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 34 毫秒
1.
Using 2H- and 31P-NMR techniques the effects of temperature variation and phenethyl alcohol addition were investigated on lipid acyl chain order and on the macroscopic lipid organization of membrane systems derived from cells of the Escherichia coli fatty acid auxotrophic strain K1059, which was grown in the presence of [11,11-2H2]oleic acid. Membranes of intact cells showed a gel to liquid-crystalline phase transition in the range of 4-20 degrees C, which was similar to that observed for the total lipid extract and for the dominant lipid species phosphatidylethanolamine (PE). Phosphatidylglycerol (PG) remained in a fluid bilayer throughout the whole temperature range (4-70 degrees C). At 30 degrees C acyl chain order was highest in PE, followed by the total lipid extract, PG, intact cells, and isolated inner membrane vesicles. Acyl chain order in E. coli PE and PG was much higher than in the corresponding dioleoylphospholipids. E. coli PE was found to maintain a bilayer organization up to about 60 degrees C, whereas in the total lipid extract as well as in intact E. coli cells bilayer destabilization occurred already at about 42 degrees C. It is proposed that the regulation of temperature at which the bilayer-to-non-bilayer transition occurs may be important for membrane functioning in E. coli. Addition of phenethyl alcohol did not affect the macroscopic lipid organization in E. coli cells or in the total lipid extract, but caused a large reduction in chain order of about 70% at 1 mol% of the alcohol in both membrane systems. It is concluded that while both increasing temperature and addition of phenethyl alcohol can affect membrane integrity, in the former case this is due to the induction of non-bilayer lipid structures, whereas in the latter case this is caused by an increase in membrane fluidity.  相似文献   

2.
A mutant of Escherichia coli with a thermosensitive defect, possibly in the outer membrane (omsA mutant), was isolated from E. coli K-12 by mutagenization and selection for thermosensitivity and beta-lactam supersensitivity of growth. The mutant also showed very high sensitivity to other antibiotics, such as macarbomycin, midecamycin, rifampin, and bacitracin. The mutation was recessive to the wild type and was mapped at about 4 min on the E. coli chromosome between fhuA and metD. The mutation caused rapid release into the medium of periplasmic enzymes such as RTEM penicillinase but practically no cytoplasmic enzyme when cells grown at 30 degrees C were transferred to 37 or 42 degrees C. Electron microscopic observations showed many large double-layered vesicles attached to the surface of cells incubated at 42 degrees C. We conclude that the mutant had a mutation that caused a temperature-dependent defect in the outer membrane structure or its assembly (named an oms mutation). The omsA mutant may be useful for production of periplasmic proteins, which it releases into the culture medium on shift up of temperature.  相似文献   

3.
Proton motive force is not obligatory for growth of Escherichia coli.   总被引:13,自引:6,他引:7       下载免费PDF全文
When 50 microM carbonyl cyanide-m-chlorophenyl hydrazone (CCCP), a protonophore, was added to growth medium containing glucose at pH 7.5, Escherichia coli TK1001 (trkD1 kdpABC5) started exponential growth after 30 min; the generation time was 70 min at 37 degrees C. Strain AS1 (acrA), another strain derived from E. coli K-12, also grew in the presence of 50 microM CCCP under the same conditions, except that the lag period was ca. 3 h. When this strain was grown in the presence of 50 microM CCCP and then transferred to fresh medium containing 50 microM CCCP, cells grew without any lag. Neither a membrane potential nor a pH gradient was detected in strain AS1 cells growing in the presence of CCCP. When either succinate or lactate was substituted for glucose, these strains did not grow in the presence of 50 microM CCCP. Thus, it is suggested that E. coli can grow in the absence of a proton motive force when glucose is used as an energy source at pH 7.5.  相似文献   

