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 972 毫秒
1.
2.
Previously, we investigated ubisemiquinone (SQ) EPR spectra associated with NADH-ubiquinone oxidoreductase (complex I) in the tightly coupled bovine heart submitochondrial particles (SMP). Based upon their widely differing spin relaxation rate, we distinguished SQ spectra arising from three distinct SQ species, namely SQ(Nf) (fast), SQ(Ns) (slow), and SQ(Nx) (very slow). The SQ(Nf) signal was observed only in the presence of the proton electrochemical gradient (deltamu(H)(+)), while SQ(Ns) and SQ(Nx) species did not require the presence of deltamu(H+). We have now succeeded in characterizing the redox and EPR properties of SQ species in the isolated bovine heart complex I. The potentiometric redox titration of the g(z,y,x)=2.00 semiquinone signal gave the redox midpoint potential (E(m)) at pH 7.8 for the first electron transfer step [E(m1)(Q/SQ)] of -45 mV and the second step [E(m2)(SQ/QH(2))] of -63 mV. It can also be expressed as [E(m)(Q/QH(2))] of -54 mV for the overall two electron transfer with a stability constant (K(stab)) of the SQ form as 2.0. These characteristics revealed the existence of a thermodynamically stable intermediate redox state, which allows this protein-associated quinone to function as a converter between n=1 and n=2 electron transfer steps. The EPR spectrum of the SQ species in complex I exhibits a Gaussian-type spectrum with the peak-to-peak line width of approximately 6.1 G at the sample temperature of 173 K. This indicates that the SQ species is in an anionic Q(-) state in the physiological pH range. The spin relaxation rate of the SQ species in isolated complex I is much slower than the SQ counterparts in the complex I in situ in SMP. We tentatively assigned slow relaxing anionic SQ species as SQ(Ns), based on the monophasic power saturation profile and several fold increase of its spin relaxation rate in the presence of reduced cluster N2. The current study also suggests that the very slowly relaxing SQ(Nx) species may not be an intrinsic complex I component. The functional role of SQ(Ns) is further discussed in connection with the SQ(Nf) species defined in SMP in situ.  相似文献   

3.
(20S)-[7,7,21,21-2H(4)]-3beta-(tert-Butyldimethylsilanyloxy)-20-methyl-pregn-5-en-21-ol, an intermediate for the preparation of deuterated isotopomers of sterols to be used as standards for biomedical studies, was prepared by reduction with dichloroaluminum deuteride of ethyl (20S)-3beta-(tert-butyldimethylsilanyloxy)-7-oxo-pregn-5-en-20-carboxylate. Using controlled experimental conditions, it has also been shown that the dichloroaluminum hydride reduction of a 7-keto steroid affords the corresponding 7beta-hydroxy derivative in a highly stereoselective manner.  相似文献   

4.
Activation of the O(2)(-) generating NADPH oxidase of phagocytes results from the assembly of the membrane-bound flavocytochrome b(558) with cytosolic proteins, p67(phox), p47(phox), and Rac. However, it has been recently reported that the arachidonic acid- and calcium-binding heterodimer S100A8/A9, abundant in neutrophil cytosol, influences the activation process. In a semi-recombinant system comprising neutrophil membranes, recombinant proteins, p67(phox), p47(phox), GTPgamma S-loaded Rac2, and arachidonic acid (AA), both the rate and the extent of the oxidase activation were increased by S100A8/A9, provided it was preloaded with AA. Binding of [(14)C]AA to S100A8/A9 was potentiated by recombinant cytosolic phox proteins and GTPgammaS, suggesting the formation of a complex, comprising oxidase activating proteins and S100A8/A9, with a greater affinity for AA. The rate constant of oxidase activation was not increased by AA-loaded S100A8/A9, whereas the maximal oxidase activity elicited was twice as high. AA-loaded S100A8/A9 increases oxidase activation probably by decreasing the deactivation rate.  相似文献   

