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.
Mild acid hydrolysis with 1% acetic acid (100 degrees C, 15-60 min) of lipopolysaccharide (LPS) isolated from Coxiella burnetii phase I cells leads to a drastic decrease in its serological reactivity as shown by the passive hemolysis test. This decrease in reactivity occurs parallel or even prior to the cleavage of LPS into free lipid A and the polysaccharide moiety. During this mild hydrolysis two unusual sugars (X and Y) are released from the LPS, which were obtained in pure state by thin-layer chromatography. Analysis of their alditol acetate derivatives by gas chromatography/mass spectrometry revealed that sugar X is a 6-deoxy-3-C-methyl-hexose and sugar Y a 3-C-(hydroxymethyl)-pentose. Using a range of authentic standards and different thin-layer and gas chromatographic conditions, X could be recognized as 6-deoxy-3-C-methyl-gulose (virenose), very probably as the L form of this sugar (L-virenose). Y has been identified as 3-C-(hydroxymethyl)-lyxose (dihydrohydroxystreptose) by comparing it with newly synthesized 3-C-(hydroxymethyl)-pentoses (Dahlman, O., Garegg, P. J., Mayer, H., Schramek, S., unpublished results). Both branched sugars are (at least partially) in terminal positions since methylation analysis of LPS afforded (mainly) their permethylated derivatives. This analysis further showed virenose to be linked in C. burnetii phase I LPS as pyranose and dihydro-hydroxystreptose as furanose. The terminal linkage and the chemical nature of X and Y are in accordance with the observed acid-lability of the serological determinants.  相似文献   

2.
Lipopolysaccharides (LPSs) isolated from phase I and phase II Coxiella burnetii (LPS I and LPS II, respectively) were analyzed for chemical compositions, molecular heterogeneity by sodium dodecyl sulfate-polyacrylamide gel electrophoresis, and immunological properties. The yields of crude phenol-water extracts from phase I cells were roughly three to six times higher than those from phase II cells. Purification of LPSs by ultracentrifugation gave similar yields for both LPS I and LPS II. Purified LPS I and LPS II contained roughly 0.8 and 0.6% protein, respectively. The fatty acid constituents of the LPSs were different in composition and content, with branched-chain fatty acids representing about 15% of the total. beta-Hydroxymyristic acid was not detected in either LPS I or LPS II. A thiobarbituric acid-periodate-positive compound was evident in the LPSs; however, this component was not identified as 3-deoxy-D-mannooctulosonic acid by gas and paper chromatographies. LPS II contained D-mannose, D-glucose, D-glyceromannoheptose, glucosamine, ethanolamine, 3-deoxy-D-mannooctulosonic acid-like material, phosphate, and fatty acids. LPS I contained the unique disaccharide galactosaminuronyl glucosamine and nine unidentified components in addition to the components of LPS II. The hydrophobic, putative lipid A fraction of LPS I and LPS II contained the above constituents, but the hydrophilic fraction was devoid of ethanolamine. The LPS I disaccharide galactosaminuronyl glucosamine was found in both fractions of the acetic acid hydrolysates. Analysis of LPSs by sodium dodecyl sulfate-polyacrylamide gel electrophoresis followed by silver staining indicated that LPS II was composed of only one band, whereas LPS I consisted of six or more bands with irregular spacing. Ouchterlony immunodiffusion tests demonstrated that LPS I reacted with phase I but not with phase II whole-cell hyperimmune antibody, and LPS II reacted neither with phase I nor phase II hyperimmune antibody. From these results, it was concluded that the chemical structures of LPSs from C. burnetii were different from those of the LPSs of gram-negative bacteria; however, the LPS structural variation in C. burnetii may be similar to the smooth-to-rough mutational variation of saccharide chain length in gram-negative bacteria.  相似文献   

