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.
Spleen cells from mice infected with the rough Brucella melitensis strain B115 were fused with NSO myeloma cells. Hybridoma supernatants were screened in ELISA with cell walls (CW), sonicated cell extracts (CE) and rough lipopolysaccharide (R-LPS) of B. melitensis strain B115 and whole B. melitensis B115 cells. Surprisingly, 22 monoclonal antibodies (mAbs) reacting in ELISA with both CW and CE but not with R-LPS and bacterial cells were shown by immunoblot analysis and ELISA to react with smooth lipopolysaccharide (S-LPS). These mAbs also reacted in ELISA with O polysaccharides (OPS) from the smooth Brucella abortus strain 99 and the smooth B. melitensis strain 16M and thus recognize epitopes present on the O-chain. Proteinase K LPS preparations from B. melitensis B115 analysed by immunoblotting with one mAb (12G12) recognizing S-LPS of both A and M specificity displayed the typical S-LPS high-molecular-mass ladder pattern but no S-LPS was detected in the phenol/water/chloroform/light petroleum LPS preparation of the same strain. mAb 12G12, specific for S-LPS, and a mAb (A68/03F03/D05) specific for R-LPS were used to localize the O-chain and R-LPS expressed in B. melitensis strain B115 by immunoelectron microscopy. Immunogold labelling was observed at the surface of B. melitensis B115 cells with the anti-R-LPS mAb but not with the anti-S-LPS mAb. In ultrathin sections, immunogold labelling with the S-LPS specific mAb was observed in the cytoplasm and in the periphery of the cytoplasm, probably at the cytoplasmic membrane.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.

Background

Recognition of lipopolysaccharide (LPS) is required for effective defense against invading gram-negative bacteria. Recently, in vitro studies revealed that CD14 is required for activation of the myeloid differentiation factor (MyD)88-dependent Toll-like receptor (TLR)4 signaling pathway by smooth (S)-LPS, but not by rough (R)-LPS. The present study investigated the role of CD14 in induction of lung inflammation in mice by these different LPS chemotypes.

Methodology/Results

Neutrophil accumulation and tumor necrosis factor (TNF) release in bronchoalveolar lavage fluid were determined 6 hours after intranasal treatment of wild type (WT) and CD14 knock-out (KO) mice with different doses S-LPS or R-LPS. The contribution of CD14 to lung inflammation induced by S-LPS or R-LPS depended on the LPS dose. At low doses, S-LPS and R-LPS induced neutrophil influx in a CD14-dependent manner. Low dose S-LPS-induced cytokine release also depended on CD14. Strikingly, neutrophil influx and TNF release induced by high dose S-LPS or R-LPS was diminished in the presence of CD14. Intranasal administration of sCD14 to CD14 KO mice treated with S-LPS partially reversed the inflammatory response to the response observed in WT mice.

Conclusions

In conclusion, CD14 modulates effects of both S-LPS and R-LPS within the lung in a similar way. Except for R-LPS-induced TNF release, S-LPS and R-LPS at low dose induced acute lung inflammation in a CD14-dependent manner, while the inflammatory response triggered by high dose S-LPS or R-LPS was diminished by CD14.  相似文献   

3.
The chemical structure of free lipid A isolated from rough- and smooth-form lipopolysaccharides (R-LPS and S-LPS, respectively) of the human gastroduodenal pathogen Helicobacter pylori was elucidated by compositional and degradative analysis, nuclear magnetic resonance spectroscopy, and mass spectrometry. The predominant molecular species in both lipid A components are identical and tetraacylated, but a second molecular species which is hexaacylated is also present in lipid A from S-LPS. Despite differences in substitution by acyl chains, the hydrophilic backbone of the molecules consisted of beta(1,6)-linked D-glucosamine (GlcN) disaccharide 1-phosphate. Because of microheterogeneity, nonstoichiometric amounts of ethanolamine-phosphate were also linked to the glycosidic hydroxyl group. In S-LPS, but not in R-LPS, the hydroxyl group at position 4' was partially substituted by another phosphate group. Considerable variation in the distribution of fatty acids on the lipid A backbone was revealed by laser desorption mass spectrometry. In tetraacyl lipid A, the amino group of the reducing GlcN carried (R)-3-hydroxyoctadecanoic acid (position 2), that of the nonreducing GlcN carried (R)-3-(octadecanoyloxy)octadecanoic acid (position 2'), and ester-bound (R)-3-hydroxyhexadecanoic acid was attached at position 3. Hexaacyl lipid A had a similar substitution by fatty acids, but in addition, ester-bound (R)-3-(dodecanoyloxy)hexadecanoic acid or (R)-3(tetradecanoyloxy)hexadecanoic acid was attached at position 3'. The predominant absence of ester-bound 4'-phosphate and the presence of tetraacyl lipid A with fatty acids of 16 to 18 carbons in length differentiate H. pylori lipid A from that of other bacterial species and help explain the low endotoxic and biological activities of H. pylori LPS.  相似文献   

