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.
Twelve of sixteen different cell types including fibroblasts and tumor cells were able to attach and spread on substrates of pepsin-solubilized or intact collagen VI, and on its triple helical domain. Attachment and spreading were independent of soluble mediator proteins (fibronectin, laminin) and collagen VI was distinct from collagens I, IV and V in the cells with which it interacted. Many of the same cells bound and spread on substrates prepared from unfolded α2(VI) and α3(VI) chains but not on the α1(VI) chain. The interactions with the chains were inhibited by low concentrations (10–100 μM) of synthetic RGDS and RGDT but not RGES peptides while the binding of cells to pepsin-solubilized collagen VI was more than 20-fold less sensitive to these peptides. The data incidate that cells have the ability to bind to collagen VI in a specific manner suggesting a similar function for collagen VI in situ.  相似文献   

2.
The mechanism of chain selection and trimerization of fibril-associated collagens with interrupted triple helices (FACITs) differs from that of fibrillar collagens that have special C-propeptides. We recently showed that the second carboxyl-terminal non-collagenous domain (NC2) of homotrimeric collagen XIX forms a stable trimer and substantially stabilizes a collagen triple helix attached to either end. We then hypothesized a general trimerizing role for the NC2 domain in other FACITs. Here we analyzed the NC2 domain of human heterotrimeric collagen IX, the only member of FACITs with all three chains encoded by distinct genes. Upon oxidative folding of equimolar amounts of the α1, α2, and α3 chains of NC2, a stable heterotrimer with a disulfide bridge between α1 and α3 chains is formed. Our experiments show that this heterotrimerization domain can stabilize a short triple helix attached at the carboxyl-terminal end and allows for the proper oxidation of the cystine knot of type III collagen after the short triple helix.  相似文献   

3.
Acid-soluble and pepsin-treated collagen V were prepared from fetal human bones or human placenta, respectively, to be tested for potential cell adhesion promoting activity. Out of 14 different collagen I-adhering cell lines, 10 showed distinct adhesion to collagen V. In all cases adhesion was followed by spreading. The activities of intact and pepsin-solubilized collagen V were similar, suggesting that the cell binding sites are restricted to the triple-helical domain of the molecules. Cell adhesion was also induced by the unfolded form of collagen V and after separation of the α chains by heparin affinity chromatography. Isolated α2(V) chains, rich in RGD sequences, were more efficient than isolated α1(V) chains. However, cell adhesion to native or denatured collagen V did not proceed by the same molecular mechanisms as shown by cell adhesion inhibition experiments. Cell adhesion to native collagen V was insensitive to the presence of RGD-containing synthetic peptides while adhesion to denatured collagen V was inhibited by the peptides. Furthermore, the results strongly suggested a major role for α1α1 and α2β1 integrins in the RGD-independent cell adhesion to native collagen V. These data indicate that collagen V is a specific adhesive substrate for different cell types. It also suggests that distinct sets of RGD-dependent and RGD-independent receptors mediate cell attachment to unfolded and native collagen V, respectively. This mechanism is shared by at least the interstitial collagens I and VI, which supports the hypothesis that when included in the triple-helical conformation of collagens, RGD sequences are either not accessible to cells or exhibit specific conformations recognized by different integrins.  相似文献   

4.
A neutral metal protease has been identified which cleaves native type V collagen under conditions where pepsinized type IV collagen or the interstitial collagens are not significantly degraded. The enzyme is secreted into the media of cultured M50-76 reticulum cell sarcoma (malignant macrophages) and leiomyosarcoma tumor cells. Biosynthetically labeled type V collagen prepared from organ cultures of human amnion membrane is used for a routine assay of type V collagenolytic activity. The partially purified enzyme a) exists in a latent form requiring trypsin activation for maximum activity; b) has a molecular weight estimated by molecular sieve chromatography of approximately 80,000 daltons; c) is inhibited by EDTA but not phenylmethylsulfonyl fluoride; and d) produces specific cleavage products of both A and B collagen chains.  相似文献   

