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 419 毫秒
1.
In vivo, collagen I, the major structural protein in human body, is found assembled into fibrils. In the present work, we study a high concentrated collagen sample in its soluble, fibrillar, and denatured states using one and two dimensional {1H}-13C solid-state NMR spectroscopy. We interpret 13C chemical shift variations in terms of dihedral angle conformation changes. Our data show that fibrillogenesis increases the side chain and backbone structural complexity. Nevertheless, only three to five rotameric equilibria are found for each amino acid residue, indicating a relatively low structural heterogeneity of collagen upon fibrillogenesis. Using side chain statistical data, we calculate equilibrium constants for a great number of amino acid residues. Moreover, based on a 13C quantitative spectrum, we estimate the percentage of residues implicated in each equilibrium. Our data indicate that fibril formation greatly affects hydroxyproline and proline prolyl pucker ring conformation. Finally, we discuss the implication of these structural data and propose a model in which the attractive force of fibrillogenesis comes from a structural reorganization of 10 to 15% of the amino acids. These results allow us to further understand the self-assembling process and fibrillar structure of collagen.  相似文献   

2.
Relative intensities of main resonance high-resolution 13C NMR spectra of acid hydrolysis products of bovine skin and human aortic collagens were measured. The contents of Gly, Pro, Hyp amino acid residues and Gly-Pro-Y triplets in collagen peptides was estimated. A good agreement between the experimental and theoretical data of the analysis of amino acid sequences of bovine skin collagen was shown.  相似文献   

3.
Interaction between cartilage proteoglycan and the collagen(s) composed of 1 alpha, 2 alpha, and 3 alpha chains was studied in vitro. Most of the collagen was insoluble under the conditions of assay (0.15 M NaCl, 0.008 M phosphate buffer, pH 7.4; 4 degrees C) and was in the form of fibrils 20 nm in diameter or thinner. The larger fibrils had 60-70 nm periodicity, characteristic of native collagens. Proteoglycan monomers which had been labeled by incubating cartilage slices in vitro with Na2 35SO4 were used to assay the interaction. The insoluble collagen fraction bound proteoglycan from solution. At proteoglycan:collagen ratios lower than 1:2, binding was rapid and linear, and the dissociation constant was 1.7 X 10(-9) M. At higher proteoglycan:collagen ratios, more proteoglycan was bound, but at a slower rate. Binding of proteoglycan to collagen did not require fibrils, since soluble 1 alpha, 2 alpha, and 3 alpha containing collagen also bound to proteoglycan and formed an insoluble complex. Denatured collagens did not bind proteoglycan or compete for binding with normal collagen. Optimum binding occurred with intact proteoglycan, but proteoglycan which had been treated with protease was also bound at low levels. Both protease-treated proteoglycan and free chondroitin sulfate competed with intact proteoglycan in the binding assays, but neither chondroitinase ABC-treated proteoglycan nor the oligosaccharides produced by digestion of chondroitin sulfate with testicular hyaluronidase altered the binding of proteoglycan to collagen. Hyaluronic acid did not compete with radioactive proteoglycan, but heparin and dextran sulfate were extremely effective inhibitors of binding. These data suggest a relatively nonspecific interaction between sulfated polyanions and 1 alpha, 2 alpha, and 3 alpha containing collagens. However, given the location of these collagens near the chondrocyte surface, the interaction of fibrillar 1 alpha, 2 alpha, 3 alpha collagen with proteoglycan is likely to occur and to be of biological importance.  相似文献   

