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 31 毫秒
1.
The kinetics of the decomposition reactions of the CO(py)3(CO3)(H2O)+ ion have been investigated in aqueous perchloric acid solutions over a range of hydrogen ion concentrations (0.10 to 5.0 M) and at two ionic strengths (I = 1.0 and 5.0 M). At the lower ionic strength, plots of ln (AtA versus time show a nonlinearity that is consistent with that expected for consecutive first-order reactions. The rates of the faster reaction are similar to those reported for the spontaneous reduction of aquopyridine-cobalt(III) cations. At the higher ionic strength, the above noted curvature is not apparent and the decarboxylation kinetics of the title complex may be described by a pseudo-first-order rate constant: kobs = k[H3O+]. At 20°C, k = (1.75−+0.09) s−1 M−1 with activation parameters ofΔH = (97 −+ 4) kJ mol−1 and ΔS = −(54 −+ 32) J deg−1 mol−1. These kinetic parameters are compared with those previously reported for the similar complexes, Co(py)4CO3+ and Co(py)2(CO3)(H2O)2+.  相似文献   

2.
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   

3.
The kinetics of formation of the complex ion, μ-carbonato-di-μ-hydroxo-bis((1,5-diamino-3-aza-pentane) cobalt(III), from the tri-μ-hydroxo-bis((1,5-diamino-3-aza-pentane(III)cobalt(III)) ion in aqueous buffered carbonate solution have been studied spectrophotometrically at 295 nm over the ranges 20.0θ°C34.8, 8.03pH9.44, 5 mM [CO32−35 mM and at an ionic strength of 0.1 M (LiClO4). On the basis of the kinetic results a mechanism, involving rapid cleavage of an hydroxo bridge followed by carbon dioxide uptake with subsequent bridge formation, has been proposed. At 25 °C, the rate of the carbon dioxide uptake is 0.58 M−1 s−1 with ΔH≠ = (13.2±0.7) kcal mol−1 and ΔS≠ = (−15.1 ± 0.7) cal deg−1 mol−1. The results are composed with those obtained for several mononuclear cobalt(III) and one dinuclear cobalt(III) complexes.  相似文献   

4.
31P NMR has been employed to study the interaction between zinc(II) bis(O,O′-di-iso-butyldithiophosphate), Zn[S2P(OiBu)2]2, and four multidentate amines (diethylenetriamine, triethylenetetramine, tetraethylenepentamine and pentaethylenehexamine) in chloroform at 294 K. The major interaction of Zn[S2P(OiBu)2]2 and these polyamines involves displacement of the {S2P(OiBu)2} ligands from the zinc giving [Zn(amine)]2+ and [S2P(OiBu)2] ions in solution. The magnitudes of the equilibrium constants, K1 (=[{Zn(amine)}2+][{DDP}]2/[Zn(DDP)2][amine]), have been evaluated in the cases of triethylenetetramine (20.0 l mol−1), tetraethylenepentamine (19.1 l mol−1) and pentaethylenehexamine (1.58 l mol−1). Crystalline 1:1 ionic complexes have also been isolated from these systems and characterised.  相似文献   

5.
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   

6.
The rate constant for the hydrolysis of prostacyclin (PGI2) to 6-keto-PGF was measured by monitoring the UV spectral change, over a pH range 6 to 10 at 25°C and the total ionic strength of 0.5 M. The first-order rate constant (kobs) extrapolated to zero buffer concentration follows an expression, kobs = kH+ (H+), where kH+ is a second-order rate constant for the specific acid catalyzed hydrolysis. The value of kH+ obtained (3.71 × 104 sec−1 M−1) is estimated approximately 700-fold greater than a kH+ value expected from the hydrolysis of other vinyl ethers. Such an unusually high reactivity of PGI2 even for a vinyl ether is attributed to a possible ring strain release that would occur upon the rate controlling protonation of C5. A Brønsted slope (α) of 0.71 was obtained for the acid (including H3O+) catalytic constants, from which a pH independent first-order rate constant for the spontaneous hydrolysis (catalyzed by H2O as a general acid) was estimated to be 1.3 × 10−6 sec−1. An apparent activation energy (Ea) of 11.85 Kcal/mole was obtained for the hydrolysis at pH 7.48, from which a half-life of PGI2 at 4°C was estimated to be approximately 14.5 min. when the total phosphate concentration is 0.165 M (cf. 3.5 min. at 25°C).  相似文献   