4.
The binding of serum C3 to the O-antigen capsule (OAg Cap), lipopolysaccharide (LPS), and outer membrane proteins (OMP) of Escherichia coli 0111B4 was examined. Bacteria were intrinsically labeled with [3H] or [14C]galactose (*gal) in the OAg Cap and LPS moieties or with [14C]leucine (*leu) to label proteins. Organisms were then incubated in serum containing differentially labeled C3, the above fractions were separated, and the proportion of each binding to a column containing anti-C3 was measured. The OAg Cap fraction bound 72 to 82% of the C3, which bound to E. coli 0111B4 during incubation in absorbed 10% pooled normal human serum (10% PNHS) or absorbed 40% C8-deficient serum (C8D). This distribution did not change when the organism was presensitized with immune IgG before serum incubation. A total of 2.93% +/- 0.48 of OAg Cap and 0.52% +/- 0.16 of LPS *gal bound specifically to Sepharose-containing antibodies to C3 (A:C3-Seph) after incubation in 10% PNHS; these values increased to 10.1% +/- 4.5 and 1.8% +/- 0.3, respectively, when C3 deposition was increased fourfold by incubation in 40% C8D. When encapsulated E. coli 0111B4 was incubated in 10% PNHS containing biotinylated C3, specific attachment of OAg Cap *gal to avidin-Sepharose was demonstrated in 1% sodium dodecyl sulfate (SDS), and complete release of bound *gal but not C3 occurred with 1 M NH2OH. When a mutant of E. coli 0111B4 lacking OAg Cap was incubated in 40% C8D, the outer membrane (OM) bound 85% of C3. Five percent of OM *gal from the unencapsulated organism bound to A:C3-Seph in 0.05% SDS, indicating that the fraction of LPS molecules with bound C3 increased threefold in the absence of OAg Cap. OAg Cap does not contain protein, and no net specific binding of *leu from OAg Cap fractions to A:C3 was detectable; 2.4 to 3.6% of OM *leu bound to A:C3-Seph. Immunoprecipitation of 82.9% of OAg Cap *gal with antisera that were directed to E. coli 0111B4 was associated with co-precipitation of 69.5% of C3 in the capsular fraction. Therefore, the majority of C3 bound to E. coli 0111B4 was covalently attached to OAg Cap and LPS. As corroboration of experiments with whole bacteria, purified OAg Cap and purified LPS consumed C3 when incubated in serum in the fluid phase. These results are the first to evaluate the acceptor site for C3 deposition on a Gram-negative organism incubated in serum, and show that LPS, OAg Cap, and OMP are all major acceptor sites for C3 in nonimmune serum.  相似文献   

5.
Outer membrane permeability of Escherichia coli O157:H7 was determined by an in vivo kinetic model with the periplasmic enzyme alkaline phosphatase [Martinez et al. (1996) Biochemistry 35, 1179-1186]. p-Nitrophenyl phosphate (PNPP) substrate, added to intact bacteria, must diffuse through the outer membrane to reach the enzyme. At low substrate concentration the bacterium was in the perfectly reactive state where all molecules that entered the periplasm were captured and converted to product. Transmembrane diffusion was rate limiting, and the permeability of the outer membrane was determined from kinetic properties. The O157:H7 strain grown at 30 degrees C showed one-sixth the permeability of wild-type E. coli grown at 30 degrees C. Wild-type bacteria grown at >/=37 degrees C show a physiological response with a shift in expression of outer membrane porins that lowered permeability to PNPP by approximately 70%. The O157:H7 strain did not display this temperature-sensitive shift in permeability even though a change in porin expression could be visualized by staining intensity of Omp F and Omp C on acrylamide gels. Altered behavior of the O157:H7 membrane was also indicated by a several thousand-fold lower response to transformation relative to wild-type E. coli. Matrix-assisted laser desorption ionization time of flight mass spectrometry and electrospray ionization mass spectrometry confirmed the expression of the Omp F and Omp C variants that are unique to E. coli O157:H7. This reduced outer membrane permeability can contribute to enhanced resistance of O157:H7 to antimicrobial agents.  相似文献   