3.
Coxiella burnetii, the etiologic agent of human Q fever, is a gram-negative and naturally obligate intracellular bacterium. The O-specific polysaccharide chain (O-PS) of the lipopolysaccharide (LPS) of C. burnetii is considered a heteropolymer of the two unusual sugars β-D-virenose and dihydrohydroxystreptose and mannose. We hypothesize that GDP-D-mannose is a metabolic intermediate to GDP-β-D-virenose. GDP-D-mannose is synthesized from fructose-6-phosphate in 3 successive reactions; Isomerization to mannose-6-phosphate catalyzed by a phosphomannose isomerase (PMI), followed by conversion to mannose-1-phosphate mediated by a phosphomannomutase (PMM) and addition of GDP by a GDP-mannose pyrophosphorylase (GMP). GDP-D-mannose is then likely converted to GDP-6-deoxy-D-lyxo-hex-4-ulopyranose (GDP-Sug), a virenose intermediate, by a GDP-mannose-4,6-dehydratase (GMD). To test the validity of this pathway in C. burnetii, three open reading frames (CBU0671, CBU0294 and CBU0689) annotated as bifunctional type II PMI, as PMM or GMD were functionally characterized by complementation of corresponding E. coli mutant strains and in enzymatic assays. CBU0671, failed to complement an Escherichia coli manA (PMM) mutant strain. However, complementation of an E. coli manC (GMP) mutant strain restored capsular polysaccharide biosynthesis. CBU0294 complemented a Pseudomonas aeruginosa algC (GMP) mutant strain and showed phosphoglucomutase activity (PGM) in a pgm E. coli mutant strain. Despite the inability to complement a manA mutant, recombinant C. burnetii PMI protein showed PMM enzymatic activity in biochemical assays. CBU0689 showed dehydratase activity and determined kinetic parameters were consistent with previously reported data from other organisms. These results show the biological function of three C. burnetii LPS biosynthesis enzymes required for the formation of GDP-D-mannose and GDP-Sug. A fundamental understanding of C. burnetii genes that encode PMI, PMM and GMP is critical to fully understand the biosynthesic pathway of GDP-β-D-virenose and LPS structure in C. burnetii.  相似文献   

4.
Lipopolysaccharides (LPSs) isolated from three Kanagawa-positive and three negative strains of Vibrio parahaemolyticus were characterized by using electrophoretic, immunochemical, and chemical methods. The results of this study indicated that the LPSs of all six strains of V. parahaemolyticus examined did not have an O-specific side chain. These V. parahaemolyticus LPSs appeared to have molecular weights similar to that of the rough-type (Ra) LPS of Salmonella typhimurium TV-119 and might just contain lipid A and a core region. However, the microheterogeneity of V. parahaemolyticus LPS observed was greater than that of S. typhimurium LPS. The profile of V. parahaemolyticus LPS consisted of closely spaced triplet or quadruplet bands, but that of S. typhimurium consisted of doublet bands. Slower-moving bands appeared on sodium dodecyl sulfate-polyacrylamide gel electrophoresis gels only when large amounts of V. parahaemolyticus LPS were loaded. These bands were proven to be the aggregates of the fastest-moving low-molecular-weight bands by re-electrophoresis. The banding pattern of V. parahaemolyticus LPSs produced on nitrocellulose membranes by immunoblotting indicated that the V. parahaemolyticus LPSs did not have an O-specific side chain. The low ratio of total carbohydrate to lipid A of V. parahaemolyticus LPSs also suggested that they were like rough-type LPS. The mobility and profile of V. parahaemolyticus LPS on sodium dodecyl sulfate-polyacrylamide gel electrophoresis gel and its chemical composition were closely related to the serotype of a specific strain but not with the Kanagawa phenomenon.  相似文献   

