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 54 毫秒
1.
Hyperpolarization enhances the intensity of the NMR signals of a molecule, whose in vivo metabolic fate can be monitored by MRI with higher sensitivity. SABRE is a hyperpolarization technique that could potentially be used to image nitric oxide (NO) production in vivo. This would be very important, because NO dysregulation is involved in several pathologies, including cardiovascular ones. The nitric oxide synthase (NOS) pathway leads to NO production via conversion of l-arginine into l-citrulline. NO is a free radical gas with a short half-life in vivo (≈5 s), therefore direct NO quantification is challenging. An indirect method – based on quantifying conversion of an l-Arg- to l-Cit-derivative by 1H NMR spectroscopy – is herein proposed. A small library of pyridyl containing l-Arg derivatives was designed and synthesised. In vitro tests showed that compounds 4aj and 11ac were better or equivalent substrates for the eNOS enzyme (NO2? production = 19–46 μM) than native l-Arg (NO2? production = 25 μM). Enzymatic conversion of l-Arg to l-Cit derivatives could be monitored by 1H NMR. The maximum hyperpolarization achieved by SABRE reached 870-fold NMR signal enhancement, which opens up exciting future perspectives of using these molecules as hyperpolarized MRI tracers in vivo.  相似文献   

2.
Carbon fiber microelectrodes and carbon fiber composite minielectrodes (CFM/CFCM) have been generally used for measurements of nitric oxide (NO) concentration in chemical and biological systems. The response time of a CFM/CFCM is usually from milliseconds to seconds depending on the electrode size, the thickness of coating layers on the electrode, and NO diffusion coefficients of the coating layers. As a result, the time course of recoded current changes (It curves) by the CFM/CFCM may be different from the actual time course of NO concentration changes (ct curves) if the half-life of NO decay is close to or shorter than the response time of the electrode used. This adds complexity to the process for determining rate constants of NO decay kinetics from the recorded current curves (It curves). By computer simulations based on a mathematical model, an approximation method was developed for determining rate constants of NO decay from the recorded current curves. This method was first tested and valuated using a commercial CFCM in several simple reaction systems with known rate constants. The response time of the CFCM was measured as 4.7 ± 0.7 s (n = 5). The determined rate constants of NO volatilization and NO autoxidation in our measurement system at 37 °C are (1.9 ± 0.1) × 10?3 s?1 (n = 4) and (2.0 ± 0.3) × 103 M?1 s?1 (n = 7), which are close to the reported rate constants. The method was then applied to determine the rate of NO decay in blood samples from control and smoking exposed mice. It was observed that the NO decay rate in the smoking group is >20% higher than that in control group, and the increased NO decay rate in the smoking group was reversed by 10 μM diphenyleneiodonium chloride (DPI), an inhibitor of flavin enzymes such as leukocyte NADPH oxidase.  相似文献   

3.
PurposeIt has been suggested that nitric oxide (NO) has a role in ischemic retinopathies. Since retinal ischemia may develop in retinal vein occlusion, we investigated the presence of nitric oxide in the pathogenesis of central retinal vein occlusion (CRVO).MethodsEighteen consecutive patients with CRVO were included in this study. Aqueous humor specimens were obtained within 21 days of diagnosis. Samples of aqueous humor were also collected from 20 control patients undergoing cataract surgery. For each sample after reduction of nitrate to nitrite with vanadium chloride (VCl3), we used spectrophotometric method for simultaneous detection of nitrate and nitrite (NOx).ResultsMean level of aqueous humor NOx in CRVO and control group was 94.1 ± 23.2 μmol/l and 55.6 ± 11.0 μmol/l, respectively. The difference between two groups was statistically significant (p < 0.0001).ConclusionsOur results may support involvement of nitric oxide in the pathogenesis of CRVO.  相似文献   