6.
Rabbits were immunized with the enterobacterial common antigen (ECA)-immunogenic strain Escherichia coli F470. ECA-specific antiserum was obtained by absorbing the resulting antisera with the genetically closely related ECA-negative strain E. coli F1283. These two strains also served as positive and negative controls in the localization study of ECA in Yersinia enterocolitica strain 75, smooth and rough forms (Ye75S and Ye75R), by the indirect immunoferritin technique. Cells of Ye75S grown at 22 degrees C showed no labeling with ferritin after treatment with the ECA-specific antiserum and subsequent ferritin-conjugated goat anti-rabbit antibodies. If the cells were grown at 40 degrees C, however, most of the cells showed weak ferritin labeling. At this higher growth temperature, the lipopolysaccharide of this strain contains less O-specific chains (6-deoxy-L-altrose), as was shown in a previous study. The rough mutant Ye75R, which lacks O-specific chains completely, showed denser labeling with ferritin. These results indicate that ECA on the cell surface of Ye75S is covered by O-specific chains of the lipopolysaccharide if grown at 22 degrees C and is therefore not accessible to ECA antibodies. It becomes accessible, however, when O-chains are lacking (R mutants) or when they are reduced in size or amount (growth at 40 degrees C).  相似文献   

7.
We have analyzed the relationship between expression of the transformed phenotype and thyroid hormone (triiodothyronine, T3) inducibility of Na,K-ATPase and binding of 125I-epidermal growth factor (EGF) to cell membrane receptors in wild-type (wt) and mutant type 5 adenovirus (Ad5)-transformed CREF cells displaying a cold-sensitive (cs) expression of the transformed phenotype. CREF cells respond to thyroid hormone treatment with increased Na,K-ATPase activity and bind similar levels of 125I-EGF at 32 degrees C, 37 degrees C and 39.5 degrees C. In contrast, CREF cells transformed by wt Ad5 or the E1a plus E1b-transforming genes of wt Ad5 are refractile to T3 treatment and bind lower levels of 125I-EGF than CREF cells at all three temperatures. By employing a series of cloned CREF cell lines transformed by a host-range cold-sensitive mutant virus, H5hr1 or H5dl101, or the E1a or E1a plus E1b genes from these viruses, we have investigated expression of the transformed state and its relationship with hormone inducibility and EGF binding. When cs virus, cs E1a- or cs E1a plus E1b-transformed CREF clones were grown at 32 degrees C, a nonpermissive transforming temperature in which cs-transformed cells exhibit properties similar to untransformed CREF cells, T3 induced Na,K-ATPase activity and these cells bound similar levels of 125I-EGF as CREF cells. However, when cs virus- and cs Ela plus E1b-transformed CREF clones were incubated at 37 degrees C or 39.5 degrees C, temperatures at which cs-transformed cells exhibit properties similar to wt Ad5-transformed CREF cells, they did not respond to T3 and bound lower levels of 125I-EGF than CREF cells. In the case of cs E1a-transformed CREF clones, thyroid hormone responsiveness was observed at both 32 degrees C and 37 degrees C, but not at 39.5 degrees C. By performing temperature shift experiments--i.e. 32 degrees C to 37 degrees C, 32 degrees C to 39.5 degrees C, 37 degrees C to 32 degrees C, and 39.5 degrees C to 32 degrees C, it was demonstrated that after a shift from lower to higher temperature a 24-hr lag period was required for cs-transformed CREF cells to lose T3 inducibility and exhibit reduced EGF binding, whereas 96 hr after a shift from higher to lower temperature a 96-hr lag period was required for cs-transformed cells to regain T3 inducibility and increased 125I-EGF binding.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

