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 38 毫秒
1.
Can elevated CO(2) improve salt tolerance in olive trees?   总被引:2,自引:0,他引:2  
We compared growth, leaf gas exchange characteristics, water relations, chlorophyll fluorescence, and Na+ and Cl concentration of two cultivars (‘Koroneiki’ and ‘Picual’) of olive (Olea europaea L.) trees in response to high salinity (NaCl 100 mM) and elevated CO2 (eCO2) concentration (700 μL L−1). The cultivar ‘Koroneiki’ is considered to be more salt sensitive than the relatively salt-tolerant ‘Picual’. After 3 months of treatment, the 9-month-old cuttings of ‘Koroneiki’ had significantly greater shoot growth, and net CO2 assimilation (ACO2) at eCO2 than at ambient CO2, but this difference disappeared under salt stress. Growth and ACO2 of ‘Picual’ did not respond to eCO2 regardless of salinity treatment. Stomatal conductance (gs) and leaf transpiration were decreased at eCO2 such that leaf water use efficiency (WUE) increased in both cultivars regardless of saline treatment. Salt stress increased leaf Na+ and Cl concentration, reduced growth and leaf osmotic potential, but increased leaf turgor compared with non-salinized control plants of both cultivars. Salinity decreased ACO2, gs, and WUE, but internal CO2 concentrations in the mesophyll were not affected. eCO2 increased the sensitivity of PSII and chlorophyll concentration to salinity. eCO2 did not affect leaf or root Na+ or Cl concentrations in salt-tolerant ‘Picual’, but eCO2 decreased leaf and root Na+ concentration and root Cl concentration in the more salt-sensitive ‘Koroneiki’. Na+ and Cl accumulation was associated with the lower water use in ‘Koroneiki’ but not in ‘Picual’. Although eCO2 increased WUE in salinized leaves and decreased salt ion uptake in the relatively salt-tolerant ‘Koroneiki’, growth of these young olive trees was not affected by eCO2.  相似文献   

2.
A novel β-glucosidase from Fusarium proliferatum ECU2042 (FPG) was successfully purified to homogeneity with a 506-fold increase in specific activity. The molecular mass of the native purified enzyme (FPG) was estimated to be approximately 78.7 kDa, with two homogeneous subunits of 39.1 kDa, and the pI of this enzyme was 4.4, as measured by two-dimensional electrophoresis. The optimal activities of FPG occurred at pH 5.0 and 50 °C, respectively. The enzyme was stable at pH 4.0–6.5 and temperatures below 60 °C, and the deactivation energy (Ed) for FPG was 88.6 kJ mo1−1. Moreover, it was interesting to find that although the purified enzyme exhibited a very low activity towards p-nitrophenyl β-d-glucoside (pNPG), and almost no activity towards cellobiose, a relatively high activity was observed on ginsenoside Rg3. The enzyme hydrolyzed the 3-C, β-(1 → 2)-glucoside of ginsenoside Rg3 to produce ginsenoside Rh2, but did not sequentially hydrolyze the β-d-glucosidic bond of Rh2. The Km and Vmax values of FPG for ginsenoside Rg3 were 2.37 mM and 0.568 μmol (h mg protein)−1, respectively. In addition, this enzyme also exhibited significant activities towards various alkyl glucosides, aryl glucosides and several natural glycosides.  相似文献   

3.
α1-Antitrypsin, α2-macroglobulin and low-molecular weight kininogen were isolated from human serum and kallikreins from human urine and saliva.α1-Antitrypsin and α2-macroglobulin inhibited the activity of trypsin in releasing kinin from low-molecular weight kininogen, due to their binding with the enzyme, but did non inhibit or bind with urinary and salivary kallikreins.  相似文献   