4.
The reactions of NO2 with both oxidized and reduced cytochrome c at pH 7.2 and 7.4, respectively, and with N-acetyltyrosine amide and N-acetyltryptophan amide at pH 7.3 were studied by pulse radiolysis at 23 °C. NO2 oxidizes N-acetyltyrosine amide and N-acetyltryptophan amide with rate constants of (3.1±0.3)×105 and (1.1±0.1)×106 M−1 s−1, respectively. With iron(III)cytochrome c, the reaction involves only its amino acids, because no changes in the visible spectrum of cytochrome c are observed. The second-order rate constant is (5.8±0.7)×106 M−1 s−1 at pH 7.2. NO2 oxidizes iron(II)cytochrome c with a second-order rate constant of (6.6±0.5)×107 M−1 s−1 at pH 7.4; formation of iron(III)cytochrome c is quantitative. Based on these rate constants, we propose that the reaction with iron(II)cytochrome c proceeds via a mechanism in which 90% of NO2 oxidizes the iron center directly—most probably via reaction at the solvent-accessible heme edge—whereas 10% oxidizes the amino acid residues to the corresponding radicals, which, in turn, oxidize iron(II). Iron(II)cytochrome c is also oxidized by peroxynitrite in the presence of CO2 to iron(III)cytochrome c, with a yield of ~60% relative to peroxynitrite. Our results indicate that, in vivo, NO2 will attack preferentially the reduced form of cytochrome c; protein damage is expected to be marginal, the consequence of formation of amino acid radicals on iron(III)cytochrome c.  相似文献   

5.
The role of endogenous nitric oxide (NO) in modulating myocardial contractility is still unclear, in part because of unknown, secondary effects of blocking NO release. We hypothesized that the nonspecific inhibition of nitric oxide synthase (NOS) enhances endothelin-1 (ET-1) effects, which can play a role in ET-A receptor-dependent myocardial contractile responses. The myocardial contractility was estimated from the slope of the left ventricular end-systolic pressure–diameter relationship in closed-chest, pentobarbital-anesthetized dogs. Group 1 (n = 7) was the saline-treated control, while in groups 2 (n = 7) and 3 (n = 7) N-nitro-l-arginine (NNA, 4 mg kg?1), a nonselective NOS blocker, was administered with or without pretreatment with the ET-A receptor antagonist ETR-P1/fl peptide (100 nmol kg?1 iv). Plasma ET-1, nitrite/nitrate (NOx) and blood superoxide levels were measured, and myocardial ET-1 content and xanthine oxidoreductase (XOR) activity were determined from myocardial biopsies. The infusion of NNA over 120 min decreased the plasma NOx, significantly elevated the plasma ET-1 and blood superoxide levels, and in parallel greatly increased the left ventricular contractility as compared with the untreated controls [47.5 vs 30 mm Hg mm?1]. The myocardial ET-1 content decreased simultaneously, while the XOR activity and blood superoxide level were significantly elevated. These effects, including NNA-induced positive inotropy, were significantly suppressed by pretreatment with ETR-P1/fl peptide. These results demonstrate that a diminished NO synthesis leads to a preponderant ET-1 effect, which increases myocardial contractility through an ET-A receptor-dependent mechanism.  相似文献   

6.
While endogenous nitric oxide (NO) may be relevant to the beneficial hemodynamic effects produced by sildenafil during acute pulmonary embolism (APE), huge amounts of inducible NO synthase (iNOS)-derived NO may contribute to lung injury. We hypothesized that iNOS inhibition with S-methylisothiourea could attenuate APE-induced increases in oxidative stress and pulmonary hypertension and, therefore, could improve the beneficial hemodynamic and antioxidant effects produced by sildenafil during APE. Hemodynamic evaluations were performed in non-embolized dogs treated with saline (n = 4), S-methylisothiourea (0.01 mg/kg followed by 0.5 mg/kg/h, n = 4), sildenafil (0.3 mg/kg, n = 4), or S-methylisothiourea followed by sildenafil (n = 4), and in dogs that received the same drugs and were embolized with silicon microspheres (n = 8 for each group). Plasma nitrite/nitrate (NOx) and thiobarbituric acid reactive substances (TBARS) concentrations were determined by Griess and a fluorometric assay, respectively. APE increased mean pulmonary arterial pressure (MPAP) and pulmonary vascular resistance index (PVRI) by 25 ± 1.7 mm Hg and by 941 ± 34 dyn s cm?5 m?2, respectively. S-methylisothiourea neither attenuated APE-induced pulmonary hypertension, nor enhanced the beneficial hemodynamic effects produced by sildenafil after APE (>50% reduction in pulmonary vascular resistance). While sildenafil produced no change in plasma NOx concentrations, S-methylisothiourea alone or combined with sildenafil blunted APE-induced increases in NOx concentrations. Both drugs, either alone or combined, produced antioxidant effects. In conclusion, although iNOS-derived NO may play a key role in APE-induced oxidative stress, our results suggest that the iNOS inhibitor S-methylisothiourea neither attenuates APE-induced pulmonary hypertension, nor enhances the beneficial hemodynamic effects produced by sildenafil.  相似文献   