8.
Beverly Wolf 《Genetics》1972,72(4):569-593
A temperature sensitive strain of E. coli K12 has been isolated in which residual DNA synthesis occurs at the 40 degrees C restrictive temperature; syntheses of RNA, protein and DNA precursors are not directly affected. The mutation has been designated dna-325 and is located at 89 min on the E. coli map in the same region where the dnaC locus is found. dnaC mutants are considered to be defective in DNA initiation. Some of the data are consistent with the view that the dna-325 mutation is temperature sensitive in the process of DNA initiation rather than DNA chain elongation: (1) more than two cell divisions occur after a shift to 40 degrees C; (2) upon a shift down to 30 degrees C, cell division occurs again only after the DNA content of the cells has doubled; (3) 80% more DNA is made at 30 degrees C in the presence of chloramphenicol after prior inhibition of DNA synthesis at 40 degrees C. These three observations indicate that rounds of DNA replication were completed at 40 degrees C. Also (4) infective lambda particles can be made at 40 degrees C long after bacterial DNA replication has ceased. It appears however that some DNA initiation can occur at 40 degrees C since (1) a limited amount of DNA synthesis does occur at 40 degrees C after prior alignment of the chromosomes by amino acid starvation at 30 degrees C, and (2) after incubation in bromouracil at the restrictive temperature, heavy DNA is found with both strands containing bromouracil.  相似文献   

9.
The coupling of membrane-bound glucose dehydrogenase (EC 1.1.99.17) to the respiratory chain has been studied in whole cells, cell-free extracts, and membrane vesicles of gram-negative bacteria. Several Escherichia coli strains synthesized glucose dehydrogenase apoenzyme which could be activated by the prosthetic group pyrrolo-quinoline quinone. The synthesis of the glucose dehydrogenase apoenzyme was independent of the presence of glucose in the growth medium. Membrane vesicles of E. coli, grown on glucose or succinate, oxidized glucose to gluconate in the presence of pyrrolo-quinoline quinone. This oxidation led to the generation of a proton motive force which supplied the driving force for uptake of lactose, alanine, and glutamate. Reconstitution of glucose dehydrogenase with limiting amounts of pyrrolo-quinoline quinone allowed manipulation of the rate of electron transfer in membrane vesicles and whole cells. At saturating levels of pyrrolo-quinoline quinone, glucose was the most effective electron donor in E. coli, and glucose oxidation supported secondary transport at even higher rates than oxidation of reduced phenazine methosulfate. Apoenzyme of pyrrolo-quinoline quinone-dependent glucose dehydrogenases with similar properties as the E. coli enzyme were found in Acinetobacter calcoaceticus (var. lwoffi) grown aerobically on acetate and in Pseudomonas aeruginosa grown anaerobically on glucose and nitrate.  相似文献   

10.
Escherichia coli strain NP2907 was isolated as a spontaneous mutant of strain NP29, which possesses a thermolabile valyl-transfer ribonucleic acid (tRNA) synthetase. The valyl-tRNA synthetase of the new mutant, unlike that of its immediate parent, retains enzymatic activity in vitro but differs from the wild-type enzyme in stability and apparent K(m) for adenosine triphosphate. The new mutant locus, valS-102, cotransduces with pyrB at the same frequency as does the parental locus, valS-1. Cultures of strain NP29 cease growth immediately in any medium when shifted from 30 to 40 C. The new mutant grows normally at 30 C, and upon a shift to 40 C growth quickly accelerates exactly as for normal cells. Exponential growth, however, cannot be sustained at 40 C. At a point characteristic for each medium, growth becomes linear with time. This transition occurs almost immediately in rich media and after 1.5 generations in glucose minimal medium. Net synthesis of valyl-tRNA synthetase ceases in the new mutant as soon as the temperature is raised to 40 C, irrespective of the growth medium. We conclude that it is the amount of valyl-tRNA synthetase activity that limits the rate of growth in the linear phase at 40 C. This property of the mutant makes it possible to evaluate the in vivo efficiency of this enzyme at different growth rates and thereby to determine the concentration that is necessary for a given rate of protein synthesis. The results of our measurements indicate that cells of E. coli growing in minimal medium normally possess a functional excess of valyl-tRNA synthetase with respect to protein synthesis and to repression of threonine deaminase.  相似文献   

