首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 609 毫秒
1.
The natural occurrence of 22-hydroxylated steroids in cultured Catharanthus roseus cells and in Arabidopsis seedlings was investigated. Using full-scan gas chromatography-mass spectrometry analysis, (22S)-22-hydroxycampesterol (22-OHCR), (22S,24R)-22-hydroxyergost-4-en-3-one (22-OH-4-en-3-one), (22S,24R)-22-hydroxy-5alpha-ergostan-3-one (22-OH-3-one), 6-deoxocathasterone (6-deoxoCT), 3-epi-6-deoxoCT, 28-nor-22-OHCR, 28-nor-22-OH-4-en-3-one, 28-nor-22-OH-3-one, 28-nor-6-deoxoCT, and 3-epi-28-nor-6-deoxoCT were identified. Metabolic experiments with deuterium-labeled 22-OHCR were performed in cultured C. roseus cells and Arabidopsis seedlings (wild type and det2), and the metabolites were analyzed by gas chromatography-mass spectrometry. In both C. roseus cells and wild-type Arabidopsis seedlings, [(2)H(6)]22-OH-4-en-3-one, [(2)H(6)]22-OH-3-one, [(2)H(6)]6-deoxoCT, and [(2)H(6)]3-epi-6-deoxoCT were identified as metabolites of [(2)H(6)]22-OHCR, whereas the major metabolite in det2 seedlings was [(2)H(6)]22-OH-4-en-3-one. Analysis of endogenous levels of these brassinosteroids revealed that det2 accumulates 22-OH-4-en-3-one. The levels of downstream compounds were remarkably reduced compared with the wild type. Exogenously applied 22-OH-3-one and 6-deoxoCT were found to rescue det2 mutant phenotypes, whereas 22-OHCR and 22-OH-4-en-3-one did not. These results substantiate the existence of a new subpathway (22-OHCR --> 22-OH-4-en-3-one --> 22-OH-3-one --> 6-deoxoCT) and reveal that the det2 mutant is defective in the conversion of 22-OH-4-en-3-one to 22-OH-3-one, which leads to brassinolide biosynthesis.  相似文献   

2.
The total fatty acids in human endothelial cells include approximately 5% each of 22:4(n-6), 22:5(n-3) and 22:6(n-3), whereas 22:5(n-6) is present only in trace amounts. This study evaluates the effect of three of these fatty acids bound to albumin on lipid composition and prostacyclin (prostaglandin I2) synthesis in primary cultures of endothelial cell monolayers. 22:4(n-6), 22:5(n-6) and 22:6(n-3) were all incorporated into total phospholipids. 20:4(n-6) was reduced in phospholipids in all cells incubated with the three different docosaenoic fatty acids. This reduction was abolished when equimolar concentrations of 20:4(n-6) and the separate docosaenoic fatty acid were added to the medium simultaneously. 22:4(n-6) incorporation into the free fatty acids was associated with an increase of 20:4(n-6) in this fraction. 22:4(n-6), 22:5(n-6) and 22:5(n-3) all reduced the synthesis of prostacyclin measured as 6-ketoprostaglandin F1 alpha. These effects were reversed by simultaneous incubation with 20:4(n-6). This study shows that three of the docosaenoic fatty acids present in human endothelial cells of the (n-6) and (n-3) family were all incorporated into endothelial cells with a simultaneous reduction in 20:4(n-6). The three fatty acids reduced the synthesis of prostacyclin.  相似文献   

