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 21 毫秒
1.
Guanine (Gua) modification by nitrating and hydroxylating systems was investigated in DNA. In isolated calf thymus DNA, 8-NO(2)-Gua and 8-oxo-Gua were dose-dependently formed with peroxynitrite, and 8-NO(2)-Gua was released in substantial amounts. Myeloperoxidase (MPO) with H(2)O(2) and NO(2)(-) reacted with calf thymus DNA to form 8-NO(2)-Gua dose dependently without release of 8-NO(2)-Gua. The frequency of strand breaks was higher than the sum of 8-NO(2)-Gua and 8-oxo-Gua, particularly in the MPO-treated DNA, indicating the importance of other types of damage. The activation of human neutrophils and lymphocytes with phorbol ester did not induce 8-NO(2)-Gua and 8-oxo-Gua in their nuclear DNA. However, 8-NO(2)-Gua was found in calf thymus DNA co-incubated with activated neutrophils in the presence of NO(2)(-). No significant formation of 8-NO(2)-Gua was found in liver DNA from mice treated with Escherichia coli lipopolysaccharide. The incubation of peroxynitrite or MPO-H(2)O(2)-NO(2)(-)-treated DNA with formamidopyrimidine glycosylase (Fpg) released 8-oxo-Gua, but not 8-NO(2)-Gua, indicating that 8-NO(2)-Gua is not a substrate for Fpg. Although 8-NO(2)-Gua was generated in isolated DNA by different nitrating systems, other types of damage were formed in abundance, and the lesion could not be found reliably in nuclear DNA, suggesting that the biological importance is limited.  相似文献   

2.
INTRODUCTION: Prostaglandins have an anabolic effect on bone. Possible mediation of this effect is via calcitriol. This study determines in vivo and in vitro effects of PGE(1) on calcitriol synthesis. METHODOLOGY: In vivo: rabbits received intravenous vehicle or prostaglandin E(1) (50 microg/day) for 20 days before measurements of serum total and ionic calcium, magnesium and phosphorus levels, total and bone-specific alkaline phosphatases, 25(OH)D(3), calcitriol, parathyroid hormone and calcitonin. In vitro: rabbit proximal renal tubules were incubated with 25(OH)D(3) (8 microM) together with PGE(1) (2.82 x 10(-6) M) and the prostaglandin receptor inhibitor AH6809 (10(-4) M) in selected samples. After 5 or 30 min incubation, calcitriol production was measured by radioimmunoassay and data analysed statistically. RESULTS: In vivo, in groups receiving PGE(1), levels of total Ca, Mg and calcitriol increased significantly and 25 dihydroxyvitamin D(3), parathyroid hormone and calcitonin remained unchanged. In vitro, PGE(1) increased calcitriol biosynthesis and the prostaglandin inhibitor AH6809 reduced calcitriol levels significantly after prolonged incubation. CONCLUSIONS: In vivo and in vitro results demonstrate that PGE(1) stimulates calcitriol synthesis. This study represent a major advancement in knowledge of bone metabolism.  相似文献   

3.
A significant proportion of P-glycoprotein (P-gp) and caveolin was co-localized in caveolae isolated from resistant (CH(R)C5) cells overexpressing P-gp and from drug-sensitive Chinese hamster ovary cells (AuxB1). The proportion of P-gp and caveolin associated with caveolar microdomains was higher in CH(R)C5 cells grown in the presence of P-gp substrates (cyclosporin A or colchicine) than in untreated CH(R)C5 cells. Coimmunoprecipitation of P-gp and caveolin from CH(R)C5 lysates suggests that there is a physical interaction between them. Furthermore, co-localization of P-gp and caveolin was found in caveolae from brain capillaries, indicating that this association also takes place in vivo.  相似文献   

