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.
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   

2.
These studies defined the expression patterns of genes involved in fatty acid transport, activation and trafficking using quantitative PCR (qPCR) and established the kinetic constants of fatty acid transport in an effort to define whether vectorial acylation represents a common mechanism in different cell types (3T3-L1 fibroblasts and adipocytes, Caco-2 and HepG2 cells and three endothelial cell lines (b-END3, HAEC, and HMEC)). As expected, fatty acid transport protein (FATP)1 and long-chain acyl CoA synthetase (Acsl)1 were the predominant isoforms expressed in adipocytes consistent with their roles in the transport and activation of exogenous fatty acids destined for storage in the form of triglycerides. In cells involved in fatty acid processing including Caco-2 (intestinal-like) and HepG2 (liver-like), FATP2 was the predominant isoform. The patterns of Acsl expression were distinct between these two cell types with Acsl3 and Acsl5 being predominant in Caco-2 cells and Acsl4 in HepG2 cells. In the endothelial lines, FATP1 and FATP4 were the most highly expressed isoforms; the expression patterns for the different Acsl isoforms were highly variable between the different endothelial cell lines. The transport of the fluorescent long-chain fatty acid C1-BODIPY-C12 in 3T3-L1 fibroblasts and 3T3-L1 adipocytes followed typical Michaelis–Menten kinetics; the apparent efficiency (kcat/KT) of this process increases over 2-fold (2.1 × 106–4.5 × 106 s−1 M−1) upon adipocyte differentiation. The Vmax values for fatty acid transport in Caco-2 and HepG2 cells were essentially the same, yet the efficiency was 55% higher in Caco-2 cells (2.3 × 106 s−1 M−1 versus 1.5 × 106 s−1 M−1). The kinetic parameters for fatty acid transport in three endothelial cell types demonstrated they were the least efficient cell types for this process giving Vmax values that were nearly 4-fold lower than those defined form 3T3-L1 adipocytes, Caco-2 cells and HepG2 cells. The same cells had reduced efficiency for fatty acid transport (ranging from 0.82 × 106 s−1 M−1 to 1.35 × 106 s−1 M−1).  相似文献   

3.
Bisphenol A is used as a monomer in the production of polycarbonate plastic products. The widespread use of bisphenol A has raised concerns about its effects in humans. Since there is little information on the mutagenic potential of the chemical, the mutagenicity of bisphenol A was tested using human RSa cells, which has been utilized for identification of novel mutagens. In genomic DNA from cells treated with bisphenol A at concentrations ranging from 1×10−7 to 1×10−5 M, base substitution mutations at K-ras codon 12 were detected using PCR and differential dot-blot hybridization with mutant probes. Mutations were also detected using the method of peptide nucleic acid (PNA)-mediated PCR clamping. The latter method enabled us to detect the mutation in bisphenol A-treated cells at a dose (1×10−8 M) equivalent to that typically found in the environment. Induction of ouabain-resistant (OuaR) phenotypic mutation was also found in cells treated with 1×10−7 and 1×10−5 M of bisphenol A. The induction of K-ras codon 12 mutations and OuaR mutations was suppressed by pretreating RSa cells with human interferon (HuIFN)-α prior to bisphenol A treatment. The cells treated with bisphenol A at the concentration of 1×10−6 M elicited unscheduled DNA synthesis (UDS). These findings suggested that bisphenol A has mutagenicity in RSa cells as well as mutagens that have been tested in these cells, and furthermore, that a combination of the PNA-mediated PCR clamping method with the human RSa cell line may be used as an assay system for screening the mutagenic chemicals at very low doses.  相似文献   

