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 343 毫秒
1.
We studied the process of conversion of microcrystalline-cellulose into fermentable glucose in the formic acid reaction system using cross polarization/magic angle spinning 13C-nuclear magnetic resonance, X-ray diffraction and Fourier transform infrared spectroscopy. The results indicated that formic acid as an active agent was able to effectively penetrate into the interior space of the cellulose molecules, thus collapsing the rigid crystalline structure and allowing hydrolysis to occur easily in the amorphous zone as well as in the crystalline zone. The microcrystalline-cellulose was hydrolyzed using formic acid and 4% hydrochloric acid under mild conditions. The effects of hydrochloric acid concentration, the ratio of solid to liquid, temperature (55–75 °C) and retention time (0–9 h), and the concentration of glucose were analyzed. The hydrolysis velocities of microcrystalline-cellulose were 6.14 × 10− 3 h− 1 at 55 °C, 2.94 × 10− 2 h− 1 at 65 °C, and 6.84 × 10− 2 h− 1 at 75 °C. The degradation velocities of glucose were 0.01 h− 1 at 55 °C, 0.14 h− 1 at 65 °C, 0.34 h− 1 at 75 °C. The activation energy of microcrystalline-cellulose hydrolysis was 105.61 kJ/mol, and the activation energy of glucose degradation was 131.37 kJ/mol.  相似文献   

2.
The kinetic properties of a microsomal gill (Na+,K+)-ATPase from the freshwater shrimp, Macrobrachium olfersii, acclimated to 21‰ salinity for 10 days were investigated using the substrate p-nitrophenylphosphate. The enzyme hydrolyzed this substrate obeying cooperative kinetics at a rate of 123.6 ± 4.9 U mg− 1 and K0.5 = 1.31 ± 0.05 mmol L− 1. Stimulation of K+-phosphatase activity by magnesium (Vmax = 125.3 ± 7.5 U mg− 1; K0.5 = 2.09 ± 0.06 mmol L− 1), potassium (Vmax = 134.2 ± 6.7 U mg− 1; K0.5 = 1.33 ± 0.06 mmol L− 1) and ammonium ions (Vmax = 130.1 ± 5.9 U mg− 1; K0.5 = 11.4 ± 0.5 mmol L− 1) was also cooperative. While orthovanadate abolished p-nitrophenylphosphatase activity, ouabain inhibition reached 80% (KI = 304.9 ± 18.3 μmol L− 1). The kinetic parameters estimated differ significantly from those for freshwater-acclimated shrimps, suggesting expression of different isoenzymes during salinity adaptation. Despite the ≈2-fold reduction in K+-phosphatase specific activity, Western blotting analysis revealed similar α-subunit expression in gill tissue from shrimps acclimated to 21‰ salinity or fresh water, although expression of phosphate-hydrolyzing enzymes other than (Na+,K+)-ATPase was stimulated by high salinity acclimation.  相似文献   

3.
The kinetic properties of a microsomal gill (Na+,K+)-ATPase from the blue crab Callinectes danae were analyzed using the substrate p-nitrophenylphosphate. The (Na+,K+)-ATPase hydrolyzed PNPP obeying cooperative kinetics (n=1.5) at a rate of V=125.4±7.5 U mg−1 with K0.5=1.2±0.1 mmol l−1; stimulation by potassium (V=121.0±6.1 U mg−1; K0.5=2.1±0.1 mmol l−1) and magnesium ions (V=125.3±6.3 U mg−1; K0.5=1.0±0.1 mmol l−1) was cooperative. Ammonium ions also stimulated the enzyme through site–site interactions (nH=2.7) to a rate of V=126.1±4.8 U mg−1 with K0.5=13.7±0.5 mmol l−1. However, K+-phosphatase activity was not stimulated further by K+ plus NH4+ ions. Sodium ions (KI=36.7±1.7 mmol l−1), ouabain (KI=830.3±42.5 μmol l−1) and orthovanadate (KI=34.0±1.4 nmol l−1) completely inhibited K+-phosphatase activity. The competitive inhibition by ATP (KI=57.2±2.6 μmol l−1) of PNPPase activity suggests that both substrates are hydrolyzed at the same site on the enzyme. These data reveal that the K+-phosphatase activity corresponds strictly to a (Na+,K+)-ATPase in C. danae gill tissue. This is the first known kinetic characterization of K+-phosphatase activity in the portunid crab C. danae and should provide a useful tool for comparative studies.  相似文献   