4.
Shipbuilding workers are exposed to a variety of genotoxic compounds including polycyclic aromatic hydrocarbons (PAHs). A limited number of studies have been conducted to evaluate biomarkers related to PAH exposure in painters in the shipyard industry. We examined this in 208 workers recruited from a shipyard located in South Korea. Employees were grouped into three exposure groups: (1) 111 painters using coal tar paints, (2) 70 painters using general paints, and (3) 27 on-site controls using no paints. Levels of urinary 1-hydroxypyrene glucuronide (1-OHPG), as internal dose of PAH exposure, were measured by synchronous fluorescence spectroscopy. Glutathione S-transferase (GST) M1 and T1 genotypes were assessed by a multiplex polymerase chain reaction (PCR)-based method, aromatic-DNA adducts in peripheral white blood cells were measured by 32P-postlabeling, and glycophorin A (GPA) variant frequencies in red blood cells were assessed by flow cytometry. Information on demographic characteristics, smoking habits, diet, job title and use of personal protective equipment (e.g. respiratory and dermal) were collected by self-administered questionnaire. Average urinary 1-OHPG levels in coal tar paint (2.24 micromol/mol creatinine) and general paint (1.38 micromol/mol creatinine) users were significantly higher than in on-site controls (0.62 micromol/mol creatinine) (P<0.001). Paint use, irrespective of the type of paints, and smoking (yes/no) were positively associated with urinary 1-OHPG levels, whereas green tea consumption (yes/no) was negatively associated with the 1-OHPG levels. No significant effect in the 1-OHPG levels were observed for the GSTM1 and GSTT1 genotypes. Aromatic-DNA adduct levels tended to be higher in coal tar paint users (P = 0.06) and painters (P = 0.07) compared to on-site controls. No differences in adduct levels were observed, between the two groups of painters, and the combined group showed greater adduct levels than on-site controls (P = 0.05). GPA mutation frequencies measured in 55 individuals with MN heterozygote genotypes were not significantly different among the three exposure groups, and no correlation was observed between urinary 1-OHPG levels and aromatic-DNA adducts or GPA mutation frequency. These results suggest that painters in the shipyard were exposed to significant amounts of PAHs and possibly to other genotoxic aromatic compounds, and that urinary 1-OHPG may be a potential biomarker of PAH exposure in this population.  相似文献   

5.
The detailed mechanisms underlying morphine-signaling pathways in platelets remain obscure. Therefore, we systematically examined the influence of morphine on washed human platelets. In this study, washed human platelet suspensions were used for in vitro studies. Furthermore, platelet thrombus formation induced by irradiation of mesenteric venules with filtered light in mice pretreated with fluorescein sodium was used for an in vivo thrombotic study. Morphine concentration dependently (0.6, 1, and 5 microM) potentiated platelet aggregation and the ATP release reaction stimulated by agonists (i.e., collagen and U46619) in washed human platelets. Yohimbine (0.1 microM), a specific alpha(2)-adrenoceptor antagonist, markedly abolished the potentiation of morphine in platelet aggregation stimulated by agonists. Morphine also potentiated phosphoinositide breakdown and intracellular Ca(2+) mobilization in human platelets stimulated by collagen (1 microg/ml). Moreover, morphine (0.6-5 microM) markedly inhibited prostaglandin E(1) (10 microM)-induced cyclic AMP formation in human platelets, while yohimbine (0.1 microM) significantly reversed the inhibition of cyclic AMP by morphine (0.6 and 1 microM) in this study. The thrombin-evoked increase in pH(i) was markedly potentiated in the presence of morphine (1 and 5 microM). Morphine (2 and 5 mg/g) significantly shortened the time require to induce platelet plug formation in mesenteric venules. We concluded that morphine may exert its potentiation in platelet aggregation by binding to alpha(2)-adrenoceptors in human platelets, with a resulting inhibition of adenylate cyclase, thereby reducing intracellular cyclic AMP formation followed by increased activation of phospholipase C and the Na(+)/H(+) exchanger. This leads to increased intracellular Ca(2+) mobilization, and finally potentiation of platelet aggregation and of the ATP release reaction.  相似文献   

