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 15 毫秒
1.
The mitochondrial respiratory cytochrome contents of dormant and germinating conidia of Botryodiplodia theobromae were examined. Oxidized versus reduced difference spectra at 77 degrees K of whole mitochondria from physiologically mature germinated spores showed a typical a-band pattern for cytochromes c, b, and a, with absorption maxima at 549, 554 + 559, and 604 nm, respectively, whereas the difference spectrum of the counterpart mitochondrial fraction from dormant spores showed no cytochrome a bands. However, a fraction prepared from dormant spore mitochondria by detergent extraction and (NH4)2SO4 fractionation contained readily detectable quantities of cytochromes c and b (as shown by the a and Soret absorption bands), but it did not contain the a or Soret bands of cytochrome a observed in a counterpart preparation from germinated spores. The pyridine hemochromogen preparation from the dormant spore mitochondria contained no material that is spectroscopically characteristic of a-type heme and protoheme. These results suggest that cytochrome a is not present as a functional molecule in dormant spores. The first spectroscopically detectable cytochromes were observed in whole mitochondria at 210 min of spore germination, and the amount of each of the cytochromes increased with cell growth. A precursor of the heme porphyrin, delta-[4-14C]aminolevulinic acid, was first incorporated (at accelerating rates) into acid-insoluble spore material at 180 min of germination, which appears to be the approximate time of organization of new mitochondria in these spores.  相似文献   

2.
The carboxy-terminal region of dystrophin has been suggested to be crucially important for its function to prevent muscle degeneration. We have previously shown that this region is the locus that interacts with the sarcolemmal glycoprotein complex, which mediates membrane anchoring of dystrophin, as well as with the cytoplasmic peripheral membrane protein, A0 and beta 1-syntrophin (Suzuki, A., M. Yoshida, K. Hayashi, Y. Mizuno, Y. Hagiwara, and E. Ozawa. 1994. Eur. J. Biochem. 220:283- 292). In this work, by using the overlay assay technique developed previously, we further analyzed the dystrophin-syntrophin/A0 interaction. Two forms of mammalian syntrophin, alpha 1- and beta 1- syntrophin, were found to bind to very close but discrete regions on the dystrophin molecule. Their binding sites are located at the vicinity of the glycoprotein-binding site, and correspond to the amino acid residues encoded by exons 73-74 which are alternatively spliced out in some isoforms. This suggests that the function of syntrophin is tightly linked to the functional diversity among dystrophin isoforms. Pathologically, it is important that the binding site for alpha 1- syntrophin, which is predominantly expressed in skeletal muscle, coincides with the region whose deletion was suggested to result in a severe phenotype. In addition, A0, a minor component of dystrophin- associated proteins with a molecular mass of 94 kD which is immunochemically related to syntrophin, binds to the same site as beta 1-syntrophin. Finally, based on our accumulated evidence, we propose a revised model of the domain organization of dystrophin from the view point of protein-protein interactions.  相似文献   

3.

Background

In dystrophic mdx skeletal muscle, aberrant Ca2+ homeostasis and fibre degeneration are found. The absence of dystrophin in models of Duchenne muscular dystrophy (DMD) has been connected to altered ion channel properties e.g. impaired L-type Ca2+ currents. In regenerating mdx muscle, ‘revertant’ fibres restore dystrophin expression. Their functionality involving DHPR-Ca2+-channels is elusive.