4.
Leaf and stem explants of Cleome rosea formed calluses when cultured on MS medium supplemented with different concentrations of 2,4-dichlorophenoxyacetic acid (2,4-D) or 4-amino-3,5,6-trichloropicolinic acid (PIC). The highest biomass accumulation was obtained in the callus cultures initiated from stem explants on medium supplemented with 0.90 μM 2,4-D. Reddish-pink regions were observed on callus surface after 6–7 months in culture and these pigments were identified as anthocyanins. Anthocyanins production was enhanced by reducing temperature and increasing light irradiation. Pigmented calluses transferred to MS1/2 with a 1:4 ratio NH4+/NO3, 70 g L−1 sucrose and supplementation with 0.90 μM 2,4-D maintained a high biomass accumulation and showed an increase of 150% on anthocyanin production as compared with the initial culture conditions. Qualitative analysis of calluses was performed by high performance liquid chromatography coupled to diode array detector and electrospray ionization mass spectrometry (HPLC-DAD/ESIMS). Eleven anthocyanins were characterized and the majority of them were identified as acylated cyanidins, although two peonidins were also detected. The major peak was composed by two anthocyanins, whose proposed identity were cyanidin 3-(p-coumaroyl) diglucoside-5-glucoside and cyanidin 3-(feruloyl) diglucoside-5-glucoside.  相似文献   

5.
Soluble and alginate immobilized urease was utilized for detection and quantitation of mercury in aqueous samples. Urease from the seeds of pumpkin, being a vegetable waste, was extracted and purified to apparent homogeneity (sp. activity 353 U/mg protein; A280/A260 = 1.12) by heat treatment at 48 ± 0.1 °C and gel filtration through Sephadex G-200. Homogeneous enzyme preparation was immobilized in 3.5% alginate leading to 86% immobilization, no leaching of enzyme was found over a period of 15 days at 4 °C. Urease catalyzed urea hydrolysis by soluble and immobilized enzyme revealed a clear dependence on the concentration of Hg2+. Inhibition caused by Hg2+ was non-competitive (Ki = 1.2 × 10−1 μM for soluble and 1.46 × 10−1 μM for alginate immobilized urease.). Time-dependent inhibition both in presence and in absence of Hg2+ ion revealed a biphasic inhibition in activity. For optimization of this process response surface methodology (RSM) was utilized where two-level-two-full factorial (22) central composite design (CCD) has been employed. The regression equation and analysis of variance (ANOVA) were obtained using MINITAB® 15 software. Predicted values thus obtained were closed to experimental value indicating suitability of the model. 3D response surface plot, iso-response contour plot and process optimization curve were helpful to predict the results by performing only limited set of experiments.  相似文献   