5.
Precise mapping and unraveling the mechanism of interaction or degradation of a certain type of collagen triple helix requires the generation of short and stable collagenous fragments. This is a great challenge especially for hetero-trimeric collagens, where chain composition and register (stagger) are important factors. No system has been reported that can be efficiently used to generate a natural collagenous fragment with exact chain composition and desired chain register. The NC2 domain (only 35–50 residues) of FACIT collagens is a potent trimerization domain. In the case of type IX collagen it provides the efficient selection and hetero-trimerization of three distinct chains. The ability of the NC2 domain to determine the chain register of the triple helix is studied. We generated three possible sequence combinations (α1α1α2, α1α2α1, α2α1α1) of a type I collagen fragment (the binding region for the von Willebrand factor A3 domain) attached to the NC2 domain. In addition, two control combinations were produced that constitute homo-trimers of (α1)3 or (α2)3. For the hetero-trimeric constructs, α1α1α2 demonstrated a higher melting temperature than the other two. Binding experiments with the von Willebrand factor A3 domain revealed the homo-trimer of (α1)3 as the strongest binding construct, whereas the homo-trimer of (α2)3 showed no binding. For hetero-trimers, α1α1α2 was found to be the strongest binding construct. Differences in thermal stability and binding to the A3 domain unambiguously demonstrate that the NC2 domain of type IX collagen determines not only the chain composition but also the chain register of the adjacent triple helix.  相似文献   

6.
Aberrant expression of the type V collagen α1(V) chain can underlie the connective tissue disorder classic Ehlers-Danlos syndrome, and autoimmune responses against the α1(V) chain are linked to lung transplant rejection and atherosclerosis. The α1(V) collagenous COL1 domain is thought to contain greater numbers of post-translational modifications (PTMs) than do similar domains of other fibrillar collagen chains, PTMs consisting of hydroxylated prolines and lysines, the latter of which can be glycosylated. These types of PTMs can contribute to epitopes that underlie immune responses against collagens, and the high level of PTMs may contribute to the unique biological properties of the α1(V) chain. Here we use high resolution mass spectrometry to map such PTMs in bovine placental α1(V) and human recombinant pro-α1(V) procollagen chains. Findings include the locations of those PTMs that vary and those PTMs that are invariant between these α1(V) chains from widely divergent sources. Notably, an unexpectedly large number of hydroxyproline residues were mapped to the X-positions of Gly-X-Y triplets, contrary to expectations based on previous amino acid analyses of hydrolyzed α1(V) chains from various tissues. We attribute this difference to the ability of tandem mass spectrometry coupled to nanoflow chromatographic separations to detect lower-level PTM combinations with superior sensitivity and specificity. The data are consistent with the presence of a relatively large number of 3-hydroxyproline sites with less than 100% occupancy, suggesting a previously unknown mechanism for the differential modification of α1(V) chain and type V collagen properties.  相似文献   

7.
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   

8.
Approximately half the proline residues in fibrillar collagen are hydroxylated. The predominant form is 4-hydroxyproline, which helps fold and stabilize the triple helix. A minor form, 3-hydroxyproline, still has no clear function. Using peptide mass spectrometry, we recently revealed several previously unknown molecular sites of 3-hydroxyproline in fibrillar collagen chains. In fibril-forming A-clade collagen chains, four new partially occupied 3-hydroxyproline sites were found (A2, A3, A4 and (GPP)n) in addition to the fully occupied A1 site at Pro986. The C-terminal (GPP)n motif has five consecutive GPP triplets in α1(I), four in α2(I) and three in α1(II), all subject to 3-hydroxylation. The evolutionary origins of this substrate sequence were investigated by surveying the pattern of its 3-hydroxyproline occupancy from early chordates through amphibians, birds and mammals. Different tissue sources of type I collagen (tendon, bone and skin) and type II collagen (cartilage and notochord) were examined by mass spectrometry. The (GPP)n domain was found to be a major substrate for 3-hydroxylation only in vertebrate fibrillar collagens. In higher vertebrates (mouse, bovine and human), up to five 3-hydroxyproline residues per (GPP)n motif were found in α1(I) and four in α2(I), with an average of two residues per chain. In vertebrate type I collagen the modification exhibited clear tissue specificity, with 3-hydroxyproline prominent only in tendon. The occupancy also showed developmental changes in Achilles tendon, with increasing 3-hydroxyproline levels with age. The biological significance is unclear but the level of 3-hydroxylation at the (GPP)n site appears to have increased as tendons evolved and shows both tendon type and developmental variations within a species.  相似文献   