Methods and Results

We developed a novel ‘in-situ’ confocal immuno-fluorescence and imaging technique that allows, for the first time, quantitative subcellular dystrophin-DHPR colocalization in individual, non-fixed, muscle fibres. Tubular DHPR signals alternated with second harmonic generation signals originating from myosin. Dystrophin-DHPR colocalization was substantial in wt fibres, but diminished in most mdx fibres. Mini-dystrophin (MinD) expressing fibres successfully restored colocalization. Interestingly, in some aged mdx fibres, colocalization was similar to wt fibres. Most mdx fibres showed very weak membrane dystrophin staining and were classified ‘mdx-like’. Some mdx fibres, however, had strong ‘wt-like’ dystrophin signals and were identified as ‘revertants’. Split mdx fibres were mostly ‘mdx-like’ and are not generally ‘revertants’. Correlations between membrane dystrophin and DHPR colocalization suggest a restored putative link in ‘revertants’. Using the two-micro-electrode-voltage clamp technique, Ca2+-current amplitudes (imax) showed very similar behaviours: reduced amplitudes in most aged mdx fibres (as seen exclusively in young mdx mice) and a few mdx fibres, most likely ‘revertants’, with amplitudes similar to wt or MinD fibres. Ca2+ current activation curves were similar in ‘wt-like’ and ‘mdx-like’ aged mdx fibres and are not the cause for the differences in current amplitudes. imax amplitudes were fully restored in MinD fibres.

Conclusions

We present evidence for a direct/indirect DHPR-dystrophin interaction present in wt, MinD and ‘revertant’ mdx fibres but absent in remaining mdx fibres. Our imaging technique reliably detects single isolated ‘revertant’ fibres that could be used for subsequent physiological experiments to study mechanisms and therapy concepts in DMD.  相似文献   

4.
SYNOPSIS. The process of conjugation in a strain of Stylonychia muscorum Kahl is described. It follows the general pattern reported earlier in other Oxytrichidae, but the following peculiarities can be noted: 1) mating never occurs between individuals possessing only two micronuclei, 2)exchange of small macronuclear lobes can be occasionally observed, and 3) the reorganization of the exconjugants involves the quick, successive passage of a series of reorganization bands.  相似文献   

5.
α-Dystrobrevin is both a dystrophin homologue and a component of the dystrophin protein complex. Alternative splicing yields five forms, of which two predominate in skeletal muscle: full-length α-dystrobrevin-1 (84 kD), and COOH-terminal truncated α-dystrobrevin-2 (65 kD). Using isoform-specific antibodies, we find that α-dystrobrevin-2 is localized on the sarcolemma and at the neuromuscular synapse, where, like dystrophin, it is most concentrated in the depths of the postjunctional folds. α-Dystrobrevin-2 preferentially copurifies with dystrophin from muscle extracts. In contrast, α-dystrobrevin-1 is more highly restricted to the synapse, like the dystrophin homologue utrophin, and preferentially copurifies with utrophin. In yeast two-hybrid experiments and coimmunoprecipitation of in vitro–translated proteins, α-dystrobrevin-2 binds dystrophin, whereas α-dystrobrevin-1 binds both dystrophin and utrophin. α-Dystrobrevin-2 was lost from the nonsynaptic sarcolemma of dystrophin-deficient mdx mice, but was retained on the perisynaptic sarcolemma even in mice lacking both utrophin and dystrophin. In contrast, α-dystrobrevin-1 remained synaptically localized in mdx and utrophin-negative muscle, but was absent in double mutants. Thus, the distinct distributions of α-dystrobrevin-1 and -2 can be partly explained by specific associations with utrophin and dystrophin, but other factors are also involved. These results show that alternative splicing confers distinct properties of association on the α-dystrobrevins.  相似文献   

6.
γ-Sarcoglycan is a transmembrane, dystrophin-associated protein expressed in skeletal and cardiac muscle. The murine γ-sarcoglycan gene was disrupted using homologous recombination. Mice lacking γ-sarcoglycan showed pronounced dystrophic muscle changes in early life. By 20 wk of age, these mice developed cardiomyopathy and died prematurely. The loss of γ-sarcoglycan produced secondary reduction of β- and δ-sarcoglycan with partial retention of α- and ε-sarcoglycan, suggesting that β-, γ-, and δ-sarcoglycan function as a unit. Importantly, mice lacking γ-sarco- glycan showed normal dystrophin content and local- ization, demonstrating that myofiber degeneration occurred independently of dystrophin alteration. Furthermore, β-dystroglycan and laminin were left intact, implying that the dystrophin–dystroglycan–laminin mechanical link was unaffected by sarcoglycan deficiency. Apoptotic myonuclei were abundant in skeletal muscle lacking γ-sarcoglycan, suggesting that programmed cell death contributes to myofiber degeneration. Vital staining with Evans blue dye revealed that muscle lacking γ-sarcoglycan developed membrane disruptions like those seen in dystrophin-deficient muscle. Our data demonstrate that sarcoglycan loss was sufficient, and that dystrophin loss was not necessary to cause membrane defects and apoptosis. As a common molecular feature in a variety of muscular dystrophies, sarcoglycan loss is a likely mediator of pathology.  相似文献   

