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 36 毫秒
1.
There are two possible mechanisms (co- or post-translational) for incorporation of Se into glutathione peroxidase in which selenocysteine presents at the active site of the enzyme and corresponds to UGA on the mRNA. We studied the above mechanisms using opal suppressor tRNA in mammals. Opal suppressor tRNA did not accept any selenocysteine and phosphoseryl-tRNA did not change to selenocysteyl-tRNA. Meanwhile, phosphoprotein changed to a protein containing selenocysteine by the incubation with H2Se and some enzymes. From these results, we propose that phosphoserine on glutathione peroxidase (apo-enzyme), which is synthesized with phosphoseryl-tRNA, is converted to selenocysteine in the mature enzyme, by a posttranslational mechanism. Opal suppressor tRNA may play a role to synthesize the apo-enzyme of glutathione peroxidase.  相似文献   

2.
Under aerobic conditions the addition of (C2N5)2N(N[O]NO) · Na+(DEA/NO), S-nitroso-N-macetyl penicillamine and nitric oxide (NO)-saturated buffer, but not S-nitroso- -glutathione, to dopamine solutions resulted in dopamine o-semiquinone formation that was dependent on the formation of a NO/oxygen intermediate. High pressure liquid chromatography (HPLC) electrochemical analysis of dopamine demonstrated that the DEA/NO-induced oxidation of dopamine was abrogated in the presence of the antioxidants, ascorbate and glutathione. NO spontaneously released from DEA/NO decreased [3H]dopamine accumulation in primary cultures of mesencephalic neurons in a dose-dependent fashion. In contrast, [3H]γ-aminobutyric acid uptake by mesencephalic neurons tested under the same conditions was unchanged. When DEA/NO was added to incubation buffer that contained [3H]dopamine and the antioxidant, ascorbate or glutathione, [3H]dopamine uptake was also inhibited. These data excluded that oxidation of extracellular [3H]dopamine by the intermediates of the NO/O2 reaction could have caused this decrease. Instead, NO may have acted directly on a not yet identified target operative in the regulation of dopamine storage and release. Analysis of the rate constants for the NO reaction with ascorbate, glutathione and dopamine revealed that dopamine quinone formation was delayed by the presence of antioxidants. Since the formation of NO as well as neurotransmitter release are activated during ischemia reperfusion injury, it is possible that prolonged NO exposure could deplete antioxidants and facilitate the oxidation of dopamine and thereby cause neurotoxicity.  相似文献   

3.
Binding characteristics of the selective V2 antagonist radioligand [3H]desGly-NH29-d(CH2)5[D-Ileu2,Ileu4]AVP to rat kidney were determined. Binding was specific, saturable and reversible. The peptide bound to a single class of high-affinity binding sites with Bmax 69.4±6.8 fmol/mg protein and KD 2.8±0.3 nM. AVP and other related peptides displaced [3H]desGly-NH29-d(CH2)5[D-Ileu2,Ileu4]AVP binding. The order of potency of inhibition was desamino-D-AVP > AVP > d(CH2)5[D-Ileu2,Ileu4]AVP > oxytocin > d(CH2)3[Tyr(Me)2]AVP > d(CH2)5[sarcosine7]AVP, which is typical of a selective V2 radioligand. Autoradiographic localization of [3H]desGly-NH29-d(CH2)5[D-Ileu2,Ileu4]AVP binding sites in kidney showed dense binding in the inner and outer medulla with less binding in the cortex, which is consistent with known renal V2 receptor distribution.  相似文献   

4.
The mechanisms of cholinergic stimulation of gastrin cells were studied in the rat pancreatic cell line B6 RIN. Carbachol induced an increase in intracellular Ca2+ and stimulated gastrin release in a dose-dependent manner over the range 10−5-10−3 M. These effects were completely abolished by atropine, suggesting the implication of muscarinic cholinergic receptors. The binding properties of these receptors were investigated. [N-Methyl-3H]scopolamine ([3h]nms) binding on cell homogenates was time-dependent, saturable and consistent with a single high-affinity binding class (Kd = 39.5 pM, and Bmax = 7.9 fmol/mg DNA). Carbachol competitively inhibited [3H]NMS binding. The potency of inhibition of [3H]NMS binding by subtype selective antagonists was hexahydrodifenidol> pirenzepine> AF-DX 116. These results suggest the M3, muscarinic receptors may be involved in the carbachol-induced gastrin release from B6 RIN cells.  相似文献   