6.
Three types of ionic current essentially determine the firing pattern of nerve cells: the persistent Na+ current, the M current and the low-voltage-activated Ca2+ current. The present article summarizes recent experiments concerned with the basic properties of these currents. Keynes and Meves (Proc R Soc Lond B (1993) 253, 61–68) studied the persistent or steady-state Na+ current on dialysed squid axons and measured the probability of channel opening both for the peak and the steady-state Na+ current (PFpeak and PFss) as a function of voltage. Whereas PFpeak starts to rise at −50 mV and reaches a maximum at +40 to +50 mV, PFss only begins to rise appreciably at around 0 mV and is still increasing at +100 mV. This differs from observations on vertebrate excitable tissues where the persistent Na+ current turns on in the threshold region and saturates at around 0 mV. Schmitt and Meves (Pflügers Arch (1993) 425, 134–139) recorded M current, a non-inactivating K+ current, from NG108-15 neuroblastoma × glioma hybrid cells, voltage-clamped in the whole-cell mode, and studied the effects of phorbol 12,13-dibutyrate (PDB), an activator of protein kinase C (PKC), and arachidonic acid (AA). PDB and AA both decreased IM, the effective concentrations being 0.1–1 μM and 5–25 μM, respectively; while the PDB effect was regularly observed, the M current depression by AA was highly variable from cell to cell. The PKC 19–31 peptide, an effective inhibitor of PKC, in a concentration of 1 μM almost totally prevented the effects of PDB and AA on M current, suggesting that both are mediated by PKC. Schmitt and Meves (Pflügers Arch (1994a) 426, Suppl R 59) measured low-voltage-activated (l-v-a) and high-voltage-activated (h-v-a) Ca2+ currents on NG108-15 cells and investigated the effect of AA and PDB on both types of current. At pulse potentials > −20 mV, AA (25–100 μM) decreased l-v-a and h-v-a ICa. The decrease was accompanied by a small negative shift and a slight flattening of the activation and inactivation curves of the l-v-a ICa. The AA effect was not prevented by 50 μM eicosa-5,8,11,14-tetraynoic acid (ETYA), an inhibitor of AA metabolism, or PKC 19–31 peptide and not mimicked by 0.1–1 μM PDB. Probably, AA acts directly on the channel protein or its lipid environment. The physiological relevance of these three sets of observations is briefly discussed.  相似文献   

7.
Exogenous (phorbol ester) and endogenous (diacylglycerol) activators of protein kinase C (PKC) inhibited sodium efflux across the gills of Atlantic cod Gadus morhua and inhibited sodium-plus-potassium-stimulated adenosine triphosphatase (Na+-K+-ATPase) in isolated chloride cells. The branchial sodium efflux measured in a perfused whole-body preparation was inhibited by 47% on administration of 10−6 mol.L−1 phorbol 12, 13-dibutyrate (PDB). The branchial perfusion pressure was increased by 46% by 10−6 mol.L−1 PDB. In contrast the synthetic diacylglycerol, 1-oleoyl-2-acetyl gycerol (OAG) did not alter significantly perfusion pressure but did reduce sodium efflux by 13% at a concentration of 4 × 10−6 mol.L−1. The effects of these agents on Na+-K+-ATPase activity were determined in isolated chloride cells with a control activity of 30.9 ± 1.9 μmol Pi mg protein−1 hour−1. PDB and OAG both inhibited enzyme activity in a dose-dependent manner, with 10−5 mol.L−1 causing 45% and 26% inhibition, respectively. These results suggest that PKC is involved in regulating sodium efflux in the gills of cod by modulating Na+-K+ATPase activity.  相似文献   

8.
A highly sensitive fluorimetric assay using 3-O-methylfluorescein phosphate as substrate was used in the determination of K+-dependent phosphatase activity in preparations of rat skeletal muscle. The gastrocnemius muscle was chosen because of mixed fibre composition. Crude, detergent treated homogenate was used so as to avoid loss of activity during purification. K+-dependent phosphatase activities in the range 0.19–0.37 μmol · (g wet weight)−1 · min−1 were obtained, the value decreasing with age and K+-deficiency. Complete inhibition of the K+-dependent phosphatase was obtained with 10−3 M ouabain. Using a KSCN-extracted muscle enzyme the intimate relation between K+-dependent phosphatase activity and (Na+ + K+)-activated ATP hydrolysis could be demonstrated. A molecular activity of 620 min−1 was estimated from simultaneous determination of K+-dependent phosphatase activity and [3H]ouabain binding capacity using the partially purified enzyme preparation. The corresponding enzyme concentration in the crude homogenates was calculated and corresponded well with the number of [3H]ouabain binding sites measured in intact muscles or biopsies hereof.  相似文献   

