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 750 毫秒
1.
The catalytic properties of the rotenone-sensitive NADH:ubiquinone reductase (Complex I) in bovine heart submitochondrial particles and in inside-out vesicles derived from Paracoccus denitrificans and Rhodobacter capsulatus were compared. The prokaryotic enzymes catalyze the NADH oxidase and NADH:quinone reductase reactions with similar kinetic parameters as those for the mammalian Complex I, except for lower apparent affinities for the substrates--nucleotides. Unidirectional competitive inhibition of NADH oxidation by ADP-ribose, previously discovered for submitochondrial particles, was also evident for tightly coupled P. denitrificans vesicles, thus suggesting that a second, NAD(+)-specific site is present in the simpler prokaryotic enzyme. The inhibitor sensitivity of the forward and reverse electron transfer reactions was compared. In P. denitrificans and Bos taurus vesicles different sensitivities to rotenone and Triton X-100 for the forward and reverse electron transfer reactions were found. In bovine heart preparations, both reactions showed the same sensitivity to piericidin, and the inhibition was titrated as a straight line. In P. denitrificans, the forward and reverse reactions show different sensitivity to piericidin and the titrations of both activities were curvilinear with apparent I(50) (expressed as mole of inhibitor per mole of enzyme) independent of the enzyme concentration. This behavior is explained by a model involving two different sites rapidly interacting with piericidin within the hydrophobic phase.  相似文献   

2.
Fungal respiration: a fusion of standard and alternative components   总被引:22,自引:0,他引:22  
In animals, electron transfer from NADH to molecular oxygen proceeds via large respiratory complexes in a linear respiratory chain. In contrast, most fungi utilise branched respiratory chains. These consist of alternative NADH dehydrogenases, which catalyse rotenone insensitive oxidation of matrix NADH or enable cytoplasmic NADH to be used directly. Many also contain an alternative oxidase that probably accepts electrons directly from ubiquinol. A few fungi lack Complex I. Although the alternative components are non-energy conserving, their organisation within the fungal electron transfer chain ensures that the transfer of electrons from NADH to molecular oxygen is generally coupled to proton translocation through at least one site. The alternative oxidase enables respiration to continue in the presence of inhibitors for ubiquinol:cytochrome c oxidoreductase and cytochrome c oxidase. This may be particularly important for fungal pathogens, since host defence mechanisms often involve nitric oxide, which, whilst being a potent inhibitor of cytochrome c oxidase, has no inhibitory effect on alternative oxidase. Alternative NADH dehydrogenases may avoid the active oxygen production associated with Complex I. The expression and activity regulation of alternative components responds to factors ranging from oxidative stress to the stage of fungal development.  相似文献   

3.
Superoxide production by inside-out coupled bovine heart submitochondrial particles, respiring with succinate or NADH, was measured. The succinate-supported production was inhibited by rotenone and uncouplers, showing that most part of superoxide produced during succinate oxidation is originated from univalent oxygen reduction by Complex I. The rate of the superoxide (O2*-)) production during respiration at a high concentration of NADH (1 mM) was significantly lower than that with succinate. Moreover, the succinate-supported O2*- production was significantly decreased in the presence of 1 mM NADH. The titration curves, i.e., initial rates of superoxide production versus NADH concentration, were bell-shaped with the maximal rate (at 50 microM NADH) approaching that seen with succinate. Both NAD+ and acetyl-NAD+ inhibited the succinate-supported reaction with apparent Ki's close to their Km's in the Complex I-catalyzed succinate-dependent energy-linked NAD+ reduction (reverse electron transfer) and NADH:acetyl-NAD+ transhydrogenase reaction, respectively. We conclude that: (i) under the artificial experimental conditions the major part of superoxide produced by the respiratory chain is formed by some redox component of Complex I (most likely FMN in its reduced or free radical form); (ii) two different binding sites for NADH (F-site) and NAD+ (R-site) in Complex I provide accessibility of the substrates-nucleotides to the enzyme red-ox component(s); F-site operates as an entry for NADH oxidation, whereas R-site operates in the reverse electron transfer and univalent oxygen reduction; (iii) it is unlikely that under the physiological conditions (high concentrations of NADH and NAD+) Complex I is responsible for the mitochondrial superoxide generation. We propose that the specific NAD(P)H:oxygen superoxide (hydrogen peroxide) producing oxidoreductase(s) poised in equilibrium with NAD(P)H/NAD(P)+ couple should exist in the mitochondrial matrix, if mitochondria are, indeed, participate in ROS-controlled processes under physiologically relevant conditions.  相似文献   

