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.
Di-deuterated and di-tritiated 16,16-dimethyl-trans-Δ2-PGE1 has been synthesized and used for development of a GC-MS method for quantitation of corresponding unlabelled drug in patient plasma. Although these carrier/internal standard molecules only contain 2 deuterium atoms the lower limit of detection at each injection is as low as about 40 pg.The maximum plasma levels of this drug following administration of vaginal suppositories used in clinical studies (1 mg 16,16-dimethyl-trans-Δ2-PGE1 methyl ester in 0.8 g Witepsol S-52) were 100–350 pg/ml i.e. in the same order of magnitude as earlier seenf ro 16,16-dimethyl-PGE2.  相似文献   

2.
Serum relaxin was estimated in 11 women during termination of first-trimester pregnancy with 16,16-dimethyl-trans-Δ2-PGE1 methyl ester (16 DM-PGE1). Vaginal administration of 16 DM-PGE1 was associated with a significant increase in serum relaxin.  相似文献   

3.
A method is described for measuring a trimethyl prostaglandin E2 analog, TM-PGE2, in human plasma. Trideuterated and monofluorinated analogs of TM-PGE2 are added to plasma as internal standard and carrier, respectively. The plasma is adjusted to pH 3.0 and is extracted with a mixture of benzene—dichloromethane (9:1). The residue, following removal of the extracting solvent, is reacted consecutively with pentafluorobenzyl bromide and bistrimethylsilyltrifluoroacetamide. The excess derivatizing reagents are removed by evaporation, and an aliquot of the reconstituted residue is analyzed by capillary column gas chromatography using methane as the carrier gas. A quadrupole mass spectrometer is set to monitor in the gas chromatographic effluent the (M − C7H2F5) fragmention of TM-PGE2 (m/e 449) and trideuterated TM-PGE2 (m/e 452) generated by methane negative chemical ionization. Quantitation of unknowns is based on a comparison of the m/e 449 to m/e 452 ion ratio in each unknown to that obtained from the analysis of control plasma spiked with known amounts of TM-PGE2 and fixed amounts of internal standard and carrier. The sensitivity limit of the assay is approximately 100 pg ml−1, which is equivalent to 1 pg injected. The assay was used to measure the concentration of TM-PGE2 in the plasma of two subjects following a single 10 μg kg−1 oral dose of the drug.  相似文献   

4.
Utilizing Corey's synthesis, a variety of prostaglandins (PGs) with a modified ω-side chain were prepared. The 16,16-dimethyl-oxa-alkyl analogues of PGA2 had potent antihypertensive activity. HR 466 (16,16-dimethyl-18-oxa-PGA2), the best compound out of this series was active for 5–6 hours after oral administration of 0,1 mg/kg to conscious renal hypertensive dogs. The corresponding analogues of PGE2 were also potent antihypertensive compounds, but were much more spasmogenic. Structural variations within the trans-Δ2-11-deoxy-PGE1-series, in both side chains, gave HR 601 (trans-Δ2-15α-acetoxy-16,16-dimethyl-18-oxa-11-deoxy-PGE1-methylester) which was orally active in the hypertensive dog with similar activity to HR 466.  相似文献   

5.
A monoclonal antibody against cis-3-hexen-1-ol was prepared and used to separate and/or concentrate Δ17-6-keto-prostaglandin F1α (PGF1α) in the human sera. cis-3-Hexen-1-ol was conjugated with the human serum albumin (HSA) according to the N-succinimidylester method and hyperimmunized to BALB/c mouse. The monoclonal afntibodies were obtained from hybridoma clones established by a fusion between SP2/0-Ag14-k13 mouse myeloma cells and splenocytes of a mouse. A monoclonal antibody, named 4G9-12B, recognized the epitope characteristic for ω3-olefin structure. The 4G9-12B antibody became more specific for Δ17-6-keto-PGF1α than 6-keto-PGF1α by applying inhibition ELISA using amino-residue coating plates. Using the prepared immunoaffinity columns of this antibody, Δ17-6-keto-PGF1α was clearly detected in 6 pg/ml of the human blood sera by GC/MS analysis. These results suggest that the monoclonal antibody to the partial structure of trienoic prostanoid, ω3-olefin unit, and that its immunoaffinity columns are useful in separating and concentrating Δ17-6-keto-PGF1α in the human blood or urine.  相似文献   

