, which is stimulated by insulin. The NADP/NADPH ratio does not change during 15 min incubation with the oxidants. The total pyridine nucleotide pool of HeLa cells is not affected. Incubation with apotransferrin and ferrocyanide, which cannot act as oxidants for transmembrane electron transport, does not change the pyridine nucleotide concentrations in the cells. Our results show that NADH can act as the internal electron donor for the reduction of external oxidants by the transmembrane reductase. It appears that oxidation of NADH by the transmembrane electron transport using ferricyanide or iron transferrin as external electron acceptors is sufficient to stimulate growth in HeLa cells.  相似文献   

19.
The Na+-translocating NADH:ubiquinone oxidoreductase (Na+-NQR) from Vibrio harveyi was purified and studied by EPR and visible spectroscopy. Two EPR signals in the NADH-reduced enzyme were detected: one, a radical signal, and the other a line around g = 1.94, which is typical for a [2Fe-2S] cluster. An E(m) of -267 mV was found for the Fe-S cluster (n = 1), independent of sodium concentration. The spin concentration of the radical in the enzyme was approximately the same under a variety of redox conditions. The time course of Na+-NQR reduction by NADH indicated the presence of at least two different flavin species. Reduction of the first species (most likely, a FAD near the NADH dehydrogenase site) was very rapid in both the presence and absence of sodium. Reduction of the second flavin species (presumably, covalently bound FMN) was slower and strongly dependent on sodium concentration, with an apparent activation constant for Na+ of approximately 3.4 mM. This is very similar to the Km for Na+ in the steady-state quinone reductase reaction catalyzed by this enzyme. These data led us to conclude that the sodium-dependent step within the Na+-NQR is located between the noncovalently bound FAD and the covalently bound FMN.  相似文献   

20.
Conditions for the reversible dissociation of flavin mononucleotide (FMN) from the membrane-bound mitochondrial NADH:ubiquinone oxidoreductase (complex I) are described. The catalytic activities of the enzyme, i.e. rotenone-insensitive NADH:hexaammineruthenium III reductase and rotenone-sensitive NADH:quinone reductase decline when bovine heart submitochondrial particles are incubated with NADH in the presence of rotenone or cyanide at alkaline pH. FMN protects and fully restores the NADH-induced inactivation whereas riboflavin and flavin adenine dinucleotide do not. The data show that the reduction of complex I significantly weakens the binding of FMN to protein thus resulting in its dissociation when the concentration of holoenzyme is comparable with K(d ( approximately 10(-8)M at pH 10.0).  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