4.
β-Amyloid peptide (Aβ) 1–42, involved in the pathogenesis of Alzheimer’s disease, binds copper ions to form Aβ · Cun complexes that are able to generate H2O2 in the presence of a reductant and O2. The production of H2O2 can be stopped with chelators. More reactive than H2O2 itself, hydroxyl radicals HO (generated when a reduced redox active metal complex interacts with H2O2) are also probably involved in the oxidative stress that creates brain damage during the disease. We report in the present work a method to monitor the effect of chelating agents on the production of hydrogen peroxide by metallo-amyloid peptides. The addition of H2O2 associated to a pre-incubation step between ascorbate and Aβ · Cun allows to study the formation of H2O2 but also, at the same time, its transformation by the copper complexes. Aβ · Cun peptides produce but do not efficiently degrade H2O2. The reported analytic method, associated to precipitation experiments of copper-containing amyloid peptides, allows to study the inhibition of H2O2 production by chelators. The action of a ligand such as EDTA is probably due to the removal of the copper ions from Aβ · Cun, whereas bidentate ligands such as 8-hydroxyquinolines probably act via the formation of ternary complexes with Aβ · Cun. The redox activity of these bidentate ligands can be modulated by the incorporation or the modification of substituents on the quinoline heterocycle.  相似文献   

5.
The synthesis and pharmacology of 15 1-deoxy-Δ8-THC analogues, several of which have high affinity for the CB2 receptor, are described. The deoxy cannabinoids include 1-deoxy-11-hydroxy-Δ8-THC (5), 1-deoxy-Δ8-THC (6), 1-deoxy-3-butyl-Δ8-THC (7), 1-deoxy-3-hexyl-Δ8-THC (8) and a series of 3-(1′,1′-dimethylalkyl)-1-deoxy-Δ8-THC analogues (2, n=0–4, 6, 7, where n=the number of carbon atoms in the side chain−2). Three derivatives (1719) of deoxynabilone (16) were also prepared. The affinities of each compound for the CB1 and CB2 receptors were determined employing previously described procedures. Five of the 3-(1′,1′-dimethylalkyl)-1-deoxy-Δ8-THC analogues (2, n=1–5) have high affinity (Ki=<20 nM) for the CB2 receptor. Four of them (2, n=1–4) also have little affinity for the CB1 receptor (Ki=>295 nM). 3-(1′,1′-Dimethylbutyl)-1-deoxy-Δ8-THC (2, n=2) has very high affinity for the CB2 receptor (Ki=3.4±1.0 nM) and little affinity for the CB1 receptor (Ki=677±132 nM).
Scheme 3. (a) (C6H5)3PCH3+ Br, n-BuLi/THF, 65°C; (b) LiAlH4/THF, 25°C; (c) KBH(sec-Bu)3/THF, −78 to 25°C then H2O2/NaOH.  相似文献   

6.
Potato plants (Solanum tuberosum L. cv. Bintje) were grown to maturity in open-top chambers under three carbon dioxide (CO2; ambient and 24 h d−1 seasonal mean concentrations of 550 and 680 μmol mol−1) and two ozone levels (O3; ambient and an 8 h d−1 seasonal mean of 50 nmol mol−1). Chlorophyll content, photosynthetic characteristics, and stomatal responses were determined to test the hypothesis that elevated atmospheric CO2 may alleviate the damaging influence of O3 by reducing uptake by the leaves. Elevated O3 had no detectable effect on photosynthetic characteristics, leaf conductance, or chlorophyll content, but did reduce SPAD values for leaf 15, the youngest leaf examined. Elevated CO2 also reduced SPAD values for leaf 15, but not for older leaves; destructive analysis confirmed that chlorophyll content was decreased. Leaf conductance was generally reduced by elevated CO2, and declined with time in the youngest leaves examined, as did assimilation rate (A). A generally increased under elevated CO2, particularly in the older leaves during the latter stages of the season, thereby increasing instantaneous transpiration efficiency. Exposure to elevated CO2 and/or O3 had no detectable effect on dark-adapted fluorescence, although the values decreased with time. Analysis of the relationships between assimilation rate and intercellular CO2 concentration and photosynthetically active photon flux density showed there was initially little treatment effect on CO2-saturated assimilation rates for leaf 15. However, the values for plants grown under 550 μmol mol−1 CO2 were subsequently greater than in the ambient and 680 μmol mol−1 treatments, although the beneficial influence of the former treatment declined sharply towards the end of the season. Light-saturated assimilation was consistently greater under elevated CO2, but decreased with time in all treatments. The values decreased sharply when leaves grown under elevated CO2 were measured under ambient CO2, but increased when leaves grown under ambient CO2 were examined under elevated CO2. The results obtained indicate that, although elevated CO2 initially increased assimilation and growth, these beneficial effects were not necessarily sustained to maturity as a result of photosynthetic acclimation and the induction of earlier senescence.  相似文献   

