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.
A systemic study has been made of copper and heme a binding to subunits of beef heart cytochrome c oxidase. Copper and heme a were readily mobilized by ionic detergents, high ionic strengths, temperatures above 0 degrees C, thiol compounds, and gel-bound peroxides and free radicals when the subunits of the oxidase were dissociated from one another during polyacrylamide gel electrophoresis. Most subunits showed some affinity for heme a and copper under these conditions. However, in the presence of specific mixtures of ionic and nonionic detergents (e.g. 0.1% sodium dodecyl sulfate, 0.025% Triton X-100) at temperatures below 0 degrees C and in buffers of low ionic strength using 10 to 12% polyacrylamide gels preelectrophoresed for 3 days with thioglycolate, about 90% of the Cu was found on subunit II (Mr = 24,100), and heme a was found in equal amounts of subunits I (Mr = 35,800) and II. The oxidized-reduced and reduced-CO absorption spectra of these subunits resembled those of cytochrome c oxidase. It appears probable that in the native enzyme, subunit I contains heme a and subunit II contains copper and heme a. A relationship of mammalian cytochrome c oxidase to the two-subunit microbial cytochrome oxidase systems appears to exist.  相似文献   

2.
We report that a macrocyclic ligand enables transfer of a protein from an aqueous phase to ionic liquids. The extraction behavior of heme protein cytochrome c (Cyt-c) from an aqueous phase into ionic liquids was investigated with crown ethers. A hydroxyl-group-containing ionic liquid with dicyclohexano-18-crown-6 was found to be capable of quantitative partitioning of Cyt-c, whereas the protein transfer using conventional organic solvents was negligibly small. Furthermore, we clarified that Cyt-c solubilized in ionic liquids caused a structural transformation of Cyt-c, which triggers its functional conversion from an electron-transfer protein to peroxidase.  相似文献   

3.
The solubilization and mineralization of (14)C-phenanthrene in soil-water systems was examined with several commercially available surface-active agents, viz., an alkyl ethoxylate C(12)E(4); two alkylphenol ethoxylate surfactants: C(8)PE(9.5) and C(9)PE(10.5); two sorbitan ethoxylate surfactants: the sorbitan monolaurate (Tween 20) and the sorbitan monooleate (Tween 80); two pairs of nonionic ethoxylate surfactant mixtures: C(12)E(4)/C(12)E(23) at a 1:1 ratio, and C(12-15)E(3)/C(12-15)E(9) at a 1:3 ratio; and two surfactants possessing relatively high critical micelle concentration (CMC) values and low aggregation numbers: CHAPS and octyglucoside. Surface tension experiments were performed to evaluate surfactant sorption onto soil and the surfactant doses required to attain the CMC in the soil-water systems. Surfactant solubilization of (14)C-phenanthrene commenced with the onset of micellization. The addition of surface-active agents was observed not to be beneficial to the microbial mineralization of phenanthrene in the soil-water systems and, for supra-CMC surfactant doses, phenanthrene mineralization was completely inhibited for all the surfactants tested. A comparison of solubilization, surface tension, and mineralization data confirms that the inhibitory effect on microbial degradation of phenanthrene is related to the CMC of the surfactant in the presence of soil. Additional tests demonstrated the recovery of mineralization upon dilution of surfactant concentration to sub-CMC levels, and a relatively high exit rate for phenanthrene from micelles. These tests suggest that the inhibitory effect is probably related to a reversible physiological surfactant micelle-bacteria interaction, possibly through partial complexing or release of membrane material with disrupting membrane lamellar structure. This study indicates that nonionic surfactant solubilization of sorbed hydrophobic organic compounds from soil may not be beneficial for the concomitant enhancement of soil bioremediation. Additional work is needed to address physicochemical processes for bioavailability enhancement, and effects of solubilizing agents on microorganisms for remediation and treatment of hydrophobic organic compounds and nonaqueous phase liquids. (c) 1992 John Wiley & Sons Inc.  相似文献   