3.
The concentration-dependent metabolism of 1-(14)C-labelled precursors of 22:5n-6 and 22:6n-3 was compared in rat testis cells. The amounts of [(14)C]22- and 24-carbon metabolites were measured by HPLC. The conversion of [1-(14)C]20:5n-3 to [3-(14)C]22:6n-3 was more efficient than that of [1-(14)C]20:4n-6 to [3-(14)C]22:5n-6. At low substrate concentration (4 microM) it was 3.4 times more efficient, reduced to 2.3 times at high substrate concentration (40 microM). The conversion of [1-(14)C]22:5n-3 to [1-(14)C]22:6n-3 was 1.7 times more efficient than that of [1-(14)C]22:4n-6 to [1-(14)C]22:5n-6 using a low, but almost equally efficient using a high substrate concentration. When unlabelled 20:5n-3 was added to a cell suspension incubated with [1-(14)C]20:4n-6 or unlabelled 22:5n-3 to a cell suspension incubated with [1-(14)C]22:4n-6, the unlabelled n-3 fatty acids strongly inhibited the conversion of [1-(14)C]20:4n-6 or [1-(14)C]22:4n-6 to [(14)C]22:5n-6. In the reciprocal experiment, unlabelled 20:4n-6 and 22:4n-6 only weakly inhibited the conversion of [1-(14)C]20:5n-3 and [1-(14)C]22:5n-3 to [(14)C]22:6n-3. The results indicate that if both n-6 and n-3 fatty acids are present, the n-3 fatty acids are preferred over the n-6 fatty acids in the elongation from 20- to 22- and from 22- to 24-carbon atom fatty acids. In vivo the demand for 22-carbon fatty acids for spermatogenesis in the rat may exceed the supply of n-3 precursors and thus facilitate the formation of 22:5n-6 from the more abundant n-6 precursors.  相似文献   

4.
The integral membrane protein p22(phox) forms a heterodimeric enzyme complex with NADPH oxidases (Noxs) and is required for their catalytic activity. Nox4, a Nox linked to cardiovascular disease, angiogenesis, and insulin signaling, is unique in its ability to produce hydrogen peroxide constitutively. To date, p22(phox) constitutes the only identified regulatory component for Nox4 function. To delineate structural elements in p22(phox) essential for formation and localization of the Nox4-p22(phox) complex and its enzymatic function, truncation and point mutagenesis was used. Human lung carcinoma cells served as a heterologous expression system, since this cell type is p22(phox)-deficient and promotes cell surface expression of the Nox4-p22(phox) heterodimer. Expression of p22(phox) truncation mutants indicates that the dual tryptophan motif contained in the N-terminal amino acids 6-11 is essential, whereas the C terminus (amino acids 130-195) is dispensable for Nox4 activity. Introduction of charged residues in domains predicted to be extracellular by topology modeling was mostly tolerated, whereas the exchange of amino acids in predicted membrane-spanning domains caused loss of function or showed distinct differences in p22(phox) interaction with various Noxs. For example, the substitution of tyrosine 121 with histidine in p22(phox), which abolished Nox2 and Nox3 function in vivo, preserved Nox4 activity when expressed in lung cancer cells. Many of the examined p22(phox) mutations inhibiting Nox1 to -3 maturation did not alter Nox4-p22(phox) association, further accenting the differences between Noxs. These studies highlight the distinct interaction of the key regulatory p22(phox) subunit with Nox4, a feature which could provide the basis for selective inhibitor development.  相似文献   

5.
We compared the polyunsaturated fatty acid (PUFA) status of Dutch vegans and omnivores to investigate whether disparities can be explained by different diets and long chain PUFA (LCP) synthesis rates. Dietary intakes and fatty acid compositions of erythrocytes (RBC), platelets (PLT), plasma cholesterol esters (CE) and plasma triglycerides (TG) of 12 strict vegans and 15 age- and sex-matched omnivores were determined. Vegans had higher omega 6 (CE, TG), 18:2 omega 6 (RBC, CE, TG), 18:3 omega 6 (TG), 20:3 omega 6 (TG), 22:4 omega 6 (TG), 22:5 omega 3 (RBC, PLT), 22:5 omega 3/22:6 omega 3 (RBC, PLT) and 22:5 omega 6/22:6 omega 3 (RBC, PLT), and lower 22:4 omega 6 (RBC, PLT), 22:4 omega 6/22:5 omega 6 (RBC, PLT), omega 3 (CE), LCP omega 3 (CE, TG), 20:5 omega 3 (RBC, PLT, CE), 22:5 omega 3 (TG) and 22:6 omega 3 (all compartments). Vegans had lower 20:4 omega 6 (TG) after normalization of PUFA to 100%, and normalization of eicosanoid precursors to 100% revealed similar 20:4 omega 6 (all), higher 20:3 omega 6 (TG) and lower 20:5 omega 3 (all). High omega 6 (notably 18:2 omega 6) and low omega 3 (notably 20:5 omega 3, 22:6 omega 3) status in Dutch vegans derives from low dietary LCP omega 3 and 18:3 omega 3/18:2 omega 6 ratio. Higher 18:3 omega 6 and 20:3 omega 6 in their TG may reflect higher hepatic 20:4 omega 6 production rate, whereas higher 20:4 omega 6 and 22:4 omega 6 in omnivores indicates 20:4 omega 6 intake from meat.  相似文献   