9.
Laccase-catalyzed oxidation of phenolic compounds in organic media   总被引:1,自引:0,他引:1  
Rhus vernificera laccase-catalyzed oxidation of phenolic compounds, i.e., (+)-catechin, (−)-epicatechin and catechol, was carried out in selected organic solvents to search for the favorable reaction medium. The investigation on reaction parameters showed that optimal laccase activity was obtained in hexane at 30 °C, pH 7.75 for the oxidation of (+)-catechin as well as for (−)-epicatechin, and in toluene at 35 °C, pH 7.25 for the oxidation of catechol. Ea and Q10 values of the biocatalysis in the reaction media of the larger log p solvents like isooctane and hexane were relatively higher than those in the reaction media of lower log p solvents like toluene and dichloromethane. Maximum laccase activity in the organic media was found with 6.5% of buffer as co-solvent. A wider range of 0–28 μg protein/ml in hexane than that of 0–16.7 μg protein/ml in aqueous medium was observed for the linear increasing conversion of (+)-catechin. The kinetic studies revealed that in the presence of isooctane, hexane, toluene and dichloromethane, the Km values were 0.77, 0.97, 0.53 and 2.9 mmol/L for the substrate of (+)-catechin; 0.43, 0.34, 0.14 and 3.4 mmol/L for (−)-epicatechin; 2.9, 1.8, 0.61 and 1.1 mmol/L for catechol, respectively, while the corresponding Vmax values were 2.1 × 10−2, 2.3 × 10−2, 0.65 × 10−2 and 0.71 × 10−2 δA/μg protein min); 1.8 × 10−2, 0.88 × 10−2, 0.19 × 10−2 and 1.0 × 10−2 δA/μg protein min); 0.48 × 10−2, 0.59 × 10−2, 0.67 × 10−2 and 0.54 × 10−2 δA/μg protein min), respectively. FT-IR indicated the formation of probable dimer from (+)-catechin in organic solvent. These results suggest that this laccase has higher catalytic oxidation capacity of phenolic compounds in suitable organic media and favorite oligomers could be obtained.  相似文献   

10.
In this study, the hydraulic conductivity (Lp), Me2SO permeability ( Me2SO), and the reflection coefficients (ς) and their activation energies were determined for Metaphase II (MII) mouse oocytes by exposing them to 1.5 M Me2SO at temperatures of 30, 20, 10, 3, 0, and −3°C. These data were then used to calculate the intracellular concentration of Me2SO at given temperatures. Individual oocytes were immobilized using a holding pipette in 5 μl of an isosmotic PBS solution and perfused with precooled or prewarmed 1.5 M Me2SO solutions. Oocyte images were video recorded. The cell volume changes were calculated from the measurement of the diameter of the oocytes, assuming a spherical shape. The initial volume of the oocytes in the isoosmotic solution was considered 100%, and relative changes in the volume of the oocytes after exposure to the Me2SO were plotted against time. Mean (means ± SEM) Lpvalues in the presence of Me2SO ( Me2SOp) at 30, 20, 10, 3, 0, and −3°C were determined to be 1.07 ± 0.03, 0.40 ± 0.02, 0.18 ± 0.01, 7.60 × 10−2± 0.60 × 10−2, 5.29 × 10−2± 0.40 × 10−2, and 3.69 × 10−2± 0.30 × 10−2μm/min/atm, respectively. The Me2SOvalues were 3.69 × 10−3± 0.3 × 10−3, 1.07 × 10−3± 0.1 × 10−3, 2.75 × 10−4± 0.15 × 10−4, 7.83 × 10−5± 0.50 × 10−5, 5.24 × 10−5± 0.50 × 10−5, and 3.69 × 10−5± 0.40 × 10−5cm/min, respectively. The ς values were 0.70 ± 0.03, 0.77 ± 0.04, 0.81 ± 0.06, 0.91 ± 0.05, 0.97 ± 0.03, and 1 ± 0.04, respectively. The estimated activation energies (Ea) for Me2SOp, Me2SO, and ς were 16.39, 23.24, and −1.75 Kcal/mol, respectively. These data may provide the fundamental basis for the development of more optimal cryopreservation protocols for MII mouse oocytes.  相似文献   

