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 328 毫秒
1.
A full-length cDNA of the Type I procollagen alpha1 [pro-alpha1(I)] chain (4388 bp), coding for 1463 amino acid residues in the total length, was determined by RACE PCR using a cDNA library constructed from 4-week embryo of the skate Raja kenojei. The helical region of the skate pro-alpha1(I) chain consisted of 1014 amino acid residues - the same as other fibrillar collagen alpha chains from higher vertebrates. Comparison on denaturation temperatures of Type I collagens from the skate, rainbow trout (Oncorhynchus mykiss) and rat (Rattus norvegicus) revealed that the number of Gly-Pro-Pro and Gly-Gly in the alpha1(I) chains could be directly related to the thermal stability of the helix. The expression property of the skate pro-alpha1(I) chain mRNA and phylogenetic analysis with other vertebrate pro-alpha1(I) chains suggested that skate pro-alpha1(I) chain could be a precursor form of the skate Type I collagen alpha1 chain. The present study is the first evidence for the primary structure of full-length pro-alpha1(I) chain in an elasmobranch.  相似文献   

2.
3.
E Crouch  P Bornstein 《Biochemistry》1978,17(25):5499-5509
Second trimester human amniotic fluid cells synthesize and secrete a variety of collagenous proteins in culture. F cells (amniotic fluid fibroblasts) are the most active biosynthetically and synthesize predominantly type I with smaller amounts of type III procollagen. Epithelioid AF cells (the predominating clonable cell type) synthesize a type IV-like procollagen and a procollagen with three identical proalpha chains, structurally and immunologically related to the proalpha1 chains of type I procollagen. The latter procollagen, when cleaved with pepsin and denatured, yields a single non-disulfide-bonded alpha chain that migrates more slowly than F cell or human skin alpha1(I) on sodium dodecyl sulfate-polyacrylamide gel electrophoresis but coelutes with these chains from carboxymethyl-cellulose. The major cyanogen bromide produced peptides demonstrate a similar behavior relative to peptides derived from alpha1(I). The collagen is characterized by an increased solubility at neutral pH and high ionic strength, relative to type I collagen. The amino acid composition of the pepsin-resistant alpha chain is essentially identical with that of human alpha1(I), except for marked increases in the content of 3- and 4-hydroxyproline and hydroxylysine. Preliminary experiments suggest that these increased posttranslational modifications are responsible for the unusually slow migration of this collagen and its cyanogen bromide peptides on sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The procollagen has, therefore, been assigned the chain composition [proalpha1(I)]3. Like type I procollagen, [proalpha1(I)]3 undergoes a time-dependent conversion, in the medium and cell layer, to procollagen intermediates and alpha chains. The production of [proalpha1(I)]3 probably reflects the state of differentiation and/or embryologic derivation of AF cells rather than a characteristic of the fetal phenotype, since F cells do not synthesize significant amounts of the procollagen.  相似文献   

4.
The oim mouse is a model of human Osteogenesis Imperfecta (OI) that has deficient synthesis of proalpha2(I) chains. Cells isolated from oim mice synthesize alpha1(I) collagen homotrimers that accumulate in tissues. To explore the feasibility of gene therapy for OI, a murine proalpha2(I) cDNA was inserted into an adenovirus vector and transferred into bone marrow stromal cells isolated from oim mice femurs. The murine cDNA under the control of the cytomegalovirus early promoter was expressed by the transduced cells. Analysis of the collagens synthesized by the transduced cells demonstrated that the cells synthesized stable type I collagen comprised of alpha1(I) and alpha2(I) heterotrimers in the correct ratio of 2:1. The collagen was efficiently secreted and also the cells retained the osteogenic potential as indicated by the expression of alkaline phosphatase activity when the transduced cells were treated with recombinant human bone morphogenetic protein 2. Injection of the virus carrying the murine proalpha2(I) cDNA into oim skin demonstrated synthesis of type I collagen comprised of alpha1 and alpha2 chains at the injection site. These preliminary data demonstrate that collagen genes can be transferred into bone marrow stromal cells as well as fibroblasts in vivo and that the genes are efficiently expressed. These data encourage further studies in gene replacement for some forms of OI and use of bone marrow stromal cells as vehicles to deliver therapeutic genes to bone.  相似文献   

