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 0 毫秒
1.
C. Indiveri  F. Palmieri  F. Bisaccia  R. Kr  mer 《BBA》1987,890(3):310-318
The 2-oxoglutarate carrier from the inner membrane of bovine heart mitochondria was purified by chromatography on hydroxyapatite / celite and reconstituted with egg yolk phospholipid vesicles by the freeze-thaw-sonication technique. In the reconstituted system the incorporated 2-oxoglutarate carrier catalyzed a first-order reaction of 2-oxoglutarate / 2-oxoglutarate exchange. The substrate affinity for 2-oxoglutarate was determined to be 65 ± 18 μM (15 determinations) and the maximum exchange rate at 25°C reaches 4000–22000 μmol / min per g protein, in dependence of the particular reconstitution conditions. The activation energy of the exchange reaction is 54.3 kJ / mol. The transport is independent of pH in the range between 6 and 8. When the first fraction of the hydroxyapatite / celite column eluate was used for reconstitution, besides the 2-oxoglutarate / 2-oxoglutarate exchange, a significant activity of unidirectional uptake was observed. This activity may be due to a population of the carrier protein which is in a different state.  相似文献   

2.
The alpha-oxoglutarate carrier from pig heart mitochondria has been solubilized with Triton X-114 and purified by chromatography on hydroxyapatite and celite in the presence of cardiolipin. When applied to SDS gel electrophoresis, the purified protein consists of only a single protein band with an apparent Mr of 31.5 kDa. It corresponds to band 4 of the five protein bands previously identified in the hydroxyapatite pass-through of Triton X-114 solubilized heart mitochondria (Bisaccia, F. and Palmieri, F. (1984) Biochim. Biophys. Acta 766, 386-394). When reconstituted into liposomes the alpha-oxoglutarate transport protein catalyzes a phthalonate-sensitive alpha-oxoglutarate/alpha-oxoglutarate exchange. It is purified 250-fold with a recovery of 62% and a protein yield of 0.1% with respect to the mitochondrial extract. The properties of the reconstituted carrier, i.e., the requirements for a counteranion, the substrate specificity and the inhibitor sensitivity, are similar to those described for alpha-oxoglutarate transport in mitochondria.  相似文献   

3.
4.
Folding type-specific secondary structure propensities of 20 naturally occurring amino acids have been derived from α-helical, β-sheet, α/β, and α+β proteins of known structures. These data show that each residue type of amino acids has intrinsic propensities in different regions of secondary structures for different folding types of proteins. Each of the folding types shows markedly different rank ordering, indicating folding type-specific effects on the secondary structure propensities of amino acids. Rigorous statistical tests have been made to validate the folding type-specific effects. It should be noted that α and β proteins have relatively small α-helices and β-strands forming propensities respectively compared with those of α+β and α/β proteins. This may suggest that, with more complex architectures than α and β proteins, α+β and α/β proteins require larger propensities to distinguish from interacting α-helices and β-strands. Our finding of folding type-specific secondary structure propensities suggests that sequence space accessible to each folding type may have differing features. Differing sequence space features might be constrained by topological requirement for each of the folding types. Almost all strong β-sheet forming residues are hydrophobic in character regardless of folding types, thus suggesting the hydrophobicities of side chains as a key determinant of β-sheet structures. In contrast, conformational entropy of side chains is a major determinant of the helical propensities of amino acids, although other interactions such as hydrophobicities and charged interactions cannot be neglected. These results will be helpful to protein design, class-based secondary structure prediction, and protein folding. © 1998 John Wiley & Sons, Inc. Biopoly 45: 35–49, 1998  相似文献   

5.
In anterior pituitaries from male rats, it appeared that 5α-androstane-3β, 17β-diol was quickly metabolized into 5α-androstane-3β,6α-17β-triol and 5α-androstane-3β,7α, 17β-triol by action of 6α- and 7α-hydroxylases. Hydroxysteroid hydroxylases were located in endoplasmic reticulum and were dependent on NADPH+. Their optimum pH was 8.0, optima temperature, 37°C, and their apparent Km was 2.7 μM. Hydroxylative reactions were not reversible and not modified by gonadectomy. Hydroxylation seemed an efficient control of the pituitary level of 5α-andros-tane-3β, 17β-diol.  相似文献   