4.
The effects of Tinopals (cationic benzoxazoles) AMS-GX and 5BM-GX on NADH-oxidase, NADH:ferricyanide reductase, and NADH APAD+ transhydrogenase reactions and energy-linked NAD+ reduction by succinate, catalyzed by NADH:ubiquinone oxidoreductase (Complex I) in submitochondrial particles (SMP), were investigated. AMS-GX competes with NADH in NADH-oxidase and NADH:ferricyanide reductase reactions (K i = 1 M). 5BM-GX inhibits those reactions with mixed type with respect to NADH (K i = 5 M) mechanism. Neither compound affects reverse electron transfer from succinate to NAD+. The type of the Tinopals' effect on the NADH APAD+ transhydrogenase reaction, occurring with formation of a ternary complex, suggests the ordered binding of nucleotides by the enzyme during the reaction: AMS-GX and 5BM-GX inhibit this reaction uncompetitively just with respect to one of the substrates (APAD+ and NADH, correspondingly). The competition between 5BM-GX and APAD+ confirms that NADH is the first substrate bound by the enzyme. Direct and reverse electron transfer reactions demonstrate different specificity for NADH and NAD+ analogs: the nicotinamide part of the molecule is significant for reduced nucleotide binding. The data confirm the model suggesting that during NADH APAD+ reaction, occurring with ternary complex formation, reduced nucleotide interacts with the center participating in NADH oxidation, whereas oxidized nucleotide reacts with the center binding NAD+ in the reverse electron transfer reaction.  相似文献   

5.
Studies on submitochondrial particles (SMP) preparation showed that in the sourse of the redox reactions at the octane-water interface, catalyzed by SMP enzymes, the charges are transferred from the aqueous to the octane phase. The effects were detected by a shift of the Volta potential, using the vibrating electrode method. In the presence of 2-N-methyl-amino-1,4-naphthoquinone in octane, acting as electron acceptor, the negative charges were transferred from water to octane following the oxidation of NADH, succinate and ascorbate. The charging of the octane phase was sensitive to the inhibitors of the respiratory chain, e. g. rotenone, antimycin and cyanide. In the presence of 2,4-DNP in octane, acting as a proton acceptor, the oxidation of NADH and succinate by ferricyanide, catalyzed by CMP in the presence of antimycin and cyanide correspondingly, was followed by a transfer of positive charges from water to octane. The positive charging of the octane phase, coupled with NADH oxidation, was found insensitive to rotenone, and that coupled with succinate oxidation, was completely inhibited by antimycin. The positive charging of the octane phase was also observed during the reverse transhydrogenase reaction, catalyzed by SMP at the division of the phases. The effect was inhibited by palmitoyl-CoA.  相似文献   

6.
Study of the uncoupling effect of various saturated fatty acids (from caprylic to palmitic) revealed that the glutamate recoupling effect was more pronounced in the case of short chain fatty acids, whereas recoupling of mitochondria by carboxyatractylate was more effective in the case of long chain fatty acids. The overall recoupling effect, however, did not depend on the fatty acid chain length. Besides carboxyatractylate, glutamate and aspartate also exhibited a recoupling effect under uncoupling by lauryl sulfate. The uncoupling effect of lauryl sulfate was markedly weaker in the presence of DNP or laurate (but not FCCP) which were added in concentrations causing twofold increase in mitochondrial respiration. In the presence of lauryl sulfate the uncoupling action of laurate and DNP was insensitive to carboxyatractylate and glutamate. With laurate and DNP as uncouplers increasing the pH from 7.0 to 7.8 potentiated the recoupling effect of carboxyatractylate and attenuated the recoupling effect of glutamate. In the case of uncoupling by lauryl sulfate similar changes in the recoupling effect of carboxyatractylate and glutamate were observed only in the presence of 10 microM tetraphenylphosphonium. Thus, when uncoupling is induced by fatty acids, DNP, and lauryl sulfate, the ADP/ATP and aspartate/glutamate antiporters function as two parallel and independent pathways for mitochondrial membrane potential dissipation. We suggest that the role of the ADP/ATP antiporter in uncoupling includes proton capture from the intermembrane space with subsequent protonation of uncoupler anions, their transport as neutral molecules on the internal side, and deprotonation followed by proton release into the matrix and transfer of the uncoupler anion in the reverse direction. During uncoupling the aspartate/glutamate antiporter cyclically carries the uncoupler anion with simultaneous proton transfer from the intermembrane space into the matrix.  相似文献   