7.
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   

8.
This study compared the mass-specific routine metabolic rate (RMR) of similar sized mulloway (Argyrosomus japonicus), a sedentary species, and yellowtail kingfish (Seriola lalandi), a highly active species, acclimated at one of several temperatures ranging from 10–35 °C. Respirometry was carried out in an open-top static system and RMR corrected for seawater–atmosphere O2 exchange using mass-balance equations. For both species RMR increased linearly with increasing temperature (T). RMR for mulloway was 5.78T − 29.0 mg O2 kg− 0.8 h− 1 and for yellowtail kingfish was 12.11T − 39.40 mg O2 kg− 0.8 h− 1. The factorial difference in RMR between mulloway and yellowtail kingfish ranged from 2.8 to 2.2 depending on temperature. The energetic cost of routine activity can be described as a function of temperature for mulloway as 1.93T − 9.68 kJ kg− 0.8 day− 1 and for yellowtail kingfish as 4.04T − 13.14 kJ kg− 0.8 day− 1. Over the full range of temperatures tested Q10 values were approximately 2 for both species while Q10 responses at each temperature increment varied considerably with mulloway and yellowtail kingfish displaying thermosensitivities indicative of each species respective niche habitat. RMR for mulloway was least thermally dependent at 28.5 °C and for yellowtail kingfish at 22.8 °C. Activation energies (Ea) calculated from Arrhenius plots were not significantly different between mulloway (47.6 kJ mol− 1) and yellowtail kingfish (44.1 kJ mol− 1).  相似文献   

9.
Net ecosystem exchange of CO2 (NEE) was measured during 2005 using the eddy covariance (EC) technique over a reed (Phragmites australis (Cav.) Trin. ex Steud.) wetland in Northeast China (121°54′E, 41°08′N). Diurnal NEE patterns varied markedly among months. Outside the growing season, NEE lacked a diurnal pattern and it fluctuated above zero with an average value of 0.07 mg CO2 m−2 s−1 resulting from soil microbial activity. During the growing season, NEE showed a distinct V-like diel course, and the mean daily NEE was −7.48 ± 2.74 g CO2 m−2 day−1, ranging from −13.58 g CO2 m−2 day−1 (July) to −0.10 g CO2 m−2 day−1 (October). An annual cycle was also apparent, with CO2 uptake increasing rapidly in May, peaking in July, and decreasing from August. Monthly cumulative NEE ranged from −115 ± 24 g C m−2 month−1 (the reed wetland was a CO2 sink) in July to 75 ± 16 g C m−2 month−1 (CO2 source) in November. The annual CO2 balance suggests a net uptake of −65 ± 14 g C m−2 year−1, mainly due to the gains in June and July. Cumulative CO2 emission during the non-growing season was 327 g C m−2, much greater than the absolute value of the annual CO2 balance, which proves the importance of the wintertime CO2 efflux at the study site. The ratio of ecosystem respiration (Reco) to gross primary productivity (GPP) for this reed ecosystem was 0.95, indicating that 95% of plant assimilation was consumed by the reed plant or supported the activities of heterotrophs in the soil. Daytime NEE values during the growing season were closely related to photosynthetically active radiation (PAR) (r2 > 0.63, p < 0.01). Both maximum ecosystem photosynthesis rate (Amax) and apparent quantum yield (α) were season-dependent, and reached their peak values in July (1.28 ± 0.11 mg CO2 m−2 s−1, 0.098 ± 0.027 μmol CO2 μmol−1 photon, respectively), corresponding to the observed maximum NEE in July. Ecosystem respiration (Reco) relied on temperature and soil water content, and the mean value of Q10 was about 2.4 with monthly variation ranging from 1.8 to 4.1 during 2005. Annual methane emission from this reed ecosystem was estimated to be about 3 g C m−2 year−1, and about 5% of the net carbon fixed by the reed wetland was released to the atmosphere as CH4.  相似文献   