6.
The maturation of connective tissue involves the organization of collagen fibres by resident fibroblasts. Fibroblast attachment to collagen has been demonstrated to involve cell surface receptors, integrins of the β1 family. Integrins are associated with cytoplasmic actin of microfilaments either directly or through focal adhesions. The major actin isoform of fibroblast microfilaments is β actin and to a lesser extent α smooth muscle (α SM) actin. Cultured human dermal fibroblasts derived from adult dermis, newborn foreskin or keloid scar were grown on either uncoated or collagen-coated surfaces. The expression and synthesis of both α2β1 integrin and α SM actin were followed by immunohistology and immunoprecipitation. Fibroblasts on uncoated surfaces expressed little α2β1 integrin on their surface, while 20 per cent of them demonstrated α SM actin within microfilaments. Fibroblasts grown on a collagen-coated surface minimally expressed α SM actin in microfilament structures and a majority of the cells were positive for α2β1 integrin on their membranes. Using [35S]-methionine incorporation and immunoprecipitation, it was shown that fibroblasts grown in uncoated dishes synthesized more α SM actin than fibroblasts grown on collagen-coated dishes. In contrast, fibroblasts grown on collagen coated dishes synthesized more α2β1 integrin compared to the same cells grown on uncoated dishes. Fibroblasts maintained on a type I collagen upregulate the expression and synthesis of α2β1 integrin, and downregulate the expression and synthesis of α SM actin. © 1998 John Wiley & Sons, Ltd.  相似文献   

7.
The metabolism of the illegal growth promoter ethylestrenol (EES) was evaluated in bovine liver cells and subcellular fractions of bovine liver preparations. Incubations with bovine microsomal preparations revealed that EES is extensively biotransformed into norethandrolone (NE), another illegal growth promoter. Furthermore, incubations of monolayer cultures of hepatocytes with NE indicated that NE itself is rapidly reduced to 17α-ethyl-5β-estrane-3α,17β-diol (EED). In vivo tests confirmed that, after administration of either EES or NE, EED is excreted as a major metabolite. Therefore, it was concluded that, both in urine and faeces samples, EED can be used as a biological marker for the illegal use of EES and/or NE. Moreover, by monitoring EED in urine or faeces samples, the detection period after NE administration is significantly prolonged. These findings were further confirmed by three cases of norethandrolone abuse in a routine screening program for forbidden growth promoters.  相似文献   

8.
Inflammatory cytokines are closely related to pigmentary changes. In this study, the effects of IFN‐γ on melanogenesis were investigated. IFN‐γ inhibits basal and α‐MSH‐induced melanogenesis in B16 melanoma cells and normal human melanocytes. MITF mRNA and protein expressions were significantly inhibited in response to IFN‐γ. IFN‐γ inhibited CREB binding to the MITF promoter but did not affect CREB phosphorylation. Instead, IFN‐γ inhibited the association of CBP and CREB through the increased association between CREB binding protein (CBP) and STAT1. These findings suggest that IFN‐γ inhibits both basal and α‐MSH‐induced melanogenesis by inhibiting MITF expression. The inhibitory action of IFN‐γ in α‐MSH‐induced melanogenesis is likely to be associated with the sequestration of CBP via the association between CBP and STAT1. These data suggest that IFN‐γ plays a role in controlling inflammation‐ or UV‐induced pigmentary changes.  相似文献   