11.
The nitrogen uptake and growth capabilities of the potentially harmful, raphidophycean flagellate Heterosigma akashiwo (Hada) Sournia were examined in unialgal batch cultures (strain CCMP 1912). Growth rates as a function of three nitrogen substrates (ammonium, nitrate and urea) were determined at saturating and sub-saturating photosynthetic photon flux densities (PPFDs). At saturating PPFD (110 μE m−2 s−1), the growth rate of H. akashiwo was slightly greater for cells grown on NH4+ (0.89 d−1) compared to cells grown on NO3 or urea, which had identical growth rates (0.82 d−1). At sub-saturating PPFD (40 μE m−2 s−1), both urea- and NH4+-grown cells grew faster than NO3-grown cells (0.61, 0.57 and 0.46 d−1, respectively). The N uptake kinetic parameters were investigated using exponentially growing batch cultures of H. akashiwo and the 15N-tracer technique. Maximum specific uptake rates (Vmax) for unialgal cultures grown at 15 °C and saturating PPFD (110 μE m−2 s−1) were 28.0, 18.0 and 2.89 × 10−3 h−1 for NH4+, NO3 and urea, respectively. The traditional measure of nutrient affinity—the half saturation constants (Ks) were similar for NH4+ and NO3 (1.44 and 1.47 μg-at N L−1), but substantially lower for urea (0.42 μg-at N L−1). Whereas the α parameter (α = Vmax/Ks), which is considered a more robust indicator for substrate affinity when substrate concentrations are low (<Ks), were 19.4, 12.2 and 6.88 × 10−3 h−1/(μg-at N L−1) for NH4+, NO3 and urea, respectively. These laboratory results demonstrate that at both saturating and sub-saturating N concentrations, N uptake preference follows the order: NH4+ > NO3 > urea, and suggests that natural blooms of H. akashiwo may be initiated or maintained by any of the three nitrogen substrates examined.  相似文献   

12.
The shuttle system that mediates the transport of fatty acids across the mitochondrial membrane in invertebrates has received little attention. Carnitine O-palmitoyltransferase I (EC 2.3.1.21; CPT I) is a key component of this system that in vertebrates controls long-chain fatty acid β-oxidation. To gain knowledge on the acyltransferases in aquatic arthropods, physical, kinetic, regulatory and immunological properties of CPT of the midgut gland mitochondria of Macrobrachium borellii were assayed. CPT I optimum conditions were 34 °C and pH = 8.0. Kinetic analysis revealed a Km for carnitine of 2180 ± 281 μM and a Km for palmitoyl-CoA of 98.9 ± 8.9 μM, while Vmax were 56.5 ± 6.6 and 36.7 ± 4.8 nmol min− 1 mg protein− 1, respectively. A Hill coefficient, n ~ 1, indicate a Michaelis–Menten behavior. The CPT I activity was sensitive to regulation by malonyl-CoA, with an IC50 of 25.2 μM. Electrophoretic and immunological analyses showed that a 66 kDa protein with an isoelectric point of 5.1 cross-reacted with both rat liver and muscle-liver anti CPT I polyclonal antibodies, suggesting antigenic similarity with the rat enzymes. Although CPT I displayed kinetic differences with insect and vertebrates, prawn showed a high capacity for energy generation through β-oxidation of long-chain fatty acids.  相似文献   