5.
The constituent fatty acids of lipopolysaccharides (LPS) of Coxiella burnetii (phase I and II) were qualitatively and quantitatively analysed by combined gas-liquid chromatography/mass spectrometry. The total fatty acid content (per mg LPS) was determined as 90.0 nmol (2.3 wt%) for LPS of phase I cells (LPS I) and 179.1 nmol (4.8 wt%) for LPS of phase II cells (LPS II). Of the 24 different acyl residues characterized (12 to 18 carbon atoms), nine were 3-hydroxy fatty acids (normal, iso- and anteiso-branched) which quantitatively predominated. All 3-hydroxylated fatty acids were found to possess the (R)-configuration, to be exclusively amide-linked and to be acylated at their 3-hydroxyl group. Ester-linked nonhydroxylated fatty acids (normal, iso- and anteiso-branched) were present but ester-bound 3-hydroxy- or 3-acyloxyacyl residues were lacking from C. burnetii LPS I and LPS II. As the major acyl group (R)-3-(12-methyl-tetradecanoyloxy)-12-methyl-tetradecanoic acid was identified. Our results show that the complex fatty acid spectrum of C. burnetii differs considerably from that of LPS of other Gram-negative bacteria. They further suggest an enormous heterogeneity of the lipid A component of C. burnetii LPS I and LPS II.  相似文献   

6.
Comparison of lipopolysaccharides (LPS) from phase variants of different strains of Bordetella phase variants of different strains of Bordetella pertussis has shown a difference in their composition, antigenicity and reactogenicity. Phase I variants of B. pertussis, with the exception of strain 134, contain a preponderance of LPS I whereas the major component of LPS of phase IV variants is LPS II. Sera raised to LPSs of phase I strains, other than 134, cross-react with each other but not with phase IV LPSs; and similarly all sera raised to phase IV LPSs cross-react with each other and with LPS from 134 phase I. The LPSs of all phase I variants, including that of 134, are approximately ten-fold or more reactive in the limulus amoebocyte lysate assay (LAL) than phase IV LPSs. In the human mononuclear cell pyrogen assay phase IV LPSs also stimulated a lower response than phase I LPSs. The B. pertussis phase I LPSs are 10-times more reactive than Escherichia coli standard endotoxin in the LAL assay but 100-times less reactive than E. coli LPS in the monocyte test for pyrogen. The SDS-PAGE profiles of B. pertussis LPSs are quite different from those of B. parapertussis and B. bronchiseptica strains. B. pertussis LPSs produced a typical lipo-oligosaccharide (LOS) pattern. B. bronchiseptica LPS produced a similar pattern but was antigenically distinct from B. pertussis LPSs I and II. B. parapertussis in contrast produced a ladder pattern typical of smooth type LPS.  相似文献   

7.
Chlamydial hemagglutinin identified as lipopolysaccharide.   总被引:2,自引:0,他引:2       下载免费PDF全文
Chlamydial lipopolysaccharide (LPS) agglutinated mouse and rabbit erythrocytes but not human, guinea pig, or pronghorn antelope erythrocytes. Hemagglutination was not specific for Chlamydia spp., as rough LPSs from Coxiella burnetii and Escherichia coli also agglutinated erythrocytes from the same animal species. Nonagglutinated and agglutinated erythrocytes bound equivalent amounts of LPS, indicating that hemagglutination was not due to a specific interaction of chlamydial LPS with erythrocytes. Thus, hemagglutination by chlamydial LPS is not mediated by specific receptor-ligand interactions but is a property of the altered surface of the LPS-coated erythrocytes.  相似文献   

8.
In this study we compared the interleukin 1 (IL 1)-inducing capacity and the reactivity in the Limulus amoebocyte assay (LAL) of purified lipopolysaccharides (LPSs) from various bacterial strains. LPSs differed greatly in their capacities (on a weight basis) to induce IL 1 release from serum-free cultured human monocytes. LPS species that induced high levels of IL 1 release from human monocytes exhibited a high thiobarbiturate-reactive 2-keto-3-deoxy-octonic acid (KDO) content. No relationship was found between the IL 1-inducing activity and the LAL reactivity of purified LPSs. Filtration experiments in which membranes of decreasing size-exclusion limits were used demonstrated that molecular species of LPS with an apparent Mr below 3,000 may induce IL 1, whereas only species with an apparent Mr above 8,000 are recognized in the LAL assay. The latter observation suggests that the reaction with LAL requires an aggregated form of LPS. These results indicate that biologically active LPS species can cross dialysis membranes in vivo although no LAL reactive material is detected in the blood compartment. The Limulus assay is an insufficient criterion for the absence of LPS in biological fluids.  相似文献   