6.
S-Nitrosoalbumin (SNOALB) is the most abundant physiological circulating nitric oxide (NO) carrier regulating NO-dependent biological actions in humans. The mechanisms of its formation and biological actions are still incompletely understood. Nitrosation by authentic NO and S-transnitrosylation of the single sulfhydryl group located at Cys-34 of human albumin by the physiological S-nitroso compounds S-nitrosocysteine (SNOC) and S-nitrosoglutathione (GSNO) are two possible mechanisms. On a quantitative basis, we investigated by gas chromatography-mass spectrometry the contribution of these two mechanisms to SNOALB formation in human plasma and blood in vitro. GSNO and SNOC (0-100 microM) rapidly and efficiently (recovery=35%) S-transnitrosylated albumin to form SNOALB. NO (100 microM) S-nitrosated albumin to SNOALB at a considerably lower extent (recovery=5%). The putative NO-donating drugs glyceryl trinitrate and sodium nitroprusside (each 100 microM) failed completely in S-nitrosating albumin. Bubbling NO into human plasma and blood resulted in formation of SNOALB that inhibited ADP-induced platelet aggregation. Infusion of GS(15)NO in the rat resulted in formation of S(15)NOALB, [(15)N]nitrate and [(15)N]nitrite. Our results suggest that S-transnitrosylation of albumin by SNOC and GSNO could be a more favored mechanism for the formation of SNOALB in the circulation in vivo than S-nitrosation of albumin by NO itself.  相似文献   

7.
We studied the seasonal fluctuation of soil respiration (R(S)), and its root-dependent (R(R)) and basal (R(B)) components, in a Vitis vinifera (Chardonnay) vineyard. The R(S) components were estimated through independent field methods (y-intercept and trenching) and modeled on the basis of a Q(10) response to soil temperature, and fine and coarse root respiration coefficients. The effect of assimilate availability on R(R) was assessed through a trunk girdling treatment. The apparent Q(10) for R(R) was twice that of R(B) (3.5 vs 1.6) and increased linearly with increasing vine root biomass. The fastest R(R) of fine roots was during rapid fruit growth and the fastest R(R) of coarse roots was immediately following fruit development. R(S) was estimated at 32.6 kg ha(-1) d(-1) (69% as a result of R(R) ) for the hottest month and at 7.6 kg ha(-1) d(-1) (18% as a result of R(R)) during winter dormancy. Annual R(S) was low compared with other natural and cultivated ecosystems: 5.4 Mg ha(-1) (46% as a result of R(R)). Our estimates of annual vineyard R(S) are the first for any horticultural crop and suggest that the assumption that they are similar to those of annual crops or forest trees might lead to an overestimation.  相似文献   

8.
Iron-induced oxidative DNA damage in rat sperm cells in vivo and in vitro   总被引:3,自引:0,他引:3  
We investigated whether acute iron intoxication causes oxidative DNA damage, measured in terms of 7-hydro-8-oxo-2'-deoxyguanosine, 8-oxodG, in nuclear DNA in testes and epididymal sperm cells in vivo and in vitro in rats. In addition, we investigated levels of the modified nucleoside in liver and kidney and measured its urinary excretion. Sperm cells were isolated from the epididymides and the testes cells were isolated after homogenisation. In vitro, the sperm and testes cells were incubated with increasing concentrations of FeCl2 ranging from 0 to 600 microM. The median (range) levels of 8-oxodG/10(5) dG in the epididymal sperm cells increased from 0.48 (0.42-0.90) to 15.1 (11.4-17.6) (p < 0.05), whereas the level rose from 0.63 (0.22-0.81) to 8.8 (4.5-11.6) (p < 0.05) at 0 and 600 microM, respectively, in the testicular cells. In vivo groups of 7-8 rats received 0, 200 or 400 mg iron/kg as dextran i.p. After 24 h, epididymal sperm cells, testes, kidneys and liver were collected for analysis. Kidney and sperm DNA showed a significant increase in 8-oxodG in the iron-treated animals. The median (range) values of the 8-oxodG/10(5) dG in the epididymal sperm cells rose from 0.66 (0.38-1.09) to 1.12 (0.84-5.88) (p < 0.05) at 0 and 400 mg iron/kg, respectively, whereas the values in the testes and liver showed no significant change. In the kidneys the 8-oxodG/10(5) dG median (range) values were 0.98 (0.73-1.24), 1.21 (1.13-1.69) and 1.34 (1.12-1.66) after 0, 200 and 400 mg iron/kg, respectively (p < 0.05). The 8-oxodG-excretion rate was measured in 24h urine before and after iron treatment. The rate of urinary 8-oxodG excretion increased from 129 (104-179) pmol/24 h before treatment to 147 (110-239) pmol/24 h after treatment in the group receiving 400 mg iron/kg (p < 0.05). The results indicate that acute iron intoxication may increase oxidative damage to sperm and kidney DNA.  相似文献   

