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.
Rainbow trout were used to characterize the direct influence of growth hormone on hepatic lipid mobilization in vitro. Liver was removed from fish fasted 72 h, sliced into 1-mm3 pieces and incubated in Hanks' medium containing ovine or salmon growth hormone (0.28 ng·ml-1–28 g·ml-1). Lipid mobilization, resulting from triacylglycerol hydrolysis, was evaluated by fatty acid and glycerol release into culture medium. Control rates of fatty acid release and glycerol release were 0.95±0.16 (mean ± SE) and 0.88±0.28 mol·l-1·mg fresh weight, respectively. Both ovine growth hormone (28 ng·ml-1) and salmon growth hormone (28 ng·ml-1) stimulated fatty acid release into culture medium, increasing rates by 112% and 97%, respectively, during the course of a 24-h incubation. Glycerol release was similarly augmented by growth hormone treatment. Growth hormone-stimulated metabolite release became evident within 12 h and persisted for up to 72 h. The presence of amino acids in the culture medium potentiated the lipolytic action of growth hormone. Ovine growth hormone (28 ng·ml-1) in the presence of amino acids stimulated a 53% increase in fatty acid, and a 108% increase in glycerol, release over rates observed in the absence of amino acids. Salmon growth hormone (28 ng·ml-1) in the presence of amino acids stimulated a 53% increase in fatty acid, and a 44% increase in glycerol, release over rates observed in the absence of amino acids. Ovine growth hormone (28 ng·ml-1) also stimulated gluconeogenesis, as indicated by a 75% increase in phosphoenolpyruvate carboxykinase activity in liver pieces incubated in the presence of amino acids. These results indicate that growth hormone directly stimulates lipid breakdown in the liver of trout and that amino acids potentiate growth hormone-stimulated lipolysis.Abbreviations AA amino acid(s) - dGDP deoxy-guanosine diphosphate - ED 50 50% effective dose - FA fatty acid(s) - fw fesh weight - GH growth hormone - Hepes 4-(2-hydroxyethyl)-1-piperazineethanesulphonic acid - MS-222 tricaine methanesulfonate - MEM minimum essential medium - oGH ovine growth hormone - PEPCK phosphoenolpyruvate carboxykinase - PKc protein kinase C - rpm revolutions per minute - sGH salmon growth hormone - TG triacylglycerol - w/v weight per volume This paper was presented in abstract form at the Annual Meeting of the American Society of Zoologists, Dec. 26–30, 1991, Atlanta, GA  相似文献   

2.
Summary Rates of O2 uptake across isolated perfused skin of bullfrogs (Rana catesbeiana) were measured in relation to blood flow at three levels of ambient O2 tension: normoxia (O2 tension=152 torr), hypoxia (12% O2, 87 torr) and hyperoxia (42% O2, 306 torr). At bulk perfusion rates ranging from 3.4 to 10.1 l·cm-2·min-1, O2 uptake was positively correlated with hemoglobin delivery rate in both normoxia and hyperoxia, but was independent of delivery rate in hypoxia. Mean O2 uptake in normoxia was 3.8 nmol O2·cm-2·min-1 at a delivery rate of 9.8 nmol·cm-2·min-1 and 6.5 nmol O2·cm-2·min-1 at a delivery rate of 28.3 nmol·cm-2·min-1. At any given bulk perfusion rate, oxygen uptake averaged about 49% lower in hypoxia than in normoxia, decreasing in proportion to the reduction of O2 tension difference between medium and blood. In hyperoxia, O2 uptake did not increase proportionally with the difference in O2 tension between blood and medium, averaging only 50% higher at a 2.4-fold greater O2 tension difference. Cutaneous diffusing capacity for O2 averaged 0.041 nmol O2·cm-2·torr-1·min-1 during the first hour of perfusion in normoxia, and was not affected by reduction of ambient O2 tension. The results indicate that cutaneous O2 uptake in hypoxia is highly diffusion limited, and consequently, increases in cutaneous perfusion can not effectively compensate for reduction of ambient O2 tension. In hyperoxia, O2 uptake may be substantially perfusion limited because of reduced blood O2 capacitance at high O2 saturations.Abbreviations O2 capacitance - C Hb hemoglobin concentration - D diffusing capacity - PO2 medium-blood PO2 difference - Hb flow, hemoglobin delivery rate - Hepes N-[2-Hydroxyethyl]piperacine-N-[2 ethanesulfonic acid] - L diff extent of diffusion limitation - MO2 oxygen uptake rate - PO2 oxygen tension - S O2 saturation  相似文献   