5.
We previously demonstrated that insect cells were able to synthesize recombinant human procollagen I as triple-helical heterotrimers when transfected with cDNAs of both proalpha1(I) and proalpha2(I) chains. However, most of the heterotrimers were retained within the cells, unlike in the case of mammalian cells [Tomita, M., Kitajima, T., and Yoshizato, K. (1997) J. Biochem. 1061-1069]. In an attempt to improve the secretion of the heterotrimers, we introduced the putative collagen-specific chaperone HSP47 into this insect expression model. Mouse HSP47 produced by the insect cells bound intracellularly to both human proalpha1(I) and proalpha2(I) chains and enhanced the secretion of procollagen I heterotrimers. HSP47 was also coexpressed with either proalpha1(I) chains or proalpha2(I) chains, which showed that it enhanced the secretion of the former but not the latter. This selective effect of HSP47 was similarly observed in the cells treated with inhibitors of procollagen triple helix formation, indicating that HSP47 can also accelerate the secretion of non-helical procollagens. HSP47 did not change the intracellular solubility of proalpha1(I) and proalpha2(I) chains in 1% NP-40, eliminating the possibility that it prevents proalpha chains from aggregating into insoluble forms within the insect cells. We concluded that HSP47 can play a role in the secretion of alpha1(I)-procollagen chains in the insect cell model. The present study also demonstrated the dissimilarity in the mechanism of folding and secretion of the expressed procollagen I between the insect and mammalian cells.  相似文献   

6.
Overlapping cDNA clones were isolated for human type II procollagen. Nucleotide sequencing of the clones provided over 2.5 kb of new coding sequences for the human pro alpha 1(II) gene and the first complete amino acid sequence of type II procollagen from any species. Comparison with published data for cDNA clones covering the entire lengths of the human type I and type III procollagens made it possible to compare in detail the coding sequences and primary structures of the three most abundant human fibrillar collagens. The results indicated that the marked preference in the third base codons for glycine, proline and alanine previously seen in other fibrillar collagens was maintained in type II procollagen. The domains of the pro alpha 1(II) chain are about the same size as the same domains of the pro alpha chains of type I and type III procollagens. However, the major triple-helical domain is 15 amino acid residues less than the triple-helical domain of type III procollagen. Comparison of hydropathy profiles indicated that the alpha chain domain of type II procollagen is more similar to the alpha chain domain of the pro alpha 1(I) chain than to the pro alpha 2(I) chain or the pro alpha 1(III) chain. The results therefore suggest that selective pressure in the evolution of the pro alpha 1(II) and pro alpha 1(I) genes is more similar than the selective pressure in the evolution of the pro alpha 2(I) and pro alpha 1(III) genes.  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.
Type I procollagen was purified from cultured fibroblasts of a proband with a lethal variant of osteogenesis imperfecta. The protein was a mixture of normal procollagen and mutated procollagens containing a substitution of cysteine for glycine in either one pro alpha 1(I) chain or both pro alpha 1(I) chains, some or all of which were disulfide-linked through the cysteine at position alpha 1-748. The procollagen was then examined in a system for generating collagen fibrils de novo by cleavage of the pCcollagen to collagen with procollagen C-proteinase [Kadler et al. (1987) J. Biol. Chem. 262, 15696-15701]. The mutated collagens and normal collagens were found to form copolymers under a variety of experimental conditions. With two preparations of the protein that had a high content of alpha 1(I) chains disulfide-linked through the cysteine alpha 1-748, all the large structures formed had a distinctive, highly branched morphology that met one of the formal criteria for a fractal. Preparations with a lower content of disulfide-linked alpha 1(I) chains formed fibrils that were 4 times the diameter of control fibrils. The formation of copolymers was also demonstrated by the observation that the presence of mutated collagens decreased the rate of incorporation of normal collagen into fibrils. In addition, the solution-phase concentration at equilibrium of mixtures of mutated and normal collagens was 5-10-fold greater than that of normal collagen.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
We have isolated three overlapping cDNA clones encoding the pro alpha 2(XI) collagen chain from a human chondrocyte cDNA library. Together, the cDNAs code for 257 uninterrupted Gly-X-Y triplets (almost 80% of the triple helical domain) and about 200 amino acid residues of the carboxyl telopeptide and carboxyl propeptide. The identification of the clones as pro alpha 2(XI) cDNAs was based on the complete identity between the amino acid sequences of three tryptic peptides derived from human alpha 2(XI) collagen and the cDNA-derived sequence. We have also sequenced six exons within a human genomic alpha 2(XI) cosmid clone. This sequence shows that although type XI collagen belongs to the fibril-forming class of collagens, there are substantial differences in exon sizes at the 3' end of the gene when comparing the alpha 2(XI) gene with those of human types I, II, and III collagens. Finally, pro alpha 2(XI) cDNA has been used as a probe to determine the location of the gene by in situ hybridization of chromosome spreads. The results demonstrate that the gene is located close to the region p212 on chromosome 6. Northern blot analysis shows that the gene is expressed in cartilage but not in adult liver, skin, and tendon.  相似文献   