9.
Lipid extracts of bovine pulmonary surfactant, which retain many of the biophysical characteristics of natural surfactant, contain approx. 98% lipid and 2% protein, as determined by amino acid analysis. Polyacrylamide/urea gel electrophoresis reveals that lipid extract surfactant contained a major apoprotein band with apparent Mr 3500 and minor apoprotein bands with apparent Mr 15,000 and 7000. After reduction, the 15 kDa band disappears and is replaced by a prominent band with apparent Mr = 5000. Reduction also results in a relative diminution of the 7 kDa band and a relative increase in the intensity of the 3.5-kDa band. Edman degradation reveals two major peptide sequences which have been designated surfactant-associated peptide (N-terminal Phe) and surfactant-associated peptide (N-terminal Leu) and a minor sequence designated surfactant-associated peptide (N-terminal Ile). The latter surfactant-associated peptide appears to be related to the N-terminal Leu peptide but lacks the terminal Leu. N-Terminal analysis by dansylation demonstrates that the 15 and 5 kDa (reduced) apoprotein species contain N-terminal Phe, Leu and Ile. The 3.5 and 7 kDa bands contain only N-terminal Leu and Ile. Chromatography of lipid extracts on silicic acid columns gives rise to fraction I, which contains protein and phosphatidylglycerol, and fraction II, which contains protein, phosphatidylglycerol and phosphatidylethanolamine. Fraction I was primarily composed of the 15-kDa apoproteins, while fraction II contained mainly the 3.5 and 7 kDa apoproteins. Both fractions exhibited biophysical activity after reconstitution with dipalmitoylphosphatidylcholine. These results indicate that lipid extracts contain an oligomer of 15 kDa containing surfactant-associated peptide (N-terminal Phe) and surfactant-associated peptides (N-terminal Leu or Ile) which interact through sulfhydryl and perhaps other bonds. Lipid extracts also contain 3.5 kDa monomers of surfactant-associated peptides with N-terminal Leu and N-terminal Ile which can dimerize through sulfhydryl and perhaps hydrophobic interactions.  相似文献   

10.
The structure and biological properties of lipopolysaccharides (LPSs) from strains IMB 4125 (=ATCC 13525) and IMB 7769 of the bacterium Pseudomonas fluorescens (biovar I) were studied in vitro. LPSs were similar in the composition of lipid A and the core lipid but differed in the structure of O-specific polysaccharide chains, which was corroborated by the absence of serological relationships between them. The toxicity (LD50) of LPSs of P. fluorescens with respect to D-glucosamine-sensitized mice was 40-50 times lower than the toxicity of the classic endotoxins, LPSs of E. coli. The LPSs studied stimulated the production of tumor necrosis factor (TNF) and nitric oxide (NO) by mouse peritoneal macrophages. The rates of TNF and NO synthesis induced by the LPSs of interest were eight to nine and three to five times lower, respectively, than the corresponding parameters of the control LPSs of E. coli 055:B5 and 026:B6. Additionally, LPS preparations of the P. fluorescens strains induced TNF synthesis by monocytes of human whole-blood preparations. Certain differences in biological properties of these strains have been revealed, which could be due to the characteristic features of LPS structure and composition in different cultures.  相似文献   