5.
Ohnishi T  Salerno JC 《FEBS letters》2005,579(21):4555-4561
A novel mechanism for proton/electron transfer is proposed for NADH-quinone oxidoreductase (complex I) based on the following findings: (1) EPR signals of the protein-bound fast-relaxing semiquinone anion radicals (abbreviated as Q(Nf)-) are observable only in the presence of proton-transmembrane electrochemical potential; (2) Iron-sulfur cluster N2 and Q(Nf)- are directly spin-coupled; and (3) The projection of the interspin vector extends only 5A along the membrane normal [Yano, T., Dunham, W.R. and Ohnishi, T. (2005) Biochemistry, 44, 1744-1754]. We propose that the proton pump is operated by redox-driven conformational changes of the quinone binding protein. In the input state, semiquinone is reduced to quinol, acquiring two protons from the N (matrix) side of the mitochondrial inner membrane and an electron from the low potential (NADH) side of the respiratory chain. A conformational change brings the protons into position for release at the P (inter-membrane space) side of the membrane via a proton-well. Concomitantly, an electron is donated to the quinone pool at the high potential side of the coupling site. The system then returns to the original state to repeat the cycle. This hypothesis provides a useful frame work for further investigation of the mechanism of proton translocation in complex I.  相似文献   

6.
To confirm whether or not the sulfo group of estradiol 17-sulfate (ES) is removed during in vivo metabolism in rats, the doubly labeled conjugate [6,7-3H, 35S] ES was injected into rats, and its biliary and urinary metabolites were determined by reverse isotope dilution method (RIDM). In male rats, the major radioactivity was detected in biliary disulfate fraction, which was composed of mainly ES and its two minor metabolites, 2-hydroxyestradiol 17-sulfate (2-OH-ES) and 2-methoxyestradiol 17-sulfate (2-MeO-ES). In female rats, in contrast, the radioactivity was dispersed into three fractions:biliary monosulfate, biliary disulfate, and urinary monosulfate fractions (Frs.) In both monosulfate Frs., 7beta-hydroxyestradiol 17-sulfate was detected as the major metabolite followed by 6alpha-, 6beta-, and 15beta-hydroxyestradiol 17-sulfates. Like male rats, 2-OH-ES and 2-Meo-ES as the minor products were detected in biliary disulfate fraction. The isotope ratios of ES and its metabolites in both sexes were essentially the same as that of the dose except that of 6alpha-hydroxylated metabolite, which may be derived from the loss of the tritium labeled at C6. These results confirm the occurrence of the direct metabolism of ES in rats.  相似文献   

7.
Electron paramagnetic resonance (EPR) spectroscopy reveals functional and structural similarities between the reaction centres of the chlorophyll d-binding photosystem I (PS I) and chlorophyll a-binding PS I. Continuous wave EPR spectrometry at 12K identifies iron-sulphur centres as terminal electron acceptors of chlorophyll d-binding PS I. A transient light-induced electron spin echo (ESE) signal indicates the presence of a quinone as the secondary electron acceptor (Q) between P(740)(+) and the iron-sulphur centres. The distance between P(740)(+) and Q(-) was estimated within point-dipole approximation as 25.23+/-0.05A, by the analysis of the electron spin echo envelope modulation.  相似文献   

8.
Pressure-induced unfolding of 23-kDa protein from spinach photosystem II has been systematically investigated at various experimental conditions. Thermodynamic equilibrium studies indicate that the protein is very sensitive to pressure. At 20 degrees C and pH 5.5, 23-kDa protein shows a reversible two-state unfolding transition under pressure with a midpoint near 160 MPa, which is much lower than most natural proteins studied to date. The free energy (DeltaG(u)) and volume change (DeltaV(u)) for the unfolding are 5.9 kcal/mol and -160 ml/mol, respectively. It was found that NaCl and sucrose significantly stabilize the protein from unfolding and the stabilization is associated not only with an increase in DeltaG(u) but also with a decrease in DeltaV(u). The pressure-jump studies of 23-kDa protein reveal a negative activation volume for unfolding (-66.2 ml/mol) and a positive activation volume for refolding (84.1 ml/mol), indicating that, in terms of system volume, the protein transition state lies between the folded and unfolded states. Examination of the temperature effect on the unfolding kinetics indicates that the thermal expansibility of the transition state and the unfolded state of 23-kDa protein are closer to each other and they are larger than that of the native state. The diverse pressure-refolding pathways of 23-kDa protein in some conditions were revealed in pressure-jump kinetics.  相似文献   