4.
Proteins (bovine serum albumin (BSA), α‐chymotrypsin, cytochrome c, and lysozyme) were extracted from 0.5 to 2.0 g L?1 aqueous solution by adding an equal volume of isooctane solution that contained a surfactant mixture (Aerosol‐OT, or AOT, and a 1,3‐dioxolane (or cyclic ketal) alkyl ethoxylate, CK‐2,13‐E5.6), producing a three‐phase (Winsor‐III) microemulsion with a middle, bicontinuous microemulsion, phase highly concentrated in protein (5–13 g L?1) and small in volume (12–20% of entire volume). Greater than 90% forward extraction was achieved within a few minutes. Robust W‐III microemulsion systems were formulated at 40°C, or at 25°C by including a surfactant with shorter ethoxylate length, CK‐2,13‐E3, or 1.5% NaCl (aq). Successful forward extraction correlated with high partitioning of AOT in the middle phase (>95%). The driving force for forward extraction was mainly electrostatic attractions imposed by the anionic surfactant AOT, with the exception of BSA at high ionic strength, which interacted via hydrophobic interactions. Through use of aqueous stripping solutions of high ionic strength (5.0 wt %) and/or pH 12.0 (to negate the electrostatic attractive driving force), cytochrome c and α‐chymotrypsin were back extracted from the middle phase at >75% by mass, with the specific activity of recovered α‐chymotrypsin being >90% of its original value. © 2011 American Institute of Chemical Engineers Biotechnol. Prog., 2011  相似文献   

5.
The binding of ionic and nonionic, nondenaturing detergents to cytochrome c oxidase has been examined. All bind and displace part but not all of the phospholipid that is associated with the enzyme after isolation. From 6 to 10 phospholipid molecules, depending on the detergent used, do not exchange and these are mostly diphosphatidylglycerol molecules as first shown by Awasthi et al. ((1971) Biochim. Biophys. Acta 226, 42). The binding of Triton X-100 and deoxycholate to the cytochrome c oxidase complex has been studied in detail. Both bind to the enzyme above their critical micelle concentrations: Triton X-100 in the amount of 180 +/- 10 molecules per complex and deoxycholate in the amount of 80 +/- 4 molecules per complex. In nonionic detergents, cytochrome c oxidase exists as a dimer (4 heme complex). The enzyme is dissociated into the monomer or heme aa3 complex by delipidation in bile salts. Activity measurements in different detergents suggest that cytochrome c oxidase requires a flexible, hydrophobic environment for maximal activity and that the dimer or 4 heme complex may be the active species.  相似文献   

6.
The electron transfer complexes, succinate: ubiquinone reductase, ubiquinone: cytochrome c reductase, and cytochrome c: O2 oxidase were isolated from the mitochondrial membranes of Neurospora crassa by the following steps. Modification of the contents of the complexes in mitochondria by growing cells on chloramphenicol; solubilisation of the complexes by Triton X-100; affinity chromatography on immobilized cytochrome c and ion exchange and gel chromatography. Ubiquinone reductase was obtained in a monomeric form (Mr approximately 130 000) consisting of a flavin subunit (Mr 72 000) an iron-sulfur subunit (Mr 28 000) and a cytochrome b subunit (Mr probably 14 000). Cytochrome c reductase was obtained in a dimeric form (Mr approximately 550 000), the monomeric unit comprising the cytochromes b (Mr each 30 000), a cytochrome c1 (Mr 31 000), the iron-sulfur subunit (Mr 25 000), and six subunits without known prosthetic groups (Mr 9000, 11 000, 14 000, 45 000, 45 000, and 52 000). Cytochrome c oxidase was also isolated in a dimeric form (Mr approximately 320 000) comprising two copies each of seven subunits (Mr 9000, 12 000, 14 000, 18 000, 21 000, 29 000, and 40 000). The complexes were essentially free of phospholipid. Each bound one micelle of Triton X-100 (Mr approximately 90 000). After isolation, the bound Triton X-100 could be replaced by other nonionic detergents such as: alkylphenyl polyoxyethylene ethers, alkyl polyoxyethylene ethers and acyl polyoxyethylene sorbitan esters.  相似文献   