11.
The lipopolysaccharide (LPS) of Salmonella enteritidis has been implicated as a virulence factor of this organism. Therefore, the LPS from a stable virulent isolate, SE6-E21, was compared with that from an avirulent isolate, SE6-E5. The LPSs were extracted, and the high-molecular-weight (HMW) LPS was separated from the low-molecular-weight (LMW) LPS for both isolates. Both the HMW and LMW LPSs were characterized by glycosyl composition and linkage analyses. Immunochemical characterization was performed by Western blotting using factor 9 antiserum and using S. typhimurium antiserum which contains factors 1, 4, 5, and 12(2). In addition, the polysaccharides released by mild acid hydrolysis were isolated and subjected to hydrolysis by bacteriophage P22, which contains endorhamnosidase activity. The resulting oligosaccharides were purified by using Bio-Gel P4 gel permeation chromatography and characterized by nuclear magnetic resonance spectroscopy, fast atom bombardment mass spectrometry (FAB-MS), tandem MS-MS, and matrix-assisted laser desorption time of flight MS. The results show that the HMW LPS O-antigen polysaccharides from both isolates are comprised of two different repeating units, -[-->2)-[alpha-Tyvp-(1-->3)]beta-D-Manp-(1-->4)-alpha-L-R hap-(1-->3)-alpha-D-Galp-(1-->]- (structure I) and [-->2)-[alpha-Tyvp-(1-->3)]beta-D-Manp-(1-->4)-alpha--L-R hap-(1-->3)-[alpha-D-Glcp-(1-->4)]alpha-D-Galp-(1-->]- (structure II). The LMW LPSs from both isolates contains truncated O-antigen polysaccharide which is comprised of only structure I. In the virulent SE6-E21 isolate, the HMW LPS has a structure I/II ratio of 1:1, while in the avirulent SE6-E5 isolate, this ratio is 7:1. While the 7:1 ratio represents the published level of glucosylation for S. enteritidis LPS as well as for S. enteritidis LPS purchased from Sigma Chemical Co., the 1:1 ratio found for the virulent SE6-E21 is identical to the high level of glucosylation reported for S. typhi LPS. Thus, the LPS from the virulent SE6-E21 isolate produces an S. typhi-like LPS. Furthermore, the amount of O-antigen polysaccharide in SE6-E21 was twice that in SE6-E5.  相似文献   

12.
Caput epididymis proteins from control, pairfed and zinc deficient (ZD) wistar weanling albino rats after 2-, 4-, 6- and 8-weeks were examined using SDS-PAGE followed by densitometric scanning of the gels. In comparison to the control and pairfed rats, ZD rats displayed new proteins. These included a Mr 42 kDa from 2ZD, Mr 47.5, 27.5, 23.2 and 16.0 kDa from 4ZD and Mr 87 and 14.2 kDa from 6ZD group. The 8ZD group, however, revealed no additional protein bands over controls. Further, several other proteins were missing from ZD rats. These included Mr 93 and 71 kDa from 2ZD; 93, 90, 79, 67, 62, 55 and 15.3 kDa from 4ZD; 60, 45.5, 34, 30 and 24 kDa from 6ZD and 41.5, 33 and 27.5 kDa bands from 8ZD group. The results indicate that the induced Zn-deficient state may be responsible for the altered protein patterns in the caput epididymis. The duration of low Zn uptake period also appears to influence the protein pattern in caput epididymis.  相似文献   