11.
Escherichia coli O111a is a thermosensitive strain which, when grown at 40 C, accumulates large quantities of intracellular membranes. The ultrastructure of these membranes in cells which have been chemically fixed, embedded, and examined as thin sections has been compared with that of membranes in cells negatively stained or freeze-fractured. Results indicate that the extra membranes are present in the three types of preparations examined and, therefore, clearly are not artifacts of chemical fixation. Negative staining has proved also to be a valuable tool as a rapid means of monitoring cells for the accumulation of large amounts of extra membranes. Also, examination of thin sections has shown that distinct continuities between the plasma membrane and the extra membranes exist. In general, membrane surfaces in freeze-fractured cells containing extra membranes appear smooth and lack the particles associated with the plasma membranes of many cells.  相似文献   

12.
An acyltransferase induced by cold shock in Escherichia coli, designated LpxP, incorporates a palmitoleoyl moiety into nascent lipid A in place of the secondary laurate chain normally added by LpxL(HtrB) (Carty, S. M., Sreekumar, K. R., and Raetz, C. R. H. (1999) J. Biol. Chem. 274, 9677-9685). To determine whether the palmitoleoyl residue alters the properties of the outer membrane and imparts physiological benefits at low growth temperatures, we constructed a chromosomal insertion mutation in lpxP, the structural gene for the transferase. Membranes from the lpxP mutant MKV11 grown at 12 degrees C lacked the cold-induced palmitoleoyltransferase present in membranes of cold-shocked wild type cells but retained normal levels of the constitutive lauroyltransferase encoded by lpxL. When examined by mass spectrometry, about two-thirds of the lipid A molecules isolated from wild type E. coli grown at 12 degrees C contained palmitoleate in place of laurate, whereas the lipid A of cold-adapted MKV11 contained only laurate in amounts comparable with those seen in wild type cells grown at 30 degrees C or above. To probe the integrity of the outer membrane, MKV11 and an isogenic wild type strain were grown at 30 or 12 degrees C and then tested for their susceptibility to antibiotics. MKV11 exhibited a 10-fold increase in sensitivity to rifampicin and vancomycin at 12 degrees C compared with wild type cells but showed identical resistance when grown at 30 degrees C. We suggest that the palmitoleoyltransferase may confer a selective advantage upon E. coli cells growing at lower temperatures by making the outer membrane a more effective barrier to harmful chemicals.  相似文献   

13.
Escherichia coli kgtP which maps at 56.5 min codes for alpha-ketoglutarate permease (KgtP). This protein, expressed from the cloned gene using the T7 polymerase system and [35S]methionine labeling, fractionated with cell membranes. Right-side-out (RSO) membrane vesicles prepared from a kgtP negative mutant strain did not transport alpha-ketoglutarate, but RSO vesicles from the same strain expressing KgtP from a transforming plasmid transported alpha-ketoglutarate effectively as measured by uptake of the 14C-labeled substrate. E. coli JC7623 strain grown in M9 minimal medium with glucose, glycerol, or alpha-ketoglutarate as carbon source contained a 1.3-kilobase RNA which hybridized to nick-translated kgtP probe. In addition, strain MC1061 cultures grown under these same conditions were all capable of transporting alpha-ketoglutarate, demonstrating that KgtP is constitutively expressed. The Km and Vmax of KgtP assayed in strain MC1061 vesicles were 13-46 microM and 8 nmol/min/mg protein, respectively. Uncouplers that permeabilized the membrane to protons inhibited alpha-ketoglutarate transport into energized vesicles, and the addition of alpha-ketoglutarate to vesicle suspensions under non-energized conditions resulted in an increase in pH. These results indicate that KgtP is an alpha-ketoglutarate-proton symporter.  相似文献   