10.
11.
[Alpha 1(III)]3 collagen was solubilized by pepsin digestion of normal human placental membranes and was purified by differential salt precipitation and carboxymethylcellulose chromatography. This collagen was digested with CNBr, and the resultant nine peptides were isolated and characterized. The chains are cross-linked by cysteinyl residues in the COOH-terminal peptide. Isolation of peptides derived from CNBr digestion of insoluble tissues was used as an assay for the presence of [alpha 1(I)]2alpha 2 and [alpha 1(III)]3 collagens. Both types are present in human skin, intestine, liver, spleen, kidney, lung, aorta, umbilical cord, placental membranes, and myocardium. Bone and tendon contain [alpha 1(I)]2alpha 2 collagen but, unlike the other tissues, lack [alpha 1(III)]3 collagen. Both [alpha 1(I)]2alpha 2 and[alpha 1(III)]3 collagens are present in scars of human skin, myocardium, tendon, and liver and of rabbit skin. The degree of hydroxylation of proline was 4 to 5% lower in the same peptides in skin, bone, and tendon than in the other tissues. The degree of hydroxylation of lysine in the same peptides derived from different tissues varied more widely.  相似文献   

12.
We performed in vivo dimethylsulfate footprinting of the 220 bp mouse proximal proalpha1(I) collagen promoter and the 350 bp mouse proximal proalpha2(I) collagen promoter in BALB/3T3 fibroblasts, primary mouse skin fibroblasts, S-194 B cells, NMuLi liver epithelial cells and RAG renal adenocarcinoma cells and in vitro DNase I footprinting of these promoters using nuclear extracts of these different cell types. Whereas proalpha1(I) and proalpha2(I) collagen RNAs were present in BALB/3T3 fibroblasts and primary fibroblasts, these RNAs could not be detected in the three other cell lines. Comparison of in vitro DNase I footprints for each of the two proximal collagen promoters indicated that the patterns of protection were very similar with the different nuclear extracts, suggesting that the DNA binding proteins binding to these promoters were present in all cell types tested. In contrast, in vivo footprints over these proximal promoters were cell-specific, occurring only in fibroblast cells and not in the other three cell types. The in vivo footprints were generally located within the in vitro footprinted regions. Our results suggest that although all cell types tested contained nuclear proteins that can bind to the proximal proalpha1(I) and proalpha2(I) collagen promoters in vitro , it is only in fibroblasts that these proteins bind to their cognate sites in vivo . We discuss possible regulatory mechanisms in type I collagen genes that can contribute to the cell-specific in vivo protein-DNA interactions at the proximal promoters.  相似文献   

13.
Mapping of SPARC/BM-40/osteonectin-binding sites on fibrillar collagens   总被引:10,自引:0,他引:10  
The 33-kDa matrix protein SPARC (BM-40, osteonectin) binds several collagen types with moderate affinity. The collagen-binding site resides in helix alphaA of the extracellular calcium-binding domain of SPARC and is partially masked by helix alphaC. Previously, we found that the removal of helix alphaC caused a 10-fold increase in the affinity of SPARC for collagen, and we identified amino acids crucial for binding by site-directed mutagenesis. In this study, we used rotary shadowing, CNBr peptides, and synthetic peptides to map binding sites of SPARC onto collagens I, II, and III. Rotary shadowing and electron microscopy of SPARC-collagen complexes identified a major binding site approximately 180 nm from the C terminus of collagen. SPARC binding was also detected with lower frequency near the matrix metalloproteinase cleavage site. These data fit well with our analysis of SPARC binding to CNBr peptides, denaturation of which abolished binding, indicating triple-helical conformation of collagen to be essential. SPARC binding was substantially decreased in two of seven alpha2(I) mutant procollagen I samples and after N-acetylation of Lys/Hyl side chains in wild-type collagen. Synthetic peptides of collagen III were used to locate the binding sites, and we found SPARC binding activity in a synthetic triple-helical peptide containing the sequence GPOGPSGPRGQOGVMGFOGPKGNDGAO (where O indicates 4-hydroxyproline), with affinity for SPARC comparable with that of procollagen III. This sequence is conserved among alpha chains of collagens I, II, III, and V. In vitro collagen fibrillogenesis was delayed in the presence of SPARC, suggesting that SPARC might modulate collagen fibril assembly in vivo.  相似文献   