9.
Cpe(fat/fat) mice are obese, diabetic, and infertile. They have a mutation in carboxypeptidase E (CPE), an enzyme that converts prohormone intermediates to bioactive peptides. The Cpe(fat) mutation leads to rapid degradation of the enzyme. To test whether pro-thyrotropin-releasing hormone (TRH) conversion to TRH involves CPE, processing was examined in the Cpe(fat/fat) mouse. Hypothalamic TRH is depressed by at least 75% compared with wild-type controls. Concentrations of pro-TRH forms are increased in homozygotes. TRH-[Gly(4)-Lys(5)-Arg(6)] and TRH-[Gly(4)-Lys(5)] represent approximately 45% of the total TRH-like immunoreactivity in Cpe(fat/fat) mice; they constitute approximately 1% in controls. Levels of TRH-[Gly(4)] were depressed in homozygotes. Because the hypothalamus contains some TRH, another carboxypeptidase must be responsible for processing. Immunocytochemical studies indicate that TRH neurons contain CPE- and carboxypeptidase D-like immunoreactivity. Recombinant CPE or carboxypeptidase D can convert synthetic TRH-[Gly(4)-Lys(5)] and TRH-[Gly(4)-Lys(5)-Arg(6)] to TRH-[Gly(4)]. When Cpe(fat/fat) mice are exposed to cold, they cannot maintain their body temperatures, and this loss is associated with hypothalamic TRH depletion and reduction in thyroid hormone. These findings demonstrate that the Cpe(fat) mutation can affect not only carboxypeptidase activity but also endoproteolysis. Because Cpe(fat/fat) mice cannot sustain a cold challenge, and because alterations in the hypothalamic-pituitary-thyroid axis can affect metabolism, deficits in pro-TRH processing may contribute to the obese and diabetic phenotype in these mice.  相似文献   

10.
Mast cell activation, or neurogenic inflammation, is known to induce lowering of interstitial fluid pressure (P(if)) and plasma protein extravasation (PPE) in several tissues from both rats and mice. To examine a possible role of connective tissue mast cells (CTMCs) in these inflammatory responses, we used mice with dysfunctional CTMCs due to lack of the N-deacetylase/N-sulfotransferase-2 enzyme (NDST-2(-/-)). P(if) and PPE were measured after challenge with compound 48/80 (C48/80), and P(if) alone was measured after treatment either with capsaicin, substance P (SP), or calcitonin gene-related peptide (CGRP). Measurements of P(if) in anesthetized (fentanyl/fluanison and midazolam, 1:1) mice were performed in paw skin with glass capillaries connected to a servo-controlled counterpressure system. PPE was measured with microdialysis by using hollow plasmapheresis fibers (cutoff at 3,000 kDa) placed subcutaneously on the back. Intravenous administration of C48/80 lowered P(if) significantly (P < 0.05) in NDST-2(-/-) mice (-1.67 +/- 0.42 mmHg) compared with vehicle (-0.57 +/- 0.17 mmHg) but the lowering was significantly (P < 0.05) less compared with that of the NDST-2(+/+) mice (-2.31 +/- 0.47 mmHg). PPE was increased 300% after treatment with C48/80 in NDST-2(+/+) mice, whereas there was no increase in PPE in NDST-2(-/-) mice. Capsaicin, SP, and CGRP lowered P(if) significantly (P < 0.05) compared with vehicle and to the same extent in both NDST-2(+/+) and NDST-2(-/-) mice. We can conclude that although NDST-2(-/-) mice demonstrate an altered response in P(if) after mast cell activation, there was no similar alteration after neurogenic inflammation. Therefore, we suggest that neurogenic inflammation in mouse skin is not exclusively dependent on intact CTMCs.  相似文献   