7.
15N-labelled NO3? was used in a surface-flow constructed wetland in spring to examine the relative importance of competing NO3? removal processes. In situ mesocosms (0.25 m2) were dosed with 2 l of 15NO3? (NaNO3, 300 mg N l?1, 99 atom% 15N) and bromide (Br?) solution (LiBr, 4.3 g l?1, as a conservative tracer). Concentrations of NO3?, Br?, dissolved oxygen and 15N2 were monitored periodically and replicate mesocosms were destructively sampled prior to and 6 days after 15N addition. Denitrification, immobilisation, plant uptake and dissimilatory NO3? reduction to NH4+ (DNRA) accounted for 77, 11, 9 and 2% of 15NO3? transformed during the experiment. Only 6% of denitrification gases were directly measured as atmospheric or dissolved 15N2; the remainder (71%) was determined via 15N mass balance. This indicated that a large proportion of the denitrification gases were entrapped within the soil matrix and/or plant aerenchyma. The floating plant Lemna minor exhibited a significantly higher NO3? uptake rate (221 mg kg?1 d?1) than Typha orientalis (10 mg kg?1 d?1), but periodic harvest of plants would remove <3% of annual NO3? inputs. Our results suggest that this 6-year-old constructed wetland functions effectively as a sink for NO3? during the growing season with less than one-quarter of the NO3? processed sequestered into wetland plant, algal and microbial N pools and the balance permanently removed by denitrification.  相似文献   

8.
Pike-perch Sander lucioperca is currently considered as one of the most promising candidates for production in freshwater recirculation aquaculture systems (RAS). Here, due to the lack of studies on nitrite (NO2?) toxicity in pike-perch, a flow-through exposure at 0, 0.44, 0.88, 1.75, 3.5, 7, 14 and 28 mg/L NO2?–N was carried out to determine the acute and chronic toxicity over a period of 32 days. In juvenile pike-perch, 120 h LC50 was 6.1 mg/L NO2?–N and at ≥ 14 mg/L NO2?–N all fish had died within 24 h. Chronic exposure revealed a significant build up of NO2? in the plasma as well as in the muscles at ≥ 0.44 mg/L NO2?–N peaking in fish exposed to the highest concentration of 3.5 mg/L NO2?–N after 32 days. Still, due to high individual variation methemoglobin (MetHb) was only significantly increased (p < 0.01) at 3.5 mg/L NO2?–N. No adverse effects on red blood cells (RBC) and hematocrit were observed in any of the treatments. In a second experiment, compensation of NO2? toxicity at increasing chloride concentrations (40 (freshwater), 65, 90, 140, 240, 440 mg/L Cl?) was observed at a constant exposure of 10 mg/L NO2?–N for 42 days. At ≥ 240 mg/L Cl?, NO2? build-up in blood plasma and muscle was completely inhibited. At lower Cl? concentrations (≤ 140 mg/L), NO2? was significantly increased in plasma, but only insignificantly elevated in muscle due to high individual variation. MetHb was increased significantly difference only at 40 mg/L Cl? (freshwater control) compared to the control. Again, high individual variations were observed. As a conclusion, S. lucioperca is moderately sensitive towards NO2? and acceptable levels in RAS should hence not exceed 1.75 mg/L NO2?–N to avoid MetHb formation. However, based on the 120 h LC50 and a factor of 0.01 according to Sprague (1971), a NO2? concentration of ≤ 0.061 mg/L NO2?–N is considered as “safe.” Thereby, no NO2? should accumulate in the plasma or muscle tissue during chronic exposure. For 10 mg/L NO2?–N, ≥ 240 mg/L chloride compensates for NO2? uptake in plasma and muscle.  相似文献   

9.
The series of nitrosyl complexes trans-[Ru(NH3)4L(NO)]Cl3, L = caffeine, theophylline, imidazole and benzoimidazole in position trans to NO were prepared and their photochemical properties studied. All complexes showed nitric oxide (NO) release under light irradiation at 330–440 nm. Quantum yields for [Ru(NH3)4L(H2O)]3+ formation (?Ru(III)) were sensitive to the natures of L, λirr and pH. The major product of the irradiation of trans-[Ru(NH3)4L(NO+)]3+ is the trans-[RuIII(NH3)4L(Cl)]2+ and NO as suggested by UV–Vis, electrochemical, and FTIR techniques.  相似文献   