14.
A 15-residue peptide dimer G15 derived from the cell lytic protein granulysin has been shown to exert potent activity against microbes, including E. coli, but not against human Jurkat cells [Z. Wang, E. Choice, A. Kaspar, D. Hanson, S. Okada, S.C. Lyu, A.M. Krensky, C. Clayberger, Bactericidal and tumoricidal activities of synthetic peptides derived from granulysin. J. Immunol. 165 (2000) 1486-1490]. We investigated the target membrane selectivity of G15 using fluorescence, circular dichroism and 31P NMR methods. The ANS uptake assay shows that the extent of E. coli outer membrane disruption depends on G15 concentration. 31P NMR spectra obtained from E. coli total lipid bilayers incorporated with G15 show disruption of lipid bilayers. Fluorescence binding studies on the interaction of G15 with synthetic liposomes formed of E. coli lipids suggest a tight binding of the peptide at the membrane interface. The peptide also binds to negatively charged POPC/POPG (3:1) lipid vesicles but fails to insert deep into the membrane interior. These results are supported by the peptide-induced changes in the measured isotropic chemical shift and T1 values of POPG in 3:1 POPC:POPG multilamellar vesicles while neither a non-lamellar phase nor a fragmentation of bilayers was observed from NMR studies. The circular dichroism studies reveal that the peptide exists as a random coil in solution but folds into a less ordered conformation upon binding to POPC/POPG (3:1) vesicles. However, G15 does not bind to lipid vesicles made of POPC/POPG/Chl (9:1:1) mixture, mimicking tumor cell membrane. These results explain the susceptibility of E. coli and the resistance of human Jurkat cells to G15, and may have implications in designing membrane-selective therapeutic agents.  相似文献   

15.
Escherichia coli B strains that have acquired the malB region from E. coli K-12 are able to utilize maltose and to adsorb phage lambda when grown at 30 C, but when grown at 40 C they do not absorb phage lambda and are devoid of amylomaltase activity. These Mal(ts) Lam(ts) cells can be mutated or transduced to become able to grow on maltose at 40 C, but they still have no detectable amylomaltase activity nor functional lambda receptors at that temperature. This Mal(40) phenotype is governed by a gene located near or at malA. It is suggested that the temperature sensitivity of both characters results from a defect in malT. However, transduction of malA from E. coli B to E. coli K-12 results in a wild-type phenotype, whereas E. coli B cells that have acquired malA from E. coli K-12 donors are still temperature sensitive for both amylomaltase and lambda-receptor production.  相似文献   

16.
17.
Several versions of methods for the indirect detection of expression of staphylococcal protein A gene (spa) in Escherichia coli (E. coli) were devised by making use of biological properties of staphylococcal protein A (SpA). i) Hemagglutination of sheep red blood cells (SRBC) sensitized with anti-SRBC-antibodies using heat-treated spa-transformed E. coli organisms; Native spa-transformed E. coli organisms did not agglutinate the sensitized SRBC. The heat-treatment (60 C, 4 hr) of the transformants, however, caused positive hemagglutination like SpA-positive Staphylococcus aureus (S. aureus) organisms. ii) Halo formation around colonies on agar plates containing normal dog serum, which is originally used for the detection of SpA of S. aureus. A mutant strain NMJ was isolated, which showed formation of the halo of precipitate due to interaction between immunoglobulin and SpA. iii) A new version of immunodetection; After lysis of the transformants grown on a nitrocellulose membrane by alkali, SpA could be directly detected by immuno-detection procedures after inactivation of endogenous peroxidase in bacteria by phenylhydrazine and hydrogen peroxide.  相似文献   