6.
A gas chromatography–electron capture mass spectrometry assay has been developed for the histamine H3 receptor agonist, Nα-methylhistamine (Nα-MH). The assay is linear from 50 pg–10 ng, with a limit of detection of 50 pg/ml for gastric juice and plasma, and 50 pg/sample for bacteria (107–108 CFU) and gastric tissue (5–10 mg wet weight). The limits of quantification are 100 pg/ml for gastric juice (%RSD=1.4) and plasma (%RSD=9.4), and 100 pg/sample for bacteria (%RSD=3.9) and tissue (%RSD=5.8). Nα-MH was not present in human plasma, but low levels (1.4 ng/ml and 0.4 ng/ml) were detected in two samples of human gastric juice obtained from patients infected with Helicobacter pylori.  相似文献   

7.
Formation of {3H}-PGF and {3H}-13,14,dihydro-15-keto-PGF from {3H}-PGE2 by the supernatant of uterine homogenates from estrous and ovariectomized rats, was studied, using the reaction system PGE2 + NADPH + {3H}-PGE2 + supernatant. Enzymatic conversion was lower in uterine supernatants from spayed rats than in uterine homogenates of rats at natural estrus.Spayed animals were injected with progesterone (P) or with estradiol-17-β (E0) at a dose of 1.0 or 50.0 ug. Conversion of {3H}-PGF to {3H}-PGE2 or to {3H}-13,14,dihydro-15-keto-PGF did not differ in control ovariectomized or ovariectomized rats receiving P or 1.0 ug E0. However, 50.0 ug E0 induced a significant oversion after 30 (P < 0.01) and 60 (P < 0.001) min of incubation.It is concluded that E0, at the 50.0 ug dose, but not the 1.0 ug dose of E0, nor progesterone, stimulated conversion of {3H}-PGE2 into {3H}-PGF or {3H}-13, 14,dihydro-15-keto-PGF, presumably through the activity of the enzyme PGE2-9-keto-reductase.  相似文献   

8.
The synthesis and pharmacology of 15 1-deoxy-Δ8-THC analogues, several of which have high affinity for the CB2 receptor, are described. The deoxy cannabinoids include 1-deoxy-11-hydroxy-Δ8-THC (5), 1-deoxy-Δ8-THC (6), 1-deoxy-3-butyl-Δ8-THC (7), 1-deoxy-3-hexyl-Δ8-THC (8) and a series of 3-(1′,1′-dimethylalkyl)-1-deoxy-Δ8-THC analogues (2, n=0–4, 6, 7, where n=the number of carbon atoms in the side chain−2). Three derivatives (1719) of deoxynabilone (16) were also prepared. The affinities of each compound for the CB1 and CB2 receptors were determined employing previously described procedures. Five of the 3-(1′,1′-dimethylalkyl)-1-deoxy-Δ8-THC analogues (2, n=1–5) have high affinity (Ki=<20 nM) for the CB2 receptor. Four of them (2, n=1–4) also have little affinity for the CB1 receptor (Ki=>295 nM). 3-(1′,1′-Dimethylbutyl)-1-deoxy-Δ8-THC (2, n=2) has very high affinity for the CB2 receptor (Ki=3.4±1.0 nM) and little affinity for the CB1 receptor (Ki=677±132 nM).
Scheme 3. (a) (C6H5)3PCH3+ Br, n-BuLi/THF, 65°C; (b) LiAlH4/THF, 25°C; (c) KBH(sec-Bu)3/THF, −78 to 25°C then H2O2/NaOH.  相似文献   

9.
Yeast cytochrome c peroxidase (CCP) efficiently catalyzes the reduction of H2O2 to H2O by ferrocytochrome c in vitro. The physiological function of CCP, a heme peroxidase that is targeted to the mitochondrial intermembrane space of Saccharomyces cerevisiae, is not known. CCP1-null-mutant cells in the W303-1B genetic background (ccp1Δ) grew as well as wild-type cells with glucose, ethanol, glycerol or lactate as carbon sources but with a shorter initial doubling time. Monitoring growth over 10 days demonstrated that CCP1 does not enhance mitochondrial function in unstressed cells. No role for CCP1 was apparent in cells exposed to heat stress under aerobic or anaerobic conditions. However, the detoxification function of CCP protected respiring mitochondria when cells were challenged with H2O2. Transformation of ccp1Δ with ccp1W191F, which encodes the CCPW191F mutant enzyme lacking CCP activity, significantly increased the sensitivity to H2O2 of exponential-phase fermenting cells. In contrast, stationary-phase (7-day) ccp1Δ-ccp1W191F exhibited wild-type tolerance to H2O2, which exceeded that of ccp1Δ. Challenge with H2O2 caused increased CCP, superoxide dismutase and catalase antioxidant enzyme activities (but not glutathione reductase activity) in exponentially growing cells and decreased antioxidant activities in stationary-phase cells. Although unstressed stationary-phase ccp1Δ exhibited the highest catalase and glutathione reductase activities, a greater loss of these antioxidant activities was observed on H2O2 exposure in ccp1Δ than in ccp1Δ-ccp1W191F and wild-type cells. The phenotypic differences reported here between the ccp1Δ and ccp1Δ-ccp1W191F strains lacking CCP activity provide strong evidence that CCP has separate antioxidant and signaling functions in yeast.  相似文献   

