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 495 毫秒
1.
Dose-response studies were performed in 6 human volunteer subjects to determine the threshold and optimal doses of intravenous bombesin for stimulation of gastric acid secretion and gastrin release. A significant stimulation of both acid and gastrin was obtained with a very low dose, 3 pmol · kg?1 · h?1. Peak stimulation of acid secretion (67% of pentagastrin PAO) was obtained at 12.5 pmol · kg?1 · h?1. Serum gastrin response to this dose of bombesinn was similar to that obtained after a high protein meal. Higher doses of bombesin caused further increases in serum gastrin but not in acid secretion. Since very low doses of bombesin, too small to produce detectable increases in immunoreactive serum bombesim, caused parallel increases in gastrin and acid secretion, it is possible that the bombesin-like peptides present in human gastrointestinal tissues contribute to regulation of human gastric secretion.  相似文献   

2.
Maximum quantum yield (φmB) and maximum photosynthetic rate (PmB) of light-saturation curves of phytoplankton photosynthesis were determined for nannoplankton (< 20 μm) and netplankton (>20 μm) from the subsurface chlorophyll-maximum layer at 14 stations in the tropical North Pacific Ocean in the spring of 1976. The maximum quantum yield mB ± s.e.) was significantly higher for nannoplankton (0.056 ± 0.006 moles CO2·Einstein?1 absorbed) than netplankton (0.039 ± 0.002 moles CO2·Einstein?1 absorbed). The importance of nannoplankton in the maximum photosynthetic rate (PmB) appears to be less consistent. At least 60% of the theoretical maximum quantum yield (0.12 moles CO2·Einstein?1 absorbed) was probably incorporated into the particulate fraction at the subsurface chlorophyll-maximum layer.  相似文献   

3.
Blood glucose, gastric inhibitory polypeptide (GIP), vasoactive intestinal polypeptide (VIP) and gastrin secretions were measured over a three-hour period following the ingestion by normal subjects of a mixed meal with two different caloric levels (1055 Kcal and 1192 Kcal). No VIP secretion was observed after either meal. Gastrin release was not modified by the increase of caloric intake (mainly carbohydrates and lipids), whereas GIP secretion was significantly more important after the meal with the highest caloric value (peak at 30 mn: 499.5±250.4 vs. 273.4±128.7 pg/ml and integrated response 53.3±20.5 vs. 28.2±9.9 ng×ml?1×180 min?1?p<0.05). This difference could not be attributed to glucose since the blood glucose levels were not significantly different. It is more probably related to the total amount of ingested food. This suggests the existence of rapid mechanisms of adaptation to the incoming load of the GIP-producing cells.  相似文献   

4.
The effect of intravenously administered calcitonin and secretin on bombesin-stimulated serum gastrin and gastric acid secretion was studied in 7 volunteers. Secretin G.I.H. (1 C.U./kg per h) and calcitonin (0.5 I.U./kg per h) significantly (P < 0.05) inhibited the serum gastrin and gastric acid responses to bombesin-14 (90 pmol/kg per h). Inhibition of gastrin release could not fully account for the inhibition of gastric acid secretion.  相似文献   

5.
The binding of chlorpromazine · HCl at the human erythrocyte surface has been detected through its effect on cellular electrophoretic mobility. Incubation of erythrocytes (approx. 5 · 106/ml) in 23 μM chlorpromazine · HCl resulted in a reduction of negative electrophoretic mobility from the control value of ?1.11 ± 0.01 (μm · s?1)/(V · cm?1) to ?1.00 ± 0.02 (μm · s?1)/(V · cm?1) (pH 7.2, ionic strength 0.155). This mobility change was completely reversed when chlorpromazine · HCl was removed by centrifugal washing. Increasing the drug concentration to 70μM did not affect the mobility change, indicating saturation of the electrophoretically detectable drug binding sites over chlorpromazine · HCl concentration range studied here. The effect of the 23 μM chlorpromazine · HCl on electrophoretic mobility was also measured in isotonic media of reduced ionic strength. The drug-induced reduction in negative surface charge density was found to be independent of ionic strength over the range 0.155 (Debye length, 0.8 nm) to 0.00310 (Debye length, 5.7 nm).Fixation of erythrocytes with glutaraldehyde affected neither the normal electrophoretic mobility of discocytes nor the reduced electrophoretic mobility of chlorpromazine · HCl-induced stomatocytes. When these stomatocytes were first fixed with glutaraldehyde, then washed free of chlorpromazine · HCl, they retained the stomatocyte form while regaining a normal control electrophoretic mobility. Conversely, when discocytes fixed in that form were treated with chlorpromazine · HCl, they showed the same mobility change as did fixed or unfixed stomatocytes. The drug-induced mobility change is therefore independent of the shape change, but reflects a contribution to cellular surface charge density from the membrane-bound chlorpromazine · HCl molecules. From the charge reduction, it is estimated that about 106 chlorpromazine · HCl molecules are bound at the electrokinetic cell surface and occupy approximately 0.4% of the total surface area.  相似文献   