4.
Calf skin and rat tendon type I, bovine cartilage type II, and human amnion type III collagens have been radiolabeled by reaction with [3H]acetic anhydride, [3H]formaldehyde, and succinimidyl 2,3-[3H]propionate. All three reactions produce collagens with high specific activities that are suitable for use as substrates in collagenase assays. The identity of the radiolabel and the labeling indices do not alter the molecular weights or thermal stabilities of the collagens or the solubilities of the collagens or gelatins in dioxane-water mixtures at 4 degrees C. However, in contrast to native or sparsely labeled collagens, those with 40 or more lysine + hydroxylysine residues labeled per molecule do not undergo fibrillogenesis in the presence of 0.2-0.4 M NaCl in the 4-35 degree C temperature range. Thus, the modification reactions not only serve to introduce the radiolabel, but also to keep the collagens soluble over a wide range of temperatures and concentrations. The TCA, TCB fragments produced on partial reaction of each collagen type with tissue collagenases can be selectively denatured by a 10-minute incubation under specific conditions and the intact collagens selectively precipitated by addition of 50% v/v dioxane. This serves as the basis for soluble collagenase assays. The effect of labeling index on the properties of the collagens has been investigated and the results establish the range of conditions over which these collagens can be used as substrates for soluble versus fibrillar collagenase assays.  相似文献   

5.
Various collagens were extracted and purified from human placenta after partial pepsin digestion. We prepared type III + I (57:43), enriched type I, type III, and type IV collagens on an industrial level, and studied their biological properties with MRC5 fibroblast cells. Using the process of contraction of a hydrated collagen lattice described by Bell, we found tha the contraction rate was dependent on collagen type composition. The contraction was faster and more pronounced with pepsinized type I collagen than with pepsinized type III + I (57:43) collagen; the lowest rate was obtained with the pepsinized type III collagen. Using a new technique of collagen cross-linking, a gel was made with type IV collagen. This cross-linking procedure, based on partial oxidation of sugar residues and hydroxylysine by periodic acid, followed by neutralization, resulted in an increased number of natural cross-link bridges between oxidized and nonoxidized collagen molecules, without internal toxic residues. The fibroblasts were unable to contract type IV/IVox collagen gels. The type IV/IVox collagen gel was transparent and its amorphous ultrastructure lacked any visible striated fibrils. Fibroblast cells exhibited atypical behavior in these type IV/IVox collagen gels as evidenced by optical and electron microscopy. The penetration of fibroblasts could be measured. Fibroblasts penetrated faster in type IV/IVox collagen gels than in untreated type III + I collagen gels. The lowest rate of penetration was obtained with cross-linked type III + I gels. Fibroblast proliferation was similar on untreated or cross-linked type III + I collagen gels and slightly increased on type IV/IVox collagen gels, suggesting that this cross-linking procedure was not toxic.  相似文献   

6.
Aggregation of 42-residue amyloid β-protein (Aβ42) plays a pivotal role in the etiology of Alzheimer's disease (AD). Curcumin, the yellow pigment in the rhizome of turmeric, attracts considerable attention as a food component potentially preventing the pathogenesis of AD. This is because curcumin not only inhibits the aggregation of Aβ42 but also binds to its aggregates (fibrils), resulting in disaggregation. However, the mechanism of interaction between curcumin and the Aβ42 fibrils remains unclear. In this study, we analyzed the binding mode of curcumin to the Aβ42 fibrils by solid-state NMR using dipolar-assisted rotational resonance (DARR). To improve the quality of 2D spectra, 2D DARR data were processed with the covariance NMR method, which enabled us to detect weak cross peaks between carbons of curcumin and those of the Aβ42 fibrils. The observed (13)C-(13)C cross peaks indicated that curcumin interacts with the 12th and 17-21st residues included in the β-sheet structure in the Aβ42 fibrils. Interestingly, aromatic carbons adjacent to the methoxy and/or hydroxy groups of curcumin showed clear cross peaks with the Aβ42 fibrils. This suggested that these functional groups of curcumin play an important role in its interaction with the Aβ42 fibrils.  相似文献   