7.
The enzyme glucose-6-phosphate dehydrogenase (G6PD) plays an important role in maintaining the level of NADPH and in producing pentose phosphates for nucleotide biosynthesis. It is also of great value as an analytical reagent, being used in various quantitative assays. In searching for new strategies to purify this enzyme, the partitioning of G6PD in two-phase aqueous mixed (nonionic/cationic) micellar systems was investigated both experimentally and theoretically. Our results indicate that the use of a two-phase aqueous mixed micellar system composed of the nonionic surfactant C(10)E(4) (n-decyl tetra(ethylene oxide)) and the cationic surfactant C(n)TAB (alkyltrimethylammonium bromide, n = 8, 10, or 12) can improve significantly the partitioning behavior of G6PD relative to that obtained in the two-phase aqueous C(10)E(4) micellar system. This improvement can be attributed to electrostatic attractions between the positively charged mixed (nonionic/cationic) micelles and the net negatively charged enzyme G6PD, resulting in the preferential partitioning of G6PD to the top, mixed micelle-rich phase of the two-phase aqueous mixed micellar systems. The effect of varying the cationic surfactant tail length (n = 8, 10, and 12) on the denaturation and partitioning behavior of G6PD in the C(10)E(4) /C(n)TAB/buffer system was investigated. It was found that C(8)TAB is the least denaturing to G6PD, followed by C(10)TAB and C(12)TAB. However, the C(10)E(4)/C(12)TAB/buffer system generated stronger electrostatic attractions with the net negatively charged enzyme G6PD than the C(10)E(4)/C(10)TAB/buffer and the C(10)E(4)/C(8)TAB/buffer systems, when using the same amount of cationic surfactant. Overall, the two-phase aqueous mixed (C(10)E(4)/C(10)TAB) micellar system yielded the highest G6PD partition coefficient of 7.7, with a G6PD yield in the top phase of 71%, providing the optimal balance between the denaturing effect and the electrostatic attractions for the three cationic surfactants examined. A recently developed theoretical framework to predict protein partition coefficients in two-phase aqueous mixed (nonionic/ionic) micellar systems was implemented, and the theoretically predicted G6PD partition coefficients were found to be in reasonable quantitative agreement with the experimentally measured ones.  相似文献   

8.
9.
The effect of pH and salt concentration on the partitioning behavior of bovine serum albumin (BSA) and cytochrome c in an aqueous two-phase polymer system containing a novel pH-responsive copolymer that mimics the structure of proteins and poly(ethylene glycol) (PEG) was investigated. The two-phase system has low viscosity. Depending on pH and salt concentration, the cytochrome c was found to preferentially partition into the pH-responsive copolymer-rich (bottom) phase under all conditions of pH and salt concentrations considered in the study. This was caused by the attraction between the positively charged protein and negatively charged copolymer. BSA partitioning showed a more complex behavior and partitioned either to the PEG phase or copolymer phase depending on the pH and ionic strength. Extremely high partitioning levels (partition coefficient of 0.004) and very high separation ratios of the two proteins (up to 48) were recorded in the new systems. This was attributed to strong electrostatic interactions between the proteins and the charged copolymer.  相似文献   