7.
Receptors for α2-macroglobulin-proteinase complexes have been characterized in rat and human liver membranes. The affinity for binding of 125I-labelled α2-macroglobulin · trypsin to rat liver membranes was markedly pH-dependent in the physiological range with maximum binding at pH 7.8–9.0. The half-time for association was about 5 min at 37°C in contrast to about 5 h at 4°C. The half-saturation constant was about 100 pM at 4°C and 1 nM at 37°C (pH 7.8). The binding capacity was approx. 300 pmol per g protein for rat liver membranes and about 100 pmol per g for human membranes. Radiation inactivation studies showed a target size of 466 ± 71 kDa (S.D., n = 7) for α2-macroglobulin · trypsin binding activity. Affinity cross-linking to rat and human membranes of 125I-labelled rat α1-inhibitor-3 · chymotrypsin, a 210 kDa analogue which binds to the α2-macroglobulin receptors in hepatocytes (Gliemann, J. and Sottrup-Jensen, L. (1987) FEBS Lett. 221, 55–60), followed by SDS-polyacrylamide gel electrophoresis, revealed radioactivity in a band not distinguishable from that of cross-linked α2-macroglobulin (720 kDa). This radioactivity was absent when membranes with bound 125I-α1-inhibitor-3 complex were treated with EDTA before cross-linking and when incubation and cross-linking were carried out in the presence of a saturating concentration of unlabelled complex. The saturable binding activity was maintained when membranes were solubilized in the detergent 3-[(3-cholamidopropyl)dimethylammonio]profane sulfonate (CHAPS) and the size of the receptor as estimated by cross-linking experiments was shown to be similar to that determined in the membranes. It is concluded that liver membranes contain high concentrations of an approx. 400–500 kDa α2-macroglobulin receptor soluble in CHAPS. The soluble preparation should provide a suitable material for purification and further characterization of the receptor.  相似文献   

8.
Plants were regenerated from root explants of Arabidopsis halleri (L.) O’Kane and Al-Shehbaz via a three-step procedure callus induction, induction of somatic embryos and shoot development. Callus was induced from root segments, leaflets and petiole segments after incubation for 2 weeks in Murashige and Skoog medium (MS) supplemented with 0.5 mg/l−1 (2.26 μM) 2,4-D (2,4-dichlorophenoxyacetic acid) and 0.05 mg/l−1 (0.23 μM) kinetin. Only calli developed from root segments continued to grow when transferred to a regeneration medium containing 2.0 mg/l−1 (9.8 μM) 6-γ-γ-(dimethylallylamino)-purine (2ip) and 0.05 mg/l−1 (2.68 μM) α-naphthalenacetic acid (NAA) and eventually 40 of them developed embryogenic structures. On the same medium 38 of these calli regenerated shoots. Rooting was achieved for 50 of the shoots subcultured in MS medium without hormones. The regeneration ability of callus derived from root cuttings, observed in this study, makes this technique useful for genetic transformation experiments and in vitro culture studies.  相似文献   