14.
Collagen synthesis was examined in skin fibroblasts from a patient with a variant of Ehlers-Danlos syndrome. The relative rate of collagen synthesis to total protein synthesis in the patient's fibroblasts was always one-half of that in fibroblasts from normal controls. Total collagen synthesis, as assessed by quantification of total hydroxyproline, was also significantly lower than that of controls, indicating that the rate of collagen synthesis by the patient's fibroblasts was decreased compared with that by normal fibroblasts. Analysis of procollagen and collagen components showed the absence of the pro alpha 2(I) chain and its derivatives. Dot-blot and Northern-blot analyses showed the patient's fibroblasts to contain less than 10% of the mRNAs for pro alpha 2(I) found in control fibroblasts. In spite of these results, Southern blot analysis of genomic DNA indicated the presence of the same number of genes for the pro alpha 2(I) collagen chain in the patient's fibroblasts as in control fibroblasts, suggesting malfunctioning pro alpha 2(I) collagen genes as the cause for failure of the patient's fibroblasts to synthesize pro alpha 2(I) collagen chains.  相似文献   

15.
We have examined the collagenous proteins extracted from skin and produced by skin fibroblast cultures from the members of a family with mild dominant osteogenesis imperfecta (OI type I). The two affected patients, mother and son, produce two populations of alpha 1(I) chains of type I collagen, one chain being normal, the other containing a cysteine within the triple-helical domain. Both forms can be incorporated into triple-helical molecules with an alpha 2(I) chain. When two mutant alpha (I) chains are incorporated into the same molecule, a disulfide bonded dimer is produced. We have characterized these chains by sodium dodecyl sulfate-gel electrophoresis and CNBr-peptide mapping and by measuring a number of biosynthetic and physical variables. The cysteine was localized to the COOH-terminal peptide alpha (I) CB6. Molecules containing the mutant chains are stable, have a normal denaturation temperature, are secreted normally, and have normal levels of post-translational modification of lysyl residues and intracellular degradation. We have compared and contrasted these observations with those made in a patient with lethal osteogenesis imperfecta in which there was a cysteine substitution in alpha 1(I) CB6 (Steinmann, B., Rao, V. H., Vogel, A., Bruckner, P., Gitzelmann, R., and Byers, P. H. (1984) J. Biol. Chem 259, 11129-11138) and have concluded that the mutation in the present family occurs in the X or Y position of a Gly-X-Y repeating unit of collagen and not in the glycine position shown for the previous patient (Cohn, D. H., Byers, P. H., Steinmann, B, and Gelinas, R. E. (1986) Proc. Natl. Acad. Sci. U. S. A., in press.  相似文献   

16.
We have isolated several overlapping cDNA clones encoding alpha 1(XI) collagen chains from human and rat cDNA libraries. Together the human cDNAs code for 335 uninterrupted Gly-X-Y triplets, and a 264-amino acid C-propeptide, while the rat cDNAs cover the entire C-propeptide and about a third of the triple-helical domain. Comparison of the human and rodent nucleotide sequences showed a 95% sequence similarity. The identification of the clones as alpha 1(XI) cDNAs was based on the complete identity between the amino acid sequences of three human alpha 1(XI) cyanogen bromide peptides and the cDNA-derived sequence. Examination of and the cDNA-derived amino acid sequence showed a variety of structural features characteristic of fibrillar-forming collagens. In addition, nucleotide sequence analysis of a selected portion of the corresponding human gene revealed the characteristic 54-base pair exon motif. We conclude therefore that pro-alpha 1 (XI) collagen belongs to the group of fibrillar collagen genes. We also suggest that the expression of this gene is not restricted to cartilage, as previously thought, since the cDNA libraries from which the clones were isolated, originated from both cartilagenous and noncartilaginous tissues.  相似文献   