13.
1. (1) VO3 combines with high affinity to the Ca2+-ATPase and fully inhibits Ca2+-ATPase and Ca2+-phosphatase activities. Inhibition is associated with a parallel decrease in the steady-state level of the Ca2+-dependent phosphoenzyme.
2. (2) VO3 blocks hydrolysis of ATP at the catalytic site. The sites for VO3 also exhibit negative interactions in affinity with the regulatory sites for ATP of the Ca2+-ATPase.
3. (3) The sites for VO3 show positive interactions in affinity with sites for Mg2+ and K+. This accounts for the dependence on Mg2+ and K+ of the inhibition by VO3. Although, with less effectiveness, Na+ substitutes for K+ whereas Li+ does not. The apparent affinities for Mg2+ and K+ for inhibition by VO3 seem to be less than those for activation of the Ca2+-ATPase.
4. (4) Inhibition by VO3 is independent of Ca2+ at concentrations up to 50 μM. Higher concentrations of Ca2+ lead to a progressive release of the inhibitory effect of VO3.
Keywords: Ca2+-ATPase; Vanadate inhibition; K+; Li+; (Red cell membrane)  相似文献   

14.
Summary The pH-stat technique has been used to measure H+ fluxes in gastric mucosa and urinary bladder in vitro while keeping mucosal pH constant. We now report application of this method in renal tubules. We perfused proximal tubules with double-barreled micropipettes, blocked luminal fluid columns with oil and used a double-barreled Sb/reference microelectrode to measure pH, and Sb or 1n HC1-filled microelectrodes to inject OH or H+ ions into the tubule lumen. By varying current injection, pH was kept constant at adjustable levels by an electronic clamping circuit. We could thus obtain ratios of current (nA) to pH change (apparent H+-ion conductance). These ratios were reduced after luminal 10–4 m acetazolamide, during injection of OH, but they increased during injection of H+. The point-like injection source causes pH to fall off with distance from the injecting electrode tip even in oil-blocked segments. Therefore, a method analogous to cable analysis was used to obtain H+ fluxes per cm2 epithelium. The relation betweenJ H + and pH gradient showed saturation kinetics of H fluxes, both during OH and H+ injection. This kinetic behavior is compatible with inhibition ofJ H by luminal H+. It is also compatible with dependence on Na+ and H+ gradients of a saturable Na/H exchanger. H+-ion back-flux into the tubule lumen also showed saturation kinetics. This suggests that H+ flow is mediated by a membrane component, most likely the Na+–H+ exchanger.  相似文献   

15.
The soybean (Glycine max) urease was immobilized on alginate and chitosan beads and various parameters were optimized and compared. The best immobilization obtained were 77% and 54% for chitosan and alginate, respectively. A 2% chitosan solution (w/v) was used to form beads in 1N KOH. The beads were activated with 1% glutaraldehyde and 0.5 mg protein was immobilized per ml of chitosan gel for optimum results. The activation and coupling time were 6 h and 12 h, respectively. Further, alginate and soluble urease were mixed to form beads and final concentrations of alginate and protein in beads were 3.5% (w/v) and 0.5 mg/5 ml gel. From steady-state kinetics, the optimum temperature for urease was 65 °C (soluble), 75 °C (chitosan) and 80 °C (alginate). The activation energies were found to be 3.68 kcal mol−1, 5.02 kcal mol−1, 6.45 kcal mol−1 for the soluble, chitosan- and alginate-immobilized ureases, respectively. With time-dependent thermal inactivation studies, the immobilized urease showed improved stability at 75 °C and the t1/2 of decay in urease activity was 12 min, 43 min and 58 min for soluble, alginate and chitosan, respectively. The optimum pH of urease was 7, 6.2 and 7.9 for soluble, alginate and chitosan, respectively. A significant change in Km value was noticed for alginate-immobilized urease (5.88 mM), almost twice that of soluble urease (2.70 mM), while chitosan showed little change (3.92 mM). The values of Vmax for alginate-, chitosan-immobilized ureases and soluble urease were 2.82 × 102 μmol NH3 min−1 mg−1 protein, 2.65 × 102 μmol NH3 min−1 mg−1 protein and 2.85 × 102 μmol NH3 min−1 mg−1 protein, respectively. By contrast, reusability studies showed that chitosan–urease beads can be used almost 14 times with only 20% loss in original activity while alginate–urease beads lost 45% of activity after same number of uses. Immobilized urease showed improved stability when stored at 4 °C and t1/2 of urease was found to be 19 days, 80 days and 121 days, respectively for soluble, alginate and chitosan ureases. The immobilized urease was used to estimate the blood urea in clinical samples. The results obtained with the immobilized urease were quite similar to those obtained with the autoanalyzer®. The immobilization studies have a potential role in haemodialysis machines.  相似文献   