3.
Germlings of Phytophthora palmivora possess at least two systems for the uptake of inorganic phosphate (Pi). The first is synthesized on germination in medium containing 50 M Pi and has a Km of approx. 30 M (Vmax=7–9 nmol Pi/h·106 cells). The second is synthesized under conditions of Pi-deprivation and has a higher affinity for Pi (Km=1–2 M), but a lower Vmax (0.5–2 nmol Pi/h·106 cells). The fungicide phosphite likewise enters the germlings via two different transport systems, the synthesis of which also depends on the concentration of Pi in the medium. The Km of the lower affinity system is 3 mM (Vmax=20 nmol phosphite/h·106 cells) and that of the higher affinity system is 0.6 mM (Vmax=12 nmol/h·106 cells). Pi and phosphite are competitive inhibitors for each other's transport in both systems. However, whereas mM concentrations of phosphite are necessary to inhibit Pi transport, only M concentrations of Pi are required to inhibit phosphite transport. A third system of uptake for Pi also exists, since when phosphate-deprived cells are presented with mM concentrations of Pi, they transport the anion at a very high rate (around 100 nmol/h·106 cells). High rates of transport of phosphite are also observed when these cells are presented with mM concentrations of this anion.  相似文献   

4.
To obtain insight into the relation between the release of heart-type fatty acid-binding protein (H-FABPc) and of long-chain fatty acids (FA) from injured cardiac tissue, rat hearts were Langendorff perfused according to the following scheme: 30 min normoxia, 60 min ischemia, 30 min reperfusion, 10 min Ca2+ free perfusion and finally 10 min Ca2+ repletion. During this protocol right ventricular (Q rv ) and interstitial effluent samples (Q i ) were collected at regular intervals. During reperfusion a total of 0.8±0.1 nmol H-FABPc but no FA were detected in the effluents. However, during Ca2+ readmission, 45±4 nmol H-FABPc (80–90% of total tissue content) was released with an initial (first 3 min) simultaneous release of FA (FA/H-FABPc ratio 0.90±0.07 mol/mol). Thereafter, FA release continued at 10–15 nmol per min mainly inQ rv while the rate of H-FABPc release decreased. During Ca2+ repletion, tissue FA content raised rapidly from 168±20 to 1918±107 nmol/g dry weight. These findings suggest that after severe cardiac damage initially FA is released bound to H-FABPc, whereas further FA release occurs in a non-protein bound manner.  相似文献   

5.
Bovine chromaffin secretory granules were purified by isopycnic Metrizamide gradient centrifugation and their Ca2+ sequestration pathways were characterized. The rate of Ca2+ sequestration at 37°C was first order, with a maximal uptake of 26.9 ±0.46 (mean ± S.D., n = 3) nmol Ca2+/mg protein and a first order rate constant (k) of 0.046 ± 0.002 min–1. At 4°C the rate of uptake was substantially attenuated, with only 2.47 ± 0.2 (mean ± S.D, n = 3) nmol Ca2+/mg protein sequestered in 60 min. Ca2+ sequestration was 93% inhibited by 180 mM NaCl [I50% of 78.7 ± 9.3 mM NaCl (mean ± S.D., n = 11)] but only slightly inhibited by KCl or MgCl2. Ca 2+ sequestration was not stimulated by incubation with MgATP but was inhibited by 57% after incubation with 30 M monensin. Ca 2+ sequestration was dependent on extravesicular Ca 2+ with half-maximal sequestration at pCa2+ 6.81 ± 0.028 (mean ± S.D., n = 3). Sequestered Ca2+ could be exchanged with external 45Ca2+, the exchange rate was first order (k of 0.042 ± 0.004: mean ± S.D., n = 3) and saturated at 27.7 ± 1.1 nmol Ca2+/mg (mean ± S.D., n = 3). The Ca2+/Ca2+ exchange system was totally inhibited by NaCl or KCl but only slightly by MgCl2. About 75% of sequestered 45Ca2+ could be released by incubation with NaCl, but only 8% was released by incubation with KCI. Half-maximal release of sequestered 45Ca2+ required 69.3 ± 12.2 mM NaCl (mean ± S.D., n = 3). The Na+-induced release of sequestered 45Ca2+ was rapid, t0.5 of 2.80 ± 0.63 min (mean ± S.D., n = 3) and inhibited at 4°C. The concurrent incubation of chromaffin granules with 45Ca2+ and either annexin proteins V or VI resulted in attenuated uptake of 45Ca2+. These results suggest that Ca2+ uptake in adrenal chromaffin granules is regulated by Na+ and Ca2+ gradients and also possibly by annexins V and VI.Abbreviations EGTA ethylene glycol bis (-aminoethyl ether)-N,-N,N,N-tetraacetic acid - SDS Sodium dodecyl sulphate - PAGE Polyacrylamide gel electrophoresis - BSA bovine serum albumin - AI Annexin I - AIIt Annexin II tetramer - AIII Annexin III - AIV Annexin IV - AV Annexin V - AVI Annexin VI - k first order rate constant - AT total extent of Ca2+ uptake (nmol) - BufferA 300 mM sucrose, 10 mM potassium phosphate (pH 7.0), 5 mM EGTA - Buffer B 300 mM sucrose, 10 mM potassium phosphate (pH 7.0) and 1 mM EGTA - Buffer C 300 mM sucrose, 10 mM potassium phosphate (pH 7.0) - Buffer D 300 mM sucrose, 10 mM potassium phosphate (pH 7.0), 0.5 mM EGTA and 0.65 MM CaCl2 - Buffer E 300 mM sucrose, 10 mM potassium phosphate (pH 7.0), 0.25 mM EGTA and 0.325 mM CaCl2  相似文献   