6.
Robert F. Anderson 《BBA》1983,723(1):78-82
The bimolecular decay rates (2k) of the flavosemiquinones (FH·F?) of riboflavin, FMN and FAD have been determined using pulse radiolysis. The rates (defined as d[FH·F?]dt = ?2k[FH·F?]2) for the neutral flavosemiquinones at zero ionic strength and pH 5.9 are (in units of mol?1·dm3·s?1): (1.2 ± 0.1)·109, (5.0 ± 0.2)·108 and (1.4 ± 0.1)·108; and for the anionic flavosemiquinones at pH 11.2 (5.4 ± 0.9)·108, (4.5 ± 0.3)·107 and (8.5 ± 1.3)·106, respectively. The kinetic salt effect has been used to formulate rate equations for each flavin to adjust for ionic strength effects.  相似文献   

7.
This study is the first to examine the circadian rhythms of melatonin in Eriocheir sinensis and Palaemonetes sinensis, two economically important crustaceans. We collected haemolymph and optic lobes from both species every 4 h over a whole day cycle. Melatonin content was measured with high-performance liquid chromatography. E. sinensis haemolymph exhibited significant (p < 0.05) peaks in melatonin at 16:00 (0.180 ± 0.020 μg·mL?1) and 24:00 (0.244 ± 0.055 μg·mL?1), while eyestalks had significant peaks at 16:00 (72.377 ± 18.100 μg·eyestalk?1) and 24:00 (98.756 ± 30.271 μg·eyestalk?1). In P. sinensis, melatonin peaked significantly only at 16:00 in optic lobes (12.493 ± 1.475 μg·eyestalk?1) (p < 0.05); no significant peaks were present in haemolymph. Thus, both E. sinensis and P. sinensis exhibit species-specific melatonin rhythms. Time of day should therefore be considered when examining the physiological status of both crustaceans, given the potential influence of fluctuating daily melatonin concentrations.  相似文献   

8.
Leakage of the entrapped anionic fluorophore carboxyfluorescein was used as a measure of the permeability of liposomes to several different acids. Carboxyfluorescein leakage increased with increasing buffer concentration at a given pH and depended on its chemical nature: apolar weak acids such as acetic or pyruvic acids induced fast leakage at relatively high pH (4 to 5), while glycine, aspartic, citric and hydrochloric acids induced leakage only at lower pH. Fluorescence leakage measurements reflected the acidification of the liposomes' aqueous spaces, which was primarily caused by the diffusion of undissociated acid molecules across the lipid bilayer. A simple mathematical model in accord with this hypothesis and assuming that carboxyfluorescein leakage was directly related to the proportion of its neutral lactone form, described satisfactorily the carboxyfluorescein leakage kinetics and allowed rough estimation of permeability coefficients for carboxyfluorescein (neutral lactone form; 9 · 10?9 cm · s?1), acetic acid (>1 · 10?7cm · s?1) and glycine (cation: 6 · 10?9 cm · s?1). These results are consistent with low effective proton permeability of liposomes (<5 · 10?12cm · s?1) and with the permeability coefficient of HCl (3 · 10?3 cm · s?1) reported by Nozaki and Tanford ((1981) Proc. Natl. Acad. Sci. U.S.A. 78, 4324–4328). Diffusion of weak acid molecules across lipid membranes has implications for drug encapsulation and delivery, and may be of biological significance.  相似文献   