7.
Previous observations with type I collagen from a proband with lethal osteogenesis imperfecta demonstrated that type I collagen containing a substitution of cysteine for glycine alpha 1-748 copolymerized with normal type I collagen (Kadler, K. E., Torre-Blanco, A., Adachi, E., Vogel, B. E., Hojima, Y., and Prockop, D. J. (1991) Biochemistry 30, 5081-5088). Here, three preparations containing normal type I procollagen and type I procollagen with a substitution of cysteine for glycine alpha 1-175, glycine alpha 1-691, or glycine alpha 1-988 were purified from cultured skin fibroblasts from probands with osteogenesis imperfecta. The procollagens were then used as substrates in a system for assaying the self-assembly of type I collagen into fibrils. The cysteine-substituted collagens in all three preparations were incorporated into fibrils. The cysteine alpha 1-175 and cysteine alpha 1-691 collagens were shown to increase the lag time and decrease the propagation rate constant for fibril assembly. All three preparations containing cysteine-substituted collagens formed fibrils with diameters that were two to four times the diameter of fibrils formed under the same conditions by normal type I collagen. Also, the three preparations containing cysteine substituted collagens had higher solubilities than normal type I collagen. The results, therefore, demonstrated that the three cysteine-substituted collagens copolymerized with normal type I collagen. The effects of the mutated collagens on fibril assembly can be understood in terms of a recently proposed model of fibril growth from symmetrical tips by assuming that the mutated monomers partially inhibit tip growth but not lateral growth of the fibrils. Of special interest was the observation that the Cys alpha 1-175 collagen from a proband with a non-lethal variant of osteogenesis imperfecta had quantitatively less effect on several parameters of fibril assembly at 37 degrees C than cysteine-substituted collagens from three probands with lethal variants of the disease.  相似文献   

8.
The periodical D-band pattern is generally considered a unique ultrastructural feature shared by all fibril-forming collagens, which correlates with the intrafibril, paracrystalline array of tropocollagen monomers. Distinct band patterns have been reported, however, for collagen stained long-spacing (SLS) crystallites of genetic types I, II, and III. Moreover, D-band patterns of negatively stained, native type II collagen fibrils were found to be not identical to those of type I in our previous research. Because of (a) these distinctive features, (b) tropocollagen heterotrimeric conditions (type I) vs homotrimeric conditions (type II), and (c) different lengths and poor homology between extrahelical telopeptides, the molecular array or telopeptide conformation within the extensively studied type I collagen fibrils could be not the same as those in the very much less intensively studied type II collagen fibrils. In this investigation, a distinctive positive-staining D-band pattern was found for type II collagen fibrils obtained from human cartilages. A fibril model was developed by analyzing actual D-band patterns, and matching them against simulated patterns based on the primary structure of extrahelical and helical domains in human type II tropocollagen. In particular, a more prominent b(1) band was apparent in native type II collagen fibrils than in type I. This distinctive feature was also observed for native-type collagen fibrils reconstituted from purified type II collagen, i.e., free from associated minor type XI collagen. On modeling possible monomer arrays, the best fit between microdensitograms and simulation traces was found for 234 amino acid staggering, as is also the case for type I collagen fibrils. On comparing this model with an analogous one for type I collagen fibrils, there was a higher intraband distribution of charged residues for band b(1), consistent with the higher electrondensity observed for this band in type II collagen fibrils. N- and C-telopeptide displacement in the model corresponded to D-locations of a c(2) subband, which we named c(2.0), and band a(3), respectively. In simulation profiles, c(2.0) -like and a(3) -like peaks mimicked the corresponding peaks in microdensitograms when molecular reversals were adopted at positions 10N-12N, 12C-14C, and 17C-19C for N- and C-telopeptides. Hydrophobic interactions and algorithmic predictions of protein secondary structure, according to Chou and Fasman and Rost and Sander criteria, were consistent with these conformational models, and suggest that an additional molecular reversal may occur at positions 3N-5N. These telopeptide "S-fold" conformations, interpreted as axial projections of tridimensional conformation, may represent starting points for further investigation into the still unresolved tridimensional conformation of telopeptides in monomers arrayed within type II collagen fibrils.  相似文献   