6.
Summary The decreasing effect of -adrenergic blockade on skin resistance to vapor diffusion and the onset of cutaneous water evaporation in the pigeon (Columba livia) was investigated. Oral administration of 1, 2.3 and 5 mg propranolol to pigeons (268±53 g) initiated intensive trans-cutaneous water evaporation (CWE) up to 29.1 mg H2O·cm–2·h–1 in resting birds at 30°C air temperature (Ta), but had only a slight effect on CWE of birds exposed to 50 °C Ta.After 7 h of effective -adrenergic blockade (oral administration of 5 mg propranolol), skin and body temperature stabilized at 39.0±0.5 °C and 41.0±0.7 °C, compared to 40.2±0.8 °C and 41.9±0.6 °C in the control group, respectively. A slight hypothermia was accompanied by feather fluffing.Intradermal injection of 0.001, 0.01 and 0.12 mg propranolol also caused intensive CWE. Local -adrenergic blockade in relatively low blocker doses (0.001 and 0.01 mg propranolol) decreased skin resistance from a high value of 44.5 s·cm–1 to about 6.0 s·cm–1, and caused a sharp increase in CWE from a control value of about 4 to a high of 26.4 mg H2O·cm–2·h–1 during the first two hours of exposure to 30°C Ta.The possible role of -adrenergic blockade in regulation of trans-cutaneous water evaporation of latent heat dissipation is discussed.  相似文献   

7.
Addition of luteinizing hormone releasing hormone (LHRH) in vitro (10–5–5×10–9 M) to murine pituitary membranes resulted in a dose-related decrease in Ca2+-ATPase activity within 15 min. Inhibitory effects of LHRH (10–7 M) occurred after 90 sec, and appeared maximal by 120 sec. Eadie-Hofstee analysis at 10–7 M LHRH, at varying [Ca2+]free, resulted in aK m=0.89±0.06 M and aV max=18.8±0.71 nmol/mg per 2 min, compared to aK m=0.69±0.06 M and aV max=32.8±1.21 nmol/mg per 2 min for controls. Pre-incubation for 5 min with LHRH antagonist (10–8 M) significantly attenuated (50%) the inhibitory effects of 10–7 M LHRH on pituitary Ca2+ ATPase activity with aK m=0.97±0.24 M and aV max=28.1±2.8 nmol/mg per 2 min. The addition of LHRH (10–7 M) to pituitary homogenates significantly increased luteinizing hormone (LH) release already at 10 and up to 40 sec compared to basal LH release. Systemic administration of 50 ng LHRH (i.p.), significantly (P<0.05) reduced pituitary Ca2+-ATPase after 30, 60 and 90 min, with a return to control levels by 120 min. Pituitary LH content was reduced slightly at 15 min, but was increased significantly at 90 and 120 min post-treatment. Plasma LH levels were elevated by 5 min, reached a peak by 15 min and returned to control within 60 min. The present findings indicate that LHRH receptor activation may influence cytosolic Ca2+ transport through effects on membrane Ca2+-ATPase activity. These actions may regulate LHRH-induced synthesis, storage and release of LH from pituitary gonadotropes.  相似文献   