6.
In this study various C27 sterols with a 3-oxo-4-ene structure were incubated with adrenal cortex mitochondrial preparations. (22R)-22-Hydroxy-4-cholesten-3-one and (20R,22R)-20,22-dihydroxy-4-cholesten-3-one were found to be converted into progesterone. This suggests the existence of a pathway for adrenal progesterone formation analogous to the normal 3 beta-hydroxy-5-ene pathways. (20S)-20-Hydroxy-4-cholesten-3-one was hydroxylated at C25. 4-Cholesten-3-one, 25-hydroxy-4-cholesten-3-one and (22S)-22-hydroxy-4-cholesten-3-one were not converted to a measurable extent. With 3-oxo-4-ene C27 sterols as substrates, the cholesterol side-chain cleaving enzyme system seems to require the presence of a 22R-hydroxyl group in the substrate. The clinical relevance of these observations is discussed.  相似文献   

7.
The intracellular localization of the oxidation of [2-14C]adrenic acid (22:4(n-6)) and [1-14C]docosahexaenoic acid (22:6(n-3)) was studied in isolated liver cells. The oxidation of 22:4(n-6) was 2-3-times more rapid than the oxidation of 22:6(n-3), [1-14C]arachidonic acid (20:4(n-6)) or [1-14C]oleic acid (18:1). (+)-Decanoylcarnitine and lactate, both known to inhibit mitochondrial beta-oxidation, reduced the oxidation of 18:1 distinctly more efficiently than with 22:4(n-6) and 22:6(n-3). In liver cells from rats fed a diet containing partially hydrogenated fish oil, the oxidation of 22:6(n-6) and 22:6(n-3) was increased by 30-40% compared with cells from rats fed a standard pellet diet. With 18:1 as substrate, the amount of fatty acid oxidized was very similar in cells from animals fed standard pellets or partially hydrogenated fish oil. Shortened fatty acids were not produced from [5,6,8,9,11,12,14,15-3H]arachidonic acid. In hepatocytes from rats starved and refed 20% fructose, a large fraction of 14C from 22:4 was recovered in 14C-labelled C14-C18 fatty acids. Oxidation of 22:4 thus caused a high specific activity of the extramitochondrial pool of acetyl-CoA. The results suggest that 22:4(n-6) and to some extent 22:6(n-3) are oxidized by peroxisomal beta-oxidation and by this are retroconverted to arachidonic acid and eicosapentaenoic acid.  相似文献   

8.
Asymmetric dihydroxylation of (20(22)E)-cholesta-5,20(22)-dien-3beta-ol acetate (2a), prepared from pregnenolone, gave a 1:1 mixture (67% yield) of (20R,22R)-cholest-5-ene-3beta,20,22-triol 3-acetate (3a) and its 20S,22S isomer 3b. Highly purified 3a and 3b were obtained by semipreparative silver ion high performance liquid chromatography. Saponification of 3a and 3b gave (20R,22R)-cholest-5-ene-3beta,20,22-triol (4a) and its 20S,22S isomer 4b. This simple approach provided the natural isomer 4a more efficiently than previously described chemical or enzymatic syntheses. Full 1H and 13C nuclear magnetic resonance data were presented for triols 4a and 4b and their synthetic precursors. Side-chain conformations of 2a, its 20(22)Z isomer, 4a, and 4b were studied by molecular mechanics and nuclear Overhauser effect difference spectroscopy.  相似文献   