10.
Water-in-oil microemulsion systems have been studied in recent years for a number of applications in protein separation and enzymology. Although it is well established that reversed micelle systems provide an excellent medium for nonaqueous biocatalytic studies, there is still much speculation as to the interaction of the enzyme with the surfactant interface. Polyoxyethylene sorbitan trioleate (Tween 85) is a nonionic surfactant which has some interesting properties for microemulsion formation and protein solubilization. In conjunction with a separate article describing the structural features of Tween 85 reversed micelles in hexane with isopropanol as a cosurfactant, this work describes the activity of an enzyme, organophosphorus hydrolase, for degrading organophosphorus pesticides in this microemulsion system. Ternary phase diagrams were constructed to outline the phase boundaries at different temperatures and isopropanol concentrations, which elucidate the role of the cosurfactant alcohol, as well as some features of micelle structure. Kinetic and stability studies with organophosphorus hydrolase show the effect of enzyme partitioning between the micelle surfactant layer and aqueous core. (c) 1994 John Wiley & Sons, Inc.  相似文献   

11.
We discuss the utilization of a novel two-phase aqueous nonionic micellar system for the purification and concentration of biomolecules, such as proteins and viruses, by liquid–liquid extraction. The nonionic surfactant n-decyl tetra(ethylene oxide), C10E4, phase separates in water into two coexisting aqueous micellar phases by increasing temperature. The mild interactions of the C10E4 nonionic surfactant with biomolecules, combined with the high water content of the two coexisting micellar phases, suggest the potential utility of two-phase aqueous C10E4 micellar systems for the purification and concentration of biomolecules. In this paper, we review our recent experimental and theoretical studies involving the partitioning of several water-soluble proteins, including cytochrome c, soybean trypsin inhibitor, ovalbumin, bovine serum albumin, and catalase, in the two-phase aqueous C10E4 micellar system. In addition, we present results of our preliminary experimental investigation on the partitioning of bacteriophages, including φX174, P22, and T4.  相似文献   

12.
Protein extraction and activity in reverse micelles of a nonionic detergent   总被引:2,自引:0,他引:2  
We describe, for the first time, the ability of a polyoxyethylene sorbitan trioleate-isopropanol microemulsion in hexane to solubilize pure proteins. The dependences of cytochrome c extraction and buffer solubilization by the reverse micellar system on ionic strength of the aqueous phase, detergent concentration, and cosurfactant concentration result in increased extraction. In addition, subtilisin (a serine protease) is shown to be active in this microemulsion. Further the activity of the enzyme can be regulated by the water content of the micelles, enabling control of enzyme activity by "solvent engineering."  相似文献   

13.
14.
The kinetics of dithionite reduction of the oxidized heme nonapeptide fragment of horse heart cytochrome c have been measured as a function of ionic strength at pH 7 and pH 9 by the stopped-flow technique. Dithionite concentration dependences indicate that the radical anion monomer, SO2-., is the active reductant. The pH 7 ionic strength dependence suggests that the heme peptide is reacting as a negatively charged molecule (its overall charge is calculated to be -1). Comparison of these results with the known rate of dithionite reduction of cytochrome c indicates that the heme nonapeptide has substantially greater inherent reactivity than cytochrome c, perhaps due to the greater accessibility of the heme.  相似文献   

15.
The effect of ionic strength on the one-electron reduction of oxidized bovine cytochrome c oxidase by reduced bovine cytochrome c has been studied by using flavin semiquinone reductants generated in situ by laser flash photolysis. In the absence of cytochrome c, direct reduction of the heme a prosthetic group of the oxidase by the one-electron reductant 5-deazariboflavin semiquinone occurred slowly, despite a driving force of approximately +1 V. This is consistent with a sterically inaccessible heme a center. This reduction process was independent of ionic strength from 10 to 100 mM. Addition of cytochrome c resulted in a marked increase in the amount of reduced oxidase generated per laser flash. Reduction of the oxidase at the heme a site was monophasic, whereas oxidation of cytochrome c was multiphasic, the fastest phase corresponding in rate constant to the reduction of the heme a. During the fast kinetic phase, 2 equiv of cytochrome c was oxidized per heme a reduced. We presume that the second equivalent was used to reduce the Cua center, although this was not directly measured. The first-order rate-limiting process which controls electron transfer to the heme a showed a marked ionic strength effect, with a maximum rate constant occurring at mu = 110 mM (1470 s-1), whereas the rate constant obtained at mu = 10 mM was 630 s-1 and at mu = 510 mM was 45 s-1. There was no effect of "pulsing" the enzyme on this rate-limiting one-electron transfer process. These results suggest that there are structural differences in the complex(es) formed between mitochondrial cytochrome c and cytochrome c oxidase at very low and more physiologically relevant ionic strengths, which lead to differences in electron-transfer rate constants.  相似文献   