10.
Chung JY  Kim JE  Yoon HJ  Song SY  Kim SO  Roh WS 《Cryobiology》2012,65(1):33-40
Moderate hypothermia (25–31 °C) may have a significant influence on vascular tone. We investigated the cellular mechanisms by which moderate hypothermia alters α-adrenoceptor-mediated contraction in rat thoracic aortae. Cyclooxygenase inhibition by indomethacin; nitric oxide (NO) synthase inhibition by l-NAME; potassium channel and endothelium-derived hyperpolarizing factor (EDHF) inhibition by glibenclamide and TEA; G protein inhibition by pertussis toxin; α2-adrenergic inhibition by yohimbine; and β-adrenergic inhibition by propranolol were assessed for their effect on the contractile response to the α1-adrenoceptor agonist phenylephrine (Phe) in combination with moderate hypothermia (25 °C). Moderate hypothermia produced a shift to the right for the Phe concentration–response curves in endothelium-intact (E+) and endothelium-denuded (E?) aortic rings. The maximal response to Phe in E+ rings was significantly decreased (P < 0.05) at 25 °C compared to 38 °C, whereas there was no significant difference in E? rings. Hypothermia-induced vasorelaxation in E+ rings was attenuated (P < 0.05) following combined pretreatment with l-NAME (10?4 M) and indomethacin (10?5 M), whereas other inhibitors had no significant effect. Importantly, the addition of TEA to rings that were pretreated with l-NAME and indomethacin exhibited no further attenuation (P > 0.05) of hypothermia-induced vasorelaxation. The concentrations of cGMP and cAMP, as measured by radioimmunoassay, were significantly increased (P < 0.05) in E+ rings at 25 °C compared to those at 38 °C, whereas there were no significant differences (P > 0.05) in E? rings. The present study demonstrated that rat aortic endothelium is stimulated during moderate hypothermia and that the NO–cGMP and prostacyclin (PGI2)–cAMP pathways represent endothelium-dependent mechanisms of hypothermia-induced vasorelaxation. In contrast, EDHF may not be associated with hypothermia-induced vasorelaxation.  相似文献   

11.
A microbial consortium attached onto a polyethylene support was used to evaluate the simultaneous oxidation of sulfide and phenol by denitrification. The phenol, sulfide and nitrate loading rates applied to an inverse fluidized bed reactor were up to 168 mg phenol–C/(l d), 37 mg S2?/(l d) and 168 mg NO3?–N/(l d), respectively. Under steady state operation the consumption efficiencies of phenol, sulfide and nitrate were 100%. The N2 yield (g N2/g NO3?–N) was 0.89. The phenol was mineralized resulting in a yield of 0.82 g bicarbonate–C/g phenol–C and sulfide was completely oxidized to sulfate with a yield of 0.99 g SO42?–S/g S2?. 16S rRNA gene-based microbial community analysis of the denitrifying biofilm showed the presence of Thauera aromatica, Thiobacillus denitrificans, Thiobacillus sajanensis and Thiobacillus sp. This is the first work reporting the simultaneous oxidation of sulfide and phenol in a denitrifying biofilm reactor.  相似文献   

12.
Proteins containing reactive cysteine residues (protein-Cys) are receiving increased attention as mediators of hydrogen peroxide signaling. These proteins are mainly identified by mining the thiol proteomes of oxidized protein-Cys in cells and tissues. However, it is difficult to determine if oxidation occurs through a direct reaction with hydrogen peroxide or by thiol–disulfide exchange reactions. Kinetic studies with purified proteins provide invaluable information about the reactivity of protein-Cys residues with hydrogen peroxide. Previously, we showed that the characteristic UV–Vis spectrum of horseradish peroxidase compound I, produced from the oxidation of horseradish peroxidase by hydrogen peroxide, is a simple, reliable, and useful tool to determine the second-order rate constant of the reaction of reactive protein-Cys with hydrogen peroxide and peroxynitrite. Here, the method is fully described and extended to quantify reactive protein-Cys residues and micromolar concentrations of hydrogen peroxide. Members of the peroxiredoxin family were selected for the demonstration and validation of this methodology. In particular, we determined the pKa of the peroxidatic thiol of rPrx6 (5.2) and the second-order rate constant of its reactions with hydrogen peroxide ((3.4 ± 0.2) × 107 M? 1 s? 1) and peroxynitrite ((3.7 ± 0.4) × 105 M? 1 s? 1) at pH 7.4 and 25 °C.  相似文献   