9.
The reasons why most cellular lipids preferentially accumulate 22:6(n-3) rather than 22:5(n-6) are poorly understood. In the present work the metabolisms of the precursor fatty acids, [1-(14)C]20:4(n-6), [1-(14)C]22:4(n-6) versus [1-(14)C]20:5(n-3), [1-(14)C]22:5(n-3) in isolated rat hepatocytes were compared. The addition of lactate and L-decanoylcarnitine increased the formation of [(14)C]24 fatty acid intermediates and the final products, [(14)C]22:5(n-6) and [(14)C]22:6(n-3). In the absence of lactate and L-decanoylcarnitine, no [(14)C]24 fatty acids and [(14)C]22:5(n-6) were detected when [1-(14)C]22:4(n-6) was the substrate, whereas small amounts of the added [1-(14)C]22:5(n-3) was converted to [(14)C]22:6(n-3). Lactate reduced the oxidation of [1-(14)C]22:4(n-6) and [1-(14)C]22:5(n-3) while L-decanoylcarnitine did not. No significant differences between the total oxidation or esterification of the two substrates were observed. By fasting and fructose refeeding the amounts of [(14)C]24:4(n-6) and [(14)C]24:5(n-3) were increased by 2.5- and 4-fold, respectively. However, the levels of [(14)C]22:5(n-6) and [(14)C]22:6(n-3) were similar in hepatocytes from fasted and refed versus fed rats. With hepatocytes from rats fed a fat free diet the levels of [(14)C]24 fatty acid intermediates were low while the further conversion of the n-6 and n-3 substrates was high and more equal, approx. 33% of [1-(14)C]22:4(n-6) was converted to [(14)C]22:5(n-6) and 43% of [1-(14)C]22:5(n-3) was converted to [(14)C]22:6(n-3). The moderate differences found in the conversion of [1-(14)C]22:4(n-6) versus [1-(14)C]22:5(n-3) to [(14)C]22:5(n-6) and [(14)C]22:6(n-3), respectively, and the equal rates of oxidation of the two substrates could thus not explain the abundance of 22:6(n-3) versus the near absence of 22:5(n-6) in cellular membranes.  相似文献   

10.
Three 14C-labeled 22-carbon polyunsaturated fatty acids, 7,10,13,16-[14C]docosatetraenoic acid (22:4(n-6)), 7,10,13,16,19-[14C]docosapentaenoic acid (22:5(n-3)), and 4,7,10,13,16,19-[14C]docosahexaenoic acid (22:6(n-3)), were compared with [3H]arachidonic acid (20:4(n-6] and [14C]linoleic acid (18:2(n-6)) to characterize their incorporation into the lipids of Ehrlich ascites cells. The relatively rapid incorporation of the labeled 22-carbon acids into phosphatidic acid indicated that substantial amounts of these acids may be incorporated through the de novo pathway of phospholipid synthesis. In marked contrast to 20:4(n-6), the 22-carbon acids were incorporated much less into choline glycerophospholipids (CGP) and inositol glycerophospholipids (IGP). No selective preference was apparent for the (n-3) or (n-6) type of fatty acids. The amounts of the acids incorporated into diacylglycerophosphoethanolamine were in the order of: 22:6(n-3) greater than 20:4(n-6) much greater than 22:5(n-3) greater than or equal to 22:4(n-6) greater than 18:2(n-6), whereas for alkylacylglycerophosphoethanolamine they were in the order of: 22:4(n-6) greater than 22:6(n-3) greater than 22:5(n-3) much greater than 20:4(n-6) greater than 18:2(n-6). Of the mechanisms possibly responsible for the selective entry of 22-carbon acids into ethanolamine glycerophospholipids, the most reasonable explanation was that the cytidine-mediated ethanolamine phosphotransferase may have a unique double selectivity: for hexaenoic species of diacylglycerol and for 22-carbon polyunsaturated fatty acid-containing species of alkylacylglycerol. The relative distribution of fatty acids between newly incorporated and already maintained lipid classes suggested that IGP may function in Ehrlich cells as an intermediate pool for the retention of polyunsaturated fatty acids in glycerolipids.  相似文献   