9.
The kinetics of substitution reactions of [η-CpFe(CO)3]PF6 with PPh3 in the presence of R-PyOs have been studied. For all the R-PyOs (R = 4-OMe, 4-Me, 3,4-(CH)4, 4-Ph, 3-Me, 2,3-(CH)4, 2,6-Me2, 2-Me), the reactions yeild the same product [η5-CpFe(CO)2PPh3]PF6, according to a second-order rate law that is first order in concentrations of [η5-CpFe(CO)3]PF6 and of R-PyO but zero order in PPh3 concentration. These results, along with the dependence of the reaction rate on the nature of R-PyO, are consistent with an associative mechanism. Activation parameters further support the bimmolecular nature of the reactions: ΔH = 13.4 ± 0.4 kcal mol−1, ΔS = −19.1 ± 1.3 cal k−1 mol−1 for 4-PhPyO; ΔH = 12.3 ± 0.3 kcal mol−1, ΔS = 24.7 ±1.0 cal K−1 mol−1 for 2-MePyO. For the various substituted pyridine N-oxides studied in this paper, the rates of reaction increase with the increasing electron-donating abilities of the substituents on the pyridine ring or N-oxide basicities, but decrease with increasing 17O chemical shifts of the N-oxides. Electronic and steric factors contributing to the reactivity of pyridine N-oxides have been quantitatively assessed.  相似文献   

10.
One and a half year-old Ginkgo saplings were grown for 2 years in 7 litre pots with medium fertile soil at ambient air CO2 concentration and at 700 μmol mol−1 CO2 in temperature and humidity-controlled cabinets standing in the field. In the middle of the 2nd season of CO2 enrichment, CO2 exchange and transpiration in response to CO2 concentration was measured with a mini-cuvette system. In addition, the same measurements were conducted in the crown of one 60-year-old tree in the field. Number of leaves/tree was enhanced by elevated CO2 and specific leaf area decreased significantly.CO2 compensation points were reached at 75–84 μmol mol−1 CO2. Gas exchange of Ginkgo saplings reacted more intensively upon CO2 than those of the adult Ginkgo. On an average, stomatal conductance decreased by 30% as CO2 concentration increased from 30 to 1000 μmol mol−1 CO2. Water use efficiency of net photosynthesis was positively correlated with CO2 concentration levels. Saturation of net photosynthesis and lowest level of stomatal conductance was reached by the leaves of Ginkgo saplings at >1000 μmol mol−1 CO2. Acclimation of leaf net CO2 assimilation to the elevated CO2 concentration at growth occurred after 2 years of exposure. Maximum of net CO2 assimilation was 56% higher at ambient air CO2 concentration than at 700 μmol mol−1 CO2.  相似文献   

11.
An N-acetyl-β-d-hexosaminidase has been purified from primary wheat leaves (Triticum aestivum L.) by freeze-thawing, (NH4)2SO4 precipitation, methanol precipitation, gel filtration, cation exchange chromatography and affinity chromatography on concanavalin A-Sepharose. The activity of the purified preparations could be stabilised by addition of Triton X-100 and the enzyme was stored at -20°C without significant loss of activity. The enzyme hydrolysed pNP-β-d-GlcNAc (optimum pH 5.2, Km 0.29 mM, Vmax 2.56 μkat mg−1) and pNP-β-d-GalNAc (optimum pH 4.4, Km 0.27 mM, Vmax 2.50 μkat mg−1). Five major isozymes were identified, with isoelectric points in the range 5.13–5.36. All five isozymes possessed both N-acety-β-d-glucosaminidase and N-acetyl-β-d-galactosaminidase activity. Inhibition studies and mixed substrate analysis suggested that both substrates are catalysed by the same active site. Both activities were inhibited by GlcNAc, 2-acetamido-2-deoxygluconolactone, GalNAc and the ions of mercury, silver and copper. The Kis for inhibition of N-acetyl-β-d-glucosaminidase activity were: GlcNAc (15.3 mM) and GalNAc (3.4mM). For inhibition of N-acety-β-d-galactosaminidase activity the corresponding values were: GlcNAc (18.2 mM) and GalNac (2.5 mM). The enzyme was considerably less active at releasing pNP from pNP-β-d-(GlcNAc)2 and pNP-β-d-(GlcNAc)3 than from pNP-β-d-GlcNAc. The ability of the N-acetyl-β-d-hexosaminidase to relase GlcNAc from chitin oligomers (GlcNAc)2 (optimum pH 5.0) and (GlcNAc)3−6 (optimum pH 4.4) was also low. Analysis of the reaction products revealed that the initial products from the hydrolysis of (GlcNAc)n were predominantly (GlcNAc)n−1 and GlcNAc.  相似文献   