7.
1. Thirty-eight steroids were tested as substrates for a 7 alpha-hydroxy steroid dehydrogenase preparation from a strain of Escherichia coli; an improved method of making the crude enzyme is described. 2. Steroids having a 7 alpha-hydroxyl group in the molecule were substrates except (a) when the 5 beta-cholan-24-oic acid side chain was shortened to less than four carbon atoms and (b) in certain cases when sulphate ester groups were present in the molecule. 3. For testing with the enzyme, a new specimen of 7 alpha-hydroxy-3,12-dioxo-5 beta-cholan-24-oic acid was made, which had properties different from those previously described.  相似文献   

8.
Fukutin-related protein (FKRP) is a protein involved in the glycosylation of cell surface molecules. Pathogenic mutations in the FKRP gene cause both the more severe congenital muscular dystrophy Type 1C and the milder Limb-Girdle Type 2I form (LGMD2I). Here we report muscle histological alterations and the analysis of 11 muscle proteins: dystrophin, four sarcoglycans, calpain 3, dysferlin, telethonin, collagen VI, α-DG, and α2-laminin, in muscle biopsies from 13 unrelated LGMD2I patients with 10 different FKRP mutations. In all, a typical dystrophic pattern was observed. In eight patients, a high frequency of rimmed vacuoles was also found. A variable degree of α2-laminin deficiency was detected in 12 patients through immunofluorescence analysis, and 10 patients presented α-DG deficiency on sarcolemmal membranes. Additionally, through Western blot analysis, deficiency of calpain 3 and dystrophin bands was found in four and two patients, respectively. All the remaining proteins showed a similar pattern to normal controls. These results suggest that, in our population of LGMD2I patients, different mutations in the FKRP gene are associated with several secondary muscle protein reductions, and the deficiencies of α2-laminin and α-DG on sections are prevalent, independently of mutation type or clinical severity. (J Histochem Cytochem 56:995–1001, 2008)  相似文献   

9.
10.
Caveolin, a 20-24 kDa integral membrane protein, is a principal component of caveolar domains. Caveolin-1 is expressed predominantly in endothelial cells, fibroblasts, and adipocytes, while the expression of caveolin-3 is confined to muscle cells. However, their localization in various muscles has not been well documented. Using double-immunofluorescence labeling and confocal laser microscopy, we examined the localization of caveolins-1 and 3 in adult monkey skeletal, cardiac and uterine smooth muscles and the co-immunolocalization of these caveolins with dystrophin, which is a product of the Duchenne muscular dystrophy gene. In the skeletal muscle tissue, caveolin-3 was localized along the sarcolemma except for the transverse tubules, and co-immunolocalized with dystrophin, whereas caveolin-1 was absent except in the blood vessels of the muscle tissue. In cardiac muscle cells, caveolins-1 and -3 and dystrophin were co-immunolocalized on the sarcolemma and transverse tubules. In uterine smooth muscle cells, caveolin-1, but not caveolin-3, was co-immunolocalized with dystrophin on the sarcolemma.  相似文献   

