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 421 毫秒
1.
Membrane vesicles from Escherichia coli wild type and an otherwise isogenic dnaA mutant were used to immunize rabbits. In addition, a membrane protein fraction, containing the material found deficient in dnaA mutants, was purified by preparative polyacrylamide gel electrophoresis in sodium dodecylsulfate, and used for immunization. The antisera produced were analyzed by immunoelectrophoresis and immunofluorescence microscopy. The antisera obtained by immunization with membrane vesicles from either wild type or dnaA mutant membrane preparations were qualitatively similar in the precipitin bands seen after immunoelectrophoresis. The antisera obtained by immunization with the purified protein fraction contained a subset of the antibodies seen when whole vesicles were used for immunization. In a semiquantitative precipitin assay, the antisera prepared against whole membrane vesicles or the isolated protein fraction both caused the precipitation of more protein from sodium dodecylsulfate-solubilized membranes of wild type than of dnaA mutants. No difference was seen by immunoelectrophoresis between the protein composition of wild type or dnaA membrane preparations. Thus, the dnaA mutant appears to differ from the wild type in the quantitative composition of its membrane proteins, whereas no qualitative differences were detected.Fluorescein-conjugated antiserum preparations were employed to assess the reactivity of intact cells, spheroplasts and membrane vesicles with the antisera studied above. Wild type cells of E. coli have a barrier to reaction with the antisera; this barrier is removed when the cells are converted to spheroplasts or to membrane vesicle. Similarly, a highly permeable mutant of E. coli permits reaction of the antisera with unaltered cells. Antisera to both whole membrane vesicles and to the isolated protein fraction react identically with the cellular and subcellular preparations. Thus, antisera prepared from membrane proteins isolated after sodium dodecylsulfate-polyacrylamide gel electrophoresis can still recognize some antigens present in membrane vesicle preparations.  相似文献   

2.
Extraction of Escherichiacoli ML 308-225 membrane vesicles with cholate yields a particulate fraction containing 10 to 15% of the phospholipid and about 70% of the protein of intact vesicles. Addition of phospholipid to the particulate fraction in the presence of cholate, followed by sonication and removal of detergent by gel filtration yields a vesicular preparation that exhibits lac carrier function as judged by transient increases in 6′-(N-dansyl)aminohexyl-1-thio-β-D-galactopyranoside fluorescence in the presence of either a lactose diffusion gradient or an artificially-generated membrane potential (interior negative). Activity is not observed in the absence of phospholipid, in the presence of N-ethylmaleimide or in analogous preparations from ML 30 vesicles that do not contain the β-galactoside transport system.  相似文献   

3.
The oxidation of NADPH and NADH was studied in the light and in the dark using sonically derived membrane vesicles and osmotically shocked spheroplasts. These two types of cell-free membrane preparations mostly differ in that the cell and thylakoid membranes are scrambled in the former type and that they are more or less separated in the latter type of preparations. In the light, using both kinds of preparations, each of NADPH and NADH donates electrons via the plastoquinone-cytochrome bc redox complex (Qbc redox complex) to the thylakoid membrane-bound cytochrome c-553 preoxidized by a light flash and to methylviologen via Photosystem I. NADPH donates electrons to the thylakoid membrane via a weakly rotenone-sensitive dehydrogenase to a site that is situated beyond the 3(3′,4′-dichlorophenyl)-1,1-dimethylurea sensitive site and before plastoquinone. Ferredoxin and easily soluble cytoplasmic proteins are presumably not involved in light-mediated NADPH oxidation. Inhibitors of electron transfer at the Qbc redox complex as the dinitrophenylether of 2-iodo-4-nitrothymol, 2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone and 2-n-heptyl-4-hydroxy-quinone-N-oxide are effective, but antimycin A and KCN are not. The oxidation of NADH showed comparable sensitivity to these inhibitors. However, the oxidation of NADH is antimycin-A-sensitive regardless of the kind of membrane preparation used, indicating that in this case electrons are donated to a different site on the thylakoid membrane. In the dark, NADPH and NADH donate electrons at sites that behave similar to those of light-mediated oxidation, indicating that the initial steps of electron transfer are situated at the thylakoid membranes. However, NADPH oxidation is in some cases not sensitive to inhibitors active at the Qbc redox complex. It is concluded that O2 reduction takes place at two different sites, one partly developed in vitro, situated near the rotenone-sensitive NADPH dehydrogenase, and another, highly KCN-sensitive one, situated beyond the Qbc redox complex and used in vivo. The terminal oxygen-reducing step of NADPH and NADH oxidation in the dark showed a preparation-dependent sensitivity for KCN, more than 80% inhibition in sonically derived membrane vesicles and less than 30% inhibition in osmotically shocked spheroplasts. From this result we tentatively conclude that the highly KCN-sensitive oxidase is not necessarily located at the thylakoid membrane and could be located at the cytoplasmic membrane.  相似文献   