12.
Yeast cytochrome c peroxidase (CCP) efficiently catalyzes the reduction of H2O2 to H2O by ferrocytochrome c in vitro. The physiological function of CCP, a heme peroxidase that is targeted to the mitochondrial intermembrane space of Saccharomyces cerevisiae, is not known. CCP1-null-mutant cells in the W303-1B genetic background (ccp1Δ) grew as well as wild-type cells with glucose, ethanol, glycerol or lactate as carbon sources but with a shorter initial doubling time. Monitoring growth over 10 days demonstrated that CCP1 does not enhance mitochondrial function in unstressed cells. No role for CCP1 was apparent in cells exposed to heat stress under aerobic or anaerobic conditions. However, the detoxification function of CCP protected respiring mitochondria when cells were challenged with H2O2. Transformation of ccp1Δ with ccp1W191F, which encodes the CCPW191F mutant enzyme lacking CCP activity, significantly increased the sensitivity to H2O2 of exponential-phase fermenting cells. In contrast, stationary-phase (7-day) ccp1Δ-ccp1W191F exhibited wild-type tolerance to H2O2, which exceeded that of ccp1Δ. Challenge with H2O2 caused increased CCP, superoxide dismutase and catalase antioxidant enzyme activities (but not glutathione reductase activity) in exponentially growing cells and decreased antioxidant activities in stationary-phase cells. Although unstressed stationary-phase ccp1Δ exhibited the highest catalase and glutathione reductase activities, a greater loss of these antioxidant activities was observed on H2O2 exposure in ccp1Δ than in ccp1Δ-ccp1W191F and wild-type cells. The phenotypic differences reported here between the ccp1Δ and ccp1Δ-ccp1W191F strains lacking CCP activity provide strong evidence that CCP has separate antioxidant and signaling functions in yeast.  相似文献   

13.
The aim of this work was to investigate whether Fe reduction and antioxidant mechanisms were expressed differently in five Prunus rootstocks (‘Peach seedling,’ ‘Barrier,’ ‘Cadaman,’ ‘Saint Julien 655/2’ and ‘GF-677’). These rootstocks differ in their tolerance to Fe deficiency when grown in the absence of Fe (−Fe) or in presence of bicarbonate (supplied as 5 or 10 mM NaHCO3). Fe deficiency conditions, especially bicarbonate, were shown to decrease Fe and total chlorophyll (CHL) concentration. In the (−Fe)-treated roots of all rootstocks and in the 5 mM NaHCO3-treated ones of the tolerant ‘GF-677’ the Fe(III)-chelate reductase (FCR) activity was stimulated. On the contrary, apart from the ‘GF-677,’ FCR activity was greatly inhibited by the 10 mM NaHCO3. From the results obtained with decapitated rootstocks, it is not entirely clear whether or not the presence of shoot apex was a prerequisite to induce FCR function in all rootstocks tested. In the leaves of rootstocks exposed to the (−Fe) treatment, superoxide dismutase (SOD), peroxidase (POD) and catalase (CAT) activities were enhanced whereas the levels of the non-enzymatic antioxidants (FRAP values) were increased in the Fe-deprived leaves, irrespective of the rootstock. Except for ‘Peach seedling,’ foliar SOD activity was stimulated by the presence of NaHCO3. Furthermore, POD activity was increased in the ‘Saint Julien 655/2’ and ‘GF-677,’ but was depressed in the ‘Barrier’ rootstocks exposed to 10 mM NaHCO3. As a result of 10 mM NaHCO3, the expression of a Cu/Zn-SOD and a POD isoform was diminished in the leaves of ‘Peach seedling’ and ‘Barrier,’ respectively. By contrast, an additional isoform of both POD and Mn–SOD were expressed in the leaves of ‘GF-677’ exposed to 10 mM NaHCO3 suggesting that the tolerance of rootstocks to Fe deficiency is associated with induction of an antioxidant defense mechanism. Although CAT activity was increased in the 5 mM NaHCO3-treated leaves of ‘GF-677,’ specifically the 10 mM NaHCO3 treatment resulted in a decrease of CAT activity and an accumulation of H2O2, indicating that bicarbonate-induced Fe deficiency may cause more severe oxidative stress in the rootstocks, than the absence of Fe. A general link between Fe deficiency-induced oxidative stress and Fe reduction-sensing mechanism is also discussed.  相似文献   