11.
Amyloid-beta peptides (Abeta), generated by proteolysis of the beta-amyloid precursor protein (APP) by beta- and gamma-secretases, play an important role in the pathogenesis of Alzheimer disease (AD). Inflammation is also believed to be integral to the pathogenesis of AD. Here we show that prostaglandin E(2) (PGE(2)), a strong inducer of inflammation, stimulates the production of Abeta in cultured human embryonic kidney (HEK) 293 or human neuroblastoma (SH-SY5Y) cells, both of which express a mutant type of APP. We have demonstrated using subtype-specific agonists that, of the four main subtypes of PGE(2) receptors (EP(1-4)), EP(4) receptors alone or EP(2) and EP(4) receptors together are responsible for this PGE(2)-stimulated production of Abeta in HEK293 or SH-SY5Y cells, respectively. An EP(4) receptor antagonist suppressed the PGE(2)-stimulated production of Abeta in HEK293 cells. This stimulation was accompanied by an increase in cellular cAMP levels, and an analogue of cAMP stimulated the production of Abeta, demonstrating that increases in the cellular level of cAMP are responsible for the PGE(2)-stimulated production of Abeta. Immunoblotting experiments and direct measurement of gamma-secretase activity suggested that PGE(2)-stimulated production of Abeta is mediated by activation ofgamma-secretase but not of beta-secretase. Transgenic mice expressing the mutant type of APP showed lower levels of Abeta in the brain, when they were crossed with mice lacking either EP(2) or EP(4) receptors, suggesting that PGE(2)-mediated activation of EP(2) and EP(4) receptors is involved in the production of Abeta in vivo and in the pathogenesis of AD.  相似文献   

12.
In free-living (FL) reindeer eating a natural mixed winter diet dominated by lichens, captive (CF) reindeer fed pure lichens ad libitum, and CF reindeer subsequently starved for 1 day (CS1 reindeer) or 4 days (CS4 reindeer), the dominant rumen anaerobic bacteria were characterized, their population densities were estimated, and ruminal pH and volatile fatty acid concentrations were determined. In the FL reindeer, the total median viable anaerobic bacterial population ranged from 18 x 10(8) to 35 x 10(8) cells per ml of rumen fluid (n = 4), compared with 26 x 10(8) to 34 x 10(8) and 0.09 x 10(8) to 0.1 x 10(8) cells per ml of rumen fluid in CF reindeer (n = 2) and CS4 reindeer (n = 2), respectively. The median bacterial population adhering to the rumen solids ranged from 260 x 10(8) to 450 x 10(8), 21 x 10(8) to 38 x 10(8), and 0.5 x 10(8) cells per g (wet weight) of rumen solids in FL, CF, and CS4 reindeer, respectively. Although there were variations in the rumen bacterial composition among the FL reindeer (n = 4), strains of Bacteroides, Fibrobacter, Streptococcus, and Clostridium dominated in the rumen fluid. Streptococcus spp. and Clostridium spp. were the dominant bacteria in the CF reindeer (n = 2), while in the CS4 reindeer (n = 2) the dominant bacteria were Fusobacterium spp., members of the family Enterobacteriaceae, and Eubacterium spp. Transmission electron micrographs of lichen particles from the rumen of one FL reindeer, one CF reindeer, and one CS4 reindeer show bacteria resembling Bacteroides spp. adhering to the lichen particles, evidently digesting the lichen hyphae from the inside.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
Pentoxifylline increases erythrocyte flexibility, reduces blood viscosity, and inhibits platelet aggregation and is thus used in the treatment of peripheral vascular disease. It is transformed into at least seven phase I metabolites, of which two, M1 and M5, are active. The reduction of the keto group of pentoxifylline to a secondary alcohol in M1 takes place chiefly in erythrocytes, is rapidly reversible, and creates a chiral center. The aims of this study were: to develop HPLC methods to separate the enantiomers of M1, to investigate the kinetics of the reversible biotransformation of pentoxifylline to (R)- and (S)-M1 in hemolysed erythrocyte suspension, and to quantify the formation of the enantiomers of M1 (as well as M4 and M5) after intravenous and oral administration of pentoxifylline to human volunteers. (R)- and (S)-M1 could be separated preparatively on a cellobiohydrolase column, while determination in blood or plasma was by HPLC after chiral derivatization with diacetyl-L-tartaric acid anhydride. The metabolism of pentoxifylline to (R)-M1 in suspensions of hemolysed erythrocytes followed simple Michaelis-Menten kinetics (K(m) = 11 mM), while that to (S)-M1 was best described by a two-enzyme model (K(m) = 1.1 and 132 mM). Studies with inhibitors indicated that the enzymes were of the carbonyl reductase type. At a therapeutic blood concentration of pentoxifylline, the calculated rate of formation of (S)-M1 is 15 times higher than that of the (R)-enantiomer. Back-conversion of M1 to pentoxifylline was 3-4 times faster for the (S)- than for the (R)-enantiomer. In vivo, the R:S plasma concentration ratio of M1 ranged from 0.010-0.025 after intravenous infusion of 300 or 600 mg of pentoxifylline, and from 0.019-0.037 after oral administration of 600 mg. The biotransformation of pentoxifylline to M1 was thus highly stereoselective in favor of the (S)-enantiomer both in vitro and in vivo.  相似文献   