13.
14.
The effect of NO between cytochromes b and c of the mitochondrial respiratory chain were studied using submitochondrial particles (SMP) from bovine heart and GSNO and SPER-NO as NO sources. Succinate-cytochrome c reductase (complex II-III) activity (222±4 nmol/min. mg protein) was inhibited by 51% in the presence of 500 μM GSNO and by 48% in the presence of 30 μM SPER-NO, in both cases at ~1.25 μM NO. Neither GSNO nor SPER-NO were able to inhibit succinate-Q reductase activity (complex II; 220±9 nmol/min. mg protein), showing that NO affects complex III. Complex II-III activity was decreased (36%) when SMP were incubated with l-arginine and mtNOS cofactors, indicating that this effect is also produced by endogenous NO. GSNO (500 μM) reduced cytochrome b562 by 71%, in an [O2] independent manner. Hyperbolic increases in O2•- (up to 1.3±0.1 nmol/min. mg protein) and H2O2 (up to 0.64±0.05 nmol/min. mg protein) productions were observed with a maximal effect at 500 μM GSNO. The O2•-/H2O2 ratio was 1.98 in accordance with the stoichiometry of the O2•- disproportionation. Moreover, H2O2 production was increased by 72–74% when heart coupled mitochondria were exposed to 500 μM GSNO or 30 μM SPER-NO. SMP incubated in the presence of succinate showed an EPR signal (g=1.99) compatible with a stable semiquinone. This EPR signal was increased not only by antimycin but also by GSNO and SPER-NO. These signals were not modified under N2 atmosphere, indicating that they are not a consequence to the effect of NOx species on complex III area. These results show that NO interacts with ubiquinone-cytochrome b area producing antimycin-like effects. This behaviour comprises the inhibition of electron transfer, the interruption of the oxidation of cytochromes b, and the enhancement of [UQH]ss which, in turn, leads to an increase in O2•- and H2O2 mitochondrial production rates.  相似文献   

15.
The purpose of the work was to provide a crystallographic demonstration of the venerable idea that CO photolyzed from ferrous heme-a3 moves to the nearby cuprous ion in the cytochrome c oxidases. Crystal structures of CO-bound cytochrome ba3-oxidase from Thermus thermophilus, determined at ~ 2.8–3.2 Å resolution, reveal a Fe–C distance of ~ 2.0 Å, a Cu–O distance of 2.4 Å and a Fe–C–O angle of ~ 126°. Upon photodissociation at 100 K, X-ray structures indicate loss of Fea3–CO and appearance of CuB–CO having a Cu–C distance of ~ 1.9 Å and an O–Fe distance of ~ 2.3 Å. Absolute FTIR spectra recorded from single crystals of reduced ba3–CO that had not been exposed to X-ray radiation, showed several peaks around 1975 cm? 1; after photolysis at 100 K, the absolute FTIR spectra also showed a significant peak at 2050 cm? 1. Analysis of the ‘light’ minus ‘dark’ difference spectra showed four very sharp CO stretching bands at 1970 cm? 1, 1977 cm? 1, 1981 cm? 1, and 1985 cm? 1, previously assigned to the Fea3–CO complex, and a significantly broader CO stretching band centered at ~ 2050 cm? 1, previously assigned to the CO stretching frequency of CuB bound CO. As expected for light propagating along the tetragonal axis of the P43212 space group, the single crystal spectra exhibit negligible dichroism. Absolute FTIR spectrometry of a CO-laden ba3 crystal, exposed to an amount of X-ray radiation required to obtain structural data sets before FTIR characterization, showed a significant signal due to photogenerated CO2 at 2337 cm? 1 and one from traces of CO at 2133 cm? 1; while bands associated with CO bound to either Fea3 or to CuB in “light” minus “dark” FTIR difference spectra shifted and broadened in response to X-ray exposure. In spite of considerable radiation damage to the crystals, both X-ray analysis at 2.8 and 3.2 Å and FTIR spectra support the long-held position that photolysis of Fea3–CO in cytochrome c oxidases leads to significant trapping of the CO on the CuB atom; Fea3 and CuB ligation, at the resolutions reported here, are otherwise unaltered. This article is part of a Special Issue entitled: Respiratory Oxidases.  相似文献   