9.
Fibronectin is one of the main components of the extracellular matrix and associates with a variety of other matrix molecules including collagens. We demonstrate that the absence of secreted type VI collagen in cultured primary fibroblasts affects the arrangement of fibronectin in the extracellular matrix. We observed a fine network of collagen VI filaments and fibronectin fibrils in the extracellular matrix of normal murine and human fibroblasts. The two microfibrillar systems did not colocalize, but were interconnected at some discrete sites which could be revealed by immunoelectron microscopy. Direct interaction between collagen VI and fibronectin was also demonstrated by far western assay. When primary fibroblasts from Col6a1 null mutant mice were cultured, collagen VI was not detected in the extracellular matrix and a different pattern of fibronectin organization was observed, with fibrils running parallel to the long axis of the cells. Similarly, an abnormal fibronectin deposition was observed in fibroblasts from a patient affected by Bethlem myopathy, where collagen VI secretion was drastically reduced. The same pattern was also observed in normal fibroblasts after in vivo perturbation of collagen VI-fibronectin interaction with the 3C4 anti-collagen VI monoclonal antibody. Competition experiments with soluble peptides indicated that the organization of fibronectin in the extracellular matrix was impaired by added soluble collagen VI, but not by its triple helical (pepsin-resistant) fragments. These results indicate that collagen VI mediates the three-dimensional organization of fibronectin in the extracellular matrix of cultured fibroblasts.  相似文献   

10.
Heat denatured type I and type III calf skin collagen were found to be substrates for guinea pig liver transglutaminase (R-glutaminyl-peptide:amine gamma-glutamyl-yltransferase, EC 2.3.2.13) but not for active plasma factor XIII (factor XIIIa). Liver transglutaminase was shown to catalyse incorporation of 14C-putrescine into subunits of denatured collagen of both types, cross-linking of the latter into high molecular weight polymers and their co-cross-linking to fibrin and fibrinogen. Factor XIIIa is inactive in these respects. None of these reactions was catalysed by liver transglutaminase and plasma factor XIIIa when nondenatured collagens both soluble or in the forms of reconstituted fibrils served as substrates. Some cross-linking of cleavage products of collagen type I (obtained by treatment with collagenase from human neutrophiles) was induced by liver transglutaminase and factor XIIIa. The results indicate that although appropriate glutamine and lysine residues for a epsilon-(gamma-glutamine) lysine cross-linked formation are present in collagen, the native conformation of collagen prevents the action of liver transglutaminase and factor XIIIa.  相似文献   

11.
1. The neutral collagenase released into the culture medium by explants of ehrumatoid synovial tissue has been purified by ultrafiltration and column chromatography, utilising Sephadex G-200, Sephadex QAE A-50 and Sephadex G-100 superfine. 2. The final collagenase preparation had a specific activity against thermally reconstituted collagen fibrils of 312 mug collagen degraded min-1 mg enzyme protein-1, representing more than a 1000-fold increase over that of the active culture medium. 3. Electrophoresis in polyacrylamide disc-gels with and without sodium dodecyl sulphate showed the enzyme to migrate as a single protein band. Elution experiments from polyacrylamide gels and chromatography columns have provided no evidence for the existence of more than one collagenase. 4. The molecular weight of the enzyme, as determined by dodecylsulphate-polyacrylamide gel electrophoresis, was 33000. 5. Data obtained from sutdies with the ion-exchange resin and from gel electrophoresis in acid and alkaline buffer systems suggested a basically charged enzyme. 6. It did not hydrolyse the synthetic collagen peptide Pz-Pro-Leu-Gly-Pro-D-Arg and non-specific protease activity was absent. 7. The collagenase attacked undenatured collagen in solution at 25 degrees C resulting in a 58% loss of viscosity and producing the two characteristic products TCA(3/4) and TCB(1/4). 8. At 37 degrees C and pH 8.0 both reconstituted collagen fibrils and gelatin were degraded to peptides of less than 10000 molecular weight. 9. As judged by the release of soluble hydroxyproline peptides and electron microscopic appearances the enzyme degraded human insoluble collagens derived from tendon and soft juxta-articular tissues although rates of attack were less than with reconstituted fibrils. 10. The data suggests that pure rheumatoid synovial collagenase at 37 degrees C and neutral pH can degrade gelatin, reconstituted fibrils and insoluble collagens without the intervention of non-specific proteases. 11. The different susceptibilities of various collagenous substrates to collagenase attack are discussed.  相似文献   