9.
S Ayad  A P Kwan  M E Grant 《FEBS letters》1987,220(1):181-186
Sequential extraction of bovine growth-plate cartilage with 4 M guanidinium chloride and pepsin was used to identify the intact and pepsinized forms respectively of type X collagen. This collagen occurs predominantly as the processed [alpha 1(X)]3 form in vivo, although the procollagen [pro alpha 1(X)]3 form can also be detected. The bovine pro alpha 1(X) and alpha 1(X) chains have Mr values identical to the corresponding chick species (Mr 59,000 and 49,000). However, the pepsinized alpha 1(X)p chains (Mr 47,000) are larger than those of the chick (Mr 45,000), and the bovine collagen type X is further distinguished by being disulphide-bonded within the triple-helical domain.  相似文献   

10.
Binding of thrombospondin (TSP) to types I-V collagen was examined by direct binding assays using 125I-TSP and by visualization of rotary-shadowed intermolecular complexes in the electron microscope. The binding of TSP was highest to type V collagen in the absence of Ca, while lower but significant levels of binding were observed to all other collagen types in the presence or absence of Ca. Unlike intact TSP, the trimeric collagen-binding domain of TSP composed of 70-kD chains showed no Ca dependence in its binding to type V collagen. Further evidence for binding of TSP to types I and III collagen was obtained by competition studies in which these soluble collagens effectively inhibited binding of 125I-TSP to immobilized type V collagen. The binding of TSP to type V collagen was inhibited by heparin and fucoidin, both high-affinity ligands of TSP's heparin-binding domain. mAb A6.1, which binds to the 70-kD domain of TSP, is also the best of a panel of anti-TSP mAbs at inhibiting the TSP-collagen interaction. Electron microscopy of rotary-shadowed replicas of TSP-collagen complexes revealed that all five types of collagen examined had a binding site for TSP at one end of the pepsinized, triple helical molecule. The specificity of this site was tested by examining the ability of BSA to form a complex with the end of the pepsinized collagens. Rotary-shadowed replicas revealed a low frequency of apparent BSA-collagen complexes, and histograms of these data showed no evidence for the preferential association of BSA with the end of the collagen molecules. In addition to the specific end site, type V collagen had an internal binding site for TSP located about two-thirds of the distance along the length of the collagen molecule from the end site. The internal binding site for TSP on type V collagen is apparently the site responsible for the higher affinity binding of TSP to that protein observed in direct binding assays. The trimeric 70-kD collagen-binding domain of TSP bound to the same sites on the collagens as did intact TSP.  相似文献   

11.
Type VIII collagen was isolated from bovine Descemet's membranes by pepsin treatment and salt fractionation, as described by Kapoor et al. [(1986) Biochemistry 25, 3930-3937]. Contaminating type IV collagen was removed by ion-exchange chromatography. Purified type VIII collagen consisted of two different polypeptide chains and, compared to the fiber forming collagens, showed a higher thermal stability. Corresponding fractions isolated from pepsinized human Ewing's sarcoma and fetal calf aorta reacted immunologically with a protein of similar molecular mass. After extraction of Descemet's membranes with guanidine hydrochloride, a peptide of about 60 kDa was obtained. This seems to be the tissue form of type VIII collagen.  相似文献   