11.
The role of the double bonds in docosahexaenoic acid (22:6(Δ4,7,10,13,16,19); DHA) in cytotoxic lipid peroxidation was studied in a superoxide dismutase-defective human colorectal tumor cell line, HT-29. In a conventional culture, DHA and other polyunsaturated fatty acids (PUFAs) were found to induce acute lipid peroxidation and subsequent cell death. PUFAs that lack one or both the terminal double bonds (Δ19 and Δ4) but share Δ7,10,13,16 such as 22:5(Δ7,10,13,16,19), 22:5(Δ4,7,10,13,16), and 22:4(Δ7,10,13,16) were more effective than DHA. Lipid peroxidation and cell death were completely inhibited, except by 22:4(Δ7,10,13,16) when radical-mediated reactions were suppressed by culturing cells in 2% O(2) in the presence of vitamin E. DHA and C22:5 PUFAs but not 22:4(Δ7,10,13,16) were efficiently incorporated in phosphatidylinositol, regardless of the culturing conditions. These and other results suggested that the internal unsaturations Δ7,10,13,16 were sensitive to lipid peroxidation, whereas the terminal ones Δ19 and Δ4 appeared to be involved in assimilation into phospholipids.  相似文献   

12.
Many Bacteroides transfer factors are mobilizable in Escherichia coli when coresident with the IncP conjugative plasmid RP4, but not F. To begin characterization and potential interaction between Bacteroides mobilizable transfer factors and the RP4 mating channel, both mutants and deletions of the DNA processing (dtr), mating pair formation (mpf) and traG coupling genes of RP4 were tested for mobilization of Bacteroides plasmid pLV22a. All 10 mpf but none of the four dtr genes were required for mobilization of pLV22a. The RP4 TraG coupling protein (CP) was also required for mobilization of pLV22a, but could be substituted by a C-terminal deletion mutant of the F TraD CP. Potential interactions of the TraG CP with relaxase protein(s) and transfer DNA of both RP4 and pLV22a were assessed. Overlay assays identified productive interactions between TraG and the relaxase proteins of both MbpB and TraI from pLV22a and RP4 respectively. The Agrobacterium Transfer-ImmunoPrecipitation (TrIP) assay also identified an interaction between TraG and both RP4 and pLV22a transfer DNA. Thus, mobilization of the Bacteroides pLV22a in E. coli utilizes both RP4 Mpf and CP functions including an interaction between the relaxosome and the RP4 CP similar to that of cognate RP4 plasmid.  相似文献   

13.
Seven steroidal glycosides, along with one known glycoside, were isolated from the rhizomes of Ruscus hypophyllum (Liliaceae). Comprehensive spectroscopic analysis, including 2D NMR spectroscopy, and the results of acid hydrolysis allowed the chemical structures of the compounds to be assigned as (23S,25R)-23-hydroxyspirost-5-en-3beta-yl O-alpha-l-rhamnopyranosyl-(1-->4)-beta-d-glucopyranoside (1), 1beta-hydroxyspirosta-5,25(27)-dien-3beta-yl O-alpha-l-rhamnopyranosyl-(1-->4)-beta-d-glucopyranoside (2), (22S)-16beta,22-dihydroxycholest-5-en-3beta-yl O-alpha-l-rhamnopyranosyl-(1-->4)-beta-d-glucopyranoside (3), (22S)-16beta-[(beta-d-glucopyranosyl)oxy]-22-hydroxycholest-5-en-3beta-yl O-alpha-l-rhamnopyranosyl-(1-->4)-beta-d-glucopyranoside (4), (22S)-16beta-[(beta-d-glucopyranosyl)oxy]-22-hydroxycholest-5-en-3beta-yl beta-d-glucopyranoside (5), (22S)-16beta-[(beta-d-glucopyranosyl)oxy]-3beta,22-dihydroxycholest-5-en-1beta-yl O-alpha-l-rhamnopyranosyl-(1-->2)-(3,4-di-O-acetyl-beta-d-xylopyranoside) (6), and (22S)-16beta-[(beta-d-glucopyranosyl)oxy]-3beta,22-dihydroxycholest-5-en-1beta-yl O-alpha-l-rhamnopyranosyl-(1-->2)-O-[beta-d-xylopyranosyl-(1-->3)]-beta-d-xylopyranoside (7), respectively. This is the first isolation of a series of cholestane glycosides from a Ruscus species.  相似文献   