10.
A silver ion-loaded microparticulate cation-exchange resin column has been used for high-performance liquid chromatographic (HPLC) separation of the p-nitrophenacyl esters of several series of closely related prostaglandins: 8-iso-PGE2, 11-epi-PGE2, 5-trans-PGE2, PGE2, PGF, PGE1, and PGF, PGA2 and PGB2; 15 (R)-methyl-PGE2 and 15 (S)-methyl-PGE2; 5-trans-PGA2 and PGA2; and 5-trans-PGF and PGF. The properties of this column are compared with those of silica-gel and reversed-phase columns.  相似文献   

11.
Prostaglandins are well known for their ability to stimulate contraction in gastrointestinal smooth muscle, yet very little information is available on how their activity affects propulsion . Thus, studies were undertaken to determine the effect of various prostaglandins on qastric emptying (GE) and small intestinal transit (SIT) in unanesthetized fasted rats. Rats were treated with intravenous, subcutaneous, or oral PGF2α, PGE2, or 16,16 dimethyl PGE2 at various doses, followed 1 (intravenous), 20 (subcutaneous) or 10 (oral) mins later by intragastric 51Cr oxide in black ink. Forty-five mins later, rats were sacrificed by CO2 asphyxiation, the pylorus clamped, and the gut excised. SIT was expressed as the percent of intestinal length traveled by the most distal portion of ink. GE was expressed as the percent of the 51Cr emptied into the intestines. If GE was affected by prostaglandin treatment, the experiments were repeated with rats pre-implanted with duodenal cannula. This preparation allowed the visual transit marker to be deposited directly into the dueodenum, thus avoiding acceleration or delay of SIT caused by fluctuations in GE. The results of these studies show that: (1) intravenous 16,16 dimethyl PGE2 (5–50 μg/kg), but not PGF2α or PGE2, accelerates GE and delays SIT; (2) oral prostaglandin administration increases SIT; (3) oral 16,16 dimethyl PGE2 delays GE; (4) subcutaneous 16,16 dimethyl PGE2 accelerates, has no effect upon, or delays GE depending upon dose, but accelerates SIT at all doses tested; and (5) subcutaneous PGE2 accelerates SIT while PGF2α does not. Thus, the effect of prostaglandins on GE and SIT depends upon the dosage and route of administration as well as type of prostaglandin used.  相似文献   

12.
We investigate the effect of the prostaglandin D2 metabolite Δ12−PGD2 (9−Deoxy−Δ9, Δ12−13,14-dihydroprostaglandin D2) on collagen synthesis in human osteoblast. Δ12-PGJ2 at 10−5M enhanced collagen synthesis in the presence of 2 mM α-glycerophosphate-2Na. The stimulative effect appeared as early as 3 days after addition and continued until 22 days. The enhancement of type I collagen synthesis was confirmed by polyacrylamide gel electrophoresis. The potency was the same as 101t-8M 1 α, 25 dihydroxy vitamine D3 (1,25(OH)2D3). Northern blot analysis showed that 10−5M Δ 12-PGD2 and 10−8M 1,25(OH)2D3 enhanced the transcribtion of type 1 procollagen (α1) mRNA levels in osteoblasts.  相似文献   