4.
Lipopolysaccharide (LPS) is the major pathogen-associated molecular pattern of Gram-negative bacterial infections, and includes smooth (S-LPS) and rough (R-LPS) chemotypes. Upon activation by LPS through CD14, TLR4/MD-2 heterodimers sequentially induce two waves of intracellular signaling for macrophage activation: the MyD88-dependent pathway from the plasma membrane and, following internalization, the TRIF-dependent pathway from endosomes. We sought to better define the role of scavenger receptors CD36 and CD204/SR-A as accessory LPS receptors that can contribute to pro-inflammatory and microbicidal activation of macrophages. We have found that CD36 differently regulates activation of mouse macrophages by S-LPS versus R-LPS. The ability of CD36 to substitute for CD14 in loading R-LPS, but not S-LPS onto TLR4/MD-2 allows CD14-independent macrophage responses to R-LPS. Conversely, S-LPS, but not R-LPS effectively stimulates CD14 binding to CD36, which favors S-LPS transfer from CD14 onto TLR4/MD-2 under conditions of low CD14 occupancy with S-LPS in serum-free medium. In contrast, in the presence of serum, CD36 reduces S-LPS binding to TLR4/MD-2 and the subsequent MyD88-dependent signaling, by mediating internalization of S-LPS/CD14 complexes. Additionally, CD36 positively regulates activation of TRIF-dependent signaling by both S-LPS and R-LPS, by promoting TLR4/MD-2 endocytosis. In contrast, we have found that SR-A does not function as a S-LPS receptor. Thus, by co-operating with CD14 in both R- and S-LPS loading onto TLR4/MD-2, CD36 can enhance the sensitivity of tissue-resident macrophages in detecting infections by Gram-negative bacteria. However, in later phases, following influx of serum to the infection site, the CD36-mediated negative regulation of MyD88-dependent branch of S-LPS-induced TLR4 signaling might constitute a mechanism to prevent an excessive inflammatory response, while preserving the adjuvant effect of S-LPS for adaptive immunity.  相似文献   

5.
Investigation of ability of Francisella tularensis S- and R-lypopolysaccharide (LPS) preparations as well as the live bacteria with different chemotypes to interact with human lypopolysaccharide-binding protein (LBP) was carried out. It was found that LPS preparations derived from virulent(S-LPS) or isogenic avirulent mutant (R-LPS) strains of F. tularensis had markedly lower affinity to LBP as compared with typical S-LPS of Salmonella abortus and R-LPS of Yersinia pestis. It was shown that R-LPS preparation from avirulent mutant binds LPB more effectively than S-LPS from F. tularensis virulent strain. Differences in S- and R-LPS affinity were also confirmed for LPS represented by the live cells. Thus, bacteria with S-chemotype of LPS (F. tularensis 15/10) bound only 20.3% of LBP, whereas cells with R-LPS (F. tularensis 543 cap(-)) bound 39.9%. Such pattern was observed in experiments with both normal non-immune human serum and sera from people immunized with live tularemia vaccine. The latter indicates that opsonization of LPS by specific antibodies does not change its affinity to LBP. The observed more efficient binding of avirulent strain R-LPS to LBP is likely determines the more intensive host response directed to destruction and rapid elimination of the causative agent. At the same time, weak affinity of the vaccine and virulent strains S-LPS to LBP probably allows the bacterium to avoid activation of host defense mechanisms thus contributing to its long-term persistence in microorganism and development of specific immunity against tularemia.  相似文献   