14.
2,2-Bis(4-hydroxyphenyl)propane (bisphenol A; BPA) is an environmental endocrine-disrupting chemical. It mimics the effects of estrogen at multiple levels by activating estrogen receptors (ERs); however, BPA also affects the proliferation of human breast cancer cells independent of ERs. Although BPA inhibits progesterone (P4) signaling, the toxicological significance of its effects remain unknown. Tripartite motif-containing 22 (TRIM22) has been identified as a P4-responsive and apoptosis-related gene. Nevertheless, it has not yet been established whether exogenous chemicals change TRIM22 gene levels. Therefore, the present study investigated the effects of BPA on P4 signaling and TRIM22 and TP53 expression in human breast carcinoma MCF-7 cells. In MCF-7 cells incubated with various concentrations of P4, TRIM22 messenger RNA (mRNA) levels increased in a dose-dependent manner. P4 induced apoptosis and decreased viability in MCF-7 cells. The knockdown of TRIM22 abolished P4-induced decreases in cell viability and P4-induced apoptosis. P4 increased TP53 mRNA expression and p53 knockdown decrease the basal level of TRIM22 and P4 increased TRIM22 mRNA expression independent of p53 expression. BPA attenuated P4-induced increases in the ratio of cell apoptosis in a concentration-dependent manner, and the P4-induced decreases in cell viability was abolished in the presence of 100 nM and higher BPA concentrations. Furthermore, BPA inhibited P4-induced TRIM22 and TP53 expression. In conclusion, BPA inhibited P4-induced apoptosis in MCF-7 cells via the inhibition of P4 receptor transactivation. TRIM22 gene has potential as a biomarker for investigating the disruption of P4 signaling by chemicals.  相似文献   

15.
Cytogenetic analysis of meningioma cells from one particular patient (MN32) displayed the stem-line karyo-type 45, XY, -1, 4p+, 22q-, 22q+, which thus had rearrangements of both chromosomes 22. The 22q+ marker appeared as a dicentric: 22 pter----q11::1p11----qter. The reciprocal product of this translocation has presumably been lost because it lacked a centromere. The 22q- chromosome also appeared to have lost sequences distal to band q11. We assumed that this marker could have been the result of a reciprocal translocation between chromosomes 4 and 22. To investigate the 4p+ and 22q- chromosomes in more detail, human-hamster somatic cell hybrids were constructed that segregated the 22q- and 4p+ chromosomes. Southern blot analysis with DNA from these hybrids showed that sequences from 22q were indeed translocated to 4p+ and that reciprocally sequences from 4p were translocated to 22q-, demonstrating a balanced t(4;22)(p16;q11). On the basis of these results we presume that in this tumor a tumor-suppressor gene is deleted in the case of the 22q+ marker and that the t(4;22) disrupts the second allele of this gene. The latter translocation was mapped between D22S1 and D22S15, a distance of 1 cM on the linkage map of this chromosome. The area in which we have located the translocation is within the region where the gene predisposing to neurofibromatosis 2 has been mapped.  相似文献   

16.