8.
Post-ischemic reperfusion causes cardiac dysfunction and radical-induced lipid peroxidation (LPO) detectable by ESR spin trapping. This study deals with the applicability of the spin trap technique to pharmacological investigations during myocardial reperfusion injury. The use of the spin trap phenylbutylnitrone (PBN, 3 mM) in isolated rat hearts demonstrated the release of alkoxyl radicals (aN = 1.39 mT, aH = 0.19 mT) formed particularly within the first 15 min of reperfusion following 30 min of ischemia. The decline of radicals, after 10 min of reperfusion, was accompanied by recovery of function in 80% of the hearts. The radical concentration in the coronary effluent (maximum after 7.5 min) was reduced by the infusion of 1 mM mercaptopropionylglycine (MPG, 2.7 ± 0.5 U/ml, p < 0.001) or 5 M vitamin E (11.7 ± 0.8 U/ml, p < 0.001), compared to the (PBN-containing) control (29.7 ± 4.3 U/ml). Moreover, functional recovery (left ventricular developed pressure, LVDP 91.6 ± 20% of pre-ischemic level, p < 0.05) was improved by the hydrophilic radical scavenger MPG, compared to the (PBN-containing) control (LVDP 50.5 ± 15.7% of baseline). PBN alone led to higher functional recovery (p < 0.05) and reduced VF (duration of ventricular fibrillation; 7.10 ± 0.36 min/30 min, p < 0.05), compared to the untreated (PBN-free) control (LVDP 26.6 ± 11.8%; VF 19.42 ± 3.64 min/30 min). The Ca antagonist verapamil (0.1 M), MPG, and the lipophilic vitamin E showed cardioprotection in the absence of PBN: post-ischemic recovery of LVDP was 25.4 ± 6.8% (p < 0.05), 39.6 ± 12.7% (p < 0.05) and 52.4 ± 2.6% (p < 0.01), respectively, compared to the corresponding untreated control (13.3 ± 6.6%). Whereas verapamil and vitamin E were able to protect the heart when present alone, they offered no additive effect in the presence of PBN. Therefore, PBN can be used to estimate the radical scavenger properties of an agent in the heart. However, because of the protective properties of PBN itself, the results of simultaneous investigations of the effects of other compounds, such as Ca antagonists or lipophilic radical scavengers, on heart function may be limited.  相似文献   

9.
Summary In vivo electrolyte transport and water absorption from the caeca of dehydrated, low-NaCl diet hens are reported. In the absence of luminal glucose or acetate, net electrolyte transport rates and water absorption are small. When physiological concentrations of acetate (40 mM) are included in the perfusate, Na+ transport and water absorption increase significantly (P<0.01): 38±7 eqNa+/caecum kg·h and 256±33 l H2O/caecum · kg · h.A similar increase in water absorption occurs with the inclusion of 15 mM glucose in the perfusate (219±30 l H2O/caecum · kg · h), however both net Na+ and Cl absorption increase: 28±6 eq Na+/caecum · kg · h and 21±5 eq Cl/caecum kg · h.These pronounced increases in electrolyte and water absorption are not accompanied by any significant increase in transmural potential difference.The data presented establish caeca as important sites in the recuperation of water and electrolytes in dehydrated, low-NaCl diet hens.Abbreviations ECPD electrochemical potential difference - PD (transmural) potential difference - PEG polyethylene glycol  相似文献   

10.
Summary Vasopressin-induced transformation of ridges to microvilli on the surface of granular cells of toad urinary bladder occurs in conjunction with induced alterations in the water permeability of the luminal membrane. This study was designed to establish the relationship between the time course for induction of microvilli and the time course for induction of increased water permeability after vasopressin stimulation. Hemibladders were examined at 2.5, 5, 10, 20 and 30 min following exposure to 20 mU/ml of vasopressin and at 5, 10, 20, 30, 40, 50 and 60 min after washout of vasopressin. Within 2.5 min, vasopressin initiated complete transformation of ridges to microvilli on approximately 13% of the granular cells, while osmotic water flow (Jv) was 0.31±0.10 l·min–1·cm–2. Five minutes following vasopressin stimulation, microvilli were present on approximately 30% of granular cells andJv was 2.27±0.13 l·min–1·cm–2. At 10 minJv was maximum at 4.03±0.15 l·min–1·cm–2 and 50% of the granular cells were covered with microvilli. This percentage increased to 70% at 20 min and was maintained at 30 min, althoughJv decreased to 3.9±0.35 l·min–1·cm–2 at 30 min. Five minutes following vasopressin washout, ridges interspersed with microvilli reappeared asJv fell to 1.10±0.30 l·min–1·cm–2. At 10 min after vasopressin washout,Jv approached basal levels, but the reversal of microvilli to ridges remained incomplete. At 60 min after vasopressin washout, the granular cells had regained their original ridgelike surface structures. Thus, these studies establish a temporal relationship between the induction and reversibility of vasopressin-induced microvillous formation and alterations in the osmotic water permeability of the apical plasmalemma.  相似文献   