5.
We measured the toxicity and mutagenicity induced in human diploid lymphoblasts by various radiation doses of X-rays and two internal emitters. [125I]iododeoxyuridine ([125I]dUrd) and [3H]thymidine ([3H]TdR), incorporated into cellular DNA. [125I]dUrd was more effective than [3H]TdR at killing cells and producing mutations to 6-thioguanine resistance (6TGR). No ouabain-resistant mutants were induced by any of these agents. Expressing dose as total disintegrations per cell (dpc), the D0 for cell killing for [125I]dUrd was 28 dpc and for [3H]TdR was 385 dpc. The D0 for X-rays was 48 rad at 37°C. The slopes of the mutation curves were approximately 75 × 10−8 6TGR mutants per cell per disintegration for [125I]dUrd and 2 × 10−8 for [3H]TdR. X-Rays induced 8 × 10−8 6TGR mutants per cell per rad. Normalizing for survival, [125I]dUrd remained much more mutagenic at low doses (high survival levels) than the other two agents. Treatment of the cells at either 37°C or while frozen at −70°C yielded no difference in cytotoxicity or mutation for [125I]dUrd or [3H]TdR, whereas X-rays were 6 times less effective in killing cells at −70°C.

Assuming that incorporation was random throughout the genome, the mutagenic efficiencies of the radionuclides could be calculated by dividing the mutation rate by the level of incorporation. If the effective target size of the 6TGR locus is 1000–3000 base pairs, then the mutagenic efficiency of [125I]dUrd is 1.0–3.0 and of [3H]TdR is 0.02–0.06 total genomic mutations per cell per disintegration. 125I disintegrations are known to produce localized DNA double-strand breaks. If these breaks are potentially lethal lesions, they must be repaired, since the mean lethal dose (D0) was 28 dpc. The observations that a single dpc has a high probability of producing a mutation (mutagenic efficiency 1.0–3.0) would suggest, however, that this repair is extremely error-prone. If the breaks need not be repaired to permit survival, then lethal lesions are a subset of or are completely different from mutagenic lesions.  相似文献   


6.
Bhargava, H. N., V. M. Villar, J. Cortijo and E. J. Morcillo. Binding of [3H][D-Ala2, MePhe4, Gly-ol5]enkephalin, [3H][D-Pen2, D-Pen5]enkephalin, and [3H]U-69,593 to airway and pulmonary tissues of normal and sensitized rats. Peptides 18(10) 1603–1608, 1997.—The role of endogenous opioid peptides in the regulation of bronchomotor tone, as well as in the pathophysiology of asthma is uncertain. We have studied the binding of highly selective [3H]labeled ligands of μ-([D-Ala2, MePhe4, Gly-ol5]enkephalin; DAMGO), δ ([D-Pen2, D-Pen5]enkephalin; DPDPE), and κ-(U-69,593) opioid receptors to membranes of trachea, main bronchus, lung parenchyma and pulmonary artery obtained from normal (unsensitized) and actively IgE-sensitized rats acutely challenged with the specific antigen. [3H]DAMGO, [3H]DPDPE and [3H]U-69,593 bound to membranes of normal and sensitized tissues at a saturable, single high-affinity site. The rank order of receptor densities in normal tissues was δ- ≥ κ- ≥ μ-, with lung parenchyma exhibiting the greatest binding capacity for δ- and μ- receptors compared to the other regions examined. The Kd values showed small differences between ligands and regions tested. The μ- and δ-opioid receptor densities were decreased in sensitized main bronchus and lung parenchyma, respectively, compared to normal tissues. By contrast, κ-opioid receptor density was augmented in sensitized lung parenchyma but an increase in Kd values was also observed. These differential changes in the density and affinity of opioid receptor types may be related to alterations in endogenous opioid peptides during the process of sensitization.  相似文献   