4.
Abrupt changes in the Arrhenius activation energy of membrane-bound enzymes have often been correlated with changes in the physical state of membrane phospholipids. Similar changes in activation energy have also been found in soluble enzymes. The possibility exists, therefore, that in some of the membrane-bound enzymes the changes might reflect intrinsic changes of the proteins independent of changes in the membrane phospholipids. This hypothesis was investigated using Drosophila mitochondria isolated from wild type and the mutant Ocdts-1. In this mutant it has been shown that succinate-cytochrome c reductase exhibits a change in Arrhenius activation energy at 18°C which is not found in the wild type (Sondergaard, L., Nielsen, N.C. and Smillie, R.M. (1975) FEBS lett. 50, 126–129). A quantitative thin-layer chromatographic analysis of mitochondrial phospholipids showed sphingomyelin to be more abundant in the wild type than in the mutant (5.2% and 4.3% of the total phospholipids, respectively). Since it was shown that the succinate-cytochrome c reductase had a lipid requirement for full activity, reciprocal rebinding experiments were done. These experiments showed that the reconstituted membranes exhibited the change in activation energy at 18°C only when the protein moiety came from mutant mitochondria, that is, the change was independent of the source of the phospholipids used.  相似文献   

5.
6.
Strain NR70, a mutant of E. coli lacking the Mg2+-adenosine triphosphatase (E.C. 3.6.1.3.) was previously shown to be defective in amino acid and sugar transport in whole cells and right-side-out membrane vesicles. It is shown here that the mutant is also deficient in the uptake of calcium into inverted membrane vesicles. Treatment of inverted vesicles from the wild-type strain with ethylenediamine tetraacetate removes the Mg2+-adenosine triphosphatase and results in an inability to transport calcium. Addition of a crude fraction containing the wild-type Mg2+-adenosine triphosphatase restores active uptake of calcium both to vesicles from the mutant and depleted vesicles from the wild-type.  相似文献   

7.
The cell envelope of Escherichia coli was examined for changes during late stages of bacteriophage T4 infection. Late events in T4 infection are shown to result in (i) a reduction in the effectiveness of membrane separation procedures employing either isopycnic sucrose gradient centrifugation or selective solubilization of inner membrane by detergent (Sarkosyl or Triton X-100), (ii) the appearance of a 54 000 dalton host protein in membrane preparations, (iii) the adventitious presence of detergent-resistant phage morphogenetic structures in membrane preparations, and (iv) a decrease in the activity of NADH oxidase and an apparent alteration in its association with inner membrane. These modifications occur regardless of the state of the e and t genes of T4.  相似文献   

8.
Phospholipase A2 present in a highly purified, potently bactericidal, fraction from rabbit granulocytes produces net bacterial phospholipid degradation during killing of a phospholipase A-less strain of Escherichia coli. In the wild-type parent strain phospholipid breakdown is caused not only by the action of phospholipase A2 but also by phospholipase A1, indicating activation of the most prominent phospholipase of E. coli. This activation occurs as soon as the bacteria are exposed to the granulocyte fraction. Phospholipid breakdown by both phospholipases A is dose dependent but reaches a plateau after 30–60 min and at higher concentrations of the fraction.Phospholipid degradation is accompanied in both strains by an increase in permeability to actinomycin D that is also dose dependent. Even though net hydrolysis of phospholipids is greater in the parent strain than in the mutant, the increase in permeability is the same in the two strains.The addition of 0.04 M Mg2+, after the effects on phospholipids and permeability have become manifest, initiates in both strains the restoration of insensitivity to actinomycin D, the net resynthesis of phospholipids, and the disappearance of monoacylphosphatides and the partial disappearance of free fatty acids that had accumulated. Loss of ability to multiply is not reversed by Mg2+ in either strain. Less than 5 μg of granulocyte fraction causes loss of viability of from 90 to 99% of 1 × 108 microorganisms of both strains. However, at lower concentrations the parent strain is considerably more sensitive to the bactericidal effect of the granulocyte fraction than the mutant strain.  相似文献   