12.
Collagen fibrils from the dermis of Sepia officinalis were processed for immunoelectron microscopy to reveal reactions to antibodies against mammalian types I, III, and V, teleost type I and cephalopod type I-like collagens, by single and double immunogold localization. The fibrils were observed: (a) in suspensions of prepared fibrils, (b) in ultrathin sections of embedded fibril preparations, and (c) in ultrathin sections of dermal tissue. Some samples were subjected to acetic acid or urea dissociation. It was found that collagen fibrils from Sepia dermis are heterotypic in that they are composed of type I-like and type V collagens. Type I-like collagen epitopes were present mainly at the periphery of the fibrils; type V collagen epitopes were present throughout the fibrils. This is the first demonstration that collagen fibrils from an invertebrate are heterotypic, suggesting that heterotypy may be an intrinsic characteristic of the fibrils of fibrillar collagens, independent of evolutionary or taxonomic status.  相似文献   

13.
Two different collagens were isolated and characterized from the body walls of the vestimentiferan tube worm Riftia pachyptila and the annelid Alvinella pompejana, both living around hydrothermal vents at a depth of 2600 m. The acid-soluble cuticle collagens consisted of a long triple helix (2.4 microns for Alvinella, 1.5 microns for Riftia) terminating into a globular domain. Molecular masses of 2600 and 1700 kDa, respectively, were estimated from their dimensions. The two cuticle collagens were also quite different in amino acid composition, in agreement with their different supramolecular organizations within tissues. Interstitial collagens corresponding to cross-striated fibrils underneath the epidermal cells could be solubilized by digestion with pepsin and consisted of a single alpha-chain. They were similar in molecular mass (340 kDa) and length (280 nm) but differed in composition and banding patterns of segment-long-spacing fibrils. This implicates significant sequence differences also in comparison to fibril-forming vertebrate collagens, although all form typical quarter-staggered fibrils. The thermal stability of the worm collagens was, with one exception (interstitial collagen of Riftia), in the range of mammalian and bird collagens (37 to 46 degrees C), and thus distinctly above that of shallow sea water annelids. Yet, their 4-hydroxyproline contents were not directly correlated to this stability. About 20% of Riftia collagen alpha-chain sequence was elucidated by Edman degradation and showed typical Gly-X-Y repeats but only a limited homology (45 to 58% identity) to fibril-forming vertebrate collagens. A single triplet imperfection and the variable hydroxylation of proline in the X position were additional unique features. It suggests that this collagen represents an ancestral form of fibril-forming collagens not directly corresponding to an individual fibril-forming collagen type of vertebrates.  相似文献   

14.
Fibronectin and collagens are major constituents of the cell matrix of fibroblasts. Fibronectin is a 220,000 dalton glycoprotein that mediates a variety of adhesive functions of cells examined in vitro. Fibronectin is secreted in a soluble form and interacts with collagen to form extracellular filaments. Fibronectin and procollage type I were localized using the peroxidase anti-peroxidase method. Under standard culture conditions, fibronectin and procollagen were localized to non-periodic 10 nm extracellular fibrils, the cell membrane and plasma membrane vesicles. Ascorbate treatment of cells leads to a new larger fibril with a diameter of approximately 40 nm. Antibodies to fibronectin and procollagen I react to these native collagen fibrils with an axial periodicity of approximately 70 nm. Fibronectin is clearly associated with native collagen fibrils produced by ascorbate treated cells and there is an asymetric distribution or segregation of fibronectin on these collagen fibrils with a 70 nm axial repeat.  相似文献   