7.
In neuroblastoma × glioma hybrid cells (NG 108-15) labelled with [32P]-trisodium phosphate, [3H]-inositol and [14C]-arachidonic acid, bradykinin stimulated the hydrolysis of phosphatidylinositol 4,5-bisphosphate (PIP2) while it had no effect on the release of [14C]-arachidonic acid (AA). The effect on PIP, was time- and dose-dependent with a maximal effect on [3H]-inositol- and [32P]-labelled cells after 10–30 s of stimulation with 10−6 M bradykinin. However, the hydrolysis of [14C]-AA labelled PIP2 was delayed compared to the effect on [3H]- and [14C]-PIP2 and was not detectable until after 60 s of stimulation. Bradykinin stimulation resulted in an increased formation of [3H]-inositol phosphates (IP) and [32P]- and [14P]- and [14C]-phosphatidic acid (PA) but the time course for PA formation did not allow the time-course for PIP2 hydrolysis. A reduced labelling of [23P]- and [14C]-phosphatidylcholine was also found in stimulated cells suggesting that PA may derive from other sources than PIP2. In conclusion, our results indicate that bradykinin activates phospholipase C, but not phospholipase A2, in NG 108-15 cells.  相似文献   

8.
Mechanisms of aneuploid induction   总被引:5,自引:0,他引:5  
D. J. Bond 《Mutation research》1987,181(2):257-266
The effects of scavengers of active oxygen species on cadmium chloride (CdCl2)-induced inhibition of cell growth and DNA synthesis and on the metal-induced clastogenesis were investigated to evaluate whether cadmium could induce a prooxidant state in cultured Chinese hamster V79 cells.

Inhibition by CdCl2 of cell growth and [3H]thymidine incorporation into the acid-insoluble fraction of cells and the metal-induced clastogenesis were suppressed in part by the presence of the diffusible radical scavenger, butylated hydroxytoluene (BHT). The action of BHT was concentration-dependent and did not affect the intracellular level of cadmium. -Mannitol, a hydroxyl radical scavenger, also significantly suppressed Cd-induced inhibition of cell growth and [3H]thymidine incorporation. Catalase was marginally suppressive on Cd-induced inhibition of cell growth. These results suggest that cadmium can induce a prooxidant state in cultured mammalian cells.

The mechanism by which cadmium induces a prooxidant state was investigated by measuring the effect of cadmium on those enzymes which constitute a cellular defense against active oxygen and on the level of the intracellular antioxidant, glutathione (GSH). 2-h treatments with CdCl2 over a concentration range of 2–10 × 10−5 M did not influence superoxide dismutase, catalase, GSH peroxidase or GSSG reductase. In contrast, the level of glutathione was decreased to approximately 40% by treatment with 2 × 10−5 M cadmium. The decrease in glutathione level may be responsible for a role by active oxygen in Cd-induced inhibition of cell growth and DNA synthesis and the metal-induced clastogenesis.  相似文献   


9.
The novel ligand 4,5-bis(diphenylthiophosphinoyl)-1,2,3-triazole, LT-S2H, has been synthesized, converted to the triethylamine salt, and to the palladium complexes Pd[LT-S2]2 and Pd[LT-S2][η3-methallyl]. Structures of LT-S2H, of its 2-acetyl derivative, of Pd[LT-S2]2 and Pd[LT-S2][η3-methallyl] were determined by X-ray crystallography. In the last two complexes the LT-S2 ligand was N,S-bonded.  相似文献   