Introduction

The aim of this study was to characterize interleukin 17 (IL-17) and interleukin 22 (IL-22) producing cells in peripheral blood (PB), skin, synovial fluid (SF) and synovial tissue (ST) in patients with psoriasis (Ps) and psoriatic arthritis (PsA).

Methods

Flow cytometry was used to enumerate cells making IL-22 and IL-17, in skin and/or SF and PB from 11 patients with Ps and 12 patients with PsA; skin and PB of 15 healthy controls and SF from rheumatoid arthritis (RA) patients were used as controls. Expression of the interleukin 23 receptor (IL-23R) and chemokine receptors CCR4 and CCR6 was examined. Secretion of IL-17 and IL-22 was measured by ELISA. ST was analysed by immunohistochemical staining of IL-17 and IL-22.

Results

Increased frequencies of IL-17+ and IL-22+ CD4+ T cells were seen in PB of patients with PsA and Ps. IL-17 secretion was significantly elevated in both PsA and Ps, whilst IL-22 secretion was higher in PsA compared to Ps and healthy controls. A higher proportion of the CD4+ cells making IL-17 or IL-22 expressed IL-23R and frequencies of IL-17+, CCR6+ and CCR4+ T cells were elevated in patients with Ps and those with PsA. In patients with PsA, CCR6+ and IL-23R + T cells numbers were elevated in SF compared to PB. Increased frequencies of IL-17+ and IL-22+ CD4+ T cells were demonstrated in Ps skin lesions. In contrast, whilst elevated frequencies of CD4+ IL-17+ cells were seen in PsA SF compared to PB, frequencies of CD4+ IL-22+ T cells were lower. Whereas IL-17 expression was equivalent in PsA, osteoarthritis (OA) and RA ST, IL-22 expression was higher in RA than either OA or PsA ST, in which IL-22 was strikingly absent.

Conclusions

Elevated frequencies of IL-17 and IL-22 producing CD4+ T cells were a feature of both Ps and PsA. However their differing distribution at disease sites, including lower frequencies of IL-22+ CD4+ T cells in SF compared to skin and PB, and lack of IL-22 expression in ST suggests that Th17 and Th22 cells have common, as well as divergent roles in the pathogenesis of Ps and PsA.  相似文献   

17.
Senescence is a tumor suppressor program characterized by a stable growth arrest while maintaining cell viability. Senescence-associated ribogenesis defects (SARD) have been shown to regulate senescence through the ability of the ribosomal protein S14 (RPS14 or uS11) to bind and inhibit the cyclin-dependent kinase 4 (CDK4). Here we report another ribosomal protein that binds and inhibits CDK4 in senescent cells: L22 (RPL22 or eL22). Enforcing the expression of RPL22/eL22 is sufficient to induce an RB and p53-dependent cellular senescent phenotype in human fibroblasts. Mechanistically, RPL22/eL22 can interact with and inhibit CDK4-Cyclin D1 to decrease RB phosphorylation both in vitro and in cells. Briefly, we show that ribosome-free RPL22/eL22 causes a cell cycle arrest which could be relevant during situations of nucleolar stress such as cellular senescence or the response to cancer chemotherapy.  相似文献   