9.
We recently found that the concentration of HCO3- in guinea-pig saliva is very similar to that of human saliva; however, the entity that regulates HCO3- transport has not yet been fully characterized. In order to investigate the mechanism of HCO3- transport, we identified, cloned, and characterized a sodium bicarbonate (Na(+)/HCO3- cotransporter found in guinea-pig parotid glands (gpNBC1). The gpNBC1 gene encodes a 1079-amino acid protein that has 95% and 96% homology with human and mouse parotid NBC1, respectively. Oocytes expressing gpNBC1 were exposed to HCO3- or Na(+)-free solutions, which resulted in a marked change in membrane potentials (V(m)), suggesting that gpNBC1 is electrogenic. Likewise, a gpNBC1-mediated pH recovery was observed in gpNBC1 transfected human hepatoma cells; however, in the presence of 4, 4-diisothiocyanostilbene-2,2-disulfonic acid, a specific NBC1 inhibitor, such changes in V(m) and pH(i) were not observed. Together, the data show that the cloned guinea-pig gene is a functional, as well as sequence homologue of human NBC1.  相似文献   

10.
Xyloglucans from seeds of Copaifera langsdorffii (XGC), Hymenaea courbaril (XGJ) and Mucuna sloanei (XGM) were obtained from milled and defatted cotyledons by aqueous extraction at 25 degrees C. The resulting fractions contained Glc, Xyl and Gal in molar ratios of 2.5: 1.5: 1.0 (XGC), 3.8: 2.6: 1.0 (XGJ) and 2.5: 1.6: 1.0 (XGM). HPSEC-MALLS/RI analysis showed that each polysaccharide fraction was homogeneous; M(w) values were 1.6 x 10(5), 2.0 x 10(5) and 1.5 x 10(5)g/mol, respectively. The effect of the xyloglucans on the production of O(2)*(-) and NO* and on the recruitment of macrophages to the mouse peritoneum was evaluated. All polysaccharides promoted an increase in the number of peritoneal macrophages in a dose-dependent manner. The largest increase, of 576% in comparison to the control group, was elicited by XGJ at 200 mg/kg. The effect of XGC, XGJ and XGM on O(2)*(-) production, in the presence or absence of phorbol 12-myristate 13-acetate (PMA), was not statistically significant. For NO(.) production, the lowest concentration of XGC (10 microg/ml) gave rise to an increase of 262% when compared to the control group; the effect was dose-dependent, reaching 307% at 50 microg/ml. On the other hand, XGJ at a concentration of 50 microg/ml enhanced NO* production by 92%. XGM did not affect NO* production significantly. The results indicate that xyloglucans from C. langsdorffii, H. courbaril and M. sloanei have immunomodulatory activity.  相似文献   

11.
We compared the thylakoid membrane composition and photosynthetic properties of non- and cold-acclimated leaves from the dgd1 mutant (lacking >90% of digalactosyl-diacylglycerol; DGDG) and wild type (WT) Arabidopsis thaliana. In contrast to warm grown plants, cold-acclimated dgd1 leaves recovered pigment-protein pools and photosynthetic function equivalent to WT. Surprisingly, this recovery was not correlated with an increase in DGDG. When returned to warm temperatures the severe dgd1 mutant phenotype reappeared. We conclude that the relative recovery of photosynthetic activity at 5 degrees C resulted from a temperature/lipid interaction enabling the stable assembly of PSI complexes in the thylakoid.  相似文献   

12.
Nahar L  Turner AB 《Steroids》2003,68(14):1157-1161
Four lithocholic acid dimers were synthesised via esterification. The ester-linked dimer, 3-oxo-5beta-cholan-24-oic acid (cholan-24-oic acid methyl ester)-3-yl ester, (3alpha,5beta), was obtained by condensation of methyl lithocholate with 3-oxo-5beta-cholan-24-oic acid. Borohydride reduction of this ester-linked dimer gave 3alpha-hydroxy-5beta-cholan-24-oic acid (cholan-24-oic acid methyl ester)-3-yl ester, (3alpha,5beta), which was acetylated to 3alpha-acetoxy-5beta-cholan-24-oic acid (cholan-24-oic acid methyl ester)-3-yl ester, (3alpha,5beta). Reaction of methyl lithocholate with oxalyl chloride yielded the oxalate dimer, bis(5beta-cholan-24-oic acid methyl ester)-3alpha-yl oxalate.  相似文献   