11.
曲霉属内黑曲霉(Aspergitlus niger)与米曲霉(A.oryzae)具有特征明显不同的可溶性蛋白质电泳图谱,其种间杂种具有双亲的部分或全部电泳带并与黑曲霉相近。来自杂种Ⅰ的多数分离子电泳带与黑曲霉相近,只有一个分离子产生米曲霉的电泳带并具有米曲霉的遗传特性。青霉属内产黄青霉(Penicillium chrysogenum)与展青霉(P.patulum)种间及种内不同菌株间的电泳图谱基本相同,种内或种间杂种具有双亲的电泳带。结果讨论了蛋白质图谱分析的意义。  相似文献   

12.
Purified membranes were prepared from seven human T-mycoplasmas. Their amino acid composition was determined and was similar to that of other biological membranes. Their proteins were examined by isoelectric focusing and 7 to 10 protein bands were detected mostly in the pH 4 to 7 range of the gel. T-mycoplasma T-McA also had one distinctive band at pH 3 while T-mycoplasma T-213 had a prominent basic band at pH 9. The proteins were denatured by storage at -25 degrees. Five to seven Periodic Schiff-positive bands also were observed but were not identified.  相似文献   

13.
We investigated proteolytic susceptibility of the central domain in dystrophin molecules from chicken smooth and skeletal muscles. Dystrophin-enriched preparations from both muscles were made as described in Pons et al. (Proc. Natl. Acad. Sci. USA (1990) 87, 7851-7855). These preparations contained other protein components in addition to dystrophin. Three enzymes (Staphylococcus aureus proteinase, chymotrypsin and trypsin) having different proteolytic specificities were used. Time-courses of proteinase degradation were examined by the Western immunoblot technique using a specific polyclonal serum directed against a fragment (residues 1173-1728) of the dystrophin central domain. We observed accumulation of some major proteinase-resistant fragments, in the 110-160 kDa range originating from that central region of the molecule. Cleavage patterns of the smooth and skeletal muscle preparations were quite similar, but molecular weights of the breakdown products differed slightly. Interpretation of the results was based on two predictive structural models of the dystrophin central domain (Koenig and Kunkel (1990) J. Biol. Chem. 265, 4560-4566 and Cross et al. (1990) FEBS Lett. 262, 87-90). Skip residues at the end of repeat 13 (around the 1740th residue of the dystrophin amino acid sequence), as hypothesized in the Cross model, constitute probably the most sensitive site within the dystrophin central domain for any exogenous (or even endogenous) proteinase. Variations observed between dystrophins from skeletal and smooth muscles also suggest that the structures of both dystrophins differ slightly even within the dystrophin central domain. This precise identification of proteinase-resistant dystrophin fragments of variable lengths is a first step towards further physicochemical studies on the very large and rare dystrophin molecule.  相似文献   

14.
The dependency of induction of T cell cytotoxicity on lymphokines was studied. 1 X 10(5) nylon wool-purified thymic lymphocytes or 10(4) spleen cells were cultured with TNP-haptenated syngeneic UV-irradiated spleen cells in the presence of a variety of lymphokine preparations. Concanavalin A-induced spleen cell supernatants mediated strong cytotoxic responses in this system. Three other preparations, namely, a partially purified IL 2 preparation from PMA-stimulated EL-4 thymoma cells, a Con A-induced spleen cell supernatant that was absorbed with an IL 2-dependent cell line, and a Con A-induced supernatant that was dialyzed at pH 2 were all ineffective in mediating a cytotoxic response. In reconstitution experiments, cytotoxic responses were only obtained when either the absorbed preparation or the pH 2-treated preparation was mixed with the IL 2 preparation from EL-4 cells. No reconstitution occurred after mixing of the absorbed with the pH 2-treated preparation. pH 2 treatment of the absorbed preparation did not abolish its synergistic effect when added to the IL 2 preparation from EL-4 cells. These results led to the conclusion that activation of cytotoxic lymphocyte precursors requires at least two other lymphokines in addition to IL 2. One T cell cytotoxicity-inducing factor (TCF1) remained in Con A-induced supernatants after absorption with IL 2 receptor-bearing T cell line cells. It was pH 2-resistant and was not found in EL-4 supernatants. A second T cell cytotoxicity-inducing factor (TCF2) was pH 2-sensitive and was found in Con A-induced spleen cell supernatants as well as in interferon-free supernatants of PMA-stimulated EL-4 cells. This activity co-purified with IL 2. It was absorbed by the IL 2-dependent T cell line together with IL 2. IL 2 differs from TCF2 since it is pH 2-resistant.  相似文献   

15.