9.
A thermodynamic characterization of the Na+-H+ exchange system in Halobacterium halobium was carried out by evaluating the relevant phenomenological parameters derived from potential-jump measurements. The experiments were performed with sub-bacterial particles devoid of the purple membrane, in 1 M NaCl, 2 M KCl, and at pH 6.5–7.0. Jumps in either pH or pNa were brought about in the external medium, at zero electric potential difference across the membrane, and the resulting relaxation kinetics of protons and sodium flows were measured. It was found that the relaxation kinetics of the proton flow caused by a pH-jump follow a single exponential decay, and that the relaxation kinetics of both the proton and the sodium flows caused by a pNa-jump also follow single exponential decay patterns. In addition, it was found that the decay constants for the proton flow caused by a pH-jump and a pNa-jump have the same numerical value. The physical meaning of the decay constants has been elucidated in terms of the phenomenological coefficients (mobilities) and the buffering capacities of the system. The phenomenological coefficients for the Na+-H+ flows were determined as differential quantities. The value obtained for the total proton permeability through the particle membrane via all available channels, LH = (?JH +pH)Δψ,ΔpNa, was in the range of 850–1150 nmol H+·(mg protein)?1·h?1·(pH unit)?1 for four different preparations; for the total Na+ permeability, LNa = (?JNa+pNa)Δψ,ΔpH, it was 1620–2500 nmol Na+·(mg protein)?1·h?1·(pNa unit)?1; and for the proton ‘cross-permeability’, LHNa = (?JH+pNa)Δψ,ΔpH, it was 220–580 nmol H+·(mg protein)?1·h?1·(pNa unit)?1, for different preparations. From the above phenomenological parameters, the following quantities have been calculated: the degree of coupling (q), the maximal efficiency of Na+-H+ exchange (ηmax), the flow and force efficacies (?) of the above exchange, and the admissible range for the values of the molecular stoichiometry parameter (r). We found q ? 0.4; ηmax ? 5%; 0.36 ? r ? 2; ?JNa+ ? 1.3 · 105μmol · (RT unit)?1 at JNa = 1 μmolNa+ · (mgprotein)?1 · h?1; and ?ΔpNa ? 5 · 104 ΔpNa · (mg protein) · h · (RT unit)?1 at ΔpNa = 1 unit, for different preparations.  相似文献   

10.
The transport of 3-O-methylglucose in white fat cells was measured under equilibrium exchange conditions at 3-O-methylglucose concentrations up to 50 mM with a previously described method (Vinten, J., Gliemann, J. and Østerlind, K. (1976) J. Biol. Chem. 251, 794–800). Under these conditions the main part of the transport was inhibitable by cytochalasin B. The inhibition was found to be of competitive type with an inhibition constant of about 2.5 · 10?7 M, both in the absence and in the presence of insulin (1μM). The cytochalasin B-insensitive part of the 3-O-methylglucose permeability was about 2 · 10?9 cm · s?1, and was not affected by insulin. As calculated from the maximum transport capacity, the half saturation constant and the volume/ surface ratio, the maximum permeability of the fat cell membrane to 3-O-methylglucose at 37°C and in the presence of insulin was 4.3 · 10?6 cm · s?1. From the temperature dependence of the maximum transport capacity in the interval 18–37°C and in the presence of insulin, an Arrhenius activation energy of 14.8 ± 0.44 kcal/mol was found. The corresponding value was 13.9 ± 0.89 in the absence of insulin. The half saturating concentration of 3-O-methylglucose was about 6 mM in the temperature interval used, and it was not affected by insulin, although this hormone increased the maximum transport capacity about ten-fold to 1.7 mmol · s?1 per 1 intracellular water at 37°C.  相似文献   