15.
Invertebrates possess unique collagen-containing connective tissue elements, the biochemistry of which is not clearly understood. We previously reported the occurrence of a novel heterotrimeric type V/XI like collagen in the cranial cartilage of the cuttlefish Sepia officinalis. We report here the purification of the three chains by ion exchange chromatography and the physicochemical characteristics of this collagen. This collagen shared substantial similarity to the collagen purified from the cornea of S. officinalis, with respect to chain composition, cyanogen bromide peptide profile and amino acid composition. The mobility of the C3 chain was retarded in the corneal collagen, which also had an increased glycine content and a smaller ratio of hydroxylysine to lysine, together with a reduction in bound carbohydrates. The cartilage collagen had a higher denaturation temperature than corneal collagen. As observed by transmission electron microscopy of reconstituted fibrils, the heterotrimeric invertebrate collagen formed fibrils of no apparent periodicities as opposed to the regular 64-nm banding pattern of milk shark (Rhizoprionodon acutus) cartilage collagen. This is also the first report on the molecular species of collagen in an invertebrate cornea. Our results strongly support the functioning of minor vertebrate collagens as major collagens in some invertebrates, close similarity of collagens in two tissues with different functions and would hold significance to our understanding of collagen polymorphism and the evolution of the extracellular matrix.  相似文献   

16.
We report the results of solid state nuclear magnetic resonance (NMR) measurements on amyloid fibrils formed by the full-length prion protein PrP (residues 23?231, Syrian hamster sequence). Measurements of intermolecular 13C?13C dipole?dipole couplings in selectively carbonyl-labeled samples indicate that β-sheets in these fibrils have an in-register parallel structure, as previously observed in amyloid fibrils associated with Alzheimer’s disease and type 2 diabetes and in yeast prion fibrils. Two-dimensional 13C?13C and 15N?13C solid state NMR spectra of a uniformly 15N- and 13C-labeled sample indicate that a relatively small fraction of the full sequence, localized to the C-terminal end, forms the structurally ordered, immobilized core. Although unique site-specific assignments of the solid state NMR signals cannot be obtained from these spectra, analysis with a Monte Carlo/simulated annealing algorithm suggests that the core is comprised primarily of residues in the 173?224 range. These results are consistent with earlier electron paramagnetic resonance studies of fibrils formed by residues 90?231 of the human PrP sequence, formed under somewhat different conditions [Cobb, N. J., Sonnichsen, F. D., McHaourab, H., and Surewicz, W. K. (2007) Proc. Natl. Acad. Sci. U.S.A. 104, 18946?18951], suggesting that an in-register parallel β-sheet structure formed by the C-terminal end may be a general feature of PrP fibrils prepared in vitro.  相似文献   

17.
Types I and III collagens were solubilized from fetal human skin by limited digestion with pepsin and precipitated by dialysis against 0.02 M Na2HPO4. Heat denaturation of the collagens in 2 M guanidine-HCl, pH 7.5, resulted in the precipitation of the contaminant pepsin which could be removed by centrifugation. Renaturation of the denatured collagens by dialysis against deionized water at 22° for 2 hours selectively precipitated the type III collagen fibrils. Type I collagen remained in solution. The simplicity and high recovery (77%) make this a suitable approach for the rapid estimation of type III collagen in small tissue samples.  相似文献   