16.
Endothelium-derived nitric oxide (NO) is critical in maintaining vascular tone. Accumulating evidence shows that NO bioavailability is regulated by oxygen concentration. However, it is unclear to what extent the oxygen concentration regulates NO bioavailability in the vascular wall. In this study, a recently developed experimental setup was used to measure the NO diffusion flux across the aortic wall at various oxygen concentrations. It was observed that for a constant NO concentration at the endothelial surface, the measured NO diffusion flux out of the adventitial surface at [O2] = 0 μM is around fivefold greater than at [O2] = 150 μM, indicating that NO is consumed in the aortic wall in an oxygen-dependent manner. Analysis of experimental data shows that the rate of NO consumption in the aortic wall is first order with respect to [NO] and first order with respect to [O2], and the rate constant k1 was determined as (4.0 ± 0.3) × 103 M?1 s?1. Computer simulations demonstrate that NO concentration distribution significantly changes with oxygen concentration and the effective NO diffusion distance at low oxygen level ([O2] ≤ 25 μM) is significantly longer than that at high oxygen level ([O2] = 200 μM). These results suggest that oxygen-dependent NO consumption may play an important role in dilating blood vessels during hypoxia by increasing the effective NO diffusion distance.  相似文献   

17.
AimTo investigate the mechanism through which the extracellular alkalinization promotes relaxation in rat thoracic aorta.MethodsThe relaxation response to NaOH-induced extracellular alkalinization (7.4–8.5) was measured in aortic rings pre-contracted with phenylephrine (Phe, 10?6 M). The vascular reactivity experiments were performed in endothelium-intact and -denuded rings, in the presence or and absence of indomethacin (10?5 M), NG-nitro-l-arginine methyl ester (L-NAME, 10?4 M), N-(6-Aminohexyl)-5-chloro-1-naphthalenesulfonamide/HCl (W-7, 10?7 M), 2,5-dimethylbenzimidazole (DMB, 2 × 10?5 M) and methyl-β-cyclodextrin (10?2 M). In addition, the effects of NaOH-induced extracellular alkalinization (pH 8.0 and 8.5) on the intracellular nitric oxide (NO) concentration was evaluated in isolated endothelial cells loaded with diaminofluorescein-FM diacetate (DAF-FM DA, 5 μM), in the presence and absence of DMB (2 × 10?5 M).ResultsThe extracellular alkalinization failed to induce any change in vascular tone in aortic rings pre-contracted with KCl. In rings pre-contracted with Phe, the extracellular alkalinization caused relaxation in the endothelium-intact rings only, and this relaxation was maintained after cyclooxygenase inhibition; completely abolished by the inhibition of nitric oxide synthase (NOS), Ca2+/calmodulin and Na+/Ca2+ exchanger (NCX), and partially blunted by the caveolae disassembly.ConclusionsThese results suggest that, in rat thoracic aorta, that extracellular alkalinization with NaOH activates the NCX reverse mode of endothelial cells in rat thoracic aorta, thereby the intracellular Ca2+ concentration and activating the Ca2+/calmodulin-dependent NOS. In turn, NO is released promoting relaxation.  相似文献   

18.
Cai Z P  Huang W W  An M  Duan S S 《农业工程》2009,29(5):297-301
Effects of irradiance and iron on the growth of a typical harmful algal blooms (HABs) causative dinoflagellate, Scrippsiella trochoidea, were investigated under various irradiances (high light: 70 μmol m?2 s?1 and low light: 4 μmol m?2 s?1) and iron concentrations (low iron: 0.063 mg L?1, medium iron: 0.63 mg L?1 and high iron: 6.3 mg L?1), and evaluated by the parameters of algal cell density, specific growth rate, optical density and chlorophyll a content. The results indicated that there was significant difference in the cell density of dinoflagellate S. trochoidea between high light and low light intensity treatments across the entire experiments, 7-fold higher at high irradiance as compared with low irradiance, which was further enhanced by the iron concentration. It was found that the maximum cell density of 25 × 104 cell mL?1 occurred under the combination of high light intensity and high iron concentration, followed by 23 × 104 cell mL?1 in the combination of high light and medium iron, and 20 × 104 cell mL?1 in the combination of high light and low iron. There was no significant effect of iron concentration on the cell density under low light intensity. The cell density maintained about 3 × 104 cell mL?1 across all combinations of iron concentrations and low light in the end of experiments. Such interactive effects of light intensity and iron level dependent were also observed for the specific growth rate, OD680 and chlorophyll a content of S. trochoidea. The maximum values of specific growth rate, OD680 and chlorophyll a content peaked at the condition of high irradiance and high iron, which were 0.22 d?1, 0.282 and 0.673 mg L?1, respectively. In general, their values increased significantly with the increasing of iron concentration at high irradiance, whereas no significant difference was observed among three iron concentrations at low irradiance, all remaining approximately 0.06 d?1, 0.03 and 0.050 mg L?1, respectively. Those results suggest that there may be a strong interactive effect between irradiance and iron on microalgal growth and their physiological characteristics. The combination of high light and high iron concentration may accelerate algal cell growth and pigment biosynthesis, thus leading to massive occurrence of HABs.  相似文献   