4.
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   

5.
• Beta-adrenergic receptor identification and properties are probed in rat parotid membranes utilizing the high affinity β-adrenergic antagonist(−)-[3H]dihydroalprenolol.
• The binding of (−)-[3H]dihydroalprenolol to membrane preparations of parotid is rapid, equilibrium being reached in 5 min. Strict stereospecificity is observed, (−)-propanolol being 100 times more potent than (+)-propranolol in competing with (−)-[3H]dihydroalprenolol for binding sites. Beta-adrenergic agonists compete for the binding sites with (−)-[3H]dihydroalprenolol with the same characteristics, i.e., much higher concentrations of the (+)-stereoisomers than the (−)-stereoisomers are required to produce 50% inhibition, the range varies from 14-fold for epinephrine to 300-fold for isoproterenol. Among the (−)-stereoisomers, the relative potency of inhibitory action is (−)-propranolol > (−)-isoproterenol > (−)-epinephrine ≡ (−)-norepinephrine. (−)-Isoproterenol is about 20 times as potent as norepinephrine, the least potent agonist among all the catecholamine (−)-stereoisomers.
• The binding of (−)-[3H]dihydroalprenolol is saturable, with a maximum number of binding sites equalling 450 fmol/mg protein and a dissociation constant of 7.9 nM. The Scatchard plots show no significant curvilinear character. Hill plots consistently give a Hill coefficient close to unity (0.92–1.05). Both pieces of evidence suggest a single-component system with no significant cooperativity.
• Dissociation kinetics study after the method of De Metys et al. (1973) Biochem. Biophys. Res. Commun. 155, 154, indicates a lack of site-to-site interactions among the binding sites. The rate of dissociation of bound (−)-[3H]dihydroalprenolol is the same in the presence and absence of 1 · 10−5 M (±)-alprenolol.
• Based on the binding of (−)-[3H]dihydroalprenolol, it is concluded that the beta-adrenergic receptors can be identified in rat parotid and that these binding sites display β1 character. Results of the study indicate a one-component system with no observable site-to-site interactions.
Abbreviations: DHA; dihydroalprenolol  相似文献   

6.
The effects of prostaglandin F (PGF) on propulsive activity in segments of isolated colon and on isolated strips of guinea-pig colon were investigated.Using experimental conditions under which spontaneous propulsive activity was negligible, PGF (5×10−8×1×10−6M), added to the bathing medium, increased propulsive activity in a concentration dependent manner. This increase of propulsive activity was abolished in the presence of atropine or tetrodotoxin (1×10−7g/ml).The contractions produced by PGF(5×10−7 − 1×10−5M) in isolated longitudinal and circular smooth muscle strips of guinea-pig colon were unaffected in the presence of atropine or tetrodotoxin (1×10−7 g/ml).From these results it is concluded that under the conditions employed in this study propulsive activity stimulated by PGF may depend on the contractions of both muscle layers and stimulation of the peristalic reflex.  相似文献   

7.
Sympathetic nerve stimulation of the perfused mesenteric arterial bed of the rabbit, , increase the secretion of prostaglandin (PG)I2 and PGE2. Prazosin (4.8 × 10−6), and α1 adrenergic receptor antagonist, inhibited this inrease in release of PGI2 but not of PGE2 whereas rauwolsin (10−7 M), an α2 adrenergic receptor antagonist, inhibited the increase in release of PGE2 but not of PGI2. Prazosin (10−6 M) completely blocked the vasoconstrictor response to nerve stimulation, and to norepinephrine and phenylephrine administration, suggesting there to be little of an α2 adrenergic receptor component in this response. It is concluded that the increase in PGI2 release follows the activation of α1 adrenergic receptors and is therefore post-junctional in origin, whereas the increase in PGE2 release follows the activation of α2 adrenergic receptors and may be pre- and/or post-junctional in origin.Indomethacin (2.8 × 10−7, 5.6 × 10−7 and 1.12 × 10−6 M did not affect the vasoconstrictor responses to nerve stimulation at 10 Hz, whereas rauwolsin (10−7 M) in the presence of indomethacin substantially increased them. These results indicate that PGE2 does not regulate norepinephrine release following nerve stimulation at 10 Hz to rabbit mesenteric arteries, and that the inhibition of norepinephrine release following stimulation of α2 pre-junctional receptors is independent of PG involvement.  相似文献   