7.
Electron donors such as thiosulfate, sulfite, and ascorbate have been shown to enter the respiratory chain ofT. neapolitanus at the level of cytochromec. The enzymatic oxidation of these substrates catalyzed by the cytochrome oxidase (E. C. 1.9.3.1.) ofT. neapolitanus cell-free extracts was coupled to the generation of energy which could be utilized to drive the reverse electron flow from cytochromec to pyridine nucleotides.The reduction of endogenous or added flavin by thiosulfate or ascorbate has been shown to be ATP-dependent; likewise the reduction of cytochromeb by these electron donors also required energy. The rate of ATP-driven reversal of electron transfer from cytochromec to the pyridine nucleotides was much faster compared with the rate of electron reversal catalyzed by the substrate-linked generated energy. The pathway of energy-linked reversal of electron transfer from cytochrome c to pyridine nucleotides involved cytochromeb and flavoproteins.NADH oxidation byT. neapolitanus cell-free extracts is mediated by the flavoprotein and cytochrome systems and this process also appears to be coupled with energy generation. The NADH oxidase (NADH2: cytochromec oxidoreductase) was partially inhibited by amytal or rotenone, antimycin A or HOQNO, and was relatively insensitive to cyanide or azide.This investigation was supported in part by a National Science Foundation Grant No. GB 6649 and in part by the Department of Interior, Office of Water Resources Research No. A-016-KY.  相似文献   

8.
NADH:ubiquinone oxidoreductase (complex I) was purified from bovine heart mitochondria by solubilization with n-dodecyl beta-D-maltoside (lauryl maltoside), ammonium sulfate fractionation, and chromatography on Mono Q in the presence of the detergent. Its subunit composition was very similar to complex I purified by conventional means. Complex I was dissociated in the presence of N,N-dimethyldodecylamine N-oxide and beta-mercaptoethanol, and two subcomplexes, I alpha and I beta, were isolated by chromatography. Subcomplex I alpha catalyzes electron transfer from NADH to ubiquinone-1. It is composed of about 22 different and mostly hydrophilic subunits and contains 2.0 nmol of FMN/mg of protein. Among its subunits is the 51-kDa subunit, which binds FMN and NADH and probably contains a [4Fe-4S] cluster also. Three other potential Fe-S proteins, the 75- and 24-kDa subunits and a 23-kDa subunit (N-terminal sequence TYKY), are also present. All of the Fe-S clusters detectable by EPR in complex I, including cluster 2, are found in subcomplex I alpha. The line shapes of the EPR spectra of the Fe-S clusters are slightly broadened relative to spectra measured on complex I purified by conventional means, and the quinone reductase activity is insensitive to rotenone. Similar changes were found in samples of the intact chromatographically purified complex I, or in complex I prepared by the conventional method and then subjected to chromatography in the presence of lauryl maltoside. Subcomplex I beta contains about 15 different subunits. The sequences of many of them contain hydrophobic segments that could be membrane spanning, including at least two mitochondrial gene products, ND4 and ND5. The role of subcomplex I beta in the intact complex remains to be elucidated.  相似文献   

9.
A very potent and specific inhibitor of mitochondrial NADH:ubiquinone oxidoreductase (complex I), a derivative of NADH (NADH-OH) has recently been discovered (Kotlyar, A. B., Karliner, J. S., and Cecchini, G. (2005) FEBS Lett. 579, 4861-4866). Here we present a quantitative analysis of the interaction of NADH-OH and other nucleotides with oxidized and reduced complex I in tightly coupled submitochondrial particles. Both the rate of the NADH-OH binding and its affinity to complex I are strongly decreased in the presence of succinate. The effect of succinate is completely reversed by rotenone, antimycin A, and uncoupler. The relative affinity of ADP-ribose, a competitive inhibitor of NADH oxidation, is also shown to be significantly affected by enzyme reduction (KD of 30 and 500 microM for oxidized and the succinate-reduced enzyme, respectively). Binding of NADH-OH is shown to abolish the succinate-supported superoxide generation by complex I. Gradual inhibition of the rotenone-sensitive uncoupled NADH oxidase and the reverse electron transfer activities by NADH-OH yield the same final titration point (approximately 0.1 nmol/mg of protein). The titration of NADH oxidase appears as a straight line, whereas the titration of the reverse reaction appears as a convex curve. Possible models to explain the different titration patterns for the forward and reverse reactions are briefly discussed.  相似文献   