16.
Large blooms of the marine cyanobacterium Lyngbya majuscula in Moreton Bay, Australia (27°05′S, 153°08′E) have been re-occurring for several years. A bloom was studied in Deception Bay (Northern Moreton Bay) in detail over the period January–March 2000. In situ data loggers and field sampling characterised various environmental parameters before and during the L. majuscula bloom. Various ecophysiological experiments were conducted on L. majuscula collected in the field and transported to the laboratory, including short-term (2 h) 14C incorporation rates and long-term (7 days) pulse amplitude modulated (PAM) fluorometry assessments of photosynthetic capacity. The effects of L. majuscula on various seagrasses in the bloom region were also assessed with repeated biomass sampling. The bloom commenced in January 2000 following usual December rainfall events, water temperatures in excess of 24 °C and high light conditions. This bloom expanded rapidly from 0 to a maximum extent of 8 km2 over 55 days with an average biomass of 210 gdw−1 m−2 in late February, followed by a rapid decline in early April. Seagrass biomass, especially Syringodium isoetifolium, was found to decline in areas of dense L. majuscula accumulation. Dissolved and total nutrient concentrations did not differ significantly (P > 0.05) preceding or during the bloom. However, water samples from creeks discharging into the study region indicated elevated concentrations of total iron (2.7–80.6 μM) and dissolved organic carbon (2.5–24.7 mg L−1), associated with low pH values (3.8–6.7). 14C incorporation rates by L. majuscula were significantly (P < 0.05) elevated by additions of iron (5 μM Fe), an organic chelator, ethylenediaminetetra-acetic acid (5 μM EDTA) and phosphorus (5 μM PO4−3). Photosynthetic capacity measured with PAM fluorometry was also stimulated by various nutrient additions, but not significantly (P > 0.05). These results suggest that the L. majuscula bloom may have been stimulated by bioavailable iron, perhaps complexed by dissolved organic carbon. The rapid bloom expansion observed may then have been sustained by additional inputs of nutrients (N and P) and iron through sediment efflux, stimulated by redox changes due to decomposing L. majuscula mats.  相似文献   

17.
We measured Na+/K+ ATPase activity in homogenates of gill tissue prepared from field caught, winter and summer acclimatized yellow perch, Perca flavescens. Water temperatures were 2–4°C in winter and 19–22°C in summer. Na+/K+ ATPase activity was measured at 8, 17, 25, and 37°C. Vmax values for winter fish increased from 0.48±0.07 μmol P mg−1 protein h−1 at 8°C to 7.21±0.79 μmol P mg−1 protein h−1 at 37°C. In summer fish it ranged from 0.46±0.08 (8°C) to 3.86±0.50 (37°C) μmol P mg−1 protein h−1. The Km for ATP and for Na+ at 8°C was ≈1.6 and 10 mM, respectively and did not vary significantly with assay temperature in homogenates from summer fish. The activation energy for Na+/K+ ATPase from summer fish was 10 309 (μmol P mg−1 h−1) K−1. In winter fish, the Km for ATP and Na+ increased from 0.59±0.08 mM and 9.56±1.18 mM at 8°C to 1.49±0.11 and 17.88±2.64 mM at 17°C. The Km values for ATP and Na did not vary from 17 to 37°C. A single activation energy could not be calculated for Na/K ATPase from winter fish. The observed differences in enzyme activities and affinities could be due to seasonal changes in membrane lipids, differences in the amount of enzyme, or changes in isozyme expression.  相似文献   