8.
In order to study mitotic homologous recombination in somatic Drosophila melanogaster cells in vitro and to learn more on the question how recombination is influenced by mutagens, a genetic system was developed where spontaneous and drug-induced recombination could be monitored. Two recombination reporter substrates were stably introduced in multiple copies into the genome of established D. melanogaster Schneider line 2 cells: one plasmid (pSB310) contained the 5′ and 3′ deleted neomycin phosphoribosyltransferase alleles neoL and neoR as direct repeats; the other (pSB485) contained similar deletions (lacZL and lacZR) of the β-galactosidase gene (lacZ). Restoration of a functional neo gene upon mitotic recombination between homologous sequences allowed direct selection for the event, whereas recombination in single cells harbouring the integrated lacZ-based reporter plasmid was detected by histochemical staining or flow cytometric analysis (FACS). The neo-based construct in the clonal transgenic cell line 44CD4 showed a spontaneous recombination frequency of 2.9×10−4, whereas the 485AD1 cell line harbouring the lacZ-based construct exhibited a frequency of 2.8×10−4. The alkylating agents EMS and MMS and the clastogen mitomycin C were able to induce recombination in the 485AD1 cell line in a dose-dependent manner. The results obtained from these studies suggest that the transgenic cell lines are potentially useful tools for identifying agents which stimulate direct repeat recombination in somatic Drosophila cells.  相似文献   

9.
When a Euglena, in a medium containing ATP, is microinjected with 7 × 10−14 l of 0.02 M EDTA, which binds Ca2+ and Mg2+, flagellar motility stops. Flagellar arrest in Chlamydomonas occurs with the injection of 2 × 10−14 l of 0.02 M EDTA. The injection of similar amounts (7 × 10−14 l in Euglena and 3 × 10−14 l in Chlamydomonas) of 0.02 M EGTA, which preferentially binds Ca2+, did not significantly alter flagellar motility. This suggests that a decrease in the internal Ca2+ concentration in Euglena or Chlamydomonas did not stimulate flagellar beating. Further, flagellar motility decreased when internal Mg2+ was chelated. The microinjection of Zn2+ into these cells caused a decrease in flagellar frequency analogous to the decrease in frequency caused by the injection of Ca2+ and EDTA. The microinjection of 7 × 10−14 l of 0.2 M Mn2+ caused an approx. 1.5-fold increase in Euglena flagellar motility. Chlamydomonas flagella, which cease to beat upon impalement in an Mg2+-free medium, resume a flagellar frequency of 18 Hz when injected with 3 × 10−14 l of 0.2 M Mn2+. In the experiments reported here, Mn2+ acts as an analog of Mg2+.  相似文献   

10.
The concentrations of PGE, PGF, and 6-keto-PGF were increased in rat seminal vesicle tissue following mating activity. Likewise, synthesis of PGE and PGF was stimulated by epinephrine (3 × 10−7to 3 × 10−6 M) in tissues and media from incubations of intact rat seminal vesicles. The stimulation was inhibited by phentolamine, an α-adrenoreceptor blocking agent. Carbamylcholine (2 × 10−6 M) and bradykinin (1 × 10−6 M) had no effect on PGE or PGF synthesis, even though both compounds stimulated contractility of the rat seminal vesicle at these concentrations. These data suggest that mating and adrenergic stimulation increase prostaglandin synthesis in] the rat seminal vesicle, probably through an α-adrenergically mediated mechanism.  相似文献   

11.
The influence of the polyene antibiotic, amphotericin B, on the permeability of porcine and bovine erythrocytes was studied by measuring net and tracer movements of nonelectrolytes, anions and cations in these cells.
1. 1. Amphotericin B (0.5–20 μM) enhances the rates of transfer of hydrophilic nonelectrolytes (glycerol, erythritol), anions (phosphate, lactate, glycollate, Cl, SCN) and cation (Na+, K+). Different concentrations of the antibiotic are required for equal effects on the different transfer processes. Bovine erythrocytes respond much less to amphotericin than porcine cells.
2. 2. Nystatin enhances the transfer of all the permeants to a much lesser extent; gramicidin D, although producing a large increase of cation permeability, leaves unaltered anion and nonelectrolyte transfer.
3. 3. The amphotericin-induced enhancement of erythrocyte permeabability (ΔP) increases with time. It has a concentration dependence of the type ΔP = α · CnA* (n = 1.5–2.5) and becomes more pronounced at low temperatures.
4. 4. Partial depletion of membrane cholesterol, which in itself does not alter nonelectrolyte and anion permeability, reduces the effectivity of amphotericin B, indicating that in the erythrocyte membrane, too, a sterol acts as receptor for polyene antibiotics.
5. 5. The selectivity of the amphotericin-induced pathway of transfer in the erythrocyte membrane is lower than that of the normal pathways of nonelectrolyte and anion transfer in this membrane.
The results support the view that amphotericin produces the same type of molecular reorganisation of lipid constituents in biological and artificial membranes. On the other hand, the polyene-induced pathway in the erythrocyte membrane seems to differ functionally from the normal transfer pathways in this membrane.  相似文献   