10.
Adriamycin (Doxorubicin) stimulates NADH oxidase activity in liver plasma membrane, but does not cause NADH oxidase activity to appear where it is not initially present, as in erythrocyte membrane. NADH dehydrogenase from rat liver and erythrocyte plasma membranes shows similar adriamycin effects with other electron acceptors. Both NADH ferricyanide reductase and vanadate-stimulated NADH oxidation are inhibited by adriamycin, as is a cyanide insensitive ascorbate oxidase activity, whereas NADH cytochrome c reductase is not affected. The effects may contribute to the growth inhibitory (control) and/or deleterious effects of adriamycin. It is clear that adriamycin effects on the plasma membrane dehydrogenase involve more than a simple catalysis of superoxide formation.  相似文献   

11.
The steady-state kinetics of the transhydrogenase reaction (the reduction of acetylpyridine adenine dinucleotide (APAD+) by NADH, DD transhydrogenase) catalyzed by bovine heart submitochondrial particles (SMP), purified Complex I, and by the soluble three-subunit NADH dehydrogenase (FP) were studied to assess a number of the Complex I-associated nucleotide-binding sites. Under the conditions where the proton-pumping transhydrogenase (EC 1.6.1.1) was not operating, the DD transhydrogenase activities of SMP and Complex I exhibited complex kinetic pattern: the double reciprocal plots of the velocities were not linear when the substrate concentrations were varied in a wide range. No binary complex (ping-pong) mechanism (as expected for a single substrate-binding site enzyme) was operating within any range of the variable substrates. ADP-ribose, a competitive inhibitor of NADH oxidase, was shown to compete more effectively with NADH (Ki = 40 microM) than with APAD+ (Ki = 150 microM) in the transhydrogenase reaction. FMN redox cycling-dependent, FP catalyzed DD transhydrogenase reaction was shown to proceed through a ternary complex mechanism. The results suggest that Complex I and the simplest catalytically competent fragment derived therefrom (FP) possess more than one nucleotide-binding sites operating in the transhydrogenase reaction.  相似文献   

12.
Objective: A non‐silent polymorphism in the mitochondrial coding region of the ND1 gene, a subunit of reduced nicotinamide adenine dinucleotide (NADH) dehydrogenase is associated with resting metabolic rate (RMR) in 245 non‐diabetic Pima Indians. The purpose of this investigation was to determine the effect of the ND1 gene polymorphism on mitochondrial function in 14 male Pima Indians. Methods and Procedures: Seven subjects with an A at site 3547 of the ND1 gene (Ile at amino acid 81), and seven with a G at this site (Val) were studied. Mitochondria were isolated from 0.8 to 1.5 g of skeletal muscle obtained by needle biopsy of the lateral quadriceps muscle. In intact mitochondria, maximal (state‐3) and resting (state‐4) respiration rates were measured polarographically at 37 °C with a variety of single substrates or substrate combinations. Disrupted mitochondria were analyzed for maximal capacities through the entire electron transport chain (ETC) (NADH oxidase (NADHOX)), as well as through a segment of Complex I that is independent of the ND1 component (NADH‐ferricyanide (NADH‐FeCN) reductase). Results: Mitochondria were well coupled and exhibited higher respiratory control ratios (RCRs) than rodent muscle. There were no differences between the two groups for any of the measured parameters. Discussion: These results indicate that the cause of the observed association between RMR and the ND1 polymorphism is not related to in vitro mitochondrial function.  相似文献   