11.
A. Telfer  J. Barber 《BBA》1978,501(1):94-102
1. Ionophore A23187 induces uncoupling of potassium ferricyanide-dependent O2 evolution by envelope-free chloroplasts and oxaloacetate-dependent O2 evolution by intact chloroplasts. The half maximal concentration (C12) for stimulation of oxygen evolution in both cases is approximately 4 μM · 100 μg chlorophyll · ml?1.2. Ionophore A23187 also induces inhibition of CO2 and 3-phosphoglycerate-dependent O2 evolution by intact chloroplasts in the presence of 3 mM MgCl2. The half maximal concentrations (C12) for inhibition of O2 evolution are 3 μM and 5 μM respectively · 100 μg?1 chlorophyll · ml?1.3. A very high concentration of ionophore A23187 (10 μM · 20 μg?1 chlorophyll · ml?1) plus 0.1 mM EDTA lowers the fluorescence yield of intact chloroplasts suspended in a cation-free medium in the presence of 3-(3,4-dichlorophenyl)-1,1-dimethylurea, indicating loss of divalent cation from the diffuse double layers of the thylakoid membranes.4. These results are discussed in relation to ionophore A23187-induced divalent cation/proton exchange at both the thylakoid and the envelope membranes of intact chloroplasts.  相似文献   

12.
Dispersed acini from dog pancreas were used to examine the ability of dopamine to increase cyclic AMP cellular content and the binding of [3H]dopamine. Cyclic AMP accumulation caused by dopamine was detected at 1·10?8 M and was half-maximal at 7.9±3.4·10?7M. The increase at 1·10?5 M, (7.5-fold) was equal to the half-maximal increase caused by secretin at 1·10?9 M. Haloperidol, a dopaminergic receptor antagonist inhibited cyclic AMP accumulation caused by dopamine. The IC50 value for haloperidol, calculated from the inhibition of cyclic AMP increase caused by 1·10?5 M dopamine was 2.3±0.9·10?6M. Haloperidol did not alter basal or secretin-stimulated cyclic AMP content. [3H]Dopamine binding was studied on the same batch of cells as cyclic AMP accumulation. At 37°C, it was rapid, reversible, saturable and stereospecific. The Kd value for high affinity binding sites was 0.43±0.1·10?7M and 4.7±1.6·10?7M for low affinity binding sites. The concentration of drugs necessary to inhibit specific binding of dopamine by 50% was 1.2±0.4·10/t-7M noradrenaline, 2·10/t-7 M epinine, 4.1±1.8·10/t-6M fluphenazine, 8.0±1.6·10/t-6M haloperidol, 4.2±1.2·10?6Mcis-flupenthixol, 2.7±0.4·10?5Mtrans-flupenthixol, >1·10?5M apomorphine, sulpiride, naloxone and isoproterenol.  相似文献   

13.
Hemolytically active components from P. parvum and G. breve toxins   总被引:1,自引:0,他引:1  
Y S Kim  G M Padilla 《Life sciences》1977,21(9):1287-1292
Hemolytically active fractions were isolated from the toxins produced by the red-tide dinoflagellate Gymnodinium breve (GBTX) and the chrysomonad Prymnesium parvum (PPTX). High pressure liquid chromatography through bonded phase (ODS) silica columns using a gradient of methanol in chloroform yielded 6 major fractions from GBTX, 3 of which were hemolytic (HD50=0.3?0.56 μg·ml?1). None were ichthyotoxic. Of the 6 fractions obtained from PPTX, 4 were hemolytic (HD50=0.013?2.8 μg·ml?1) but only one (fraction 6) was ichthyotoxic. This fraction was ~ 2000 times more hemolytic than the crude PPTX (HD50=33.2 μg·ml?1). Analysis of their UV spectra indicates that the fractions within each group are closely related.  相似文献   

14.
The osmotic permeability coefficient (Pf) for water movement across Novikoff hepatoma cells was found to be 82 ± 3 (S.E.) · 10?5 cm · s?1 at 20°C. The corresponding diffusional permeability coefficient for 3HHO (Pd) was 97 ± 10 (S.E.) · 10?5 cm · s?1, therefore the ratio PfPd is close to unity. The apparent activation energy for water filtration was 10.4 ± 0.4 (S.E.) kcal · mol?1. This value is significantly greater than the activation energy for the self diffusion of water. The product of the hydraulic permeability coefficient and the viscosity coefficient for water was temperature-dependent. However, the product of the hydraulic permeability coefficient and the viscosity coefficient for membrane lipid did not vary with temperature. These data are interpreted as evidence for water movement across a lipid membrane barrier rather than through aqueous channels.  相似文献   