6.
The interaction of Yersinia pseudotuberculosis porin solubilized in deoxycholate with the S- and R-forms of endogenous lipopolysaccharide (LPS) was studied by the quenching of intrinsic protein fluorescence. The samples of S-LPS differed both in the length of O-specific polysaccharide (n = 1 and 4) and in the acylation degree of the 3-hydroxytetradecanoic acid residues of the lipid A moiety (12-66%). R-LPS (12%) binding to porin was found to occur with positive cooperativity on two integrated structural regions of the R-LPS macromolecule, namely, core oligosaccharide and lipid A. The mode of porin interaction with low-acylated S-LPSs (15 or 20%) coincided with a model involving three types of binding sites. The shape of Scatchard curves of binding indicates that a complex formation between porin and low-acylated S-LPS is cooperative at low and moderate ligand concentration, whereas at near-saturating LPS concentrations porin binds to LPS independently on two types of binding sites. The O-specific polysaccharide chain in the S-LPS macromolecule increases the affinity of its interaction with porin in comparison with R-LPS–porin binding. A significant increase (to 66%) in the degree of S-LPS acylation substantially changed its porin-binding character: the process becomes anti-cooperative with lowered affinity. Thus, the features of LPS–porin interaction significantly depend on the conformational changes in the LPS molecule due to expanding of its hydrophobic region.  相似文献   

7.
Water extract and salt-EDTA extract of Pseudomonas syringae, pv. atrofaciens cells were fractionated by ultracentrifugation with following salting out of ultracentrifugal supernatant by ammonium sulphate at 55% saturation (pH 4.5). The composition and distribution of LPS molecular forms were studied in the obtained fractions by means of electrophoresis in 10% polyacrylamide gel with 1% sodium dodecylsulphate when staining gels by silver nitrate and cumassi. It is shown that ultracentrifugal supernatant and a sediment as well as sulphate sediment contain S-LPS and R-LPS. SR-LPS is not differentiated. Sulphate supernatant does not contain the determinable amount of S-LPS but it is enriched by the proteins with molecular weights of 65-15 kDalton. S-LPS is localized in the gel area which corresponds to mobilities of polypeptides with molecular weights 130-45 KDalton and the number of monomeric links in O-specific chains of its molecules reaches 25-30. R-LPS migrates under electrophoresis in gel to the mobility zone of polypeptides with molecular weights 14.5-16 kDalton.  相似文献   

8.
Interaction of the pore-forming protein (porin) from Yersinia pseudotuberculosis with S- and R-forms of the endogenous lipopolysaccharide (LPS) was studied at various ionic strengths (20-600 mM NaCl), concentrations of divalent cations (5-100 mM CaCl2, MgCl2), and pH values from 3.0 to 9.0. The interaction of the R-LPS with porin has been shown in all experimental conditions to be in consensus with the model suggesting binding at independent sites of two types. S-LPS binds to interacting sites of relatively high affinity and to independent sites of low affinity at all pH values examined and at low NaCl concentration. The cooperative interaction of the S-LPS and porin is not observed at high ionic strength and in divalent cation-free medium. The number of binding sites of porin and association constants (Ka) for both LPS forms decrease significantly on increasing the solution ionic strength. The Ka values for the R- and S-LPS change oppositely on changing the pH: the Ka value for the R-LPS is maximal (Ka = 6.7 x 10(5) M-1), but that for S-LPS is minimal (Ka = 0.4 x 10(5) M(-1) at pH 5.0-5.5. The number of high-affinity and low-affinity binding sites for both LPS forms is maximal at pH 5.0-5.5. In this case, the numbers of high- and low-affinity sites for R-LPS are 3 and 10, respectively, and those for the S-LPS are 7 and 20, respectively. These data suggest an important role of electrostatic interactions on binding of LPS to porin. The contribution of conformational changes of the ligand and protein and hydrophobic interactions are discussed.  相似文献   