17.
We have previously assigned an integrin alpha(2)beta(1)-recognition site in collagen I to the sequence, GFOGERGVEGPOGPA (O = Hyp), corresponding to residues 502-516 of the alpha(1)(I) chain and located in the fragment alpha(1)(I)CB3 (Knight, C. G., Morton, L. F., Onley, D. J., Peachey, A. R., Messent, A. J., Smethurst, P. A., Tuckwell, D. S., Farndale, R. W., and Barnes, M. J. (1998) J. Biol. Chem. 273, 33287-33294). In this study, we show that recognition is entirely contained within the six-residue sequence GFOGER. This sequence, when in triple-helical conformation, readily supports alpha(2)beta(1)-dependent cell adhesion and exhibits divalent cation-dependent binding of isolated alpha(2)beta(1) and recombinant alpha(2) A-domain, being at least as active as the parent collagen. Replacement of E by D causes loss of recognition. The same sequence binds integrin alpha(1) A-domain and supports integrin alpha(1)beta(1)-mediated cell adhesion. Triple-helical GFOGER completely inhibits alpha(2) A-domain binding to collagens I and IV and alpha(2)beta(1)-dependent adhesion of platelets and HT 1080 cells to these collagens. It also fully inhibits alpha(1) A-domain binding to collagen I and strongly inhibits alpha(1)beta(1)-mediated adhesion of Rugli cells to this collagen but has little effect on either alpha1 A-domain binding or adhesion of Rugli cells to collagen IV. We conclude that the sequence GFOGER represents a high-affinity binding site in collagens I and IV for alpha(2)beta(1) and in collagen I for alpha(1)beta(1). Other high-affinity sites in collagen IV mediate its recognition of alpha(1)beta(1).  相似文献   

18.
1. Skin Type I collagen was isolated from 15 species of teleosts. 2. Chromatographic and electrophoretic analyses revealed that most of the teleost skin collagens possessed a unique subunit, alpha 3, which had not been detected in other vertebrate Type I collagens. 3. The skin collagen seems to exist as an alpha 1 alpha 2 alpha 3 heterotrimer in many teleosts and as an (alpha 1)2 alpha 2 heterotrimer in some teleosts.  相似文献   

19.
Synthesis of procollagen was examined in skin fibroblasts from a patient with a moderately severe autosomal dominant form of osteogenesis imperfecta. Proteolytic removal of the propeptide regions of newly synthesized procollagen, followed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis under nonreducing conditions, revealed the presence of type I collagen in which two alpha 1(I) chains were linked through interchain disulfide bonds. Fragmentation of the disulfide-bonded alpha 1(I) dimers with vertebrate collagenase and cyanogen bromide demonstrated the presence of a cysteine residue in alpha 1(I)CB8, a fragment containing amino acid residues 124-402 of the alpha 1(I) collagen chain. Cysteine residues are not normally found in the triple-helical domain of type I collagen chains. The heterozygous nature of the molecular defect resulted in the formation of three kinds of type I trimers: a normal type with normal pro-alpha(I) chains, a type I trimer with one mutant pro-alpha 1(I) chain and two normal chains, and a type I trimer containing two mutant pro-alpha 1(I) chains and one normal pro-alpha 2(I) chain. The presence of one or two mutant pro-alpha 1(I) chains in trimers of type I procollagen was found to reduce the thermal stability of the protein by 2.5 and 1 degree C, respectively. In addition to post-translational overmodification, procollagen containing one mutant pro-alpha 1(I) chain was also cleared more slowly from cultured fibroblasts. The most likely explanation for these disruptive changes in the physical stability and secretion of the mutant procollagen is that a cysteine residue is substituted for a glycine in half of the pro-alpha 1(I) chains synthesized by the patient's fibroblasts.  相似文献   

20.
Because of its unique physical and chemical properties, rat tail tendon collagen has long been favored for crystallographic and biochemical studies of fibril structure. In studies of the distribution of 3-hydroxyproline in type I collagen of rat bone, skin, and tail tendon by mass spectrometry, the repeating sequences of Gly-Pro-Pro (GPP) triplets at the C terminus of α1(I) and α2(I) chains were shown to be heavily 3-hydroxylated in tendon but not in skin and bone. By isolating the tryptic peptides and subjecting them to Edman sequence analysis, the presence of repeating 3-hydroxyprolines in consecutive GPP triplets adjacent to 4-hydroxyproline was confirmed as a unique feature of the tendon collagen. A 1960s study by Piez et al. (Piez, K. A., Eigner, E. A., and Lewis, M. S. (1963) Biochemistry 2, 58-66) in which they compared the amino acid compositions of rat skin and tail tendon type I collagen chains indeed showed 3-4 residues of 3Hyp in tendon α1(I) and α2(I) chains but only one 3Hyp residue in skin α1(I) and none in α2(I). The present work therefore confirms this difference and localizes the additional 3Hyp to the GPP repeat at the C terminus of the triple-helix. We speculate on the significance in terms of a potential function in contributing to the unique assembly mechanism and molecular packing in tendon collagen fibrils and on mechanisms that could regulate 3-hydroxylation at this novel substrate site in a tissue-specific manner.  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