14.
The white-throated Dipper (Cinclus cinclus) is unique among passerine birds by its reliance on diving to achieve energy gain in fast-flowing waters. Consequently, it should have evolved behavioural adaptations allowing responding directly to runoff patterns (one of the assumptions of the Natural Flow Regime Paradigm—NRFP). In this study (October 1998–August 2001), we investigated how behavioural and energy use strategies in Dippers might vary under the natural flow regime of snowmelt-dominated streams in The Pyrénées (France) where natural flow regime is highly seasonal and predictable. We recorded time spent in each of 5 behavioural activities of ringed birds to estimate time–activity budgets and derive time–energy budgets enabling the modelling of daily energy expenditure (DEE). Annual pattern in ‘foraging’ and ‘resting’ matched perfectly the annual pattern of the natural regime flow and there was a subtle relationship between water stage and time spent ‘diving’ the later increasing with rising discharge up to a point where it fell back. Thus, time–activity budgets meet the main prediction of the NRFP. For males and females Dippers, estimates of feeding rates (ratio Eobs/Ereq = observed rate of energy gain / required foraging rate) and energy stress (M = DEE / Basal Metabolic Rate) also partly matched the NFRP. Maximum value for the ratio Eobs/Ereq was registered in May whilst M peaked in spring. These ratios indicated that Pyrenean Dippers could face high energy stress during winter but paradoxically none during high snowmelt spates when food is expected to be difficult to obtain in the channel and when individual birds were observed spending ca 75% of the day ‘resting’. Annual pattern in DEE did not match the NFRP ; two phases were clearly identified, the first between January to June (with oscillating values 240–280 kJ d− 1 ind− 1) and the second between July and December (200–220 kJ d− 1 ind− 1). As total energy expenditure was higher during the most constraining season or life cycle, we suggest that energy management by Dippers in Pyrenean mountain streams may fit the ‘peak total demand’ hypothesis. At this step of the study, it is not possible to tell whether Dippers use an ‘energy-minimisation’ or an ‘energy-maximisation' strategy.  相似文献   

15.
Lysophosphatidic acid (LPA) and its ether analog alkyl-glycerophosphate (AGP) elicit arterial wall remodeling when applied intralumenally into the uninjured carotid artery. LPA is the ligand of eight GPCRs and the peroxisome proliferator-activated receptor γ (PPARγ). We pursued a gene knockout strategy to identify the LPA receptor subtypes necessary for the neointimal response in a non-injury model of carotid remodeling and also compared the effects of AGP and the PPARγ agonist rosiglitazone (ROSI) on balloon injury-elicited neointima development. In the balloon injury model AGP significantly increased neointima; however, rosiglitazone application attenuated it. AGP and ROSI were also applied intralumenally for 1 h without injury into the carotid arteries of LPA1, LPA2, LPA1&2 double knockout, and Mx1Cre-inducible conditional PPARγ knockout mice targeted to vascular smooth muscle cells, macrophages, and endothelial cells. The neointima was quantified and also stained for CD31, CD68, CD11b, and α-smooth muscle actin markers. In LPA1, LPA2, LPA1&2 GPCR knockout, Mx1Cre transgenic, PPARγfl/−, and uninduced Mx1Cre × PPARγfl/− mice AGP- and ROSI-elicited neointima was indistinguishable in its progression and cytological features from that of WT C57BL/6 mice. In PPARγ−/− knockout mice, generated by activation of Mx1Cre-mediated recombination, AGP and ROSI failed to elicit neointima and vascular wall remodeling. Our findings point to a difference in the effects of AGP and ROSI between the balloon injury- and the non-injury chemically-induced neointima. The present data provide genetic evidence for the requirement of PPARγ in AGP- and ROSI-elicited neointimal thickening in the non-injury model and reveal that the overwhelming majority of the cells in the neointimal layer express α-smooth muscle actin.  相似文献   