9.
The lipopolysaccharides (LPS) of a rough (R) and a smooth (S) strain of Pseudomonas syringae pv. phaseolicola were analysed. The S-LPS revealed markedly more rhamnose and fucose, but less glucose, than the R-LPS. The presence of 3-O-methyl-rhamnose (acofriose) in the S-LPS was confirmed by cochromatography with authentic acofriose. SDS polyacrylamide gel electrophoresis of the S-LPS demonstrated a cluster of regularly spaced high molecular weight fractions, which was almost lacking in the R-LPS. The main fatty acids of the lipid A of both LPS species were 3-OH-10:0,3-OH-12:0,2-OH-12:0, and 12:0. Two N-linked diesters were demonstrated: 3-O(12:0)-12:0 and 3-O(2-OH-12:0)-12:0. S-LPS was subjected to mild hydrolysis and the degraded polysaccharide separated into three fractions by gel permeation chromatography on a Fractogel TSK HW-50 column. Fraction I, representing nearly only the O-specific side chain, consisted of rhamnose and fucose in a molar ratio of 4:1, with 4% of the rhamnose being 3-O-methylated (acofriose). Fraction II, representing mostly core material, was composed of glucose, rhamnose, heptose, glucosamine, galactosamine, alanine, and a still unidentified amino compound, in an approximate molar ratio of 3:1:1:1:1:1:1, and KDO. Fraction III consisted of released monomers and salts. The LPS was highly phosphorylated (3.28% phosphorus in the core fraction). The thus characterized composition of the LPS O-chain seems to be unique for the pathovar phaseolicola of P. syringae, although many similarities exist to other pathovars as well as to other bacterial species.Abbreviations LPS lipopolysacchairdes - GC/MS combined gas liquid chromatography-mass spectrometry - HVE high voltage electrophoresis - KDO 2-keto-3-deoxyoctonic acid - PAGE polyacrylamide gel electrophoresis - SDS sodium dodecylsulfate P.s. pv. phaseolicola is termed P. phaseolicola in the text  相似文献   

10.
The Escherichia coli MacAB-TolC transporter has been implicated in efflux of macrolide antibiotics and secretion of enterotoxin STII. In this study, we found that purified MacA, a periplasmic membrane fusion protein, contains one tightly bound rough core lipopolysaccharide (R-LPS) molecule per MacA molecule. R-LPS was bound specifically to MacA protein with affinity exceeding that of polymyxin B. Sequence analyses showed that MacA contains two high-density clusters of positively charged amino acid residues located in the cytoplasmic N-terminal domain and the periplasmic C-terminal domain. Substitutions in the C-terminal cluster reducing the positive-charge density completely abolished binding of R-LPS. At the same time, these substitutions significantly reduced the functionality of MacA in the protection of E. coli against macrolides in vivo and in the in vitro MacB ATPase stimulation assays. Taken together, our results suggest that R-LPS or a similar glycolipid is a physiological substrate of MacAB-TolC.  相似文献   

11.
Abstract Smooth lipopolysaccharide (S-LPS) of Brucella abortus was dissociated from noncovalently attached components by differential centrifugation and exhaustive treatment with guanidinium thiocyanate. Mild alkaline hydrolysis then released ester-linked fatty acids as well as native hapten (NH)-protein complexes. These complexes, comprising 15 to 20% by weight of the S-LPS, failed to dissociate in guanidinium thiocyanate or in boiling SDS, suggesting covalent attachment. Analyses for 2-keto-3-deoxyoctonate demonstrated that the released carbohydrate was NH, rather than degraded O -polysaccharide or intact S-LPS monomers. This provides strong evidence that NH-protein complexes are covalently linked to S-LPS, most likely through ester bonds.  相似文献   