14.
The meningeal worm (Parelaphostrongylus tenuis) was found in 22 (7%) of 300 white-tailed deer (Odocoileus virginianus) (257 adults, 43 fawns) examined from Nebraska (USA) during November 1996. None of 53 mule deer (Odocoileus hemionus) (47 adults and 6 fawns) examined were infected. Twenty-two white-tailed deer from 18 counties in eastern Nebraska were infected with Parelaphostrongylus tenuis. This is the first record of P. tenuis in white-tailed deer from this state.  相似文献   

15.
Hemorrhagic disease in deer in Arizona   总被引:1,自引:0,他引:1  
Two mule deer (Odocoileus hemionus) and one white-tailed deer (Odocoileus virginianus) in Arizona (USA) were submitted for necropsy. Gross and microscopic lesions compatible with hemorrhagic disease (HD) were observed in all three deer. Epizootic hemorrhagic disease virus type 2 (EHDV-2) was isolated from two of the deer. To our knowledge, this is the first documentation of HD in deer in Arizona. Two of the mortalities were attributed to EHDV-2 infection.  相似文献   

16.
Group X secretory PLA(2) (sPLA(2)-X) is expressed in neutrophils and plays a role in the pathogenesis of neutrophil-mediated tissue inflammation and injury. This study tested the hypothesis that sPLA(2)-X in neutrophils may contribute to the pathogenesis of abdominal aortic aneurysms (AAA) using sPLA(2)-X(-/-) mice. AAA was created by application of CaCl(2) to external surface of aorta. As a result, the aortas of sPLA(2)-X(-/-) mice had smaller diameters (percent increase from baseline; 24.8 ± 3.5% vs. 49.9 ± 9.1%, respectively; P < 0.01), a reduced grade of elastin degradation, and lower activities of elastase and gelatinase (26% and 19% lower, respectively) after CaCl(2) treatment compared with sPLA(2)-X(+/+) mice. In sPLA(2)-X(+/+) mice, immunofluorescence microscopic images showed that the immunoreactivity of sPLA(2)-X was detected only in neutrophils within aortic walls 3 days, 1, 2, and 6 wk after CaCl(2) treatment, whereas the immunoreactivity was not detected in macrophages or mast cells in aortic walls. sPLA(2)-X immunoreactivity also was colocalized in cells expressing matrix metalloproteinase (MMP)-9. Neutrophils isolated from sPLA(2)-X(-/-) mice had lower activities of elastase, gelatinase, and MMP-9 in response to stimuli compared with sPLA(2)-X(+/+) mice. The attenuated release of elastase and gelatinase from sPLA(2)-X(-/-) neutrophils was reversed by exogenous addition of mouse sPLA(2)-X protein. The adoptive transfer of sPLA(2)-X(+/+) neutrophils days 0 and 3 after CaCl(2) treatment reversed aortic diameters and elastin degradation grades in the lethally irradiated sPLA(2)-X(+/+) mice reconstituted with sPLA(2)-X(-/-) bone marrow to an extent similar to that seen in sPLA(2)-X(+/+) mice. In conclusion, sPLA(2)-X in neutrophils plays a pathogenic role in AAA in a mice model.  相似文献   