9.
The dependence of the replication of several plasmids on the chromosome-determined initiation products, dna A and dna C, has been studied. The initiation of the replication of Col E1 DNA requires the chromosomal dna A product. In contrast two de-repressed transfer factors (R 1 drd 16 and Hly152) seem to determine a corresponding plasmid-specific factor. The dna C-product is necessary for the ordered initation of all plasmids studied. The addition of low concentrations of chloramphenicol leads to a relaxed replication of Col E1 DNA at the restrictive temperature in dna A-mutants, but not in dna C-mutants.  相似文献   

10.
Incubation of rabbit kidney microsomes with pig pancreatic phospholipase A2 produced residual membrane preparations with very low (Na+ + K+)-ATPase activity. The activity could be restored by recombination with lipid vesicles of negatively-charged glycerophospholipids. Vesicles of pure phosphatidylcholine and phosphatidylethanolamine were virtually inactive in this respect, but could reactivate in the presence of cholate.Incubation of the microsomes with a combination of phospholipase C (Bacillus cereus) and sphingomyelinase C (Staphylococcus aureus) resulted in 90–95% release of the phospholipids. The residual membrane contained only phosphatidylinositol and still showed 50–100% of the (Na+ + K+)-ATPase activity.  相似文献   

11.
Specific rabbit antisera against purified Hordeum vulgare seedling RNase I from two winter barley cultivars each formed a single precipitin band when reacted with the homologous crude tissue extract. RNase antigen from either cultivar was equally reactive with both antisera when evaluated by immunodiffusion and immunoelectrophoresis. A small but consistent difference in anti-RNase specificity between cultivars was shown by passive hemagglutination inhibition, suggesting that molecular differences may exist between the two RNase antigens. Immunodiffusion and rocket immunoelectrophoresis were used to qualitatively test the cross-reactivity of protein preparations from various members of the genus Hordeum and species from other related grass genera. Neither antiserum showed cross-reactivity with soluble protein preparation from species outside the genus Hordeum. A few species within the genus Hordeum were cross-reactive. A modification of rocket immunoelectrophoresis was developed to determine the amount of RNase in unpurified tissue extracts. The technique involved a template-reservoir which allowed detection of 250 ng RNase in tissue extract volumes of 50 μl. The amount of RNase in unpurified protein extracts from the two cultivars of barley was similar.  相似文献   

12.
We have determined the nucleotide sequence of a secondary λ attachment site in proAB, a site that accounts for 3% of lysogens isolated from Escherichia coli strains deleted for the primary site. Direct sequence analysis of the transducing bacteriophages carrying the left and right att junctions, as well as the recombinant pro+ phage reveals that the proAB site shares an 11-nucleotide interrupted homology with the core sequence of the primary site. We have compared the proABatt site with other secondary attachment sites to gain insights into the structural features important for λ integration.  相似文献   

13.
Phospholipase A2 activity in islet cell homogenates and dispersed islet cells of the rat was determined using an exogenous radiolabeled phospholipid substrate from E.coli membranes. Phospholipase A2 activity in islet homogenates was found to have two pH optima in acid or neutral/alkaline pH ranges. The enzyme activity at pH 7.5 was calcium dependent and responded to increasing calcium concentrations with graded increases in phospholipid hydrolysis. Preincubation of islets with a concentration of glucose known to elicit maximum rates of insulin secretion resulted in a stable activation of phospholipase A2 activity which was assayable in islet homogenates. Glucose stimulated phospholipase A2 in these preparations by as much as 220% above control. 2-Deoxy-D-glucose, a nonsecretory analogue of glucose, did not elicit a significant increase in islet phospholipase A2 activity. The glucose sensitive enzyme was associated with a membrane-enriched subcellular fraction in which the glucose-stimulated activity was greater than 2-fold higher than control activity. Glucose stimulation potentiated the phospholipase A2 activity measured in the presence of high calcium concentrations. Phospholipase A2 activity was also found in dispersed islet cell preparations where glucose stimulation of what may be a partly externalized membrane enzyme was most apparent at low calcium concentrations. These data indicate that islet cells possess phospholipase A2 activity which may be in part localized to the plasma membrane as well as other membrane systems, and which exhibits the characteristic properties of pH and calcium dependency, and sensitivity to secretagogue stimulation reported for the enzyme in other secretory systems.  相似文献   