12.
Brucella spp. are facultative intracellular pathogens that have the ability to survive and multiply in professional and non-professional phagocytes, and cause abortion in domestic animals and undulant fever in humans. Several species are recognized within the genus Brucella and this classification is mainly based on the difference in pathogenicity and in host preference. Brucella strains may occur as either smooth or rough, expressing smooth LPS (S-LPS) or rough LPS (R-LPS) as major surface antigen. This bacterium possesses an unconventional non-endotoxic lipopolysaccharide that confers resistance to anti-microbial attacks and modulates the host immune response. The strains that are pathogenic for humans (B. abortus, B. suis, B. melitensis) carry a smooth LPS involved in the virulence of these bacteria. The LPS O-chain protects the bacteria from cellular cationic peptides, oxygen metabolites and complement-mediated lysis and it is a key molecule for Brucella survival and replication in the host. Here, we review i) Brucella LPS structure; ii) Brucella genome, iii) genes involved in LPS biosynthesis; iv) the interaction between LPS and innate immunity.  相似文献   

13.
The influence of lipopolysaccharide (LPS) O-polysaccharide chain production on the colonisation ability of Helicobacter pylori in four mouse models (NMRI, C57BL/6, CBA/Ca, and BALB/cA mice) was studied. H. pylori strains that produced smooth-form LPS (S-LPS) detectable in silver-stained electrophoretic gels colonised mice. In contrast, a laboratory-passaged strain G50 and the culture collection strain CCUG 17874 did not colonise mice; the former strain produced low amounts of O-chains only detectable in immunoblotting but not in silver-stained gels, whereas the latter produced rough-form LPS (R-LPS) without O-chains. Furthermore, a galE isogenic mutant, which produced R-LPS, did not colonise mice. However, after repeated broth culture, strains G50 and CCUG 17874 produced S-LPS detectable in silver-stained gels and were capable of colonising mice. Consistent with the production of O-chains, all colonising strains produced Lewis (Le) antigens, Le(x) and/or Le(y). Except for low expression of Le(y) by non-colonising G50, reflecting low production of O-chains, all other non-colonising strains and the galE mutant lacked expression of Le antigens consistent with their production of R-LPS. Lectin typing of strains supported these findings, and also showed that lectin types did not differ before and after colonisation. The low level of O-chain production and Le antigen expression by the non-colonising G50 may not be sufficient to aid colonisation. Examination of protein profiles of H. pylori strains before inoculation showed that protein expression was not significantly different between colonising and non-colonising strains. These results show that S-LPS production with O-chain expression is required by H. pylori for colonisation in a number of mouse models and that care should be taken with inoculating H. pylori strains that loss of O-chains does not occur during subculturing.  相似文献   

14.
A monoclonal antibody (3D6) was produced which reacted only with Brucella sonicated cell extracts that had been lysozyme-treated after sonication. The monoclonal antibody (mAb) reacted with the three major outer-membrane proteins (OMPs) of B. melitensis B115 in Western blots. A large number of reactive bands ranging from 12 to 43 kDa were present in lysozyme-treated Escherichia coli and Yersinia enterocolitica sonicated cell extracts. In a latex agglutination inhibition immunoassay, mAb 3D6 showed better reactivity with purified peptidoglycan (PG) of B. melitensis B115 than with that of Escherichia coli. This mAb was also used in immunogold electron microscopy with whole Brucella cells and sections. No binding was observed on whole cells and immunogold labelling in sections was observed close to the outer membrane, in the periplasmic space and in the cytoplasm. These findings indicate that mAb 3D6 is specific for PG subunits. Immunoblot analysis of B. melitensis B115 rough sonicated cell extracts after SDS-PAGE, with or without lysozyme treatment, was performed using mAbs specific for Brucella OMPs of molecular masses of 10, 16.5, 19, 25-27, 31-34, 36-38 and 89 kDa, for PG and for rough lipopolysaccharide (R-LPS) and smooth lipopolysaccharide (S-LPS). mAbs specific for the 25-27, 31-34 and 36-38 kDa OMPs reacted with three to six bands. All of them except the band of lowest molecular mass reacted with the PG-specific mAb and not with R-LPS- and S-LPS-specific mAbs.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Fatty Acids Present in the Lipopolysaccharide of Rhizobium trifolii   总被引:4,自引:2,他引:2       下载免费PDF全文
Approximately 70% of the fatty acids recovered after acid or alkaline hydrolysis of the lipopolysaccharide of Rhizobium trifolii were hydroxy fatty acids identified as hydroxymyristic and hydroxypalmitic acids. Palmitic acid was the only saturated fatty acid found in the lipopolysaccharide of R. trifolii. Octadecenoic and a small amount of hexadecenoic acids were also identified. The results of BF(3) methanolysis and hydroxylaminolysis suggest that hydroxypalmitic acid is N-acyl bound.  相似文献   