16.
In this paper the utilization of the cyanobacteria Anabaena sp. in carbon dioxide removal processes is evaluated. For this, continuous cultures of this strain were performed at different dilution rates; alternatives for the recovery of the organic matter produced being also studied. A maximum CO2 fixation rate of 1.45 g CO2 L−1 day−1 was measured experimentally, but it can be increased up to 3.0 g CO2 L−1 day−1 outdoors. The CO2 is mainly transformed into exopolysaccharides, biomass representing one third of the total organic matter produced. Organic matter can be recovered by sedimentation with efficiencies higher than 90%, the velocity of sedimentation being 2 · 10−4 s−1. The major compounds were carbohydrates and proteins with productivities of 0.70 and 0.12 g L−1 day−1, respectively. The behaviour of the cultures of Anabaena sp. has been modelized, also the characteristics parameters requested to design separation units being reported. Finally, to valorizate the organic matter as biofertilizers and biofuels is proposed.  相似文献   

17.
Epinephrine (EPI) is thought to act by stimulating adenylyl cyclase (ACase) and cAMP production through β-adrenoceptors in the liver of more primitive vertebrates. Recent observations, however, point to an involvement of α1-adrenoceptors in EPI action, at least in some fish species. The role of the α1- and β-adrenergic transduction pathways has been investigated in rainbow trout (Oncorhynchus mykiss) hepatic tissue. Radioligand-binding assays with the β-adrenergic antagonist 3H-CGP-12177 using hepatic membranes purified on a discontinuous sucrose gradient confirmed the presence of β-adrenoceptors (Kd0.36 nM, Bmax 8.61 fmol · mg−1 protein). We provide the first demonstration of α1-adrenoceptors in these same membranes; analysis of binding data with the α1-adrenergic antagonist 3H-prazosin demonstrated a single class of binding sites with a Kdof 15.4 nM and a Bmax of 75.2 fmol · mg−1 protein. There is a straight correlation between β-adrenoceptor occupancy, ACase activation and cAMP production. On the contrary, the role of inositol 1,4,5-trisphosphate (IP3) has to be elucidated; in fact, despite the presence of specific microsomal binding sites for IP3 (Kd 6.03 nM, Bmax 90.2 fmol · mg−1 protein), its cytosolic concentration was not modulated by EPI. On the other hand, we have previously shown in American eel and bullhead hepatocytes that α1-adrenergic agonists are able to increase intracellular concentrations of IP3 and Ca2+ and to activate glycogenolysis. These data suggest a marked variation in the liver of different fish both in terms of α1-binding sites affinity and of α1-adrenoceptor/IP3/Ca2+ transduction systems.  相似文献   

18.
The immunologic cross-reactivity of the α and α+ forms of the large subunit and the β subunit of the (Na+ + K+)-ATPase from brain and kidney preparations was examined using rabbit antiserum prepared against the purified holo lamb kidney enzyme. As previously reported by Sweadner ((1979) J. Biol. Chem. 254, 6060–6067) phosphorylation of the large subunit of the (Na+ + K+)-ATPase in the presence of Na+, Mg2+, and [γ-32P]ATP revealed that dog and, very likely, rat brain contain two forms of the large subunit (designated α and α+) while dog, rat, and lamb kidney contain only one form (α). The cross-reactivity of the α and α+ forms in these preparations was investigated by resolving the subunits by SDS-polyacrylamide gel electrophoresis. The separated polypeptides were transferred to unmodified nitrocellulose paper, and reacted with rabbit anti-lamb kidney serum, followed by detection of the antigen-antibody complex with 125I-labeled protein A and autoradiography. By this method, the α and α+ forms of rat and dog brain, as well as the α form found in kidney, were shown to cross-react. In addition, membranes from human cerebral cortex were shown to contain two immunoreactive bands corresponding to the α and α+ forms of dog brain. In contrast, the brain of the insect Manduca sexta contains only one immunoreactive polypeptide with a molecular weight intermediate to the α and α+ forms of dog brain. The β subunit from lamb, dog and rat kidney and from dog and rat brain cross-reacts with anti-lamb kidney (Na+ + K+)-ATPase serum. The mobility of the β subunit from dog and rat brain on SDS-polyacrylamide electrophoresis gels is greater than the mobility of the β subunit from lamb, rat or dog kidney.  相似文献   