13.
6,7-Dimethyl-8-ribityllumazine synthase (lumazine synthase) catalyses the penultimate step in the biosynthesis of riboflavin. In Bacillus subtilis, 60 lumazine synthase subunits form an icosahedral capsid enclosing a homotrimeric riboflavin synthase unit. The ribH gene specifying the lumazine synthase subunit can be expressed in high yield. All amino acid residues exposed at the surface of the active site cavity were modified by PCR assisted mutagenesis. Polar amino acid residues in direct contact with the enzyme substrates, 5-amino-6-ribitylamino-2,4(1H,3H)-pyrimidinedione and 3,4-dihydroxy-2-butanone 4-phosphate, could be replaced with relative impunity with regard to the catalytic properties. Only the replacement of Arg127, which forms a salt bridge with the phosphate group of 3,4-dihydroxy-2-butanone 4-phosphate, reduced the catalytic rate by more than one order of magnitude. Replacement of His88, which is believed to assist in proton transfer reactions, reduced the catalytic activity by about one order of magnitude. Surprisingly, the activation enthalpy deltaH of the lumazine synthase reaction exceeds that of the uncatalysed reaction. On the other hand, the free energy of activation deltaG of the uncatalysed reaction is characterised by a large entropic term (TdeltaS) of -37.8 kJmol(-1), whereas the entropy of activation (TdeltaS) of the enzyme-catalysed reaction is -6.7 kJmol(-1). This suggests that the rate enhancement by the enzyme is predominantly achieved by establishing a favourable topological relation of the two substrates, whereas acid/base catalysis may play a secondary role.  相似文献   

14.
Cells of the coccolithophorid Emiliania huxleyi strain CS-57 grown under an atmosphere of air+0.5% CO(2) showed oxidative damage after 10 days growth with concomitant and major changes to the lipid composition. The fatty acid profile was strongly altered and lacked appreciable amounts of the polyunsaturated fatty acids (PUFA: C(18:5), C(18:3) and C(22:6)) typical of healthy cells. Oxidation products of these PUFA could not be detected, but monounsaturated fatty acids proved to be good indicators of oxidative processes. The presence (after NaBH(4)-reduction) of a high proportion of 11-hydroxyoctadec-cis-9-enoic and 8-hydroxyoctadec-cis-9-enoic acids showed that the degradation of oleic acid involved mainly free radical oxidation processes (70-75% autoxidation and 20-25% photooxidation). We also detected large amounts of degradation products of the oxidation product 9,10-epoxyoctadecanoic acid including diols, methoxyhydrins and chlorohydrins. These oxidative effects were found in all the lipid classes examined. Products included significant amounts of chlorophyll side-chain autooxidation products Z- and E-3,7,11,15-tetramethylhexadec-3-ene-1,2-diols and Z-and E-3,7,11,15-tetramethylhexadec-2-ene-1,4-diols, while phytyldiol was present in relatively low proportions. Delta(5)-3beta,7-epimeric unsaturated steroidal diols arising from the autooxidation of the Delta(5) double bond of epi-brassicasterol and minor amounts of Delta(4)-3beta,6-diols were also detected. Long-chain unsaturated ketone (alkenone) content per cell was much higher in the presence of 0.5% CO(2) likely due to carbon storage under these conditions. The proportions of di- and tri-unsaturated alkenones was relatively stable throughout the growth cycle in the absence of additional CO(2), but not when grown with 0.5% CO(2). The detection of characteristic alkenone autoxidation products in cells grown under these latter conditions allowed us to attribute the significant increase in index observed to the involvement of free radical oxidation processes.  相似文献   