14.
In a dnaCts mutant of E. coli, the reinitiation of DNA synthesis, which occurred by the shift of the culture from a restrictive temperature to a permissive temperature, was markedly prevented by habakacin, dibekacin, kanamycin, and gentamicin. On the contrary, chloramphenicol did not inhibit the reinitiation synthesis for 30 min. In a parallel experiment, leucine uptake into protein was profoundly blocked by chloramphenicol, but only slightly by habekacin. Habekacin did not significantly affect DNA elongation of the cells at a restrictive temperature. We propose that inhibition of initiation of replication by aminoglycoside antibiotics is related to their lethality.  相似文献   

15.
Purified polyhedron proteins and purified, ultrasonicated virions of four nuclear polyhedrosis viruses (NPVs), separable into two morphologic groups of singly and multiply embedded virion types (SEVs and MEVs), were investigated by immunodiffusion and immunoelectrophoresis. The four viruses were Pseudoplusia includens SEV, Trichoplusia ni SEV, T. ni MEV, and Autographa californica MEV. In immunodiffusion, SEV polyhedron proteins formed two precipitin bands with antiserum to SEV polyhedron proteins, while MEV polyhedron proteins formed only one. All four proteins formed one precipitin band with antiserum to MEV polyhedron protein, with a spur between SEV and MEV proteins. In immunoelectrophoresis, mobilities of SEV proteins were significantly different from those of MEVs. Precipitin arc patterns were similar to those in immunodiffusion when electrophoresis was carried out at 4 C; at room temperature, a single arc of precipitation formed with all four proteins. SEV virions formed five possibly identical precipitin bands in immunodiffusion with antiserum to SEV virions. MEV virions formed three possibly identical precipitin bands when reacted with antiserum to MEV virions. Little or no cross-reactions were observed between SEV and MEV virions or between virions and polyhedron proteins. In immunoelectrophoresis, SEV virions formed three precipitin arcs in reactions with SEV antisera and none with MEV antisera; MEV virions formed two arcs with MEV antisera and none with SEV antisera. When antisera were subjected to electrophoresis, five arcs were formed by SEVs and three by MEVs in homologous systems, and none were formed in heterologous systems.  相似文献   

16.
Previous communications from this laboratory have indicated that there exists a thiamine-binding protein in the soluble fraction of Saccharomyces cerevisiae which may be implicated to participate in the transport system of thiamine in vivo.In the present paper it is demonstrated that both activities of the soluble thiamine-binding protein and thiamine transport in S. cerevisiae are greatest in the early-log phase of the growth and decline sharply with cell growth. The soluble thiamine-binding protein isolated from yeast cells by conventional methods containing osmotic shock treatment appeared to be a glycoprotein with a molecular weight of 140 000 by sodium dodecyl sulfate polyacrylamide gel electrophoresis. The apparent Kd of the binding for thiamine was 29 nM which is about six fold lower than the apparent Km (0.18 μM) of thiamine transport. The optimal pH for the binding was 5.5, and the binding was inhibited reversibly by 8 M urea but irreversibly by 8 M urea containing 1% 2-mercaptoethanol. Several thiamine derivatives and the analogs such as pyrithiamine and oxythiamine inhibited to similar extent both the binding of thiamine and transport in S. cerevisiae, whereas thiamine phosphates, 2-methyl-4-amino-5-hydroxymethylpyrimidine and O-benzoylthiamine disulfide did not show similarities in the effect on the binding and transport in vivo. Furthermore, it was demonstrated by gel filtration of sonic extract from the cells that a thiamine transport mutant of S. cerevisiae (PT-R2) contains the soluble binding protein in a comparable amounts to that in the parent strain, suggesting that another protein component is required for the actual translocation of thiamine in the yeast cell membrane. On the other hand, the membrane fraction prepared from S. cerevisiae showed a thiamine-binding activity with apparent Kd of 0.17μM at optimal pH 5.0 which is almost the same with the apparent Km for the thiamine transport system. The membrane-bound thiamine-binding activity was not only repressible by exogenous thiamine in the growth medium, but as well as thiamine transport it was markedly inhibited by both pyrithiamine and O-benzoylthiamine disulfide. In addition, it was found that membrane fraction prepared frtom PT-R2 has the thiamine-binding activity of only 3% of that from the parent strain of S. cerevisiae.These results strongly suggest that membrane-bound thiamine-binding protein may be directly involved in the transport of thiamine in S. cerevisiae.  相似文献   