12.
Analyses were made of the minor collagens synthesized by cultures of chondrocytes derived from 14-day chick embryo sterna. Comparisons were made between control cultures, cultures grown for 9 days in 5-bromo-2'-deoxyuridine (BrdU) and clones of chondrocytes grown to senescence. Separation of minor collagens from interstitial collagens was achieved by differential salt precipitation in the presence of carrier collagens in acid conditions. The precipitate at 0.9 M NaCl 0.5 M acetic acid from control cultures was shown by CNBr peptide analysis to contain only the alpha 1(II) chain of type II collagen, whereas after BrdU treatment or growth to senescence synthesis of only alpha 1(I) and alpha 2(I) chains occurred. The synthesis of type III collagen was not detected. Analysis of the precipitate at 2.0 M NaCl, 0.5 M HAc from control cultures demonstrated the synthesis of 1 alpha, 2 alpha and 3 alpha chains together with the synthesis of short chain (SC) collagen of Mr 43000 after pepsin digestion. After BrdU treatment or growth to senescence alpha chains were isolated which possessed the migration positions on polyacrylamide gel electrophoresis (PAGE), or the elution positions on CM-cellulose chromatography, of the alpha 1(V) and alpha 2(V) chains of type V collagen. In addition, for BrdU-treated but not for control cultures, intracellular immunofluorescent staining was observed with a monoclonal antibody which specifically recognizes an epitope present in the triple helix of type V collagen. Synthesis of short chain (SC) collagen was not detected after BrdU treatment or growth to senescence. These results suggest that chick chondrocytes grown in conditions known to cause switching of collagen synthesis from type II to type I collagen also undergo a switch from the synthesis of 1 alpha, 2 alpha and 3 alpha chains to the synthesis of the alpha 1(V) and alpha 2(V) chains of type V collagen. It appears that there are several cartilage-specific collagens which together undergo a regulatory control to the synthesis of collagens typical of other connective tissues.  相似文献   

13.
Type IV collagen is a major component of basement membranes. We have characterized 11 mutations in emb-9, the α1(IV) collagen gene of Caenorhabditis elegans, that result in a spectrum of phenotypes. Five are substitutions of glycines in the Gly-X-Y domain and cause semidominant, temperature-sensitive lethality at the twofold stage of embryogenesis. One is a glycine substitution that causes recessive, non–temperature-sensitive larval lethality. Three putative null alleles, two nonsense mutations and a deletion, all cause recessive, non–temperature-sensitive lethality at the threefold stage of embryogenesis. The less severe null phenotype indicates that glycine substitution containing mutant chains dominantly interfere with the function of other molecules. The emb-9 null mutants do not stain with anti–EMB-9 antisera and show intracellular accumulation of the α2(IV) chain, LET-2, indicating that LET-2 assembly and/or secretion requires EMB-9. Glycine substitutions in either EMB-9 or LET-2 cause intracellular accumulation of both chains. The degree of intracellular accumulation differs depending on the allele and temperature and correlates with the severity of the phenotype. Temperature sensitivity appears to result from reduced assembly/secretion of type IV collagen, not defective function in the basement membrane. Because the dominant interference of glycine substitution mutations is maximal when type IV collagen secretion is totally blocked, this interference appears to occur intracellularly, rather than in the basement membrane. We suggest that the nature of dominant interference caused by mutations in type IV collagen is different than that caused by mutations in fibrillar collagens.  相似文献   