17.
To determine whether IL-6 increases lipolysis and fat oxidation in patients with type 2 diabetes and/or whether it exerts this effect independently of changes to the hormonal milieu, patients with type 2 diabetes (D) and healthy control subjects (CON) underwent recombinant human (rh)IL-6 infusion for 3 h. Rates of appearance (Ra) and disappearance (Rd) of [U-(13C)]palmitate and [6,6-(2H2)]glucose were determined. rhIL-6 infusion increased (P < 0.05) palmitate Ra and Rd in a similar fashion in both groups. Neither plasma glucose concentration nor glucose Ra/Rd was affected by rhIL-6 infusion in either group, whereas rhIL-6 infusion resulted in a reduction (P < 0.05) in circulating insulin in D. Plasma growth hormone (GH) was increased (P < 0.05) by IL-6 in CON, and cortisol increased (P < 0.05) in response to IL-6 in both groups. To determine whether IL-6 was exerting its effect directly or through activation of these hormones, we performed cell culture experiments. Fully differentiated 3T3-L1 adipocytes were treated with PBS (control) IL-6, or IL-6 plus dexamethasone and GH. IL-6 treatment alone increased (P < 0.05) lipolysis, but this effect was reduced by the addition of dexamethasone and GH such that IL-6 plus dexamethasone and GH had blunted (P < 0.05) lipolysis compared with IL-6 alone. To assess whether IL-6 increases fat oxidation, L6 myotubes were treated with PBS (Control), IL-6, or AICAR, a compound known to increase lipid oxidation. Both IL-6 and AICAR markedly increased (P < 0.05) oxidation of [(14)C]palmitate compared with Control. Acute IL-6 treatment increased fatty acid turnover in D patients as well as healthy CON subjects. Moreover, IL-6 appears to be activating lipolysis independently of elevations in GH and/or cortisol and appears to be a potent catalyst for fat oxidation in muscle cells.  相似文献   

18.
The effects of cadmium chloride (CdCl(2)) on oxidative stress in the skeletal muscle cell line C(2)C(12) were investigated. Myoblast cells that differentiated into myotubes were treated with CdCl(2) (1, 3, 5, 7.5, 10, and 12.5 microM) for 24, 48, and 72 h. Subsequent assay of cell homogenates for MTT (3-(4,5-dimethylthiozol-2-yl)-2,5-diphenyltetrazolium bromide) reduction, neutral red uptake and nucleic acid content showed that cadmium was toxic to C(2)C(12) cells in a concentration-dependent manner. Glutathione-S-transferase activity (nmol microg of protein(-1) min(-1)) was increased with 1 and 3 microM CdCl(2) (36.9 +/- 5.6 and 32.1 +/- 6.0, respectively) compared to control cells (21.8 +/- 1.5), but decreased at higher concentrations (7.5 microM = 15.9 +/- 3.3, 10 microM = 15.9 +/- 4.6, and 12.5 microM = 10.5 +/- 2.8). An increase in malondialdehyde content (nmol microg of protein(-1)), especially at high CdCl(2) concentrations (control = 7.3 +/- 0.5; CdCl(2): 7.5 microM = 11.2 +/- 3.1, 10 microM = 14.6 +/- 3.8, and 12.5 microM = 20.5 +/- 6.5) indicated that there was enhanced lipid peroxidation. Light and scanning electron microscopy showed that there was a concentration-dependent loss of adherent cells and the formation of vesicles indicative of cell death. These results indicated that CdCl(2) increased oxidative stress in C(2)C(12) cells, and this stress probably compromised cell adhesion and the cellular antioxidant defense mechanisms.  相似文献   