11.
Gallbladder Na+ absorption is linked to gallstone formation in prairie dogs. We previously reported Na+/H+ exchanger (NHE1-3) expression in native gallbladder tissues. Here we report the functional characterization of NHE1, NHE2 and NHE3 in primary cultures of prairie dog gallbladder epithelial cells (GBECs). Immunohistochemical studies showed that GBECs grown to confluency are homogeneous epithelial cells of gastrointestinal origin. Electron microscopic analysis of GBECs demonstrated that the cells form polarized monolayers characterized by tight junctions and apical microvilli. GBECs grown on Snapwells exhibited polarity and developed transepithelial short-circuit current, Isc, (11.6 ± 0.5 µA · cm–2), potential differences, Vt (2.1 ± 0.2 mV), and resistance, Rt (169 ± 12 · cm2). NHE activity in GBECs assessed by measuring dimethylamiloride-inhibitable 22Na+ uptake under a H+ gradient was the same whether grown on permeable Snapwells or plastic wells. The basal rate of 22Na+ uptake was 21.4 ± 1.3 nmol · mg prot–1 · min–1, of which 9.5 ± 0.7 (~45%) was mediated through apically-restricted NHE. Selective inhibition with HOE-694 revealed that NHE1, NHE2 and NHE3 accounted for ~6%, ~66% and ~28% of GBECs total NHE activity, respectively. GBECs exhibited saturable NHE kinetics (Vmax 9.2 ± 0.3 nmol · mg prot–1 · min–1; Km 11.4 ± 1.4 mM Na+). Expression of NHE1, NHE2 and NHE3 mRNAs was confirmed by RT-PCR analysis. These results demonstrate that the primary cultures of GBECs exhibit Na+ transport characteristics similar to native gallbladder tissues, suggesting that these cells can be used as a tool for studying the mechanisms of gallbladder ion transport both under physiologic conditions and during gallstone formation.  相似文献   

12.
d-Ribulose-1,5-bisphosphate carboxylase/oxygenase has been purified 80-fold from malate-grown Thiocapsa roseopersicina by salting out the enzyme from the high-speed supernatant between 68–95% saturation with respect to (NH4)2SO4, gelfiltration through Sephadex G-100, and DEAE-cellulose chromatography followed by sedimentation into a 14–34% glycerol gradient. The specific activity of enzyme for the carboxylase reaction was 2.45 mol RuBP-dependent CO2 fixed/min · mg protein (at pH 8.0 and 30° C) and for the oxygenase reaction was 0.23 mol RuBP-dependent O2 consumed/min · mg protein (at pH 8.6, and 25° C). The enzyme, which was ultracentrifugally homogeneous in the presence of 4 and 10% v/v glycerol, was stable for at least one year at-80° C in the presence of 10% glycerol. S20, w values obtained in the presence of 4 and 10% glycerol were 19.3 and 16.2, respectively. The enzyme contained both large (53,000-daltons) and mixed small subunits (15,000- and 13,500-daltons).Borate-dependent inactivation of the enzyme by 2,3-butadione, which was greatly reduced in the presence of the product 3-phosphoglycerate, suggested that one or more arginines are at the active site.Abbreviations DTT dithiotreitol - RuBP d-ribulose-1,5-bisphosphate - SDS sodium dodecylsulfate - TCA trichloroacetic acid - TEMBDG buffer (pH 8.0 at 25°C) containing 20 mM Tris, 1 mM disodium EDTA · 2 H2O, 10 mM MgCl2·6 H2O, 50 mM NaHCO3, 0.1 mM DTT and 10% glycerol (v/v)  相似文献   