14.
Collagen defects in lethal perinatal osteogenesis imperfecta.   总被引:15,自引:3,他引:12       下载免费PDF全文
Quantitative and qualitative abnormalities of collagen were observed in tissues and fibroblast cultures from 17 consecutive cases of lethal perinatal osteogenesis imperfecta (OI). The content of type I collagen was reduced in OI dermis and bone and the content of type III collagen was also reduced in the dermis. Normal bone contained 99.3% type I and 0.7% type V collagen whereas OI bone contained a lower proportion of type I, a greater proportion of type V and a significant amount of type III collagen. The type III and V collagens appeared to be structurally normal. In contrast, abnormal type I collagen chains, which migrated slowly on electrophoresis, were observed in all babies with OI. Cultured fibroblasts from five babies produced a mixture of normal and abnormal type I collagens; the abnormal collagen was not secreted in two cases and was slowly secreted in the others. Fibroblasts from 12 babies produced only abnormal type I collagens and they were also secreted slowly. The slower electrophoretic migration of the abnormal chains was due to enzymic overmodification of the lysine residues. The distribution of the cyanogen bromide peptides containing the overmodified residues was used to localize the underlying structural abnormalities to three regions of the type I procollagen chains. These regions included the carboxy-propeptide of the pro alpha 1(I)-chain, the helical alpha 1(I) CB7 peptide and the helical alpha 1(I) CB8 and CB3 peptides. In one baby a basic charge mutation was observed in the alpha 1(I) CB7 peptide and in another baby a basic charge mutation was observed in the alpha 1(I) CB8 peptide. The primary defects in lethal perinatal OI appear to reside in the type I collagen chains. Type III and V collagens did not appear to compensate for the deficiency of type I collagen in the tissues.  相似文献   

15.
Collagenase secretion was studied in cultures of rabbit articular chondrocytes. Differentiation of the cells was assessed by characterizing the type of 3H-labelled collagen produced during treatment with (1) conditioned media from rabbit peritoneal macrophages and human blood mononuclear cells, and (2) with retinol, a potent cartilage resorbing agent in tissue culture. Conditioned media stimulated collagenase secretion. Total collagen synthesis was reduced due to a decrease of synthesis of α1 chains; the amount of α2 chains synthesized was unchanged. This is thought to be due to a reduction in type II synthesis. Retinol did not stimulate collagenase secretion. Total collagen synthesis was reduced by retinol. α2 chain synthesis, however, was significantly increased, suggesting a switch of collagen synthesis in favor of type I collagen and, therefore, dedifferentiation. These results demonstrate that dedifferentiation of chondrocytes with respect to collagen synthesis is not necessarily associated with a stimulation of collagenase secretion.  相似文献   

16.
Activation of integrins upon binding to extracellular matrix proteins is believed to be a crucial step for the regulation of cell survival and proliferation. We have used integrin α1-null mice to investigate the role of this collagen receptor in the regulation of cell growth and survival in vivo. α1-deficient animals, which are viable and fertile, have a hypocellular dermis and a deficiency in dermal fibroblast proliferation as embryos. In vitro analysis of α1-null embryonic fibroblasts has revealed that their proliferation rate is markedly reduced when plated on collagenous substrata, despite normal attachment and spreading. Moreover, on the same collagenous matrices, α1-null fibroblasts fail to recruit and activate the adaptor protein Shc. The failure to activate Shc is accompanied by a downstream deficiency in recruitment of Grb2 and subsequent mitogen-activated protein kinase activation. Taken together with the growth deficiency observed on collagens, this finding indicates that the α1β1 is the sole collagen receptor which can activate the Shc mediated growth pathway. Thus, integrin α1 has a unique role among the collagen receptors in regulating both in vivo and in vitro cell proliferation in collagenous matrices.  相似文献   

17.
The methods of quantitating the relative amounts of type I and III collagens in samples containing crosslinked collagen chains were evaluated using electrophoresis of alpha chains and cyanogen bromide peptides. The densitometry areas of the alpha I(I) chains from type I collagen and the alpha I(III) chains from type III collagen were reduced because of the failure of the crosslinked chains to dissociate. However, the ratios of the unit densitometry areas of these chains (area of chain/micrograms type I or III collagen loaded) were constant for type I and III collagens prepared from the same samples of tissue. A calibration factor, which was the same for dermis and mitral valve, was derived to convert the densitometry area ratios to the weight ratios of type I to III collagens. In contrast, the densitometry areas of the alpha I(I) CB8 (type I collagen marker) and the alpha I(III) CB5 (type III collagen marker) were not reduced by crosslinked collagen chains. A calibration factor was also derived to convert the ratios of the densitometry areas of the marker peptides to weight ratios of type I to type III collagens. Almost identical results were obtained when electrophoresis of alpha chains and of cyanogen bromide peptides was used with these calibration factors to quantitate the relative amounts of type I and III collagens in tissue extracts which contained different amounts of crosslinked chains.  相似文献   