19.
A novel fish muscle serine protease named muscle soluble serine protease (MSSP) was purified from the soluble fraction of lizard fish (Saurida undosquamis: Synodontidae) muscle by ammonium sulfate fractionation followed by four steps of column chromatographies. In native-PAGE, the purified enzyme appeared as a single band with an estimated mol. mass of approximately 380 kDa by gel filtration. In SDS-PAGE under reducing conditions, the purified enzyme migrated as two protein bands at 110 and 100 kDa, named subunits A and B, respectively. The 20 residues of N-terminal amino acid sequence of subunit B showed 70% of homology to β-chain of carp α2-macroglobulin-1. Moreover, both subunits A and B showed immunoreactivity with anti carp α2-macroglobulin antibody. Purified MSSP was inactivated by Pefabloc SC, aprotinin, benzamidine and TLCK, but not by α1-antitrypsin. After acid treatment (pH 2, 24 h), however, the enzyme activity eluted at 14 kDa from Sephacryl S-200 carried out under acidic conditions was inhibited by α1-antitrypsin. Lizard fish MSSP most rapidly hydrolyzed Boc-Val-Pro-Arg-MCA and Boc-Gln-Arg-Arg-MCA, but did not hydrolyzed Suc-Leu-Leu-Val-Tyr-MCA and Suc-Ala-Ala-Pro-Phe-MCA, and was not suppressed either by E-64, pepstatin A and ethylenediaminetetraacetic acid (EDTA). These results indicate that the purified MSSP is a serine protease complexed with α2-macroglobulin, and the entrapped protease was dissociated by the acid treatment. Purified and free MSSPs were most active at pH 10.0 and 9.0, respectively. Purified MSSP degraded myofibrillar proteins and casein but time courses of degradation of these substrates by the enzyme differed.  相似文献   

20.
To quantitatively understand intracellular Na+ and Cl homeostasis as well as roles of Na+/K+ pump and cystic fibrosis transmembrane conductance regulator Cl channel (ICFTR) during the β1-adrenergic stimulation in cardiac myocyte, we constructed a computer model of β1-adrenergic signaling and implemented it into an excitation-contraction coupling model of the guinea-pig ventricular cell, which can reproduce membrane excitation, intracellular ion changes (Na+, K+, Ca2+ and Cl), contraction, cell volume, and oxidative phosphorylation. An application of isoproterenol to the model cell resulted in the shortening of action potential duration (APD) after a transient prolongation, the increases in both Ca2+ transient and cell shortening, and the decreases in both Cl concentration and cell volume. These results are consistent with experimental data. Increasing the density of ICFTR shortened APD and augmented the peak amplitudes of the L-type Ca2+ current (ICaL) and the Ca2+ transient during the β1-adrenergic stimulation. This indirect inotropic effect was elucidated by the increase in the driving force of ICaL via a decrease in plateau potential. Our model reproduced the experimental data demonstrating the decrease in intracellular Na+ during the β-adrenergic stimulation at 0 or 0.5 Hz electrical stimulation. The decrease is attributable to the increase in Na+ affinity of Na+/K+ pump by protein kinase A. However it was predicted that Na+ increases at higher beating rate because of larger Na+ influx through forward Na+/Ca2+ exchange. It was demonstrated that dynamic changes in Na+ and Cl fluxes remarkably affect the inotropic action of isoproterenol in the ventricular myocytes.  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