13.
Endogenous and maximum respiration rates of nine purple sulfur bacterial strains were determined. Endogenous rates were below 10 nmol O2 · (mg protein · min)-1 for sulfur-free cells and 15–35 nmol O2 · (mg protein · min)-1 for cells containg intracellular sulfur globules. With sulfide as electron-donating substrate respiration rates were considerably higher than with thiosulfate. Maximum respiration rates of Thiocystis violacea 2711 and Thiorhodovibrio winogradskyi SSP1 (254.8 and 264.2 nmol O2 · (mg protein · min)-1, respectively) are similar to those of aerobic bacteria. Biphasic respiration curves were obtained for sulfur-free cells of Thiocystis violacea 2711 and Chromatium vinosum 2811. In Thiocystis violacea the rapid and incomplete oxidation of thiosulfate was five times faster than the oxidation of stored sulfur. A high affinity of the respiratoty system for oxygen (K m =0.3–0.9 M O2, V max=260 nmol O2 · (mg protein · min)-1 with sulfide as substrate, K m =0.6–2.4 M O2, V max=14–40 nmol O2 · (mg protein · min)-1 with thiosulfate as substrate), for sulfide (K m =0.47 M, V max=650 nmol H2S · (mg protein × min)-1, and for thiosulfate (K m =5–6 M, V max =24–72 nmol S2O 3 2- · (mg protein · min)-1 was obtained for different strains. Respiration of Thiocystis violacea was inhibited by very low concentrations of NaCN (K i =1.7 M) while CO concentrations of up to 300 M were not inhibitory. The capacity for chemotrophic growth of six species was studied in continuous culture at oxygen concentrations of 11 to 67 M. Thiocystis violacea 2711, Amoebobacter roseus 6611, Thiocapsa roseopersicina 6311 and Thiorhodovibrio winogradskyi SSP1 were able to grow chemotrophically with thiosulfate/acetate or sulfide/acetate. Chromatium vinosum 2811 and Amoebobacter purpureus ML1 failed to grow under these conditions. During shift from phototrophic to chemotrophic conditions intracellular sulfur and carbohydrate accumulated transiently inside the cells. During chemotrophic growth bacteriochlorophyll a was below the detection limit.  相似文献   

14.
The present study tests the hypothesis that hypoxia alters the high-affinity kainate receptors in fetal guinea pig brain. Experiments were conducted in normoxic and hypoxic guinea pig fetus at preterm (45 days of gestation) and term (60 days of gestation). Hypoxia in the guinea pig fetus was induced by exposure to maternal hypoxia (FiO2=7%) for 60 min. Brain tissue hypoxia in the fetus was documented biochemically by decreased levels of ATP and phosphorreatine. [3H]-Kainate binding characteristics (Bmax=number of receptors, Kd=dissociation constant) were used as indices of kainate receptor modification. P2 membrane fractions were prepared from the cortex of normoxic and hypoxic fetuses and were washed six times prior to performing the binding assays. [3H]kainate binding was performed at 0°C for 30 min in a 500 l medium containing 50 mM Tris-HCl buffer, 0.1 mM EDTA (pH 7.4), 300 g protein and varying concentrations of radiolabelled kainate ranging from 1 to 200 nM. Non-specific binding was determined in the presence of 1.0 mM glutamate. During brain development from 45 to 60 days gestation, Bmax value increased from 330±16 to 417±10 fmoles/mg protein; however, the Kd was unchanged (8.2±0.4 vs 8.8±0.5 nM, respectively). During hypoxia at 60 days, the Kd value significantly increased as compared to normoxic control (15.5±0.7 vs 8.8±0.5 nM, respectively), whereas the Bmax was not affected (435±12 vs 417±10 fmol/mg protein, respectively). At 45 days, hypoxia also increased the Kd (11.9±0.6 vs 8.2±0.4 nM) without affecting the Bmax (290±15 vs 330±16 fmol/mg protein, respectively). The results show that the number of kainate receptors increase during gestation without change in affinity and demonstrate that hypoxia modifies the high-affinity kainate receptor sites at both ages; however the effect is much stronger at 60 days (term). The decreased affinity of the site could decrease the kainate receptor-mediated fast kinetics of desensitization and provide a longer period for increased Na+-influx, leading to increased accumulation of intracellular Ca2+ by reversal of the Na+–Ca2+ exchange mechanism. In addition, Kd values for kainate-type glutamate receptor sites are 30–40 fold lower (i.e. higher affinity) than those for NMDA-displaceable glutamate sites. The higher affinity suggests that the activation of the kainate-type glutamate receptor during hypoxia could precede initiation of NMDA receptormediated excitotoxic mechanisms. We propose that hypoxia-induced modification of the high affinity kainate receptor in the fetus is a potential mechanism of neuroexcitotoxicity.  相似文献   