9.
Simulated enzymic debranching of a β-limit dextrin model, prepared from a computed construct made by random extension and branching, and given the CCL value of w-maize amylopectin (and equal amounts of external chains with ECL values of 2 and 3) has been related to experimental chromatograms of the debranched β-limit dextrin of the amylopectin. The profile was similar to those from gel chromatograms and IEC-PAD chromatography.The equivalent lengths in glucosyl units of grid-links (g-links) of internal and external chains in constructs were calculated from the ICL and ECL values of amylopectin and models produced from the constructs with the appropriate lengths for internal and external chains. These derived models were subjected to simulated hydrolysis by Pseudomonas stutzeri amylase and the products compared with those of the experimental distribution from w-maize amylopectin. With the model the amounts of maltotetraose and maltodextrins released were similar to the experimental values but the distribution of branched maltodextrins was quite different. Unlike w-maize amylopectin – a polymer with the cluster structure – which has given a profile of molecular sizes of maltodextrins with low amounts of single and small numbers of internal chains and with a peak at a MW of about 14,000 (13 chains), in the model the proportion of maltodextrin with one internal chain was high and as d.p. increased the amounts decreased exponentially. This would be expected if the distribution of internal chains in the core was random. It is suggested that in the core of a model prepared from a construct made with alternating probabilities of extension – one in which this probability is high relative to branching, and a second in which it is low – may give clusters of branched maltodextrins with short internal chains which are joined by longer chains; more closely approximating the distribution of internal chains of different lengths in amylopectin.An arrangement for amylopectin molecules in the starch granule has been proposed. In this, they have a wafer-like, discoidal shape, composed of the amorphous zone overlain with the double helical, crystalline region. The flat macromolecules are concentrically layered with the former on the inside and the latter oriented to the outside of the granule.  相似文献   

10.
α-Ketobutyrate decarboxylase encoded in the -methionine catabolism operon of Pseudomonas putida is homologous with the E1 component of pyruvate dehydrogenase complex from gram-negative bacteria. The enzyme was purified to homogeneity from the cell extract of an Escherichia coli transformant. The purified enzyme was homodimeric with a subunit of Mr 93,000 on SDS-PAGE. The enzyme activity was activated by the addition of both thiamine pyrophosphate (TPP) and a divalent cation, such as Mg2+, Mn2+, and Co2+. The enzyme showed high activity for α-ketobutyrate and α-keto-n-valerate rather than pyruvate, but the α-keto acids with increasing length of the side chain as well as branching, such as α-keto-n-caproate and α-keto-3-methylvalerate, were not used by the enzyme. The Km values for α-ketobutyrate and pyruvate were 0.016 and 0.147 mM, respectively, and the kcat/Km value (10.69 s−1 mM−1) for α-ketobutyrate was 29-fold greater than that for pyruvate. Thus, α-ketobutyrate decarboxylase is distinguished from the pyruvate dehydrogenase E1 component with respect to the substrate specificity, although their structural and enzymological properties were similar. These results suggest that the unique substrate specificity of α-ketobutyrate decarboxylase is due to a slight difference in the highly conserved active sites of both enzymes.  相似文献   

11.
β‐dystroglycan (β‐DG) is a widely expressed transmembrane protein that plays important roles in connecting the extracellular matrix to the cytoskeleton, and thereby contributing to plasma membrane integrity and signal transduction. We previously observed nuclear localization of β‐DG in cultured cell lines, implying the existence of a nuclear targeting mechanism that directs it to the nucleus instead of the plasma membrane. In this study, we delineate the nuclear import pathway of β‐DG, characterizing a functional nuclear localization signal (NLS) in the β‐DG cytoplasmic domain, within amino acids 776–782. The NLS either alone or in the context of the whole β‐DG protein was able to target the heterologous GFP protein to the nucleus, with site‐directed mutagenesis indicating that amino acids R779 and K780 are critical for NLS functionality. The nuclear transport molecules Importin (Imp)α and Impβ bound with high affinity to the NLS of β‐DG and were found to be essential for NLS‐dependent nuclear import in an in vitro reconstituted nuclear transport assay; cotransfection experiments confirmed the dependence on Ran for nuclear accumulation. Intriguingly, experiments suggested that tyrosine phosphorylation of β‐DG may result in cytoplasmic retention, with Y892 playing a key role. β‐DG thus follows a conventional Impα/β‐dependent nuclear import pathway, with important implications for its potential function in the nucleus. J. Cell. Biochem. 110: 706–717, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