16.
Volatile alkyl formates are potential replacements for the ozone-depleting fumigant, methyl bromide, as postharvest insecticides and here we have investigated their mode of insecticidal action. Firstly, a range of alkyl esters, ethanol and formic acid were tested in mortality bioassays with adults of the rice weevil, Sitophilus oryzae (L.) and the grain borer, Rhyzopertha dominica (F.) to determine whether the intact ester or one of its components was the toxic moiety. Volatile alkyl formates and formic acid caused similar levels of mortality (LC(50) 131-165 micromol l(-1)) to S. oryzae and were more potent than non-formate containing alkyl esters and ethanol (LC(50)>275 micromol l(-1)). The order of potency was the same in R. dominica. Ethyl formate was rapidly metabolised in vitro to formic acid when incubated with insect homogenates, presumably through the action of esterases. S. oryzae and R. dominica fumigated with a lethal dose of ethyl formate had eight and 17-fold higher concentrations of formic acid, respectively, in their bodies than untreated controls. When tested against isolated mitochondria from S. oryzae, alkyl esters, alcohols, acetate and propionate salts were not inhibitory towards cytochrome c oxidase (EC 1.9.3.1), but sodium cyanide and sodium formate were inhibitory with IC(50) values of 0.0015 mM and 59 mM, respectively. Volatile formate esters were more toxic than other alkyl esters, and this was found to be due, at least in part, to their hydrolysis to formic acid and its inhibition of cytochrome c oxidase.  相似文献   

17.
The heme moieties of cytochromes P-450 and P1-450 (P-448) have been characterized. CO-binding particles, devoid of cytochrome b5, were prepared from normal or 3-methylcholanthrene-treated animals. Heme was removed by acid-acetone treatment of the CO-binding particles and crystallized. Heme isolated from hemoglobin of the corresponding animals served as a control. Reductive degradation (hydriodic acid) followed by gas chromatography/mass spectrometry analysis indicated the presence of opso-, crypto-, hemo-, and phyllopyrrole. Visible spectra of the iron-free tetrapyrroles isolated from microsomal heme and hemoglobin were identical and showed typical aetioporphyrin spectra. Finally, the mass spectra of the tetrapyrrole dimethyl esters isolated from microsomal heme and hemoglobin were identical to authentic protoporphyrin IX dimethyl ester. These data strongly suggest that the heme of cytochrome P-450 and P1-450 are identical and are the same the same as that of hemoglobin, namely protoporphyrin IX.  相似文献   

18.
The reduction of the tetraheme cytochrome c3 (from Desulfovibrio vulgaris, strains Miyazaki F and Hildenbourough) by flavin semiquinone and reduced methyl viologen follows a monophasic kinetic profile, even though the four hemes do not have equivalent reduction potentials. Rate constants for reduction of the individual hemes are obtained subsequent to incrementally reducing the cytochrome by phototitration. The dependence of each rate constant on the reduction potential difference between the heme and the reductant can be described by outer sphere electron transfer theroy. Thus, the very low reduction potentials of the cytochrome c3 hemes compensate for the very large solvent accessibility of the hemes. The relative rate constants for electron transfer to the four hemes of cytochrome c3 are consistent with the assignments of reduction potential to hemes previously made by Park et al. (Park, J.-S., Kano, K., Niki, S. and Akutsu, H. (1991) FEBS Lett. 285, 149-151) using NMR techniques. The ionic strength dependence of the observed rate constant for reduction by the methyl viologen radical cation indicates that ionic strength substantially alters the structure and/or the heme reduction potentials of the cytochrome. This result is confirmed by reduction with a neutral flavin species (5-deazariboflavin semiquinone) in which the reactivity of the highest potential heme decreases and the reactivity of the lowest potential heme increases at high (500 mM) ionic strength, and by the sensitivity of heme methyl resonances to ionic strength as observed by 1H-NMR. These unusual ionic strength-dependent effects may be due to a combination of structural changes in the cytochrome and alterations of the electrostatic fields at elevated ionic strengths.  相似文献   