15.
Michael R. Blatt 《Planta》1987,170(2):272-287
The membrane electrical characteristics of stomatal guard cells in epidermal strips from Vicia faba L. and Commelina communis L. were explored using conventional electrophysiological methods, but with double-barrelled microelectrodes containing dilute electrolyte solutions. When electrodes were filled with the customary 1–3 M KCl solutions, membrane potentials and resistances were low, typically decaying over 2–5 min to near-30 mV and <0.2 k·cm2 in cells bathed in 0.1 mM KCl and 1 mM Ca2+, pH 7.4. By contrast, cells impaled with electrodes containing 50 or 200 mM K+-acetate gave values of-182±7 mV and 16±2 k·cm2 (input resistances 0.8–3.1 G, n=54). Potentials as high as (-) 282 mV (inside negative) were recorded, and impalement were held for up to 2 h without appreciable decline in either membrane parameter. Comparison of results obtained with several electrolytes indicated that Cl- leakage from the microelectrode was primarily responsible for the decline in potential and resistance recorded with the molar KCl electrolytes. Guard cells loaded with salt from the electrodes also acquired marked potential and conductance responses to external Ca2+, which are tentatively ascribed to a K+ conductance (channel) at the guard cell plasma membrane.Measurements using dilute K+-acetate-filled electrodes revealed, in the guard cells, electrical properties common to plant and fungal cell membranes. The cells showed a high selectivity for K+ over Na+ (permeability ratio PNa/PK=0.006) and a near-Nernstian potential response to external pH over the range 4.5–7.4 (apparent PH/PK=500–600). Little response to external Ca2+ was observed, and the cells were virtually insensitive to CO2. These results are discussed in the context of primary, charge-carrying transport at the guard cell plasma membrane, and with reference to possible mechanisms for K+ transport during stomatal movements. They discount previous notions of Ca2+-and CO2-mediated transport control. It is argued, also, that passive (diffusional) mechanisms are unlikely to contribute to K+ uptake during stomatal opening, despite membrane potentials which, under certain, well-defined conditions, lie negative of the potassium equilibrium potential likely prevailing.Abbreviations and symbols EGTA ethylene glycol-bis(-aminoethyl ether)-N,N,N,N-tetraacetic acid - Hepes 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid - Mes 2-(N-morpholino) propanesulfornic acid - E equilibrium potential - Gm membrane conductance - Rin input resistance - Vm membrane potential  相似文献   

16.
Summary Isolated gills of the freshwater mussel,Ligumia subrostrata, accumulate Na from a pondwater bathing medium. The rate of Na transport by the isolated gill is 13.2±1.1 mol (g dry gill·10 min)–1 which equals or exceeds the estimated Na transport rate of intact animals. Sodium influx is saturable with aV max of 13.6±1.2 mol (g dry gill·10 min)–1 and an affinity (K s) of 0.17 mM Na/l. The isolated gills survive prolonged exposure to pondwater with a constant of 890 l O2 (g dry gill·h)–1 over a 4 h period. Sodium transport in the isolated gills is stimulated 80% above control values by 10–4 M serotonin, 60% by 0.5 mM cAMP and 60% by 12.5 g/ml nystatin. Sodium influx is inhibited by 0.5 mM amiloride and 1 mM lithium.  相似文献   

17.
The release of preloaded [3H]taurine from cultured cerebral cortical astrocytes was studied under various cell-damaging conditions, including hypoxia, ischemia, aglycemia and oxidative stress, and in the presence of free radicals. Astrocytic taurine release was enhanced by K+ (50 mM), veratridine (0.1 mM) and the ionotropic glutamate receptor agonist kainate (1.0 mM). Metabotropic glutamate receptor agonists had only weak effects on taurine release. Similarly to the swelling-induced taurine release the efflux in normoxia seems to be mediated mainly by DIDS-(diisothiocyanostilbene-2,2-disulphonate) and SITS-(4-acetamido-4-isothiocyanostilbene-2,2-disulphonate) sensitive CI channels, since these blockers were able to reduce both basal and K+ -stimulated release. The basal release of taurine was moderately enhanced in hypoxia and ischemia, whereas the potentiation in the presence of free radicals was marked. The small basal release from astrocytes signifies that taurine release from brain tissue in ischemia may originate from neurons rather than glial cells. On the other hand, the release evoked by K+ in hypoxia and ischemia was greater than in normoxia, with a very slow time-course. The enhanced release of the inhibitory amino acid taurine from astrocytes in ischemia may be beneficial to surrounding neurons, outlasting the initial stimulus and counteracting overexcitation.  相似文献   