10.
The solution of [RhCl(PPh3)3] in acidic 1-ethyl-3-methylimidazolium chloroaluminate(III) ionic liquid (AlCl3 molar fraction, xAlCl3=0.67) was investigated by 1H and 31P{1H} NMR. One triphenyl phosphine is lost from the complex and is protonated in the acidic media, and cis-[Rh(PPh3)2ClX], (2), where X is probably [AlCl4], is formed. On, standing, 2 is converted to trans-[Rh(H)(PPh3)2X], (3). The reaction of 2 and H2 also produces trans-[Rh(H)(PPh3)2X], (3). 1H and 31P{1H} NMR support the suggestion that a weak ligand such as [AlCl4], present in solution may interact with the metal centre. When [RhCl(PPh3)3] is dissolved in CH2Cl2/AlCl3/HCl for comparison, two exchanging isomers of what is probably [RhH{(μ-Cl)2AlCl2}{(μ-Cl)AlCl3}(PPh3)2], (6) and (7), are formed.  相似文献   

11.
In order to specify the source of locally synthesized prostaglandin (PG) E2 which is able to saturate the large class of low affinity PGE2 receptors in chick spinal cord, bioconversion of [1-14C]arachidonic acid into prostanoids was studied in homogenates of chick spinal cord and meninges first without addition of exogenous glutathione (GSH). Homogenates of spinal cord produced 14C-labeled PGE2, PGD2 and PGF2. Homogenates of meninges accumulated much larger amounts of [14C]PGE2 than spinal cord and surprisingly a 14C-labeled arachidonate metabolite referred to as compound Y. Compound Y generation, which was inhibited by indomethacin and enhanced by esculetin, was therefore mediated through the cyclooxygenase pathway. The fact that no labeled compound Y was detected in homogenates incubated with [3H]PGD2 or [3H]PGE2 indicated that compound Y was not degradation product of PGs. Secondly, after addition of exogenous GSH, 14C-labeled compound Y was totally converted into [14C]PGE2. The compound Y which is converted into PGFs after a strong reduction with NaBH4 and into PGE2 after a mild reduction with GSH-hemin system or SnCl2 was therefore assumed to be a 15 hydroperoxy-PGE2 (15 HP-PGE2). These results suggest that PGE2 can be synthesized in meninges either by the classical isomerization of PGH2 or by isomerization of PGG2 followed by a GSH-sensitive reaction.  相似文献   

12.
Age-related alterations in major neurotransmitter receptors and voltage dependent calcium channels were analyzed by receptor autoradiography in the gerbil brain. [3H]Quinuclidinyl benzilate (QNB). [3H]cyclohexyladenosine (CHA), [3H]muscimol, [3H]MK-801, [3H]SCH 23390, [3H]naloxone, and [3H]PN200-110 were used to label muscarinic acetylcholine receptors, adenosine A1 receptors, γ-aminobutyric acidA (GABAA) receptors, (NMDA) receptors, dopamine D1 receptors, opioid receptors, and voltage dependent calcium channels, respectively. In middle-aged gerbils (16 months old), the hippocampus exhibited a significant elevation in [3H]QNB, [3H]MK-801, [3H]SCH 23390, [3H]naloxone, and [3H]PN200-110 binding, whereas [3H]CHA and [3H]muscimol binding showed a significant reduction in this area, compared with that of young animals (1 month). On the other hand, the cerebellum showed a significant alteration in [3H]QNB, [3H]CHA, and [3H]naloxone binding and the striatum also exhibited a significant alteration in [3H]SCH 23390 and [3H]CHA binding in middle-aged gerbils. The neocortex showed a significant elevation only in [3H]CHA binding in middle-aged animals. The nucleus accumbens and thalamus also showed a significant alteration only in [3H]muscimol binding. However, the hypothalamus and substantia nigra exhibited no significant alteration in these bindings in middle-aged gerbils. These results demonstrate the age-related alterations of various neurotransmitter receptors and voltage dependent calcium channels in most brain regions. Furthermore, they suggest that the hippocampus is most susceptible to aging processes and is altered at an early stage of senescence.  相似文献   