13.
We report the electron transfer properties of the NADH:ubiquinone oxidoreductase complex of the respiratory chain (Complex I) in mitochondria of cells derived from LHON patients with two different mutations in mitochondrial DNA (mtDNA). The mutations occur in the mtDNA genes coding for the ND1 and ND4 subunits of Complex I. The ND1/3460 mutation exhibits 80% reduction in rotenone-sensitive and ubiquinone-dependent electron transfer activity, whereas the proximal NADH dehydrogenase activity of the Complex is unaffected. This is in accordance with the proposal that the ND1 subunit interacts with rotenone and ubiquinone. In contrast, the ND4/11778 mutation had no effect on electron transfer activity of the Complex in inner mitochondrial membrane preparations; also Km for NADH and NADH dehydrogenase activity were unaffected. However, in isolated mitochondria with the ND4 mutation, the rate of oxidation of NAD-linked substrates, but not of succinate, was significantly decreased. This suggests that the ND4 subunit might be involved in specific aggregation of NADH-dependent dehydrogenases and Complex I, which may result in fast ('solid state') electron transfer from the former to the latter.  相似文献   

14.
During aerobic growth of Escherichia coli, nicotinamide adenine dinucleotide (NADH) can initiate electron transport at either of two sites: Complex I (NDH-1 or NADH: ubiquinone oxidoreductase) or a single-subunit NADH dehydrogenase (NDH-2). We report evidence for the specific coupling of malate dehydrogenase to Complex I. Membrane vesicles prepared from wild type cultures retain malate dehydrogenase and are capable of proton translocation driven by the addition of malate+NAD. This activity was inhibited by capsaicin, an inhibitor specific to Complex I, and it proceeded with deamino-NAD, a substrate utilized by Complex I, but not by NDH-2. The concentration of free NADH produced by membrane vesicles supplemented with malate+NAD was estimated to be 1 μM, while the rate of proton translocation due to Complex I was consistent with a some what higher concentration, suggesting a direct transfer mechanism. This interpretation was supported by competition assays in which inactive mutant forms of malate dehydrogenase were able to inhibit Complex I activity. These two lines of evidence indicate that the direct transfer of NADH from malate dehydrogenase to Complex I can occur in the E. coli system.  相似文献   

15.
The kinetics of the NAD: artificial acceptor-oxidoreductase and delta mu H(+)-dependent succinate: NAD(+)-oxidoreductase reactions (reverse electron transfer) reactions catalyzed by the membrane-bound complex I was studied. The values of apparent rate constants of dissociation of complexes of the oxidized and reduced enzyme with NAD+ and NADH were determined. It was shown that the apparent affinity of NADH for the oxidized complex I is by nearly three orders of magnitude as high as that of the reduced one; a reverse correlation is found for NAD+. A kinetic scheme of complex I functioning in the forward and reverse reactions, according to which the free reduced enzyme is not an intermediate of the forward (NADH-oxidase) reaction and the free oxidized enzyme is not an intermediate of the reverse (NAD(+)-reductase) reaction, is proposed.  相似文献   

16.
NADH-coenzyme Q reductase from bovine heart mitochondria (complex I) was incorporated into phospholipid vesicles by the cholate dialysis procedure. Mixtures of purified phosphatidylcholine and phosphatidylethanolamine were required. Oxidation of NADH by coenzyme Q1 catalyzed by the reconstituted vesicles was coupled to proton translocation, directed inward, with an H+/2e ratio greater than 1.4. Similar experiments measuring proton translocation in submitochondrial particles gave an H+/2e ratio of 1.8. The proton translocation in both systems was not seen in the presence of uncoupling agents and was in addition to the net proton uptake from the reduction of coenzyme Q1 by NADH. Electron transfer in the reconstituted vesicles also caused the uptake of the permeant anion tetraphenylboron. The rate of electron transfer by the reconstituted vesicles was stimulated about 3-fold by uncouplers or by valinomycin plus nigericin and K+ ions. The results indicate that energy coupling can be observed with isolated NADH-coenzyme Q reductase if the enzyme complex is properly incorporated into a phospholipid vesicle.  相似文献   