13.
Heterogeneity of Rhizobium lipopolysaccharides.   总被引:23,自引:18,他引:5       下载免费PDF全文
The lipopolysaccharides ( LPSs ) from strains of Rhizobium leguminosarum, Rhizobium trifolii, and Rhizobium phaseoli were isolated and partially characterized by mild acid hydrolysis and by polyacrylamide gel electrophoresis. Mild acid hydrolysis results in a precipitate which can be removed by centrifugation or extraction with chloroform. The supernatant contains polysaccharides which, in general, are separated into two fractions ( LPS1 and LPS2 ) by Sephadex G-50 gel filtration chromatography. The higher-molecular-weight LPS1 fractions among the various Rhizobium strains are highly variable in composition and reflect the variability reported in the intact LPSs (R. W. Carlson and R. Lee, Plant Physiol. 71:223-228, 1983; Carlson et al., Plant Physiol. 62:912-917, 1978; Zevenhuizen et al., Arch. Microbiol. 125:1-8, 1980). The LPS1 fraction of R. leguminosarum 128C53 has a higher molecular weight than all other LPS1 fractions examined. All LPS2 fractions examined are oligosaccharides with a molecular weight of ca. 600. The major sugar component of all LPS2 oligosaccharides is uronic acid. The LPS2 compositions are similar for strains of R. leguminosarum and R. trifolii, but the LPS2 from R. phaseoli was different in that it contained glucose, a sugar not found in the other LPS2 fractions or found only in trace amounts. Polyacrylamide gel electrophoretic analysis shows that each LPS contains two banding regions, a higher-molecular-weight heterogeneous region often containing many bands and a lower-molecular-weight band. The lower-molecular-weight bands of all LPSs have the same electrophoretic mobility, which is greater than that of lysozyme. The banding pattern of the heterogeneous regions varies among the different Rhizobium strains. In the case of R. leguminosarum 128C53 LPS, the heterogeneous region of a higher molecular weight than is this region from all other Rhizobium strains examined and consists of many bands separated from one another by a small and apparently constant molecular weight interval. When the heterogeneous region of R. Leguminosarum 128C53 LPS was cut from the gel and analyzed, its composition was found to be that of the intact LPS, whereas the lower-molecular-weight band contains only sugars found in the LPS2 oligosaccharide. In the case of R. leguminosarum 128C63 and R. trifolii 0403 LPSs, the heterogeneous regions are similar and consist of several band s separated by a large-molecular-weight interval with a the major band of these heterogeneous regions having the lowest molecular weight with an electrophoretic mobility near that of beta-lactoglobulin. The heterogeneous region from R. phaseoli 127K14 consists of several bands with electrophoretic mobilities near that of beta-lactoglobulin, whereas this region from R. trifolii 162S7 shows a continuous staining region, indicating a great deal of heterogeneity. The results described in this paper are discussed with regard to the reported properties of Escherichia coli and Salmonella LPSs.  相似文献   

14.
A study was undertaken to discriminate the strains of Aeromonas hydrophila isolated from fish and diarrhoeal samples by SDS-PAGE analysis of outer membrane proteins (OMPs) and lipopolysaccharides (LPSs). Common bands at 47 kDa positions for OMPs and at 31–38 kDa for LPSs were observed. No strain of A. hydrophila from clinical or fish samples was found identical in either OMPs or LPSs profile.  相似文献   

15.
The lipopolysaccharide (LPS) from a Rhizobium phaseoli mutant, CE109, was isolated and compared with that of its wild-type parent, CE3. A previous report has shown that the mutant is defective in infection thread development, and sodium dodecyl sulfate-polyacrylamide gel electrophoresis shows that it has an altered LPS (K. D. Noel, K. A. VandenBosch, and B. Kulpaca, J. Bacteriol. 168:1392-1462, 1986). Mild acid hydrolysis of the CE3 LPS released a polysaccharide and an oligosaccharide, PS1 and PS2, respectively. Mild acid hydrolysis of CE109 LPS released only an oligosaccharide. Chemical and immunochemical analyses showed that CE3-PS1 is the antigenic O chain of this strain and that CE109 LPS does not contain any of the major sugar components of CE3-PS1. CE109 oligosaccharide was identical in composition to CE3-PS2. The lipid A's from both strains were very similar in composition, with only minor quantitative variations. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of CE3 and CE109 LPSs showed that CE3 LPS separated into two bands, LPS I and LPS II, while CE109 had two bands which migrated to positions similar to that of LPS II. Immunoblotting with anti-CE3 antiserum showed that LPS I contains the antigenic O chain of CE3, PS1. Anti-CE109 antiserum interacted strongly with both CE109 LPS bands and CE3 LPS II and interacted weakly with CE3 LPS I. Mild-acid hydrolysis of CE3 LPS I, extracted from the polyacrylamide gel, showed that it contained both PS1 and PS2. The results in this report showed that CE109 LPS consists of only the lipid A core and is missing the antigenic O chain.  相似文献   