10.
1. The fat mouse Steatomys pratensis natalensis (mean body mass 37.4±0.43 (se)) has a low euthermic body temperature Tb=30.1–33.8 °C and a low basal metabolic rate (BMR)=0.50 ml O2 g−1 h−1.
2. Below an ambient temperature (Ta)=15 °C, the mice were hypothermic.
3. The lowest survivable Ta=10 °C.
4. Torpor is efficient in conserving energy between Ta=15–30 °C, below Ta=15 °C, the mice arouse.
5. Euthermic and torpid mice were hyperthermic at Ta=35 °C.
6. Thermal conductance was 0.159 ml O2 g−1 h−1 °C−1, 98.8% of the expected value.
7. Non-shivering thermogenesis (NST) was 2.196 ml O2 g−1 h−1 (3.69×BMR).
8. Maximal oxygen consumption, however, was 3.83 ml O2 g−1 h−1 (6.44×BMR), indicating that other methods of heat production are additive.
9. Because fat mice conserve energy by torpor only between Ta=15–30 °C, we suggest that torpor may be a more important mechanism for surviving food shortages than for surviving cold weather.
Keywords: Steatomys pratensis natalensis; Metabolism; Torpor; Fat mouse  相似文献   

11.
Granulosa, theca and corpus luteum cells of the goat ovary were isolated and incubated separately for 6 hours, with or without various modulators. Arachidonic acid (AA, 10 ng to 100 μg/ml), the precursor for prostaglandin synthesis, produced a dose-dependent increase in progesterone (P4) and estradiol-17β (E2) productin by all the cell types. Prostaglandin synthetase inhibitors, aspirin (10−6−10−3M) and indomethacin (100 ng−1 mg/ml), produced a dose-dependent decrease in arachidonic acid-stimulated (100 μ/ml) steroid production. Prostacyclin synthetase stimulators, trapidil (1.6 μg− 1 mg/ml) and dipyridamole (10−6−10−3M), when added alone or along with AA, did not effect steroid production. Up to 100 μg/ml of U-51605 (9,11-azoprosta-5, 13-dienoic acid), a prostacyclin synthetase inhibitor, did not inhibit basal or AA-stimulated steroid production. Prostacyclin (PGl2) and its stable analog 6βPGl1(0.01–10μg/ml) produced a dose-dependent increase in P4 and E2 production in all three cell types. Increase at 1 and 10μg/ml was significant in all cases. 6-keto-PGE1 (an active metabolite of PGl2 in certain systems) produced an increase in steroid production which was significant in theca at 1μg/ml concentrations but had no significant effect on granulosa and corpus luteum cells at any dose level. 6-keto-PGf1 alpha (stable metabolite of PGl2) was without effect inthe present system. The lack of effect of PGl2 at lower concentrations was not altered by either differentiation of the cells with FSH and testosterone or addition of steroid precursors, testosterone and pregnenolene. The present results indicate that AA- stimualted steroid production in the goat ovarian cell type is mediated by prostaglandins other than PGl2 though PGl2 itself can positively modulate the steroid production.  相似文献   

12.
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   

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.
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   

15.
We have previously reported that angiotensin II (ANG II) induces oscillations in the cytoplasmic calcium concentration ([Ca2+]i) of pulmonary vascular myocytes. The present work was undertaken to investigate the effect of ANG II in comparison with ATP and caffeine on membrane currents and to explore the relation between these membrane currents and [Ca2+]i. In cells clamped at −60 mV, ANG II (10 μM) or ATP (100 μM) induced an oscillatory inward current. Caffeine (5 μM) induced only one transient inward current. In control conditions, the reversal potential (Erev) of these currents was close to the equilibrium potential for Cl ions (ECl = −2.1 mV) and was shifted towards more positive values in low-Cl solutions. Niflumic acid (10–50 μM) and DIDS (0.25-1 mM) inhibited this inward current. Combined recordings of membrane current and [Ca2+]i by Indo-1 microspectrofluorimetry revealed that ANG II- and ATP-induced currents occurred simultaneously with oscillations in [Ca2+]i, whereas the caffeine-induced current was accompanied by only one transient increase in [Ca2+]i Niflumic acid (25 μM) had no effect on agonist-induced [Ca2+]i responses, whereas thapsigargin (1 μM) abolished both membrane current and the [Ca2+]i response. Heparin (5 mg/ml in the pipette solution) inhibited both [Ca2+]i responses and membrane currents induced by ANG II and ATP, but not by caffeine. In pulmonary arterial strips, ANG II-induced contraction was inhibited by niflumic acid (25 μM) or nifedipine (1 μM) to the same extent and the two substances did not have an additive effect. This study demonstrates that, in pulmonary vascular smooth muscle, ANG II, as well as ATP, activate an oscillatory calcium dependent chloride current which is triggered by cyclic increases in [Ca2+]i and that both oscillatory phenomena are primarily IP3 mediated. It is suggested that ANG II-induced oscillatory chloride current could depolarise the cell membrane leading to activation of voltage-operated Ca2+ channels. The resulting Ca2+ influx contributes to the component of ANG II-induced contraction that is equally sensitive to chloride or calcium channel blockade.  相似文献   