15.
The purpose of this present study was to develop a method for stimulation of acid secretion by the isolated perfused rat stomach. Rat stomachs were perfused insitu via the abdominal aorta and celiac axis with Krebs-Ringer bicarbonate buffer in the presence or absence of 10% ovine erythrocytes. The gastric lumen was perfused with distilled water and gastric contents were collected at frequent intervals through a catheter at the pylorus. Sixty minute gastric acid output in response to various concentrations of pentagastrin was determined by titration of gastric contents with 0.01 N NaOH to pH 7.0. During arterial perfusion with Krebs-Ringer bicarbonate buffer in the absence of ovine erythrocytes gastric acid output was 2.50±0.58 SEM μEq H+/h, which did not increase in response to perfusion with Krebs-Ringer bicarbonate buffer containing pentagastrin. However, inclusion of 10% ovine erythrocytes in the arterial perfusate resulted in substantial stimulation of gastric acid by pentagastrin: maximal acid output, achieved with a pentagastrin dose of 0.6 μg/kg/h, was 23.5±3.73 μEq H+/h (p<0.01). The results of the present study demonstrate the capacity of the isolated vascularly perfused rat stomach to secrete acid and provide a model for studying interactions of gastrointestinal regulatory peptides and their physiologic roles in the regulation of gastric acid secretion.  相似文献   

16.
《Regulatory peptides》1987,17(5):285-293
Infusion of the neuropeptide bombesin stimulates the secretion of several gastrointestinal hormones by an unknown mechanism. We have investigated the effects of atropine (15 ng/kg as bolus followed by 2.5 ng/kg · 30 min) and somatostatin (125 μg as i.v. bolus followed by 62.5 μg/30 min) on the stimulation of 3 hormones (gastrin, cholecystokinin and pancreatic polypeptide) by 60 pmol/kg · 20 min bombesin in 6 healthy volunteers. Plasma samples for measurement of hormones by sensitive and specific radioimmunoassays were obtained at − 5, 0, 2.5, 5, 7.5, 10, 15, 20, 25 and 30 min. Bombesin induced significant increases in plasma gastrin (12 ± 2 to 34 ± 3 pM; P < 0.0005), cholecystokinin (1.2 ± 0.2 to 8.9 ± 0.7 pM; P < 0.0001) and pancreatic polypeptide (22 ± 4 to 72 ± 19 pM; P < 0.05). There were great differences between the effects of atropine and somatostatin on the hormonal responses to bombesin. Atropine slightly increased the response of gastrin by 19% and that of cholecystokinin by 15%, but strongly inhibited the bombesin-stimulated pancreatic polypeptide secretion by 97%. On the other hand, somatostatin inhibited the bombesin-induced secretion of gastrin by 48%, cholecystokinin by 82% and pancreatic polypeptide by 107%. These results point to considerable qualitative and quantitative differences in the stimulatory mechanisms of bombesin on the hormones studied.  相似文献   

17.
Corpus luteum function in the cycling and the pregnant rhesus monkey (Macaca mulatta) was evaluated through short term in vitro studies of progesterone production by suspensions of collagenase-dispersed luteal cells in the presence and absence of exogenous gonadotropin (human chortonic gonadotropin, HCG). Cells from mid-luteal phase of the menstrual cycle secreted progesterone, as measured by accumulation of this hormone in the incubation medium, and responded to the addition of 100 ng HCG/ml with a marked increase in progesterone secretion above basal level (63.7 ± 13.1 versus 24.7 ± 5.5 ng progesterone/ml/5 × 104cells/ 3 hr, X ± S.E., n = 6; p < 0.05). However, luteal cells from early pregnancy (23–26 days after fertilization) secreted significantly less progesterone than cells of the non-fertile menstrual cycle (3.6 ± 2.4 versus 24.7 ± 5.5 ng/ml/5 × 104 cells/3 hr, n = 3; p < 0.05) and did not respond to HCG with enhanced secretion. By mid-pregnancy (108–118 days gestation) luteal cells exhibited partially renewed function, and near the time of parturition (163–166 days gestation) basal and HCG-stimulated progesterone secretion (30.2 ± 5.6 and 63.0 ± 13.0 ng/ml/5 × 104 cells/3 hr, respectively; n = 3) was equivalent to that of cells from the luteal phase of the non-fertile menstrual cycle. The data suggest that following a period around the fourth week of gestation, when steroidogenic activity is markedly diminished, the corpus luteum of pregnancy progressively reacquires its functional capacity and at term exhibits gonadotropin-sensitive steroidogenesis similar to that of the corpus luteum of the menstrual cycle.  相似文献   