13.
PGJ2 and Δ12PGJ2 (1 μM to 30 μm) inhibited the growth of human astrocytoma cells (1321N1) in a time-dependent manner within 48 hrs, determined by [3H]thymidine incorporation into acid-insoluble fraction or amounts of protein. The EC50 values for PGJ2 and Δ12PGJ2 were approximately 8 μM and 6 μM, respectively. [3H]Thymidine incorporation to acid insoluble fraction was inhibited by these PGs within 1 hr, indicating that these PGs rapidly affect cell functions. Although it has been reported that an increase in cyclic AMP inhibits cell growth, PGJ2 and Δ12PGJ2, but not PGE1, reduced isoproterenol (10 μM)-induced accumulation of cyclic AMP, suggesting that PGJ2 and Δ12PGJ2 may disturb adenylate cyclase system, which might be independent on cell growth. On the other hand, these PGs inhibited the incorporation of [3H]inositol into phospholipid fraction within 6 hrs. Furthermore, PGJ2 and Δ12PGJ2 inhibited carbachol- and/or histamine-induced accumulation of inositol phosphates with a similar dose-dependency to their inhibitions of cell growth. In membrane preparations, however, PGJ2 and Δ12PGJ2 failed to inhibit GTPγS (10 μM)- nor Ca2+ (1mM)-induced accumulation of inositol phosphate. The site of PGJ2 or Δ12PGJ2 in inhibition of inositol phosphate accumulation would not be phospholipase C nor a putative GTP binding protein involved in activation of phospholipase C. The present results indicate that PGJ2 and Δ12PGJ2 inhibit cell growth in human astrocytoma cells and the inhibition of phosphoinositide turnover by these PGs might be involved in the inhibition of cell growth.  相似文献   

14.
To study the precise mechanism of cytotoxic activity of PGD2 or Δ12-PGJ2 (a biological active metabolite of PGD2), we examined the effect of various compounds on PGD2 or Δ12-PGJ2 cytottoxic, using a human neuroblastoma cell line (NCG). Cycloheximide (CHM) specifically protected PGD2 cytotoxicity on NCG cells. When Δ12-PGJ2 was tested, CHM exhibited a similar rescue effect. Puromycin, mitomycin C, and α-amanitin did not affect PGD2 or Δ12-PGJ2 cytotoxicity. Emetine showed a variable and no consistent rescue effect CHM may have been active at the primary site where PGD2 or Δ12-PGJ2 exerts its cytotoxicity. This is the first report indicating that CHM reduces the cytotoxicity induced by PGD2 or Δ12-PGJ2.  相似文献   

15.
Statistical analysis reveals that the set of differences between the secondary shifts of the α- and β-carbons for residues i of a protein (Δδ13Cαi- Δδ13Cβi) provides the means to detect and correct referencing errors for 1H and 13C nuclei within a given dataset. In a correctly referenced protein dataset, linear regression plots of Δδ13Cαi,Δδ13Cβi, or Δδ1Hαi vs. (Δδ13Cαi- Δδ13Cβi) pass through the origin from two directions, the helix-to-coil and strand-to-coil directions. Thus, linear analysis of chemical shifts (LACS) can be used to detect referencing errors and to recalibrate the 1H and 13C chemical shift scales if needed. The analysis requires only that the signals be identified with distinct residue types (intra-residue spin systems). LACS allows errors in calibration to be detected and corrected in advance of sequence-specific assignments and secondary structure determinations. Signals that do not fit the linear model (outliers) deserve scrutiny since they could represent errors in identifying signals with a particular residue, or interesting features such as a cis-peptide bond. LACS provides the basis for the automated detection of such features and for testing reassignment hypotheses. Early detection and correction of errors in referencing and spin system identifications can improve the speed and accuracy of chemical shift assignments and secondary structure determinations. We have used LACS to create a database of offset-corrected chemical shifts corresponding to nearly 1800 BMRB entries: 300 with and 1500 without corresponding three-dimensional (3D) structures. This database can serve as a resource for future analysis of the effects of amino acid sequence and protein secondary and tertiary structure on NMR chemical shifts.Supplementary material to this paper is available in electronic form at http://dx.doi.org/10.1007/s10858-005-1717-0  相似文献   

16.
The present clinical trials revealed that 16,16-Dimethyl-trans-δ2-PGE1 methyl ester in the form of vaginal suppositories is highly effective in inducing mid-trimester termination of pregnancies. It also showed that prior treatment with laminaria and metreurynter may enhance the success rate while reducing the incidence and severity of side effects. It is easy and safe to use clinically, with minimal side effects, and in our series, revealed no deleterious effects on ensuing reproductive physiology. However, the definite mechanism involved in the action of this new analogue to cause myometrial contractions is still not completely understood, and requires further intensive investigation.  相似文献   