13.
WAY–100635 is the first selective, silent 5–HT1A (5-hydroxytryptamine1A, serotonin-1A) receptor antagonist. We have investigated the use of [3H]WAY–100635 as a quantitative autoradiographic ligand in post-mortem human hippocampus, raphe and four cortical regions, and compared it with the 5–HT1A receptor agonist, [3H]8–OH–DPAT. Saturation studies showed an average Kd for [3H]WAY–100635 binding in hippocampus of 1.1 nM. The regional and laminar distributions of [3H]WAY–100635 binding and [3H]8–OH–DPAT binding were similar. The density of [3H]WAY–100635 binding sites was 60–70% more than that of [3H]8–OH–DPAT in all areas examined except the cingulate gyrus where it was 165% higher. [3H]WAY–100635 binding was robust and was not affected by the post-mortem interval, freezer storage time or brain pH (agonal state). Using [3H]WAY–100635, we confirmed an increase of 5–HT1A receptor binding sites in the frontal cortex in schizophrenia, previously demonstrated with [3H]8–OH–DPAT. Compared to [3H]8–OH–DPAT, [3H]WAY–100635 has two advantages: it has a higher selectivity and affinity for the 5–HT1A receptor, and it recognizes 5–HT1A receptors whether or not they are coupled to a G-protein, whereas [3H]8–OH–DPAT primarily detects coupled receptors. Given these considerations, the [3H]WAY–100635 binding data in schizophrenia clarify two points. First, they indicate that the elevated [3H]8–OH–DPAT binding seen in the same cases is attributable to an increase of 5–HT1A receptors rather than any other binding site. Second, the enhanced [3H]8–OH–DPAT binding in schizophrenia reflects an increased density of 5–HT1A receptors, not an increased percentage of 5–HT1A receptors which are G-protein-coupled. We conclude that [3H]WAY–100635 is a valuable autoradiographic ligand for the qualitative and quantitative study of 5–HT1A receptors in the human brain.  相似文献   

14.
[3H]Lysergic acid diethylamide (LSD) in the presence of 40 nM ketanserin labeled the 5-HT1A receptor subtype in rat hippocampal membranes. In the presence of guanosine triphosphate (GTP), the Bmax and affinity of [3H]LSD binding to the 5-HT1A binding site were significantly decreased. [3H]LSD in the presence of 40 nM WB4101 labeled the 5-HT2 receptor subtype in homogenates of rat frontal cortex. In contrast to the effect on [3H]LSD binding to the 5-HT1A binding site, GTP produced no significant effect on either the Bmax or the KD of [3H]LSD binding to the 5-HT2 binding site. Competition of 5-HT for [3H]LSD binding to the 5-HT2 binding site was best described by a computer-derived model assuming two binding sites. In the presence of GTP, the 5-HT competition curve was shifted significantly to the right with an approx. 3-fold increase in the IC50. These binding characteristics are consistent with [3H]LSD acting as an antagonist at the 5-HT2 receptor which has multiple affinity states for agonists and is coupled to a guanine nucleotide regulatory subunit. Thus, [3H]LSD has binding characteristics consistent with it acting as an agonist at the 5-HT1A receptor subtype but as an antagonist at the 5-HT2 receptor subtype in rat brain.  相似文献   

15.
T Mizutani  T Hitaka 《FEBS letters》1988,232(1):243-248
This study has been undertaken in order to elucidate the mechanisms of incorporation of Se into glutathione peroxidase (GSHPx), in which selenocysteine corresponds to the opal termination codon UGA on the mRNA. We studied the above mechanisms using an opal suppressor tRNA, prepared from bovine liver, and casein as a model protein for the GSHPx apo-enzyme which might contain phosphoserine. The results showed that opal suppressor tRNA did not accept selenocysteine (lower than 0.1 mmol/mol) under the standard conditions. A trace amount of phosphoseryl-tRNA was converted to selenocysteyl-tRNA by incubation with H2Se and some enzymes. Meanwhile, a number of phosphoserine residues in casein were converted to selenocysteine residues by incubation with H2Se and enzymes. These results suggest that opal suppressor tRNA plays a role in synthesizing GSHPx via co- and/or post-translational mechanisms.  相似文献   

16.
Rat brain cortex slices preincubated with [3H]serotonin or [3H]noradrenaline (25 100 nmol/l each) were superfused and the effects of serotonin and histamine on the electrically (0.3 or 3 Hz) evoked tritium overflow were studied.