16.
Immunogenicity and protective activity against Brucella ovis of detergent-extracted recombinant Omp31 (rOmp31 extract) from Brucella melitensis produced in Escherichia coli, purified rough lipopolysaccharide from B. ovis (R-LPS) and a mixture of rOmp31 extract and R-LPS (rOmp31 extract + R-LPS) were assessed in BALB/c mice. The experimental vaccines were compared with a hot saline extract (HS extract) from B. ovis mainly composed of outer membrane proteins (OMPs) and R-LPS, and known to be protective in mice against a B. ovis infection. Serum antibodies to Omp31 and R-LPS were detected in the corresponding mice using Western blotting with B. ovis whole-cell lysates and ELISA with purified antigens. Protection was evaluated by comparing the levels of infection in the spleens of vaccinated mice challenged with B. ovis. A significantly lower number of B. ovis colony-forming units in spleens relative to unimmunized (saline injected) controls were considered as protection. Mice immunized with rOmp31 extract or rOmp31 extract mixed with R-LPS developed antibodies that bound to the B. ovis surface with similar titers. Vaccination with rOmp31 extract plus R-LPS provided the best protection level, which was comparable with that given by HS extract. Similar protection was also obtained with rOmp31 extract alone and, to a lesser degree, with R-LPS. Comparisons between groups showed that an extract from E. coli-pUC19 (devoid of Omp31) provided no protection relative to either HS extract, rOmp31 extract or rOmp31 extract mixed with R-LPS. In conclusion, the recombinant Omp31 associated or not with B. ovis R-LPS, could be an interesting candidate for a subcellular vaccine against B. ovis infection.  相似文献   

17.
Human erythrocyte membranes which had been thoroughly extracted with organic solvents contained 20 nmol of fatty acids/mg dry wt. The major fatty acids were palmitic and stearic with their monoethenoic derivatives as minor constituents. No other fatty acids were detected. When solvent-extracted membranes were digested with Pronase about 90% of the original content of fatty acids was retained in the insoluble residue. Fatty acids were linked to membrane proteins through alkali-labile bonds of which 30% were of a thiol ester and the remainder of an O-ester type. This conclusion is based on differential liberation of fatty acids by hydroxylamine at pH 7.0 and pH 11.0. Two extracts of membranes enriched in peripheral proteins (bands 1, 2, 5 and 2.1, 4.1, 4.2, 6) were prepared and extracted with organic solvents but each contained about six times less fatty acids than the parent solvent-extracted membranes. Glycophorin A contains little if any covalently bound fatty acids. Anion transporter (band 3) contains about 1 mol of thiol ester of fatty acid. This accounts for about half of the thiol ester-linked fatty acids in the parent solvent-extracted membranes. Most of the O-ester-linked fatty acids are linked to an undisclosed membrane protein.  相似文献   