16.
A sensitive, selective, and rapid enzymatic method is proposed for the quantification of hydrogen peroxide (H2O2) using 3-methyl-2-benzothiazolinonehydrazone hydrochloride (MBTH) and 10,11-dihydro-5H-benz(b,f)azepine (DBZ) as chromogenic cosubstrates catalyzed by horseradish peroxidase (HRP) enzyme. MBTH traps free radical released during oxidation of H2O2 by HRP and gets oxidized to electrophilic cation, which couples with DBZ to give an intense blue-colored product with maximum absorbance at 620 nm. The linear response for H2O2 is found between 5 × 10−6 and 45 × 10−6 mol L−1 at pH 4.0 and a temperature of 25 °C. Catalytic efficiency and catalytic power of the commercial peroxidase were found to be 0.415 × 106 M−1 min−1 and 9.81 × 10−4 min−1, respectively. The catalytic constant (kcat) and specificity constant (kcat/Km) at saturated concentration of the cosubstrates were 163.2 min−1 and 4.156 × 106 L mol−1 min−1, respectively. This method can be incorporated into biochemical analysis where H2O2 undergoes catalytic oxidation by oxidase. Its applicability in the biological samples was tested for glucose quantification in human serum.  相似文献   

17.
Li Yang  Gary J. Stephens   《Cell calcium》2009,46(4):248-256
Voltage-dependent Ca2+ channels (VDCCs) have emerged as targets to treat neuropathic pain; however, amongst VDCCs, the precise role of the CaV2.3 subtype in nociception remains unproven. Here, we investigate the effects of partial sciatic nerve ligation (PSNL) on Ca2+ currents in small/medium diameter dorsal root ganglia (DRG) neurones isolated from CaV2.3(−/−) knock-out and wild-type (WT) mice. DRG neurones from CaV2.3(−/−) mice had significantly reduced sensitivity to SNX-482 versus WT mice. DRGs from CaV2.3(−/−) mice also had increased sensitivity to the CaV2.2 VDCC blocker ω-conotoxin. In WT mice, PSNL caused a significant increase in ω-conotoxin-sensitivity and a reduction in SNX-482-sensitivity. In CaV2.3(−/−) mice, PSNL caused a significant reduction in ω-conotoxin-sensitivity and an increase in nifedipine sensitivity. PSNL-induced changes in Ca2+ current were not accompanied by effects on voltage-dependence of activation in either CaV2.3(−/−) or WT mice. These data suggest that CaV2.3 subunits contribute, but do not fully underlie, drug-resistant (R-type) Ca2+ current in these cells. In WT mice, PSNL caused adaptive changes in CaV2.2- and CaV2.3-mediated Ca2+ currents, supporting roles for these VDCCs in nociception during neuropathy. In CaV2.3(−/−) mice, PSNL-induced changes in CaV1 and CaV2.2 Ca2+ current, consistent with alternative adaptive mechanisms occurring in the absence of CaV2.3 subunits.  相似文献   