In slices preincubated with [3H]serotonin the extent of inhibition of the electrically (3 Hz) evoked tritium overflow produced by histamine was increased when the concentration of [3H]serotonin used for incubation was decreased. The evoked overflow tended to be lower in slices from 2-year-old rats than in slices from 6-month-old animals whereas the inhibitory effect of histamine on the evoked overflow did not differ. Treatment of rats with nimodipine for at least 6 weeks did not significantly affect the evoked overflow in slices from 6-month and 2-year-old rats nor did it significantly alter the serotonin- and histamine-mediated inhibition of the evoked overflow in slices from young adult rats. The extent of histamine-mediated inhibition of the electrically evoked tritium overflow from slices (of young adult rats) preincubated with [3H]noradrenaline did not change when the concentration of [3H]noradrenaline used for incubation was decreased; the degree of inhibition markedly increased when the frequency of stimulation was lowered from 3 to 0.3 Hz. The inhibitory effect of histamine on the electrically (0.3 Hz) evoked overflow was mimicked by the H3 receptor agonist R-(−)--methylhistamine and antagonized by the H3 receptor antagonist thioperamide. The electrically evoked overflow and its inhibition by histamine were not affected by nimodipine, irrespective of whether the Ca2+ antagonist was administered in vivo (for at least 6 weeks) or added to the superfusion medium in vitro.

It is concluded that (1) the extent of the H3 receptor-mediated effect in rat brain cortex slices can be markedly increased by lowering the concentration of the tracer in slices preincubated with [3H]serotonin and by lowering the stimulation frequency in slices preincubated with [3H]noradrenaline; (2) the H3 receptor-mediated inhibition of serotonin release is not changed during aging and (3) nimodipine does not significantly influence serotonin release and noradrenaline release and their serotonin and/or histamine receptor-mediated modulation.  相似文献   


17.
The effects of subcutaneous (s.c.) oxytocin treatment have been investigated on various parameters of dopaminergic neurotransmission in basal forebrain structures (nucleus olfactorius posterior + nucleus accumbens + septum) of the mouse. Acute oxytocin treatment failed to influence dopamine utilization in the basal forebrain. Following chronic injections of oxytocin (0.2 mg/kg) for 8 8 days, the neuropeptide decreased dopamine utilization. Neither in vivo nor in vitro oxytocin treatment was capable of influencing the in vitro uptake of [3H]dopamine in basal forebrain slices. The spontaneous release of [3H]dopamine (in the presence of 4.2 mM K+) from basal forebrain tissue slices was not affected by in vitro or acute or chronic in vivo oxytocin treatment. The stimulated release of [3H]dopamine (in the presence of 30 mM K+) was significantly inhibited by chronic in vivo oxytocin administration. Chronic oxytocin treatment decreased the Bmax value of [3H]spiroperidol binding in the basal forebrain. The dissociation constant (Kd) of [3H]spiroperidol binding was not influenced by oxytocin. The data indicate that peripheral oxytocin treatment is capable of modifying dopaminergic neurotransmission in mouse basal forebrain regions.  相似文献   

18.
The formation of three [Tl(en)n]3+ complexes (n=1–3) in a pyridine solvent has been established by means of 205Tl and 1H NMR. Their stepwise stability constants based on concentrations, Kn=[Tl(en)n 3+]/{[Tl(en)n−1 3+]·[en]}, at 298 K in 0.5 M NaClO4 ionic medium in pyridine, were calculated from 205Tl NMR integrals: log K1=7.6±0.7; log K2=5.2±0.5 and log K3=2.64±0.05. Linear correlation between both the 205Tl NMR shifts and spin–spin coupling 205Tl–1H versus the stability constants has been found and discussed. A single crystal with the composition [Tl(en)3](ClO4)3 was synthesized and its structure determined by X-ray diffraction. The Tl3+ ion is coordinated by three ethylenediamine ligands via six N-donor atoms in a distorted octahedral fashion.  相似文献   