18.
Temperature-responsive microsomes of the ciliate protozoan Tetrahymena have been originally fractionated by step centrifugation on two-layered, Mg2+-containing sucrose gradients. Three fractions have been obtained, which are termed smooth I, smooth II and rough according to the appearance of the membrane vesicles upon electron-microscopy. Smooth I, smooth II, and rough microsomes exhibit RNA/protein ratios of 0.09, 0.20, and 0.34; their phospholipid/protein ratios and their neutral lipid/phospholipid ratios were 0.52, 0.43 and 0.25, and 0.17, 0.18 and 0.13, respectively. All three fractions contain equivalent, low succinic dehydrogenase and 5'-nucleotidase activities. Glucose-6-phosphatase and acid phosphatase are more concentrated in smooth I membranes than in rough membranes. The reverse is true for ATPase. The smooth II membranes occupy an intermediate position except that their ATPase activity is the lowest of the three fractions. The specific activities of these enzymes of the three microsomal fractions are compared to those of homogenates of whole cells. Thin-layer chromatography reveals a very similar polar and nonpolar lipid pattern of the three microsomal fractions. The major phospholipid compounds are phosphatidlethanolamine, glycerideaminoethylphosphonate and phosphatidylcholine, while diglycerides, an unknown NL-compound, and triglycerides are the major apolar lipids. Gas liquid chromatography shows that the fatty acids are mainly even-numbered ranging between C12 and C18. The smooth I, smooth II and rough membranes contain 65.2, 69.3 and 72.7% unsaturated fatty acids in their polar lipids, whereas only 52.7, 49.7 and 48.3% unsaturated acids are found in their apolar lipids, respectively. The fatty acids are more unevenly distributed among the individual polar lipids than in the apolar ones.  相似文献   

19.
Inhibition of Proteus mirabilis growth by cerulenin, a specific inhibitor of fatty acid biosynthesis, was reversed by exogenously supplied fatty acid mixtures containing oleic acid and palmitic or pentadecanoic acids. The growth rate of the cells treated with cerulenin in the presence of the fatty acid mixtures was slower, however, than that of untreated cells, and their lipopolysaccharide content was decreased by 30-50%, resulting in an increased sensitivity of the organisms to rifamycin and vancomycin. Polyacrylamide gel electrophoresis of the lipopolysaccharide fraction from cerulenin-treated cells revealed that of the two P. mirabilis lipopolysaccharide types, the relative amount of the higher molecular weight lipopolysaccharide was reduced from 50% to 30% of the total lipopolysaccharide. Fatty acid analysis of the phospholipid and lipopolysaccharide fractions from cells grown with cerulenin, pentadecanoate, and oleate revealed that over 60% of the native even-numbered fatty acids of the phospholipid fraction was substituted by the odd-numbered fatty acid, while no incorporation of either the pentadecanoate or oleate could be demonstrated in the lipid A moiety of the lipopolysaccharide. The only change in the lipid A observed was an increase in the content of 3-hydroxymyristic acid accompanied by a decrease in the nonhydroxylated fatty acids, supporting the highly conserved nature of this molecule.  相似文献   

20.
Brucella native haptens (NHs) extracted with hot water from smooth (S)-type B. abortus and B. melitensis were purified to high levels of serological activity and compared with the polysaccharide obtained by acid hydrolysis (PS) of the S lipopolysaccharide (S-LPS). By 13C nuclear magnetic resonance analysis, NHs showed the spectrum of a homopolymer of alpha-1,2- or alpha-1,2- plus alpha-1,3-linked 4-formamido-4,6-dideoxy-D-mannose (N-formylperosamine) previously reported for the LPS O chain. However, while PS contained up to 0.6% 3-deoxy-D-manno-2-octulosonate, this LPS-core marker was absent from NH. High performance liquid chromatography and thin-layer chromatography showed heterogeneity in NH purified from whole cells but not in PS. By immunoprecipitation, polysaccharides indistinguishable from NH were demonstrated in extracts obtained with phenol-water, saline at 60 degrees C, and ether-water treatments, and none of these treatments caused S-LPS hydrolysis detectable with antibodies to the O chain and lipid A. Two lines of evidence showed that NH was in the cell surface. First, NH became biotinylated when B. abortus live cells were labelled with biotin-hydrazide, and the examination of cell fractions and electron microscopy sections with streptavidin-peroxidase and streptavidin-coloidal gold, respectively, showed that labelling was extrinsic. Moreover, whereas only traces of NH were found in cytosols, the amount of NH was enriched in cell envelopes and in the outer membrane blebs spontaneously released by brucellae during growth. Interactions between NH and S-LPS were observed in crude cell extracts, and such interactions could be reconstituted by using purified NH and LPS. The results demonstrate that NH is not a hydrolytic product of S-LPS and suggest a model in which LPS-independent O-type polysaccharides (NH) are intertwined with the O chain in the outer membrane of S-type brucellae.  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