16.
The structure of lipid A-core region of the lipopolysaccharide (LPS) from Klebsiella pneumoniae serotype O3 was determined using NMR, MS and chemical analysis of the oligosaccharides, obtained by mild acid hydrolysis, alkaline deacylation, and deamination of the LPS: [carbohydrate structure see text] where P is H or alpha-Hep; J is H or beta-GalA; R is H or P (in the deacylated oligosaccharides).Screening of the LPS from K. pneumoniae O1, O2, O4, O5, O8, and O12 using deamination showed that they also contain alpha-Hep-(1-->4)-alpha-Kdo-(2-->6)-GlcN and alpha-Kdo-(2-->6)-GlcN fragments.  相似文献   

17.
This study describes the molecular makeup of the cell-wall lipopolysaccharides (LPSs) (O-chain polysaccharide-->core oligosaccharide-->lipid A) from five Helicobacter pylori strains: H. pylori 26695 and J99, the complete genome sequences of which have been published, the established mouse model Sydney strain (SS1), and the symptomatic strains P466 and UA915. All chemical and serological experiments were performed on the intact LPSs. H. pylori 26695 and SS1 possessed either a low-Mr semi-rough-form LPS carrying mostly a single Ley type-2 blood-group determinant in the O-chain region covalently attached to the core oligosaccharide or a high-Mr smooth-form LPS, as did strain J99, with an elongated partially fucosylated type-2 N-acetyllactosamine (polyLacNAc) O-chain polymer, terminated mainly by a Lex blood-group determinant, connected to the core oligosaccharide. In the midst of semi-rough-form LPS glycoforms, H. pylori 26695 and SS1 also expressed in the O-chain region a difucosylated antigen, alpha-L-Fucp(1-3)-alpha-L-Fucp(1-4)-beta-D-GlcpNAc, and the cancer-cell-related type-1 or type-2 linear B-blood-group antigen, alpha-D-Galp(1-3)-beta-D-Galp(1-3 or 4)-beta-D-GlcpNAc. The LPS of H. pylori strain P466 carried the cancer-associated type-2 sialyl Lex blood-group antigen, and the LPS from strain UA915 expressed a type-1 Leb blood-group unit. These findings should aid investigations that focus on identifying and characterizing genes responsible for LPS biosynthesis in genomic strains 26695 and J99, and in understanding the role of H. pylori LPS in animal model studies. The LPSs from the H. pylori strains studied to date were grouped into specific glycotype families.  相似文献   

18.
Lipopolysaccharides (LPSs) of Chlamydophila psittaci 6BC and Chlamydophila pneumoniae Kajaani 6 contain 3-deoxy-D-manno-oct-2-ulosonic acid (Kdo), GlcN, organic bound phosphate, and fatty acids in the molar ratios of approximately 3:2:2.2:4.8 and approximately 2.9:2:2.1:4.9, respectively. The LPSs were immunoreactive with a monoclonal antibody against a family-specific epitope of chlamydial LPS. This finding, together with methylation analyses of both LPSs and MALDI-TOF MS experiments on de-O-, and de-O,N-acylated LPSs, indicate the presence of a Kdo trisaccharide proximal to lipid A having a structure alpha-Kdo-(2-->8)-alpha-Kdo-(2-->4)-alpha-Kdo, which appears to be the main component of the core region in the native chlamydial LPSs. In the de-O-acylated LPSs from Chl. psittaci 6BC and Chl. pneumoniae Kajaani 6, two major molecular species are present that differ in distribution of amide-bound hydroxy fatty acids over both GlcN. It appears that either two (R)-3-hydroxy-18-methylicosanoic acids or one (R)-3-hydroxy-18-methylicosanoic acid and one (R)-3-hydroxyicosanoic acid are attached to the GlcN residues. In contrast, the de-O-acylated LPS of Chl. psittaci PK 5082 contains one major molecular species that has two (R)-3-hydroxyicosanoic acid residues attached to two GlcN residues.  相似文献   