18.
The ligand displacement reactions of oxyhemocyanin have been compared over a series of arthropods and molluscs. The arthropods (with the exception of Limulus) are found to be more reactive than the molluscs, (kcancer = .04 hr?1, kbusycon = .002 hr?1, klimulus ? 10?4hr?1 N3? reactions). Correlation of the spectral properties of the oxy sites require these to be extremely similar with small differences being associated with shifts in the dd transition energies. The met produced by ligand displacement contains variable amounts of EPR detectable, group 2 exogenous ligand damaged sites (arthropods 25–35%, molluscs 3–9%, Limulus <1%). A parallel (arthropod > mollusc ~ Limulus ~ 0) group 2 ligand damaged site is also reported for the half met derivatives.  相似文献   

19.
Background. Comparative studies of gastric acid secretion in children related to Helicobacter pylori infection are lacking. The purpose of this study was to compare acid secretion and meal‐stimulated gastrin in relation to H. pylori infection among pediatric patients. Materials and Methods. Thirty‐six children aged 10–17 years (17 with H. pylori infection) undergoing diagnostic endoscopy participated in the study. Diagnoses included gastritis only (n = 23), duodenal ulcer (n = 5) and normal histology (n = 8). Gastric acid output was studied using the endoscopic gastric secretion test before and 2–3 months after H. pylori eradication. Meal‐stimulated serum gastrin response was assessed before and 12 months after eradication. Results. H. pylori gastritis was typically antrum‐predominant. Acid secretion was greater in H. pylori‐positive patients with duodenal ulcer than in gastritis‐only patients or controls [mean ± standard error (SE): 6.56 ± 1.4, 3.11 ± 0.4 and 2.65 ± 0.2 mEq/10 minutes, respectively; p < .001]. Stimulated acid secretion was higher in H. pylori‐positive boys than girls (5.0 ± 0.8 vs. 2.51 ± 0.4 mEq/10 minutes, respectively; p < .05). Stimulated acid secretion pre‐ and post‐H. pylori eradication was similar (5.47 ± 0.8 vs. 4.67 ± 0.9 mEq/10 minutes, respectively; p = .21). Increased basal and meal‐stimulated gastrin release reversed following H. pylori eradication (e.g. basal from 134 to 46 pg/ml, p < .001 and peak from 544 to 133 pg/ml, p < .05). Conclusions. H. pylori infection in children is associated with a marked but reversible increase in meal‐stimulated serum gastrin release. Gastric acid hypersecretion in duodenal ulcer remains after H. pylori eradication, suggesting that the host factor plays a critical role in outcome of the infection.  相似文献   

20.
Unidirectional fluxes of [14C]lactose by whole cells of Escherichia coli under highly energized and partially de-energized (in the presence of CN?) conditions are analyzed kinetically.When the cells are energized, the value for V influx is 0.45 ± 0.01 mM internal concentration increment/s and Kt is 0.26 ± 0.03 mM. At an external concentration of 0.61 mM the steady-state internal concentration is 0.25 M, reached after about 1h. The maximum steady-state concentration ratio is 2 · 103.The efflux process under these conditions is non-saturable, being linearly dependent upon internal concentration over the range 25–250 mM with a first-order rate constant of 8.8 ± 0.2 · 10?4 s?1.The transport in the presence of CN? is active, with a maximum concentration ratio (internal concentration/external concentration) of 104, and the uptake is mimicked by anoxia (< 70 ppm O2).The effects of CN? are to lower the V for influx and to change the efflux from a non-saturable to a saturable process with a value for Kt (60 mM) intermediate between that for energized efflux (> 250 mM) and influxe (0.3–0.6 mM), the latter value not changing appreciably. Partial de-energization thus affects both the influx and efflux processes.  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