Background

The loss of dystrophin compromises muscle cell membrane stability and causes Duchenne muscular dystrophy and/or various forms of cardiomyopathy. Increased expression of the dystrophin homolog utrophin by gene delivery or pharmacologic up-regulation has been demonstrated to restore membrane integrity and improve the phenotype in the dystrophin-deficient mdx mouse. However, the lack of a viable therapy in humans predicates the need to explore alternative methods to combat dystrophin deficiency. We investigated whether systemic administration of recombinant full-length utrophin (Utr) or ΔR4-21 “micro” utrophin (μUtr) protein modified with the cell-penetrating TAT protein transduction domain could attenuate the phenotype of mdx mice.

Methods and Findings

Recombinant TAT-Utr and TAT-μUtr proteins were expressed using the baculovirus system and purified using FLAG-affinity chromatography. Age-matched mdx mice received six twice-weekly intraperitoneal injections of either recombinant protein or PBS. Three days after the final injection, mice were analyzed for several phenotypic parameters of dystrophin deficiency. Injected TAT-μUtr transduced all tissues examined, integrated with members of the dystrophin complex, reduced serum levels of creatine kinase (11,290±920 U versus 5,950±1,120 U; PBS versus TAT), the prevalence of muscle degeneration/regeneration (54%±5% versus 37%±4% of centrally nucleated fibers; PBS versus TAT), the susceptibility to eccentric contraction-induced force drop (72%±5% versus 40%±8% drop; PBS versus TAT), and increased specific force production (9.7±1.1 N/cm2 versus 12.8±0.9 N/cm2; PBS versus TAT).

Conclusions

These results are, to our knowledge, the first to establish the efficacy and feasibility of TAT-utrophin-based constructs as a novel direct protein-replacement therapy for the treatment of skeletal and cardiac muscle diseases caused by loss of dystrophin.  相似文献   

16.
Summary We have identified 7 patients with Becker muscular dystrophy (BMD) in whom analysis of dystrophin by immunoblotting shows a full-sized molecule produced at reduced abundance compared with controls. They have no detectable deletion in their dystrophin cDNA. One patient presented atypically with unusually severe cramps as his only symptom for 25 years. These patients were investigated using the polymerase chain reaction (PCR) with 3 sets of primers within the promoter region of the dystrophin gene, followed by dot blot and restriction analysis. In the patient with the atypical history, one of the expected fragments on PCR failed to amplify. A large deletion was excluded by the finding of normally sized fragments on amplification with the other primer sets. The mutation was localised to the 3 end of the forward primer binding site by dot blot and restriction analysis. This result supports the hypothesis that, in patients with a full-sized dystrophin molecule produced at reduced abundance, the phenotype may result from a mutation in the promoter region of the dystrophin gene. The atypical history of the patient in whom this was detected adds to the variety of phenotypes now known to exist as BMD.  相似文献   