18.
The GFOGER motif in collagens (O denotes hydroxyproline) represents a high-affinity binding site for all collagen-binding integrins. Other GxOGER motifs require integrin activation for maximal binding. The E318W mutant of the integrin α2β1 I domain displays a relaxed collagen specificity, typical of an active state. E318W binds more strongly than the wild-type α2 I domain to GMOGER, and forms a 2:1 complex with a homotrimeric, collagen-like, GFOGER peptide. Crystal structure analysis of this complex reveals two E318W I domains, A and B, bound to a single triple helix. The E318W I domains are virtually identical to the collagen-bound wild-type I domain, suggesting that the E318W mutation activates the I domain by destabilising the unligated conformation. E318W I domain A interacts with two collagen chains similarly to wild-type I domain (high-affinity mode). E318W I domain B makes favourable interactions with only one collagen chain (low-affinity mode). This observation suggests that single GxOGER motifs in the heterotrimeric collagens V and IX may support binding of activated integrins.  相似文献   

19.
The role of the non-helical regions of the collagen molecule in fibrillogenesis has been investigated by comparing the kinetics of fibril formation of pepsin-treated acid-soluble collagen, acid-soluble collagen and mixtures of the two and by comparison of the thermal stabilities of the fibrils formed. The acid-soluble collagen was found to aggregate more rapidly than the pepsin-treated collagen under physiological conditions of pH and ionic strength. Variations in ionic strength, at physiological pH, were found to have differing effects on the aggregation of these two forms of soluble collagen. Fibrils formed from the pepsinized-collagen had a lower thermal stability tha n those formed from the intact collagen. The behavior observed with mixtures of acid-soluble and pepsin-treated collagens was found to be quantitatively consistent with the pepsinized collagen being able to utilize the nuclei formed by the acid-soluble collagen for subsequent growth. However, the use of the acid-soluble nuclei by the pepsinized collagen for growth did not enhance its rate of precipitation during the growth phase, nor did it enhance the thermal stability of the fibrils formed from the pepsinized collagen.  相似文献   

20.
Kidney glomerular basement membranes (GBMs) undergo laminin and type IV collagen isoform substitutions during glomerular development, which are believed to be required for maturation of the filtration barrier. Specifically, GBMs of earliest glomeruli contain laminin α1β1γ1 and collagen α1α2α1(IV), whereas mature glomeruli contain laminin α5β2γ1 and collagen α3α4α5(IV). Here, we used confocal microscopy to simultaneously evaluate expression of different laminin and collagen IV isoforms in newborn mouse GBMs. Our results show loss of laminin α1 from GBMs in early capillary loop stages and continuous linear deposition of laminin bearing the α5 chain thereafter. In contrast, collagen α1α2α1(IV) persisted in linear patterns into late capillary loop stages, when collagen α3α4α5(IV) first appeared in discontinuous, non-linear patterns. This patchy pattern for collagen α3α4α5(IV) continued into maturing glomeruli where there were lengths of linear, laminin α5-positive GBM entirely lacking either isoform of collagen IV. Relative abundance of laminin and collagen IV mRNAs in newborn and 5-week-old mouse kidneys also differed, with those encoding laminin α1, α5, β1, β2, and γ1, and collagen α1(IV) and α2(IV) chains all significantly declining at 5 weeks, but α3(IV) and α4(IV) were significantly upregulated. We conclude that different biosynthetic mechanisms control laminin and type IV collagen expression in developing glomeruli.  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