18.
Antibody-independent C1 activation by E. coli   总被引:5,自引:0,他引:5  
Antibody-independent interactions of C1 with several E. coli strains were examined. Purified C1 was directly activated by the semi-rough mutant E. coli J-5, its parental wild-type strain, E. coli 0111:B4, and two clinical isolates, E. coli (P) and E. coli (A), in the absence of C1 inhibitor. E. coli J-5 activated C1 about 10-fold more rapidly and bound approximately threefold more C1 than the other strains. E. coli J-5, but not the other strains, also bound C1s2, provided that the subcomponent was offered to the bacteria in the presence of C1q and calcium; such binding was thus independent of the presence or absence of C1r2. After C1 activation in the absence of C1 inhibitor, activated C1s spontaneously dissociated from E. coli 0111:B4, (P), and (A), but remained associated with E. coli J-5. The regulatory protein C1 inhibitor prevented C1 activation by the weaker activators, E. coli strains 0111:B4, (P), and (A), but had no effect on C1 activation by E. coli J-5. Although C1 inhibitor thus failed to modulate C1 activation by E. coli J-5, it did block the enzymatic activity of activated C1 bound to this strain. Analyses of the molecular processes involved revealed differences with other systems. In the presence of C1 inhibitor, the C1s subunit of C1 activated by E. coli J-5 underwent further cleavage with the release into the supernatant of C1s fragments and complexes of C1 inhibitor with light chain fragments. Such fragments were not disulfide-linked to the remainder of the C1s molecule. The bulk of the heavy chain remained adherent to the surface of E. coli J-5. This finding documents the presence of a binding site for activated C1s on the surface of E. coli J-5 and localizes this site to the heavy chain. These studies thus indicate that several E. coli strains are direct C1 activators. Furthermore, E. coli J-5 provides another example of a direct C1 activator having binding sites not only for C1q but also for dimeric C1s. The studies also show that there are multiple properties of particles which determine the ability to activate C1, the rate of activation, the possibility of regulation of the activation process by C1 inhibitor, and the fate of activated C1.  相似文献   

19.
Bacillus licheniformis strain 749/C (constitutive for penicillinase formation) and uninduced cells of strain 749 (penicillinase-inducible) were examined after freezeetching. In the early stationary phase, strain 749/C organisms had clusters of vesicles (30 to 40 nm in diameter) on the outer surface of the plasma membrane. These are randomly distributed on the membrane, including the region of septum formation. The vesicles are not intimately associated with the plasma membrane, and their inner and outer surfaces are devoid of particles. Periplasmic vesicles were not detected by freeze-etching in strain 749 (uninduced) or in young cells of 749/C; however, the membrane of mid-logarithmic phase 749/C cells had a corrugated appearance. Negatively stained 749/C cells (logarithmic phase) also showed many vesicular and tubular bodies in the periplasm as well as septal and cytoplasmic mesosomes of typical morphology. The periplasmic structures appear to be formed either by evagination of plasma membrane or by migration of vesicular bodies from the membranous pockets of the cytoplasm. Stationary phase cells of 749/C still have many periplasmic vesicular bodies; however, the mesosomes are greatly reduced both in number and size. In sharp contrast, strain 749 organisms have very few structures similar to the periplasmic bodies of strain 749/C. These findings support our previous view that penicillinase-producing cells of 749/C have periplasmic membranous structures that are rare in the uninduced strain 749, though there is some lack of correspondence between freeze-etching, negative staining, and thin section data. These structures may be important for the retention or storage of penicillinase in the cell.  相似文献   

20.
Polarized light scattering for rapid observation of bacterial size changes.   总被引:3,自引:0,他引:3  
The effect of changing growth conditions on the diameter of rod-shaped bacteria was studied in vivo with the use of polarized light scattering. The value of a ratio of scattering matrix elements was measured as a function of scattering angle at various times after nutritional "upshift" for two strains of Escherichia coli cells. The peak locations of the scattering function were calibrated against the diameter for rod-shaped bacteria. The peaks moved toward smaller angles as a function of time after upshift, indicating that the diameter was increasing. Under special conditions, substantial peak shifts occurred within a few minutes of growth condition change, indicating a rapid onset of growth in diameter. The rate of increase of the diameters after upshift was obtained from the angular shift of peak location. This rate was approximately 14 nm/min for E. coli K12 and approximately 9 nm/min for E. coli B/r at 37 degrees C. The rate of diameter increase is smaller at lower temperatures. Experiments with Bacillus megaterium showed that any diameter change after nutritional upshift at 37 degrees C is limited to at most a very small increase, at least for the strain and medium tested.  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