19.
Statistical approaches have been applied to examine amino acid pairing preferences within parallel beta-sheets. The main chain hydrogen bonding pattern in parallel beta-sheets means that, for each residue pair, only one of the residues is involved in main chain hydrogen bonding with the strand containing the partner residue. We call this the hydrogen bonded (HB) residue and the partner residue the non-hydrogen bonded (nHB) residue, and differentiate between the favorability of a pair and that of its reverse pair, e.g. Asn(HB)-Thr(nHB)versus Thr(HB)-Asn(nHB). Significantly (p < or = 0.000001) favoured pairings were rationalised using stereochemical arguments. For instance, Asn(HB)-Thr(nHB) and Arg(HB)-Thr(nHB) were favoured pairs, where the residues adopted favoured chi1 rotamer positions that allowed side-chain interactions to occur. In contrast, Thr(HB)-Asn(nHB) and Thr(HB)-Arg(nHB) were not significantly favoured, and could only form side-chain interactions if the residues involved adopted less favourable chi1 conformations. The favourability of hydrophobic pairs e.g. Ile(HB)-Ile(nHB), Val(HB)-Val(nHB) and Leu(HB)-Ile(nHB) was explained by the residues adopting their most preferred chi1 and chi2 conformations, which enabled them to form nested arrangements. Cysteine-cysteine pairs are significantly favoured, although these do not form intrasheet disulphide bridges. Interactions between positively and negatively charged residues were asymmetrically preferred: those with the negatively charged residue at the HB position were more favoured. This trend was accounted for by the presence of general electrostatic interactions, which, based on analysis of distances between charged atoms, were likely to be stronger when the negatively charged residue is the HB partner. The Arg(HB)-Asp(nHB) interaction was an exception to this trend and its favorability was rationalised by the formation of specific side-chain interactions. This research provides rules that could be applied to protein structure prediction, comparative modelling and protein engineering and design. The methods used to analyse the pairing preferences are automated and detailed results are available (http://www.rubic.rdg.ac.uk/betapairprefsparallel/).  相似文献   

20.
A spontaneous Xanthomonas campestris pv. phaseoli H(2)O(2)-resistant mutant emerged upon selection with 1 mM H(2)O(2). In this report, we show that growth of this mutant under noninducing conditions gave high levels of catalase, alkyl hydroperoxide reductase (AhpC and AhpF), and OxyR. The H(2)O(2) resistance phenotype was abolished in oxyR-minus derivatives of the mutant, suggesting that elevated levels and mutations in oxyR were responsible for the phenotype. Nucleotide sequence analysis of the oxyR mutant showed three nucleotide changes. These changes resulted in one silent mutation and two amino acid changes, one at a highly conserved location (G197 to D197) and the other at a nonconserved location (L301 to R301) in OxyR. Furthermore, these mutations in oxyR affected expression of genes in the oxyR regulon. Expression of an oxyR-regulated gene, ahpC, was used to monitor the redox state of OxyR. In the parental strain, a high level of wild-type OxyR repressed ahpC expression. By contrast, expression of oxyR5 from the X. campestris pv. phaseoli H(2)O(2)-resistant mutant and its derivative oxyR5G197D with a single-amino-acid change on expression vectors activated ahpC expression in the absence of inducer. The other single-amino-acid mutant derivative of oxyR5L301R had effects on ahpC expression similar to those of the wild-type oxyR. However, when the two single mutations were combined, as in oxyR5, these mutations had an additive effect on activation of ahpC expression.  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