12.
The fracture mechanics parameters associated with the fracture of transversely oriented bovine femur compact tension specimens have been determined for a slowly propagating and stable crack, as a function of cross head speed. It was found that an increase in cross head speed from 1.7–33 × 10−6 m sec−1 produced an increase in the crack velocity from 2.1–27 × 10−5 m sec−1 and an associated increase in the critical strain energy release rate (Gc) from 920 to 2780 J m−2 and in the critical stress intensity factor (Kc) from 2.4 to .  相似文献   

13.
The present paper reports the graft copolymerization of N-vinylformamide onto sodium carboxymethylcellulose by free radical polymerization using potassium peroxymonosulphate/thiourea redox system in an inert atmosphere. The reaction conditions for maximum grafting have been optimized by varying the reaction variables, including the concentration of N-vinylformamide (12.0 × 10−2–28.0 × 10−2 mol dm−3), potassium peroxymonosulphate (4.0 × 10−3–12.0 × 10−3 mol dm−3), thiourea (1.2 × 10−3–4.4 × 10−3 mol dm−3), sulphuric acid (2.0 × 10−3–10.0 × 10−3 mol dm−3), sodium carboxymethylcellulose (0.2–1.8 g dm−3) along with time duration (60–180 min) and temperature (25–45° C). Water swelling capacity, metal ion sorption and flocculation studies of synthesized graft copolymer have been performed with respect to the parent polymer. The graft copolymer has been characterized by FTIR spectroscopy and thermogravimetric analysis.  相似文献   

14.
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   

15.
Prostaglandins (PGs) have been shown to cytoprotect various tissue types against the toxic effects of many chemicals. The mechanism of this protection is poorly understood, but the involvement of cAMP is often implied. Only one previous study examined nervous tissue and PG protection. The present study was designed to determine if PGE2 affords cytoprotection to a more specific nervous tissue (embryonic neural retina) from the toxicity of actinomycin C (AMC) using a trypan blue exclusion assay. The lowest concentration of PGE2 (2 × 10−5M) had no effect, but as the concentration increased (3 × 10−5M and 5 × 10−5M), PGE2 did afford protection against AMC in a dose dependent fashion. Theophylline treated cells were not protected, suggesting that cAMP may not be the primary mechanism of protection.  相似文献   

16.
Graded doses of Pro-Leu-Gly-NH2 (3.5 × 10−12, 3.5 × 10−11, 3.5 × 10−10 or 3.5 × 10−9 mol) were administered into the lateral cerebral ventricle of rats. The noradrenaline level of the dorsal hippocampus was increased 30 min after a dose of 3.5 × 10−10 mol Pro-Leu-Gly-NH2. The dopamine level was increased in the dorsal hippocampus and in the striatum. The serotonin level was increased in the hypothalamus, in the striatum and decreased in the dorsal hippocampus.The catecholamine disappearance following 350 mg/kg of α-methyl-p-tyrosine indicated an accelerated dopamine disappearance in the striatum for each dose studied, while the hypothalamic noradrenaline disappearance was inhibited by a dose of 3.5 × 10−11 mol of Pro-Leu-Gly-NH2.The data indicate that Pro-Leu-Gly-NH2 induces dose and region-dependent changes in the cerebral monoamine metabolism. The striatal dopamine and hypothalamic serotonin metabolism appeared to be the most sensitive for intraventricular Pro-Leu-Gly-NH2.  相似文献   