17.
The interactions of long chain (greater than C7), alkyl compounds with tightly coupled, beef heart submitochondrial particles (SMP) have been investigated with respect to their effects upon respiratory chain-linked electron transfer and energy coupling capacity. Long chain alkyl alcohols, amines, free fatty acids, and methyl esters exhibit a general uncoupling effect, with stimulation of the succinate oxidase activity but inhibition of the NADH oxidase, in SMP. The degree of effectiveness is dependent on the nature of the functional group and the length of the alkyl chain. Submitochondrial particles depleted of F1 and the F1-inhibitor protein are similarly affected. Subsequent treatment with bovine serum albumin reverses the effects of free fatty acids and results in partial recovery of activity with alkyl amines, alcohols, and methyl esters. Differences between the effects of these alkyl compounds and those of sodium dodecyl sulfate, deoxycholate, palmitoyl carnitine, and palmitoyl CoA rule out detergent-like action as the explanation for these observations. These data suggest that specific lipophilic interactions with the membrane, modulated by the nature of the functional group, are responsible for the effects of these compounds on the energy transducing system of SMP. Analyses of the reduction kinetics of the cytochromes indicate that the sites of interaction of these compounds with the inner mitochondrial membrane are associated with the primary dehydrogenase of complex I and energy coupling site 2; alkyl amines possess an additional site of interaction in the region of complex III.  相似文献   

18.
Cell-free extracts of Mycoplasma pneumoniae showed two distinct reduced nicotinamide adenine dinucleotide (NADH(2)) oxidase activities in the supernatant fraction. By ammonium sulfate fractionation and polyacrylamide gel electrophoresis, one activity not requiring flavine co-factors was precipitated by 50 to 70% ammonium sulfate concentration and identified with a slower-moving band on acrylamide gel electrophoresis; a second NADH(2) oxidase activity was flavine mononucleotide (FMN) dependent and associated with a more rapidly moving band; it could only be partially precipitated by ammonium sulfate concentrations ranging from 50 to 100%. Studies with alternate electron acceptors indicated the presence of a menadione, a 2,6-dichlorophenol indophenol and a very weak ferricyanide oxido-reductase activity, but no cytochrome c oxido-reductase, in the cell-free preparations. The NADH(2) oxidase activities of all fractions were relatively cyanide insensitive and were only minimally inhibited by flavoprotein and other respiratory chain inhibitors. H(2)O(2) formation was negligible unless FMN, but not flavine adenine dinucleotide (FAD), was added to the crude NADH(2) oxidase system; upon fractionation and electrophoresis, the H(2)O(2) formation was associated with the FMN-dependent, more rapidly moving NADH(2) oxidase band. This FMN-dependent NADH(2) oxidase-H(2)O(2) generating system may be a mechanism for the H(2)O(2) formation observed during glucose oxidation in the intact organism.  相似文献   

19.
Mitochondrial NADH:ubiquinone-reductase (Complex I) catalyzes proton translocation into inside-out submitochondrial particles. Here we describe a method for determining the stoichiometric ratio (n) for the coupled reaction of NADH oxidation by the quinone acceptors. Comparison of the initial rates of NADH oxidation and alkalinization of the surrounding medium after addition of small amounts of NADH to coupled particles in the presence of Q1 gives the value of n = 4. Thermally induced deactivation of Complex I [1,2] results in complete inhibition of the NADH oxidase reaction but only partial inhibition of the NADH:Q1-reductase reaction. N-Ethylmaleimide (NEM) prevents reactivation and thus completely blocks the thermally deactivated enzyme. The residual NADH:Q1-reductase activity of the deactivated, NEM-treated enzyme is shown to be coupled with the transmembraneous proton translocation (n = 4). Thus, thermally induced deactivation of Complex I as well as specific inhibitors of the endogenous ubiquinone reduction (rotenone, piericidin A) do not inhibit the proton translocating activity of the enzyme.  相似文献   

20.
Characterization of VDAC1 as a plasma membrane NADH-oxidoreductase   总被引:1,自引:0,他引:1  
We have recently demonstrated that voltage dependent anion selective channel~1 (porin, isoform 1) can function as a transplasma membrane NADH:ferricyanide-reductase. However, both the specific redox characteristics and the mechanism of electron transport in this enzyme presently remain unclear. Here we demonstrate that the redox capability of porin 1 is specific for ferricyanide as this same enzyme cannot reduce DCIP or cytochrome c in vitro. Furthermore, NADH-dependent ferricyanide reduction associated with VDAC1 is not sensitive to the anion channel inhibitors DIDS and dextran sulfate. However, this activity can be inhibited by thiol chelators, suggesting that at least one of the two cysteine groups present in VDAC1 are critical for electron transfer. We propose a model on how electron transport may occur in VDAC1.  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