18.
Transient glucose deprivation of the heart [GLU (-)] confers a preconditioning-like protection against subsequent ischemic/reperfusion (I/R). The mechanisms involved remain unclear. We hypothesized that GLU (-) would induce the classic ischemic preconditioning activated signaling cascade. Potential metabolic consequences and putative cell signaling events induced by transient glucose deprivation were evaluated as candidate mediators of this cardioprotection.Isolated glucose-perfused rat hearts were subjected to 30 min global ischemia followed by 30 min reperfusion (index I/R). Cardiac contractile recovery following I/R was used as the functional end-point in these studies. Metabolic preconditioning was stimulated by 15 min GLU (-) followed by 10 min glucose repletion prior to the index I/R. The potential metabolic consequences of GLU (-) were evaluated by using excess octanoate (11 mM OCT Hi) or 11 mM 2-deoxy-D-glucose (2-DG) in place of GLU (-) and by combining GLU (-) with fuels known to inhibit glycolysis supply (20 mM pyruvate or 1 mM octanoate, OCT Lo). The roles of -adrenoceptors, -adrenoceptors, adenosine receptors, protein kinase C (PKC) and mitochondrial KATP channels were investigated using inhibitors prazosin (10 M), propranolol (10 M), 8-(p-sulfophenyl) theophylline, (SPT 100 M), chelerythrine (CHEL 10 M) and 5-hydroxydecanoate (5 HD 100 M) respectively.GLU (-) increased mechanical recovery (59.8 ± 4.0 vs. 32.3 ± 4.7%; p < 0.01). Protection was abolished by pyruvate 26.6 ± 3.1; SPT 36.6 ± 3.0; CHEL 35 ± 4.8 or 5 HD 23.8 ± 3.3%. In a separate set of experiments, the specificity of SPT in this model was tested by preconditioning with adenosine (100 M) (34.7 ± 4 vs. control 16.8 ± 1.3%, p = 0.01) and blocking this protection with the same dose of SPT (16.3 ± 1.5%) used in the GLU (-) studies. Protection was unaltered by prazosin (50.2 ± 3.3%), propranolol (55.5 ± 4.0%), or OCT Lo (50.2 ± 2.5%). Protection was not mimicked by OCT Hi (35.6 ± 3.8%) or 2-DG (34 ± 4.3%).Transient glucose deprivation does not seem to achieve preconditioning-like cardioprotection by decreased glycolysis. Rather, the signal system may involve enhanced adenosine release, PKC, and activation of the mitochondrial KATP channel.  相似文献   

19.
Summary A quantitative histochemical method to determine the Km and Vmax of -glucosidases in the intestinal epithelium without disruption of the cellular structure is described. 2-Naphthyl--D-glucoside was used as substrate and hexazonium-p-rosaniline as coupling agent. Using a Leitz MPV2 microdensitometer and a field measuring 4×4 m, and reading the test samples against a blank focused on the lamina propria, we observed that the intensity of the colour was a linear function of both the incubation time up to 20 min, and the thickness of the slice up to 20 m. The ratio between the extinction at the absorption maximum and at a second wavelength was constant, whatever the intensity of the colour.By determining the relationship between the extinction and the substrate concentration under standard conditions (slice thickness of 10 m and incubation time of 10 min), we obtained a saturation curve described by a Km of 0.68±0.038 mM and a Vmax of 1.41±0.039 A480·10–2·m–1·min–1. When the hydrolysis of the same substrate by a homogenate of jejunal mucosa was examined biochemically under comparable conditions, a Km of 0.64±0.012 mM and a Vmax of 57.3±0.70 mU/mg protein were obtained. When the natural substrate, sucrose, was used in the biochemical study, a Km of 15±3.5 mM and a Vmax of 149±24.7 mU/mg protein were obtained.These experiments demonstrate that the kinetic constants of enzyme reactions can be assessed with equal accuracy on histochemical sections as in tissue homogenates.  相似文献   

20.
The effect of severe acute hypoxia (fractional concentration of inspired oxygen equalled 0.104) was studied in nine male subjects performing an incremental exercise test. For power outputs over 125 W, all the subjects in a state of hypoxia showed a decrease in oxygen consumption ( O2) relative to exercise intensity compared with normoxia (P < 0.05). This would suggest an increased anaerobic metabolism as an energy source during hypoxic exercise. During submaximal exercise, for a given O2, higher blood lactate concentrations were found in hypoxia than in normoxia (P < 0.05). In consequence, the onset of blood lactate accumulation (OBLA) was shifted to a lower O2 ( O2 1.77 l·min–1 in hypoxia vs 3.10 l·min–1 in normoxia). Lactate concentration increases relative to minute ventilation ( E) responses were significantly higher during hypoxia than in normoxia (P < 0.05). At OBLA, E during hypoxia was 25% lower than in the normoxic test. This study would suggest that in hypoxia subjects are able to use an increased anaerobic metabolism to maintain exercise performance.  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