17.
A highly sensitive, kinetically unambiguous assay for α-factor-induced delay of cell passage through the “start” step of cell division in yeast is presented. The assay employs perfusion with periodic microscopy to monitor the bud emergence kinetics on the 20% of cells within an exponentially growing population which exist prior to the α-factor execution point of start. The t1/2 for cell passage through start by this population of cells is 31 min in the absence of α-factor. The inhibition constant, KI, represents the α-factor concentration which produces a 50% inhibition of this rate and is equal to 2×10−10M. A second assay for maximal cell division arrest by α-factor on whole populations of cells is presented. This assay shows a maximum cell division arrest time of 125±5 h at saturating α-factor, and a K50 (that is, an α-factor concentration which produces a half-maximal response) of 2.5×10−8M. Both assays were performed in the effective absence of α-factor inactivation. Values of the dissociation constant KD and total number of receptors per cell which specifically mediate cell division arrest or delay were estimated to be 2.5×10−8M and 104, respectively. These estimates, along with the quantitative dose-response data for division arrest which are presented here, are consistent with each receptor·α-factor complex which is present on the cell at equilibrium producing a 43±10 s delay of cell passage through start. Surprisingly, this number is constant within twofold over the entire range of cellular division arrest responses to α-factor, that is, from a 1.9-fold inhibition of the rate of cell passage through start at 0.17 nM α-factor to a 125±5 h maximum arrest at saturating α-factor concentrations of >170 nM. The possible significance of this observation toward the mechanism of α-factor-induced cell division arrest is discussed.  相似文献   

18.
This study reports on the specific binding of [3H]heparin to human adrenocortical carcinoma cell line SW-13. Heparin binding to SW-13 cells is specific, saturable, and time- and temperature-dependent with maximum binding occurring between 90 and 120 min at 22 °C. Scatchard analysis revealed two classes of binding sites. The apparent Kd for high-affinity receptors is 2.14 × 10−8 M with 1.48 × 106 sites per cells. Six other tested mammalian cell lines also have specific binding sites for heparin.  相似文献   

19.
The effect of zinc-chelating dipeptides on osteoblastic MC3T3-E1 cells was investigated. As zinc compounds, we used zinc sulfate, AHZ, di(N-acetyl-β-alanyl-l-histidinato)zinc (AAHZ), and di(histidino)zinc (HZ). Cells were cultured for 72 h in the presence of zinc compounds (10−8–10−5M). The effect of AHZ (10−7 and 10−6M) to increase protein and deoxyribonucleic acid (DNA) contents in the cells was the greatest in comparison with those of other zinc compounds. Zinc sulfate and HZ at 10−7M did not have an effect on the cellular protein content. AHZ (10−6M) had a potent effect on cell proliferation, although zinc sulfate (10−6M) had no effect. β-Alanyl-l-histidine (10−6 and 10−5M) did not have an appreciable effect on the cells. Those effects of AHZ (10−6M) on osteoblastic cells were completely abolished by the presence of cycloheximide (10−6M). AHZ (10−8–10−5M) directly activated [3H]leucyl-tRNA synthetase in the cell homogenate, whereas the effect of zinc sulfate was seen at 10−6 and 10−5M. The present study suggests that the chemical form of zinc-chelating β-alanyl-l-histidine (AHZ) can reveal a potent anabolic effect on osteoblastic cells, and that AHZ directly stimulates protein synthesis.  相似文献   

20.
Infusions of prostacyclin (PGI2) (3 × 10−10 − 3 × 10−7M) into the coronary circulation of isolated hearts from guinea pigs or rabbits resulted in a concentration-dependent decrease in the coronary perfusion pressure (CPP). There was a slight decrease in left ventricular systolic pressure in the heart of the rabbit, whereas the heart rate remained unchanged. PGE2 was without effect on the heart of the rabbit but was as potent as PGI2 in decreasing the CPP in the guinea pig heart. 6-oxo-PGF (up to 3 × 10−6 M) did not affect any of the parameters measured.  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