17.
The aggregation of proteins is believed to be intimately connected to many neurodegenerative disorders. We recently reported an “Ockham's razor”/minimalistic approach to analyze the kinetic data of protein aggregation using the Finke–Watzky (F–W) 2-step model of nucleation (A → B, rate constant k1) and autocatalytic growth (A + B → 2B, rate constant k2). With that kinetic model we have analyzed 41 representative protein aggregation data sets in two recent publications, including amyloid β, α-synuclein, polyglutamine, and prion proteins (Morris, A. M., et al. (2008) Biochemistry 47, 2413-2427; Watzky, M. A., et al. (2008) Biochemistry 47, 10790–10800). Herein we use the F–W model to reanalyze protein aggregation kinetic data obtained under the experimental conditions of variable temperature or pH 2.0 to 8.5. We provide the average nucleation (k1) and growth (k2) rate constants and correlations with variable temperature or varying pH for the protein α-synuclein. From the variable temperature data, activation parameters ΔG, ΔH, and ΔS are provided for nucleation and growth, and those values are compared to the available parameters reported in the previous literature determined using an empirical method. Our activation parameters suggest that nucleation and growth are energetically similar for α-synuclein aggregation (ΔGnucleation = 23(3) kcal/mol; ΔGgrowth = 22(1) kcal/mol at 37 °C). From the variable pH data, the F–W analyses show a maximal k1 value at pH ~ 3, as well as minimal k1 near the isoelectric point (pI) of α-synuclein. Since solubility and net charge are minimized at the pI, either or both of these factors may be important in determining the kinetics of the nucleation step. On the other hand, the k2 values increase with decreasing pH (i.e., do not appear to have a minimum or maximum near the pI) which, when combined with the k1 vs. pH (and pI) data, suggest that solubility and charge are less important factors for growth, and that charge is important in the k1, nucleation step of α-synuclein. The chemically well-defined nucleation (k1) rate constants obtained from the F–W analysis are, as expected, different than the 1/lag-time empirical constants previously obtained. However, k2 × [A]0 (where k2 is the rate constant for autocatalytic growth and [A]0 is the initial protein concentration) is related to the empirical constant, kapp obtained previously. Overall, the average nucleation and average growth rate constants for α-synuclein aggregation as a function of pH and variable temperature have been quantitated. Those values support the previously suggested formation of a partially folded intermediate that promotes aggregation under high temperature or acidic conditions.  相似文献   

18.
19.
The kinetics of substitution reactions of [η-CpFe(CO)3]PF6 with PPh3 in the presence of R-PyOs have been studied. For all the R-PyOs (R = 4-OMe, 4-Me, 3,4-(CH)4, 4-Ph, 3-Me, 2,3-(CH)4, 2,6-Me2, 2-Me), the reactions yeild the same product [η5-CpFe(CO)2PPh3]PF6, according to a second-order rate law that is first order in concentrations of [η5-CpFe(CO)3]PF6 and of R-PyO but zero order in PPh3 concentration. These results, along with the dependence of the reaction rate on the nature of R-PyO, are consistent with an associative mechanism. Activation parameters further support the bimmolecular nature of the reactions: ΔH = 13.4 ± 0.4 kcal mol−1, ΔS = −19.1 ± 1.3 cal k−1 mol−1 for 4-PhPyO; ΔH = 12.3 ± 0.3 kcal mol−1, ΔS = 24.7 ±1.0 cal K−1 mol−1 for 2-MePyO. For the various substituted pyridine N-oxides studied in this paper, the rates of reaction increase with the increasing electron-donating abilities of the substituents on the pyridine ring or N-oxide basicities, but decrease with increasing 17O chemical shifts of the N-oxides. Electronic and steric factors contributing to the reactivity of pyridine N-oxides have been quantitatively assessed.  相似文献   

20.
To gain the structure–activity relationship of Δ1-androstenediones (Δ1-ADs) as mechanism-based inactivator of aromatase, series of 2-alkyl- and 2-alkoxy-substitiuted Δ1-ADs (6 and 9) as well as 2-bromo-Δ1-AD (14) were synthesized and tested. All of the inhibitors examined blocked aromatase in human placental microsomes in a competitive manner. In a series of 2-alkyl-Δ1-ADs (6), n-hexyl compound 6f was the most powerful inhibitor with an apparent Ki value of 31 nM. The inhibitory activities of 2-alkoxy steroids 9 decreased in relation to length of the alkyl chain up to n-hexyloxy group (Ki: 95 nM for methoxy 9a). All of the alkyl steroids 6 along with the alkoxy steroid 9, except for the ethyl and n-propyl compounds 6b and 6c, caused a time-dependent inactivation of aromatase. The inactivation rates (kinact: 0.020–0.084 min−1) were comparable to that of the parent compound Δ1-AD. The inactivation was prevented by the substrate AD, and no significant effect of l-cysteine on the inactivation was observed in each case. The results indicate that the 2-hexyl compound 6f act as the most powerful mechanism-based inactivator of aromatase among Δ1-AD analogs and may be submitted to the preclinical study in estrogen-dependent breast cancer.  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