18.
Long-lived metastable states involving multiple binding sites of a protein ligand with immobilized alkyl residues on a solid phase can be observed at high ionic strength between butyl agaroses (5.21 μ mol/ml packed gel) and phosphorylase b by perturbations enforcing either the on-reaction (adsorption) or the off-reaction (desorption). These apparent equilibrium states are suggested because the adsorption isotherms of phosphorylase b on butyl agaroses are not retraced by the desorption isotherms. In this first example of macromolecular adsorption hysteresis on immobilized alkyl residues, it can be shown that the irreversible entropy (ΔiS) produced in an adsorption-desorption cycle lies between 6 (5 μ mol/ml packed gel) and 40 (21 μ mol/ml packed gel) J mol 1 K−1. For the latter gel the apparent standard entropy of adsorption (ΔaSi0′) is 160 J mol−1 K−1. The metastable state observed during adsorption is probably due to an energy barrier which must be overcome for the nucleation of protein binding on the matrix. Other metastable states may possibly be encountered during desorption when the adsorbed enzyme resists the breakage of hydrophobic interactions. In the transition from the adsorption branch to the desorption branch of the hysteresis loop, the apparent affinity of the enzyme-matrix interaction is enhanced. For the desorption branch, the apparent association constant of half-maximal saturation corresponds to Kd,0.5′ = 4.2 × 109 ]m−1 as compared to the respective constant of adsorption Ka, 0.5′ = 1.6 × 105m−1 (gel: 21μ mol/ml packed gel). Since the area of the hysteresis loops (see also ΔiS) depends strongly on the density of butyl residues on the gel, it is concluded that the number of alkyl residues interacting with the protein molecule is crucial for the metastable states and hysteresis. It is unlikely that hysteresis is due to the pore structure of the agarose or to nearest neighbour interactions of ligand molecules. Since thermodynamic irreversibility and hysteresis may be encountered when macromolecules, such as proteins, are adsorbed to cell membranes or cell organelles: an analysis and understanding of these phenomena should be of general biological significance.  相似文献   

19.
We examined effects of protein kinase C (PKC) activation by phorbol dibutyrate (PDB) on prostaglandin production in astroglia. Astroglia were cultured from sheep fetal cortex and grown in Eagle's basal media supplemented with 10% fetal calf serum (BME-C). Prostaglandin F2a (PGF) levels in media were determined at 2–24 hours after exposure to PDB. PDB increased production of PGF at 10−8M and 10−6M. In addition, PDB increased the ratio of membrane to cytosolic PKC. Coapplication of H7 [1-(5-isoquinolinylsulfonyl)-2-methyl-piperazine] (10−4M) with PDB (10−6M) inhibited PDB-induced PGF2a production. To investigate the role of protein synthesis in increased prostaglandin production by PDB, astroglia were coincubated with actinomycin D (1 mg/ml) or cycloheximide (10 mg/ml). At 4 hrs, both actinomycin D and cycloheximide inhibited increases in PGF2a in response to PDB application. In addition, COX-2 mRNA levels and COX activity levels were examined. PDB increased COX-2 mRNA levels by 2 hours, and COX activity tripled after 12 hr exposure to PDB. In addition, the increase in COX activity was blocked by cycloheximide. In summary, PKC activation promotes enhanced prostaglandin production via an increase in COX synthesis.  相似文献   

20.
(1) In order to assess the possible role of 3′,5′-(cyclic)adenosine monophosphate (cAMP) in the control of glucose transport, the effect of the nucleotide or agents known to increase its intracellular concentration on sugar transport or 45Ca2+ washout were characterized in epididymal fat pads, free fat cells and soleus muscles of the rat. (2) When added to the incubation medium, cAMP (0.1–2.0 mM) stimulated 3-O-[14C]methylglucose washout from fat pads. This effect was abolished by cytochalasin B, and additive to that induced by submaximal (10–25 μU/ml), but not by supramaximal (10 mU/ml) concentrations of insulin. (3) cAMP (2 mM) stimulated the conversion of [U-14C]glucose into CO2 and triacylglycerols. This effect was additive to that of insulin (100 μU/ml). (4) ACTH, glucagon, adrenaline, noradrenaline and salbutamol, which are all known to increase the cAMP content of adipose tissue, stimulated the washout of 3-O-[14C]methylglucose and 45Ca2+ from preloaded fat pads. The fractional losses of the two isotopes were significantly correlated (P < 0.001, r = 0.73). (5) In free fat cells, adrenaline (10−6 M) and salbutamol (10−5 M) stimulated the uptake of 3-O-[14C]methylglucose, and salbutamol (10−5 M) did not interfere with the stimulating effect of insulin (25 μU/ml) on sugar uptake. (6) In rat soleus muscles, adrenaline and salbutamol produced a dose-dependent stimulation of the washout of 3-O-[14C]methylglucose and 45Ca2+. The effect of adrenaline on sugar efflux was abolished by propranolol. (7) It is concluded that the activation of the glucose transport system by insulin is unlikely to be mediated by a drop in the cellular concentration of cAMP. An increase in cAMP brought about by β-adrenoceptor agonists or lipolytic hormones may induce a mobilization of calcium ions from cellular pools into the cytoplasm, which in turn leads to the activation of the glucose transport system demonstrated in the present as well as in several earlier studies.  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