18.
Nox1 and Nox4, homologues of the leukocyte NADPH oxidase subunit Nox2 (gp91phox) mediate superoxide anion formation in various cell types. However, their interactions with other components of the NADPH oxidase are poorly defined. We determined whether a direct interaction of Nox1 and Nox4 with the p22phox subunit of the NADPH oxidase occurs. Using confocal microscopy, co-localization of p22phox with Nox1, Nox2, and Nox4 was observed in transiently transfected vascular smooth muscle cells (VSMC) and HEK293 cells. Plasmids coding for fluorescent fusion proteins of p22phox and the Nox proteins with cyan- and yellow-fluorescent protein (cfp and yfp, respectively) were constructed and expressed in VSMC and HEK293 cells. The cfp-tagged p22phox expression level increased upon cotransfection with Nox1 or Nox4. Protein-protein interaction between the fluorescent fusion proteins of p22phox and the Nox partners was observed using the fluorescence resonance energy transfer technique. Immunoprecipitation of native Nox1 from human VSMC revealed co-precipitation of p22phox. Immunoprecipitation from transfected HEK293 cells revealed co-precipitation of native p22phox with yfp-tagged Nox1, Nox2, and Nox4. Following mutation of a histidine (corresponding to the position 115 in human Nox2) to leucine, this interaction was abolished. Transfection of rat p22phox (but not Noxo1 and Noxa1) increased the radical generation in cells expressing Nox4. We provide evidence that p22phox directly interacts with Nox1 and Nox4, to form an superoxide-generating NADPH oxidase and demonstrate that mutation of the potential heme binding site in the Nox proteins disrupts the complex formation of Nox1 and Nox4 with p22phox.  相似文献   

19.
Cullin 4 (Cul4)-based ubiquitin ligases emerged as critical regulators of DNA replication and repair. Over 50 Cul4-specific adaptors (DNA damage-binding 1 (Ddb1)-Cul4-associated factors; DCAFs) have been identified and are thought to assemble functionally distinct Cul4 complexes. Using a live-cell imaging-based RNAi screen, we analysed the function of DCAFs and Cul4-linked proteins, and identified specific subsets required for progression through G1 and S phase. We discovered C6orf167/Mms22-like protein (Mms22L) as a putative human orthologue of budding yeast Mms22, which, together with cullin Rtt101, regulates genome stability by promoting DNA replication through natural pause sites and damaged templates. Loss of Mms22L function in human cells results in S phase-dependent genomic instability characterised by spontaneous double-strand breaks and DNA damage checkpoint activation. Unlike yeast Mms22, human Mms22L does not stably bind to Cul4, but is degraded in a Cul4-dependent manner and upon replication stress. Mms22L physically and functionally interacts with the scaffold-like protein Nfkbil2 that co-purifies with histones, several chromatin remodelling and DNA replication/repair factors. Together, our results strongly suggest that the Mms22L-Nfkbil2 complex contributes to genome stability by regulating the chromatin state at stalled replication forks.  相似文献   

20.
Hsiung HM  Smiley DL  Zhang XY  Zhang L  Yan LZ  Craft L  Heiman ML  Smith DP 《Peptides》2005,26(10):1988-1996
Human beta-MSH(1-22) was first isolated from human pituitary as a 22-amino acid (aa) peptide derived from a precursor protein, pro-opiomelanocortin (POMC). However, Bertagna et al. demonstrated that a shorter human beta-MSH(5-22), (DEGPYRMEHFRWGSPPKD), is a true endogenous peptide produced in human hypothalamus. In this report, we demonstrated that in vitro enzymatic cleavage of native human beta-MSH(5-22) with two ubiquitous dipeptidyl peptidases (DPP), DPP-I and DPP-IV, generated two potent MC3/4R peptide analogues, beta-MSH(7-22) (GPYRMEHFRWGSPPKD) and beta-MSH(9-22) (YRMEHFRWGSPPKD). In fact, the MC4R binding affinity and functional potency of beta-MSH(7-22) (Ki=4.6 nM, EC50=0.6 nM) and beta-MSH(9-22) (Ki=5.7 nM, EC50=0.6 nM) are almost an order of magnitude greater than those of their parent peptide, beta-MSH(5-22) (MC4R, Ki=23 nM, EC50= 3nM). Furthermore, the DPP-I/DPP-IV cleaved peptide, beta-MSH(9-22), when administered intracerebroventricularly (ICV) at a dose of 3 nmol/rat, potently induced an acute negative energy balance in a diet-induced obese rat model, while its parent molecule, beta-MSH(5-22), administered at the same dose did not have any effect. These data suggest that DPP-I and DPP-IV may play a role in converting the endogenous beta-MSH(5-22) to more potent peptides that regulate energy homeostasis in the hypothalamus.  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-23

京公网安备 11010802026262号