17.
Tsx mutants of Escherichia coli are fully resistant to a set of T6-like bacteriophage and are resistant to colicin K. We demonstrate that these mutants are missing an outer membrane protein (the tsx-protein) of molecular weight 32,000 as measured by SDS-polyacrylamide gel electrophoresis. Tsx mutants are receptor mutants which are unable to absorb either the bacteriophages or the colicin and the loss of receptor function can be demonstrated using outer membrane preparations.We suggest that the tsx-protein is the receptor for both the bacteriophage and colicin.  相似文献   

18.
Comparative measurements were made of the fluidity of chloroplast thylakoids, total membrane lipids and polar lipids utilizing the order parameter and motion of spin labels.No significant differences were found in the fluidity of membranes or total membrane lipids from a wild type and a mutant barley (Hordeum vulgare chlorina f2 mutant) which lacks chlorophyll b and a 25 000 dalton thylakoid polypeptide. Redistribution of intrinsic, exoplasmic face (EF) membrane particles by unstacking thylakoid membranes in low salt medium also had no effect on membrane fluidity. However, heating of isolated thylakoids decreased membrane fluidity.The fluidity of vesicles composed of membrane lipids is much greater than that of the corresponding membranes. Fluidity of the membranes, however, increased during greening indicating that the rigidity of the membranes, compared with that of total membrane lipids, is not caused by chlorophyll or its associated peptides. It is concluded that the restriction of motion in the acyl chains in the thylakoids is not caused by chlorophyll or the major intrinsic polypeptide but by some other protein components.  相似文献   

19.
Bacterial membranes are complex mixtures of lipids and proteins, the combination of which confers biophysical properties that allows cells to respond to environmental conditions. Carotenoids are sterol analogs that are important for regulating membrane dynamics. The membrane of Pantoea sp. YR343 is characterized by the presence of the carotenoid zeaxanthin, and a carotenoid-deficient mutant, ΔcrtB, displays defects in root colonization, reduced secretion of indole-3-acetic acid, and defects in biofilm formation. Here we demonstrate that the loss of carotenoids results in changes to the membrane lipid composition in Pantoea sp. YR343, including increased amounts of unsaturated fatty acids in the ΔcrtB mutant membranes. These mutant cells displayed less fluid membranes in comparison to wild type cells as measured by fluorescence anisotropy of whole cells. Studies with artificial systems, however, have shown that carotenoids impart membrane rigidifying properties. Thus, we examined membrane fluidity using spheroplasts and vesicles composed of lipids extracted from either wild type or mutant cells. Interestingly, with the removal of the cell wall and membrane proteins, ΔcrtB vesicles were more fluid than vesicles made from lipids extracted from wild type cells. In addition, carotenoids appeared to stabilize membrane fluidity during rapidly changing temperatures. Taken together, these results suggest that Pantoea sp. YR343 compensates for the loss of carotenoids by changing lipid composition, which together with membrane proteins, results in reduced membrane fluidity. These changes may influence the abundance or function of membrane proteins that are responsible for the physiological changes observed in the ΔcrtB mutant cells.  相似文献   

20.
We separated the two minor subunits (δ and ε) of the E. coli ATPase from the major subunits (α, β, and γ). The minor subunit fraction was obtained by treating purified ATPase with pyridine following the procedure that Nelson et al. (J. Biol. Chem. 348, 2049 [1973]) used to separate the subunits of chloroplast ATPase. The minor subunit fraction restored the capacity of ATPase lacking the delta subunit to recombine with ATPase-depleted membrane vesicles and to reconstitute energy coupling to the transhydrogenase and oxidative phosphorylation in the vesicles. These results clearly implicate the delta subunit in the attachment of the ATPase to the membrane.  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