15.
16.
Because oxidative stress is a component of gastrointestinal injury, we investigated the effect of H(2)O(2) on transintestinal transport using isolated rat jejunum incubated in vitro. Millimolar concentrations of H(2)O(2) inhibited all the tested parameters without inducing any cytotoxic effect. Electrophysiological experiments indicated that H(2)O(2) decreases significantly both short circuit current and transepithelial electrical potential difference without affecting transepithelial resistance. The possibility that H(2)O(2) could influence (Na+, K+) -ATPase activity was explored using isolated basolateral membranes. Besides H(2)O(2), free radicals (O(2)(*-), HO*) were generated using different iron-dependent and independent systems; (Na+, K+) -ATPase activity was inhibited after membrane exposure to all ROS tested. The inhibition was prevented by allopurinol, superoxide dismutase or desferrioxamine. Western blot analysis showed a decreased expression of the alpha(1)-subunit of (Na+, K+) -ATPase. We conclude that H(2)O(2) may be a modulator of jejunal ion and water transport by multiple mechanisms, among which a significant inhibition of the basolateral (Na+, K+) -ATPase.  相似文献   

17.
The two osteoclastogenesis pathways, receptor activator nuclear factor (NF)-kappaB ligand (RANKL)-mediated and fusion regulatory protein-1 (FRP-1)-mediated osteoclastogenesis, have recently been reported. There were significant differences in differentiation and activation mechanisms between the two pathways. When monocytes were cultured with FRP-1 without adding M-CSF, essential for the RANKL system, TRAP-positive polykaryocyte formation occurred. FRP-1-mediated osteoclasts formed larger pits on mineralized calcium phosphate plates than RANKL+M-CSF-mediated osteoclasts did. Lacunae on dentin surfaces induced by FRP-1-mediated osteoclasts were inclined to be single and isolated. However, osteoclasts induced by RANKL+M-CSF made many connected pits on dentin surfaces as if they crawled on there. Interestingly, FRP-1 osteoclastogenesis was enhanced by M-CSF/IL-1alpha, while chemotactic behavior to the dentin slices was not effected. There were differences in pH and concentration of HCO3- at culture endpoint and in adherent feature to dentin surfaces. Our findings indicate there are two types of osteoclasts with distinct properties.  相似文献   

18.
19.
This study shows the presence of all three nitric oxide synthases (NOSs) and NOS activity in H9c2 cells cultured under non-stimulated conditions. By using the 4,5 diaminofluoresceindiacetate (DAF-2DA) fluorimetric nitric oxide (NO(*)) detection system we observed NO(*) production in H9c2 cells. As revealed by confocal microscopy, NO(*) fluorescence colocalizes in mitochondria labeled with Mito-Tracker Red CM-H(2)Xros. Upon stimulation with acetylcholine (Ach), which increased NOS activity by 75%, the colocalization coefficient C(green) value, calculated as Pearson's correlation, increased from 0.07 to 0.10, demonstrating an augmented presence of NO(*) in mitochondria. Conversely, the presence of NO(*) in mitochondria decreased following cells pretreatment with l-MonoMethylArginine (L-NMMA), a competitive inhibitor of NOS activity, as indicated by the reduction of the C(green) value to 0.02. This work confirms that the presence of NO(*) in mitochondria can be modulated in response to different fluxes of NO(*).  相似文献   

20.
As typical anchorage-dependent cells myocytes must balance contractility against adequate adhesion. Skeletal myotubes grown as isolated strips from myoblasts on micropatterned glass exhibited spontaneous peeling after one end of the myotube was mechanically detached. Such results indicate the development of a prestress in the cells. To assess this prestress and study the dynamic adhesion strength of single myocytes, the shear stress of fluid aspirated into a large-bore micropipette was then used to forcibly peel myotubes. The velocity at which cells peeled from the surface, V(peel), was measured as a continuously increasing function of the imposed tension, T(peel), which ranges from approximately 0 to 50 nN/ micro m. For each cell, peeling proved highly heterogeneous, with V(peel) fluctuating between 0 micro m/s ( approximately 80% of time) and approximately 10 micro m/s. Parallel studies of smooth muscle cells expressing GFP-paxillin also exhibited a discontinuous peeling in which focal adhesions fractured above sites of strong attachment (when pressure peeled using a small-bore pipette). The peeling approaches described here lend insight into the contractile-adhesion balance and can be used to study the real-time dynamics of stressed adhesions through both physical detection and the use of GFP markers; the methods should prove useful in comparing normal versus dystrophic muscle cells.  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