18.
Residues 1-89 constitute the Asn- and Gln-rich segment of the Ure2p protein and produce the [URE3] prion of Saccharomyces cerevisiae by forming the core of intracellular Ure2p amyloid. We report the results of solid-state nuclear magnetic resonance (NMR) measurements that probe the molecular structure of amyloid fibrils formed by Ure2p1-89 in vitro. Data include measurements of intermolecular magnetic dipole-dipole couplings in samples that are 13C-labeled at specific sites and two-dimensional 15N-13C and 13C-13C NMR spectra of samples that are uniformly 15N- and 13C-labeled. Intermolecular dipole-dipole couplings indicate that the beta-sheets in Ure2p1-89 fibrils have an in-register parallel structure. An in-register parallel beta-sheet structure permits polar zipper interactions among side chains of Gln and Asn residues and explains the tolerance of [URE3] to scrambling of the sequence in residues 1-89. Two-dimensional NMR spectra of uniformly labeled Ure2p1-89 fibrils, even when fully hydrated, show NMR linewidths that exceed those in solid-state NMR spectra of fibrils formed by residues 218-289 of the HET-s prion protein of Podospora anserina [as originally reported in Siemer, A. B., Ritter, C., Ernst, M., Riek, R., and Meier, B. H. (2005) Angew. Chem., Int. Ed. 44, 2441-2444 and confirmed by measurements reported here] by factors of three or more, indicating a lower degree of structural order at the molecular level in Ure2p1-89 fibrils. The very high degree of structural order in HET-s fibrils indicated by solid-state NMR data is therefore not a universal characteristic of prion proteins, and is likely to be a consequence of the evolved biological function of HET-s in heterokaryon incompatibility. Analysis of cross peak intensities in two-dimensional NMR spectra of uniformly labeled Ure2p1-89 fibrils suggests that certain portions of the amino acid sequence may not participate in a rigid beta-sheet structure, possibly including portions of the Asn-rich segment between residues 44 and 76.  相似文献   

19.
A method is described for quantifying both reducible and mature crosslinking amino acids of collagen. The main crosslinking residue in cartilage, dentine and mature bone was the 3-hydroxypyridinium compound identified by Fujimoto et al. (1–3). Adult articular cartilage contained about one residue per collagen molecule, over forty times the content of the reducible crosslinks. We propose that hydroxypyridinium residues are formed by spontaneous interaction of two residues of hydroxylysino-5-ketonorleucine. This reaction explains the disappearance of reducible crosslinks at maturity and provides a novel mechanism for lateral crosslinking within and between fibrils which may account for some of the unique physical properties of hard tissue collagens.  相似文献   

20.
Adhesion of meningitis-associated Escherichia coli O18acK1H7 to collagens was characterized. The E. coli strain IHE 3034 adhered to type IV and type I collagens but not to type III collagen immobilized on glass. Collagens lack terminal mannosyl units, yet the bacterial adhesion was completely abolished in the presence of alpha-methyl-D-mannoside. A cat cassette was introduced into the filmA gene of IHE 3034, and the resulting mutant strain IHE 3034-2 failed to adhere to collagens. In contrast, insertion of a Gm cassette into the sfaA gene of IHE 3034, encoding the S-fimbrillin, had no significant effect on the adhesiveness. The fim cluster from IHE 3034 was cloned and expressed in trans in the fimA::cat mutant strain IHE 3034-2. The complemented strain IHE 3034-2(pRPO-1) exhibited adhesiveness to type IV and type I collagens, confirming the function of the type 1 fimbria in the adhesion. We have previously shown that the type 1 fimbria from E. coli K-12 strain PC31 does not confer bacterial adhesiveness to collagens. The fimH genes from E. coli IHE 3034 as well as from PC31 were expressed in the fimH-null strain MS4. The FimH from IHE 3034 potentiated collagen adherence, whereas the FimH from PC31 was inactive. Sequence comparison of fimH from IHE 3034 and PC31 revealed five amino-acid differences in the predicted mature FimH proteins: at residues 27, 62, 70, 78 and 201. Each of these residues in the IHE 3034-FimH were individually substituted to the corresponding amino acid in the PC31-FimH. The substitution S62-->A completely abolished collagen adhesiveness. The reverse substitution A62-->S in the PC31-FimH as well as in the FimH from another E. coli strain induced collagen adhesiveness to the level seen with IHE 3034-FimH. Out of nine fimH genes analysed from isolates of E. coli, collagen adhesiveness as well as alanine at position 62 in FimH were found only in two O18acK1H7 isolates with the isoenzyme profile ET type 1. Our results demonstrate that the amino-acid residue Ala-62 in the FimH lectin is critical for the adhesion to collagens by a highly virulent clonal group of E. coli.  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