19.
The interactions of cytochrome c1 and cytochrome c from bovine cardiac mitochondria were investigated. Cytochrome c1 and cytochrome c formed a 1:1 molecular complex in aqueous solutions of low ionic strength. The complex was stable to Sephadex G-75 chromatography. The formation and stability of the complex were independent of the oxidation state of the cytochrome components as far as those reactions studied were concerned. The complex was dissociated in solutions of ionic strength higher than 0.07 or pH exceeding 10 and only partially dissociated in 8 M urea. No complexation occurred when cytochrome c was acetylated on 64% of its lysine residues or photooxidized on its 2 methionine residues. Complexes with molecular ratios of less than 1:1 (i.e. more cytochrome c) were obtained when polymerized cytochrome c, or cytochrome c with all lysine residues guanidinated, or a "1-65 heme peptide" from cyanogen bromide cleavage of cytochrome c was used. These results were interpreted to imply that the complex was predominantly maintained by ionic interactions probably involving some of the lysine residues of cytochrome c but with major stabilization dependent on the native conformations of both cytochromes. The reduced complex was autooxidizable with biphasic kinetics with first order rate constants of 6 X 10(-5) and 5 X U0(-5) s-1 but did not react with carbon monoxide. The complex reacted with cyanide and was reduced by ascorbate at about 32% and 40% respectively, of the rates of reaction with cytochrome c alone. The complex was less photoreducible than cytochrome c1 alone. The complex exhibited remarkably different circular dichroic behavior from that of the summation of cytochrome c1 plus cytochrome c. We concluded that when cytochromes c1 and c interacted they underwent dramatic conformational changes resulting in weakening of their heme crevices. All results available would indicate that in the complex cytochrome c1 was bound at the entrance to the heme crevice of cytochrome c on the methionine-80 side of the heme crevice.  相似文献   

20.
The effect of ionic strength on the rate constant for electron transfer has been used to determine the magnitude and charge sign of the net electrostatic potential which exists in close proximity to the sites of electron transfer on various c-type cytochromes. The negatively charged ferricyanide ion preferentially reacts at the positively charged exposed heme edge region on the front side of horse cytochrome c and Paracoccus cytochrome c2. In contrast, at low ionic strength, the positively charged cobalt phenanthroline ion interacts with the negatively charged back side of cytochrome c2, and at high ionic strength at a positively charged site on the front side of the cytochrome. With horse cytochrome c, over the ionic strength range studied, cobalt phenanthroline reacts only at a positively charged site which is probably not at the heme edge. These inorganic oxidants do not react at the relatively uncharged exposed heme edge sites on Azotobacter cytochrome c5 and Pseudomonas cytochrome c-551, but rather at a negatively charged site which is away from the heme edge. The results demonstrate that at least two electron-transferring sites on a single cytochrome can be functional, depending on the redox reactant used and the ionic strength. Electrostatic interactions between charge distributions on the cytochrome surface and the other reactant, or interactions involving uncharged regions on the protein(s), are critical in determining the preferred sites of electron transfer and reaction rate constants. When unfavorable electrostatic effects occur at a site near the redox center, less optimal sites at a greater distance can become kinetically important.  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