19.
Polyclonal antibodies were raised in mice by immunization with solubilized human zonae pellucidae (ZP). Dot-blot analysis and ELISAs showed very strong immunoreactivity against human and pig ZP, and less immunoreactivity against rabbit and mouse ZP. One- and two-dimensional (1D and 2D) Western blot analysis detected immunoreactivity to three human ZP-specific bands. Under 1D nonreducing conditions, bands were detected at Mr 97,000, 61,000, and 51,000; the lower molecular weight (MW) bands showed the strongest reaction. Three acidic bands were detected under nonreducing 2D Western blot analysis, one at Mr 85,000-107,000, and two closely apposed, parallel, and more acidic bands at Mr 47,000-55,000 and Mr 55,000-62,000. The lower MW bands displayed considerable microheterogeneity, with 8-13 charge isomers detected; the lower MW form (ZP3L) appeared to be more strongly reactive than the higher MW form (ZP3H), and displayed a more intense labeling towards the basic end of the isomer charge train. The human ZP antiserum detected electrophoretically separated heterologous ZP from pigs, rabbits, and mice, and was also shown to inhibit homologous sperm-zona binding in in vitro assays. This is the first study to report that human ZP3 consists of two distinct isomer chains.  相似文献   

20.
Two classes of human cDNA encoding the insulin/mitogen-activated p70 S6 kinase have been isolated; the two classes differ only in the 5' region, such that the longer polypeptide (p70 S6 kinase alpha I; calculated Mr 58,946) consists of 525 amino acids, of which the last 502 residues are identical in sequence to the entire polypeptides encoded by the second cDNA (p70 S6 kinase alpha II; calculated Mr 56,153). Both p70 S6 kinase polypeptides predicted by these cDNAs are present in p70 S6 kinase purified from rat liver, and each is thus expressed in vivo. Moreover, both polypeptides are expressed from a single mRNA transcribed from the (longer) p70 S6 kinase alpha I cDNA through the utilization of different translational start sites. Although the two p70 S6 kinase polypeptides differ by only 23 amino acid residues, the slightly longer alpha I polypeptide exhibits anomalously slow mobility on sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE), migrating at an apparent Mr of 90,000 probably because of the presence of six consecutive Arg residues immediately following the initiator methionine. Transient expression of p70 alpha I and alpha II S6 kinase cDNA in COS cells results in a 2.5- to 4-fold increase in overall S6 kinase activity. Upon immunoblotting, the recombinant p70 polypeptides appear as a closely spaced ladder of four to five bands between 65 and 70 kDa (alpha II) and 85 and 90 kDa (alpha I). Transfection with the alpha II cDNA yields only the smaller set of bands, while transfection with the alpha I cDNA generates both sets of bands. Mutation of Met-24 in the alpha I cDNA to Leu or Thr suppresses synthesis of the alpha II polypeptides. Only the p70 alpha I and alpha II polypeptides of slowest mobility on SDS-PAGE comigrate with the 70- and 90-kDa proteins observed in purified rat liver S6 kinase. Moreover, it is the recombinant p70 polypeptides of slowest mobility that coelute with S6 kinase activity on anion-exchange chromatography. The slower mobility and higher enzymatic activity of these p70 proteins is due to Ser/Thr phosphorylation, inasmuch as treatment with phosphatase 2A inactivates kinase activity and increases the mobility of the bands on SDS-PAGE in an okadaic acid-sensitive manner. Thus, the recombinant p70 S6 kinase undergoes multiple phosphorylation and partial activation in COS cells. Acquisition of S6 protein kinase catalytic function, however, is apparently restricted to the most extensively phosphorylated recombinant polypeptides.  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