17.
Chromatin was prepared by 2 methods from isolated nuclei ofthe epicotyl, mesocotyl, and root of 4- to 5-day old maize seedlings.Extraction of the nuclei with Tris-EDTA followed by the precipitationof chromatin with 10 mM CaCl2 isolates approximately 80% ofthe DNA present in these nuclei. Similarly, the extraction ofnuclei with 2 M NaCl isolates a chromatin which can be comparedto the CaCl2-chromatin. Chromatins from all three organs hadprotein-DNA ratios between 2 and 3, and RNA-DNA ratios of approximately0.20 to 0.25. Both chromatin procedures isolate a product whichis similar in composition. Acid-soluble proteins of the chromatin were extracted with 0.2M H2SO4 and were analyzed by acid-urea polyacrylamide electrophoresis.The results demonstrated that maize histones consist of 5 majorbands which are similar to the bands found in animal systems.Some 11 minor bands were also observed in each preparation fromeach organ. Both the major and minor bands were identical inpreparations from epicotyl, mesocotyl, and root. Acid-solubleproteins prepared from the 2 M NaClchromatin were identicalto that of the CaCl2-chromatin except that the minor bands werepresent in lesser amounts. With the use of a microdensitometer,it was possible to show that the major bands were present inthe same relative proportions in each of the seedling organs.A brief survey of chromatins from mature leaves, mature internodes,young tassels, and pollen showed that they contained the samedistinct histone bands as above, and no organ-specific histonecould be demonstrated. Present address: Biology Division, Oak Ridge National Laboratory,Oak Ridge, Tennessee 37830, U.S.A. (Received April 9, 1973; )  相似文献   

18.
Activation of component C3 is central to the pathways of complement and leads directly to neutralization of pathogens and stimulation of adaptive immune responses. The convertases that catalyze this reaction assemble from fragments of complement components via multistep reactions. In the lectin pathway, mannose-binding lectin (MBL) and ficolins bind to pathogens and activate MBL-associated serine protease-2 (MASP-2). MASP-2 cleaves C4 releasing C4a and generating C4b, which attaches covalently to the pathogen surface upon exposure of its reactive thioester. C2 binds to C4b and is also cleaved by MASP-2 to form the C3 convertase (C4b2a). To understand how this complex process is coordinated, we have analyzed the interactions between MASP-2, C4, C2, and their activation fragments and have compared MASP-2-catalyzed cleavage of C4b2 and C2. The data show that C2 binds tightly to C4b but not to C4, implying that C4 and C2 do not circulate as preformed complexes but that C2 is recruited only after prior activation of C4. Following cleavage of C4, C4b still binds to MASP-2 (KD approximately 0.6 microM) and dissociates relatively slowly (koff approximately 0.06 s-1) compared with the half-life of the thioester (相似文献   

19.
Dystrophin, a component of the muscle membrane cytoskeleton, is the protein altered in Duchenne Muscular Dystrophy (DMD) and Becker Muscular Dystrophy (BMD). Dystrophin shares significant homology with other cytoskeletal proteins, such as α-actinin and spectrin. On the basis of its sequence similarity with α-actinin and spectrin, dystrophin has been proposed to function as dimer. However, the existence of both dimers and monomers have been observed by electron microscopy. To address this apparent discrepancy, we expressed dystrophin fragments composed of different domains in an in vitro translation system. The expressed fragments were tested for their ability to interact with each other and full-length dystrophin by both immunoprecipitation and blot overlay assays. These assays were successfully used to demonstrate the dimerization of α-actinin and spectrin, yet failed to detect any interaction between dystrophin fragments. Although these in vitro results do not prove that dystrophin is not a dimer in vivo, they do indicate that this interaction is not like that of the α-actinin and spectrin.  相似文献   

20.
Dystrophin is an actin binding protein that is thought to stabilize the cardiac and skeletal muscle cell membranes during contraction. Here, we investigated the contributions of each dystrophin domain to actin binding function. Cosedimentation assays and pyrene-actin fluorescence experiments confirmed that a fragment spanning two-thirds of the dystrophin molecule [from N-terminal actin binding domain (ABD) 1 through ABD2] bound actin filaments with high affinity and protected filaments from forced depolymerization, but was less effective in both assays than full-length dystrophin. While a construct encoding the C-terminal third of dystrophin displayed no specific actin binding activity or competition with full-length dystrophin, our data show that it confers an unexpected regulation of actin binding by the N-terminal two-thirds of dystrophin when present in cis. Time-resolved phosphorescence anisotropy experiments demonstrated that the presence of the C-terminal third of dystrophin in cis also influences actin interaction by restricting actin rotational amplitude. We propose that the C-terminal region of dystrophin allosterically stabilizes an optimal actin binding conformation of dystrophin.  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