19.
The rapid (2 min) nongenomic effects of aldosterone (ALDO) and/or spironolactone (MR antagonist), RU 486 (GR antagonist), atrial natriuretic peptide (ANP) and dimethyl-BAPTA (BAPTA) on the intracellular pH recovery rate (pHirr) via NHE1 (basolateral Na+/H+ exchanger isoform), after the acid load induced by NH4Cl, and on the cytosolic free calcium concentration ([Ca2+]i) were investigated in the proximal S3 segment isolated from rats, by the probes BCECF-AM and FLUO-4-AM, respectively. The basal pHi was 7.15 ± 0.008 and the basal pHirr was 0.195 ± 0.012 pH units/min (number of tubules/number of tubular areas = 16/96). Our results confirmed the rapid biphasic effect of ALDO on NHE1: ALDO (10?12 M) increases the pHirr to approximately 59% of control value, and ALDO (10?6 M) decreases it to approximately 49%. Spironolactone did not change these effects, but RU 486 inhibited the stimulatory effect and maintained the inhibitory effect. ANP (10?6 M) or BAPTA (5 × 10?5 M) alone had no significant effect on NHE1 but prevented both effects of ALDO on this exchanger. The basal [Ca2+]i was 104 ± 3 nM (15), and ALDO (10?12 or 10?6 M) increased the basal [Ca2+]i to approximately 50% or 124%, respectively. RU 486, ANP and BAPTA decreased the [Ca2+]i and inhibited the stimulatory effect of both doses of ALDO. The results suggest the involvement of GR on the nongenomic effects of ALDO and indicate a pHirr-regulating role for [Ca2+]i that is mediated by NHE1, stimulated/impaired by ALDO, and affected by ANP or BAPTA with ALDO. The observed nongenomic hormonal interaction in the S3 segment may represent a rapid and physiologically relevant regulatory mechanism in the intact animal under conditions of volume alterations.  相似文献   

20.
Storage xyloglucans from the seeds of Copaifera langsdorffii, Hymenaea courbaril and Tamarindus indica were obtained by aqueous extraction from the milled and defatted cotyledons, XGC, XGJ and XGT, respectively. The resulting fractions showed similar monosaccharide composition with Glc:Xyl:Gal molar ratios of 2.4:1.5:1.0, 3.8:1.5:1,0 and 3.6:2.4:1.0 for XGC, XGJ and XGT, respectively. High-performance size-exclusion chromatography of the polysaccharides showed unimodal profiles, and the average molar mass (Mw) was obtained for XGC (9.6 × 105 g/mol), XGJ (9.1 × 105 g/mol) and XGT (7.3 × 105 g/mol). The immunomodulatory effects of the xyloglucans on peritoneal macrophages were evaluated. Phagocytic activity was observed in macrophages treated with XGT. The effect of XGT was tested on the production of O2? and NO. At 25 μg/ml XGT caused a 100% increase in NO production when compared to the control group; however, it did not affect O2? production in the absence of PMA. The production of TNF-α, interleukins 1β and 6 by macrophages in the presence of the xyloglucans was evaluated. The polysaccharides affected the production of the cytokines by macrophages to different degrees. XGC caused an enhancement of IL-1β and TNF-α production, compared to the other xyloglucans. For IL-6 production, XGT gave greater stimulation than XGC and XGJ, reaching 87% at 50 μg/ml. XGJ promoted a statistically significant effect on all cytokine productions tested. The results indicate that the xyloglucans from C. langsdorffii, H. courbaril and T. indica can be classified as biological response modifiers (BRM).  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