19.
In order to better understand the function of aromatase, we carried out kinetic analyses to asses the ability of natural estrogens, estrone (E1), estradiol (E2), 16-OHE1, and estriol (E3), to inhibit aromatization. Human placental microsomes (50 μg protein) were incubated for 5 min at 37°C with [1β-3H]testosterone (1.24 × 103 dpm 3H/ng, 35–150 nM) or [1β-3H,4-14C]androstenedione (3.05 × 103 dpm 3H/ng, 3H/14C = 19.3, 7–65 nM) as substrate in the presence of NADPH, with and without natural estrogens as putative inhibitors. Aromatase activity was assessed by tritium released to water from the 1β-position of the substrates. Natural estrogens showed competitive product inhibition against androgen aromatization. The Ki of E1, E2, 16-OHE1, and E3 for testosterone aromatization was 1.5, 2.2, 95, and 162 μM, respectively, where the Km of aromatase was 61.8 ± 2.0 nM (n = 5) for testosterone. The Ki of E1, E2, 16-OHE1, and E3 for androstenedione aromatization was 10.6, 5.5, 252, and 1182 μM, respectively, where the Km of aromatase was 35.4 ± 4.1 nM (n = 4) for androstenedione. These results show that estrogens inhibit the process of andrigen aromatization and indicate that natural estrogens regulate their own synthesis by the product inhibition mechanism in vivo. Since natural estrogens bind to the active site of human placental aromatase P-450 complex as competitive inhibitors, natural estrogens might be further metabolized by aromatase. This suggests that human placental estrogen 2-hydroxylase activity is catalyzed by the active site of aromatase cytochrome P-450 and also agrees with the fact that the level of catecholestrogens in maternal plasma increases during pregnancy. The relative affinities and concentration of androgens and estrogens would control estrogen and catecholestrogen biosynthesis by aromatase.  相似文献   

20.
The present study was undertaken to characterize the binding activities of propiverine and its N-oxide metabolites (1-methyl-4-piperidyl diphenylpropoxyacetate N-oxide: P-4(N → O), 1-methyl-4-piperidyl benzilate N-oxide: DPr-P-4(N → O)) toward L-type calcium channel antagonist receptors in the rat bladder and brain. Propiverine and P-4(N → O) inhibited specific (+)-[3H]PN 200–110 binding in the rat bladder in a concentration-dependent manner. Compared with that for propiverine, the Ki value for P-4(N → O) in the bladder was significantly greater. Scatchard analysis has revealed that propiverine increased significantly Kd values for bladder (+)-[3H]PN 200–110 binding. DPr-P-4(N → O) had little inhibitory effects on the bladder (+)-[3H]PN 200–110 binding. Oxybutynin and N-desethyl-oxybutynin (DEOB) also inhibited specific (+)-[3H]PN 200–110 binding in the rat bladder. Propiverine, oxybutynin and their metabolites inhibited specific [N-methyl-3H]scopolamine methyl chloride ([3H]NMS) binding in the rat bladder. The ratios of Ki values for (+)-[3H]PN 200–110 to [3H]NMS were markedly smaller for propiverine and P-4(N → O) than oxybutynin and DEOB. Propiverine and P-4(N → O) inhibited specific binding of (+)-[3H]PN 200–110, [3H]diltiazem and [3H]verapamil in the rat cerebral cortex in a concentration-dependent manner. The Ki values of propiverine and P-4(N → O) for [3H]diltiazem were significantly smaller than those for (+)-[3H]PN 200–110 and [3H]verapamil. Further, their Ki values for [3H]verapamil were significantly smaller than those for (+)-[3H]PN 200–110. The Ki values of propiverine for each radioligand in the cerebral cortex were significantly (P < 0.05) smaller than those of P-4(N → O). In conclusion, the present study has shown that propiverine and P-4(N → O) exert a significant binding activity of L-type calcium channel antagonist receptors in the bladder and these effects may be pharmacologically relevant in the treatment of overactive bladder after oral administration of propiverine.  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