12.
The 90‐kDa heat shock protein (Hsp90α) has been identified on the surface of cancer cells, and is implicated in tumor invasion and metastasis, suggesting that it is a potentially important target for tumor therapy. However, the regulatory mechanism of Hsp90α plasma membrane translocation during tumor invasion remains poorly understood. Here, we show that Hsp90α plasma membrane expression is selectively upregulated upon epidermal growth factor (EGF) stimulation, which is a process independent of the extracellular matrix. Abrogation of EGF‐mediated activation of phospholipase (PLCγ1) by its siRNA or inhibitor prevents the accumulation of Hsp90α at cell protrusions. Inhibition of the downstream effectors of PLCγ1, including Ca2+ and protein kinase C (PKCγ), also blocks the membrane translocation of Hsp90α, while activation of PKCγ leads to increased levels of cell‐surface Hsp90α. Moreover, overexpression of PKCγ increases extracellular vesicle release, on which Hsp90α is present. Furthermore, activation or overexpression of PKCγ promotes tumor cell motility in vitro and tumor metastasis in vivo, whereas a specific neutralizing monoclonal antibody against Hsp90α inhibits such effects, demonstrating that PKCγ‐induced Hsp90α translocation is required for tumor metastasis. Taken together, our study provides a mechanistic basis for the role for the PLCγ1–PKCγ pathway in regulating Hsp90α plasma membrane translocation, which facilitates tumor cell motility and promotes tumor metastasis.  相似文献   

13.
The integrin α4β1 is involved in mediating exfiltration of leukocytes from the vasculature. It interacts with a number of proteins up-regulated during the inflammatory response including VCAM-1 and the CS-1 alternatively spliced region of fibronectin. In addition it binds the multifunctional protein osteopontin (OPN), which can act as both a cytokine and an extracellular matrix molecule. Here we map the region of human OPN that supports cell adhesion via α4β1 using GST fusion proteins. We show that α4β1 expressed in J6 cells interacts with intact OPN when the integrin is in a high activation state, and by deletion mapping that the α4β1 binding region in OPN lies between amino acid residues 125 and 168 (aa125–168). This region contains the central RGD motif of OPN, which also interacts with integrins αvβ3, αvβ5, αvβ1, α8β1, and α5β1. Mutating the RGD motif to RAD had no effect on the interaction with α4β1. To define the binding site the region incorporating aa125–168 was divided into 5 overlapping peptides expressed as GST fusion proteins. Two peptides supported adhesion via α4β1, aa132–146, and aa153–168; of these only a synthetic peptide, SVVYGLR (aa162–168), derived from aa153–168 was able to inhibit α4β1 binding to CS-1. These data identify the motif SVVYGLR as a novel peptide inhibitor of α4β1, and the primary α4β1 binding site within OPN.  相似文献   

14.
Gamma‐aminobutyric acid type A receptors (GABAARs) are the most important inhibitory chloride ion channels in the central nervous system and are major targets for a wide variety of drugs. The subunit compositions of GABAARs determine their function and pharmacological profile. GABAARs are heteropentamers of subunits, and (α1)2(β3)2(γ2L)1 is a common subtype. Biochemical and biophysical studies of GABAARs require larger quantities of receptors of defined subunit composition than are currently available. We previously reported high‐level production of active human α1β3 GABAAR using tetracycline‐inducible stable HEK293 cells. Here we extend the strategy to receptors containing three different subunits. We constructed a stable tetracycline‐inducible HEK293‐TetR cell line expressing human (N)–FLAG–α1β3γ2L–(C)–(GGS)3GK–1D4 GABAAR. These cells achieved expression levels of 70–90 pmol [3H]muscimol binding sites/15‐cm plate at a specific activity of 15–30 pmol/mg of membrane protein. Incorporation of the γ2 subunit was confirmed by the ratio of [3H]flunitrazepam to [3H]muscimol binding sites and sensitivity of GABA‐induced currents to benzodiazepines and zinc. The α1β3γ2L GABAARs were solubilized in dodecyl‐d ‐maltoside, purified by anti‐FLAG affinity chromatography and reconstituted in CHAPS/asolectin at an overall yield of ~30%. Typical purifications yielded 1.0–1.5 nmoles of [3H]muscimol binding sites/60 plates. Receptors with similar properties could be purified by 1D4 affinity chromatography with lower overall yield. The composition of the purified, reconstituted receptors was confirmed by ligand binding, Western blot, and proteomics. Allosteric interactions between etomidate and [3H]muscimol binding were maintained in the purified state.  相似文献   