18.
In a comparative experiment the effect of cortisol and growth hormone (GH) on the hypo-osmoregulatory ability of a landlocked and an anadromous strain of Arctic charr (Salvelinus alpinus) was investigated. Cortisol and GH were implanted either alone or in combination, and the fish were exposed to a 24 h seawater challenge test (SWT) on days 14 and 28 after implantation. Hypo-osmoregulatory ability, measured as plasma osmolality and chloride concentration after the SWTs, was better in the anadromous than in the landlocked strain, irrespective of treatment. However, cortisol provided a strong stimulation of hypo-osmoregualtory ability in both strains, and this stimulation seemed to be potentiated by GH in an additive manner. Improved hypo-osmoregulatory ability in GH + cortisol treated anadromous Arctic charr was accompanied by increased gill Na+, K+-ATPase activity and Na+–K+–2Cl cotransporter protein abundance, but no changes in gill Na+,K+-ATPase α1a and α1b mRNA levels. For landlocked charr the improved hypo-osmoregulatory ability in GH +cortisol treated fish was accompanied only with an increase in gill Na+–K+–2Cl cotransporter protein abundance. Hormone treatment caused an improvement of hypo-osmoregulatory ability that was of approximately the same magnitude in the landlocked as in the anadromous Arctic charr. This suggests that the lack of spontaneous development of hypo-osmoregulatory ability often seen in landlocked populations of Arctic charr may depend, at least partly, on a lack of the hormonal activation seen in anadromous populations.  相似文献   

19.
Oxygen isotope fractionation between human phosphate and water revisited   总被引:1,自引:0,他引:1  
The oxygen isotope composition of human phosphatic tissues (δ18OP) has great potential for reconstructing climate and population migration, but this technique has not been applied to early human evolution. To facilitate this application we analyzed δ18OP values of modern human teeth collected at 12 sites located at latitudes ranging from 4°N to 70°N together with the corresponding oxygen composition of tap waters (δ18OW) from these areas. In addition, the δ18O of some raw and boiled foods were determined and simple mass balance calculations were performed to investigate the impact of solid food consumption on the oxygen isotope composition of the total ingested water (drinking water + solid food water). The results, along with those from three, smaller published data sets, can be considered as random estimates of a unique δ18OW18OP linear relationship: δ18OW = 1.54(±0.09) × δ18OP−33.72(±1.51) (R2 = 0.87: p [H0:R2 = 0] = 2 × 10−19). The δ18O of cooked food is higher than that of the drinking water. As a consequence, in a modern diet the δ18O of ingested water is +1.05 to 1.2‰ higher than that of drinking water in the area. In meat-dominated and cereal-free diets, which may have been the diets of some of our early ancestors, the shift is a little higher and the application of the regression equation would slightly overestimate δ18OW in these cases.  相似文献   

20.
Store-operated Ca2+ entry (SOCE) is an important mechanism for Ca2+ influx in smooth muscle cells; however the activation and regulation of this influx pathway are incompletely understood. In the present study we have examined the effect of several protein kinases in regulating SOCE in pulmonary artery smooth muscle cells (PASMCs) of the rat. Inhibition of protein kinase C with chelerythrine (3 μM) potentiated SOCE by 47 ± 2%, while the tyrosine kinase inhibitors genistein (100 μM) and tyrphostin 23 (100 μM) caused a significant reduction in SOCE of 55 ± 9% and 43 ± 7%, respectively. It has been proposed that Ca2+-insensitive phospholipase A2 (iPLA2) is involved in the activation of SOCE in many different cell types. The iPLA2 inhibitor, bromoenol lactone had no effect on SOCE, suggesting that this mechanism was not involved in the activation of the pathway. The calmodulin antagonists, calmidazolium (CMZ) (10 μM) and W-7 (10 μM) appeared to potentiate SOCE in PASMCs. Further investigation established that CMZ was actually activating a Ca2+ influx pathway that was independent of the filling state of the sarcoplasmic reticulum. The CMZ-activated Ca2+ influx was blocked by Gd3+ (10 μM), but unaffected by 2-APB (75 μM), indicating a pharmacological profile distinct from the classical SOCE pathway.  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