15.
In recent years several 15β-hydroxysteroids have emerged pathognomonic of adrenal disorders in human neonates of which 3α,15β,17α-trihydroxy-5β-pregnan-20-one (2) was the first to be identified in the urine of newborn infants affected with congenital adrenal hyperplasia. In this investigation we report the synthesis of the three remaining 3ξ,5ξ-isomers, namely 3α,15β,17α-trihydroxy-5α-pregnan-20-one (3), 3β,15β,17α-trihydroxy-5α-pregnan-20-one (7) and 3β,15β,17α-trihydroxy-5β-pregnan-20-one (8) for their definitive identification in pathological conditions in human neonates. 3β,15β-Diacetoxy-17α-hydroxy-5-pregnen-20-one (11), a product of chemical synthesis was converted to the isomeric 3 and 7, while conversion of 15β,17α-dihydroxy-4-pregnen-3,20-dione (4), a product of microbiological transformation, resulted in the preparation of 8. In brief, selective acetate hydrolysis of 11 gave 15β-acetoxy-3β,17α-dihydroxy-5-pregnen-20-one (12) which on catalytic hydrogenation gave 15β-acetoxy-3β,17α-dihydroxy-5α-pregnan-20-one (13) a common intermediate for the synthesis of the 3β(and α),5α-isomers. Hydrolysis of the 15β-acetate gave 7, whereas oxidation with pyridinium chlorochromate gave 15β-acetoxy-17α-hydroxy-5α-pregnan-3,20-dione (14) which on reduction with -Selectride and hydrolysis of the 15β-acetate gave 3. Finally, hydrogenation of 4 gave 15β,17α-dihydroxy-5β-pregnan-3,20-dione (10) which on reduction with -Selectride gave 8.  相似文献   

16.
Hematopoietic homeostasis depends on the maintenance of hematopoietic stem cells (HSCs), which are regulated within a specialized bone marrow (BM) niche. When HSC sense external stimuli, their adhesion status may be critical for determining HSC cell fate. The cell surface molecule, integrin αvβ3, is activated through HSC adhesion to extracellular matrix and niche cells. Integrin β3 signaling maintains HSCs within the niche. Here, we showed the synergistic negative regulation of the pro‐inflammatory cytokine interferon‐γ (IFNγ) and β3 integrin signaling in murine HSC function by a novel definitive phenotyping of HSCs. Integrin αvβ3 suppressed HSC function in the presence of IFNγ and impaired integrin β3 signaling mitigated IFNγ‐dependent negative action on HSCs. During IFNγ stimulation, integrin β3 signaling enhanced STAT1‐mediated gene expression via serine phosphorylation. These findings show that integrin β3 signaling intensifies the suppressive effect of IFNγ on HSCs, which indicates that cell adhesion via integrin αvβ3 within the BM niche acts as a context‐dependent signal modulator to regulate the HSC function under both steady‐state and inflammatory conditions.  相似文献   

17.
A single chiral cyclic α,α‐disubstituted amino acid, (3S,4S)‐1‐amino‐(3,4‐dimethoxy)cyclopentanecarboxylic acid [(S,S)‐Ac5cdOM], was placed at the N‐terminal or C‐terminal positions of achiral α‐aminoisobutyric acid (Aib) peptide segments. The IR and 1H NMR spectra indicated that the dominant conformations of two peptides Cbz‐[(S,S)‐Ac5cdOM]‐(Aib)4‐OEt ( 1) and Cbz‐(Aib)4‐[(S,S)‐Ac5cdOM]‐OMe (2) in solution were helical structures. X‐ray crystallographic analysis of 1 and 2 revealed that a left‐handed (M) 310‐helical structure was present in 1 and that a right‐handed (P) 310‐helical structure was present in 2 in their crystalline states. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
19.
20.
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