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.
CRTH2 is a recently described chemoattractant receptor for the prostaglandin, PGD(2), expressed by Th2 cells, eosinophils and basophils, and believed to play a role in allergic inflammation. Here we describe the potency of several PGD(2) metabolites at the receptor to induce cell migration and activation. We report for the first time that the PGD(2) metabolite, 9alpha,11beta-PGF(2), and its stereoisomer, PGF(2alpha), are CRTH2 agonists. 9alpha,11beta-PGF(2) is a major metabolite produced in vivo following allergen challenge, whilst PGF(2alpha) is generated independently of PGD synthetase, with implications for CRTH2 signalling in the presence or absence of PGD(2) production.  相似文献   

2.
Tritium-labelled prostaglandin D2 (PGD2) was administered to normal volunteers by either intravenous infusion or inhalation in order to establish which metabolites of PGD2 are initially found in human plasma. Inhaled PGD2 was rapidly absorbed from the airways, as indicated by the rapid appearance of tritium in the plasma. Metabolites chromatographically similar to 9 alpha,11 beta-PGF2 and 13,14-dihydro-15-keto-9 alpha,11 beta-PGF2 were found after both routes of administration. At later time points, other unidentified compounds were present. Only after intravenous infusion was there evidence of metabolites with 9 alpha,11 alpha stereochemistry of the ring hydroxyl functions. In human lung, 9 alpha,11 beta-PGF2 was metabolized in the presence of NAD+ to compounds tentatively identified by gas chromatography/mass spectrometry (GC/MS) as 15-keto-9 alpha,11 beta-PGF2 and 13,14-dihydro-15-keto-9 alpha,11 beta-PGF2. Thus, after 11-ketoreductase-dependent metabolism of PGD2 to the biologically active compound 9 alpha,11 beta-PGF2, further metabolism probably proceeds by the combined action of 15-hydroxyprostaglandin dehydrogenase/15-ketoprostaglandin-delta 13-reductase (15-PGDH/delta 13R). Both 9 alpha,11 beta-PGF2 and its 13,14-dihydro-15-keto metabolite may be useful analytes for the measurement of PGD2 turnover, and may therefore prove to be important in understanding the pathophysiological significance of this putative mediator.  相似文献   

3.
Measurements of the prostaglandin (PGD2) metabolite 9 alpha, 11 beta-PGF2 in unextracted urine performed by enzyme immunoassay (EIA) were compared with values obtained by negative chemical ionisation gas chromatography-mass spectrometry (NCI GC-MS). Values determined by NCI GC-MS were in the same range but consistently lower than those obtained by EIA, suggesting that other endogenous compounds could be contributing to the immunoreactivity. Isoprostanes were generated by autoxidation of arachidonic acid and the 9 alpha, 11 beta-PGF2 antibody demonstrated less than 0.7% crossreactivity to the mix, making it unlikely that isoprostanes in urine interfere with quantification of 9 alpha, 11 beta-PGF2 by EIA. This was further supported by the 70% reduction in immunoreactive material measured in urine after three days treatment in a healthy volunteer with the cyclooxygenase inhibitor ibuprofen. Purification of urine samples by reverse phase high-performance liquid chromatography (HPLC) revealed the presence of two immunoreactive compounds in addition to 9 alpha, 11 beta-PGF2. The compounds were identified as dinor compounds by NCI GC-MS. One of the compounds was identical to 9 alpha, 11 beta-2,3-dinor-PGF2 which was generated by beta-oxidation of 9 alpha, 11 beta-PGF2 and identified by electron impact (EI)-GC-MS. In conclusion, urinary 9 alpha, 11 beta-PGF2 concentrations measured by EIA represent the sum of 9 alpha, 11 beta-PGF2 and two isomers of its dinor metabolite. Thus, the direct EIA is fast, sensitive and sufficiently specific to monitor activation of the PGD2 pathway, thereby providing a valuable clinical tool to assess the status of mast cell activation in vivo.  相似文献   

4.
Prostaglandin (PG) D(2) ethanolamide (prostamide D(2)) was reduced to 9alpha,11beta-PGF(2) ethanolamide (9alpha,11beta-prostamide F(2)) by PGF synthase, which also catalyzes the reduction of PGH(2) and PGD(2) to PGF(2alpha) and 9alpha,11beta-PGF(2), respectively. These enzyme activities were measured by a new method, the liquid chromatographic-electrospray ionization-mass spectrometry (LC/ESI/MS) technique, which could simultaneously detect the substrate and all products. PGF(2alpha), 9alpha,11beta-PGF(2), PGD(2), PGH(2), 9alpha,11beta-prostamide F(2), and prostamide D(2) were separated on a TSKgel ODS 80Ts column, ionized by electrospray, and detected in the negative mode. Selected ion monitoring (SIM) of m/z 353 ([M-H](-)), 353 ([M-H](-)), 351 ([M-H](-)), 333 ([M-H-H(2)O](-)), 456 ([M+59](-)), and m/z 358 ([M-37](-)) was used for quantifying PGF(2alpha), 9alpha,11beta-PGF(2), PGD(2), PGH(2), 9alpha,11beta-prostamide F(2), and prostamide D(2), respectively. The detection limit for PGF(2alpha) and 9alpha,11beta-PGF(2) was 0.01pmol; that for PGH(2) and PGD(2), 0.1pmol; and that for prostamide D(2) and 9alpha,11beta-prostamide F(2), 0.5 and 0.03pmol, respectively. The LC/ESI/MS technique for measuring PGF synthase activity showed higher sensitivity than other methods. Using this method, we found that Bimatoprost, the ethyl amide analog of 17-phenyl-trinor PGF(2alpha) and an anti-glaucoma agent, inhibited all three reductase activities of PGF synthase when used at a low concentration. These results suggest that Bimatoprost also behaves as a potent PGF synthase inhibitor in addition to having prostamide-like activity.  相似文献   

5.
Nicotinic acid (niacin) is a B vitamin which is also a potent hypolipidemic agent. However, intense flushing occurs following ingestion of pharmacologic doses of niacin which greatly limits its usefulness in treating hyperlipidemias. Previous studies have demonstrated that niacin-induced flushing can be substantially attenuated by pre-treatment with cyclooxygenase inhibitors, suggesting that the vasodilation is mediated by a prostaglandin. However, the prostaglandin that presumably mediates the flush has not been conclusively determined. In this study we report the finding that ingestion of niacin evokes the release of markedly increased quantities of PGD2 in vivo in humans. PGD2 release was assessed by quantification of the PGD2 metabolite, 9 alpha, 11 beta-PGF2, in plasma by gas chromatography mass spectrometry. Following ingestion of 500 mg of niacin in three normal volunteers, intense flushing occurred and plasma levels of 9 alpha, 11 beta-PGF2 were found to increase dramatically by 800, 430, and 535-fold. Levels of 9 alpha, 11 beta-PGF2 reached a maximum between 12 and 45 min. after ingesting niacin and subsequently declined to near normal levels by 2-4 hours. Levels of 9 alpha, 11 beta-PGF2 in plasma correlated with the intensity and duration of flushing that occurred in the 3 volunteers. Release of PGD2 was not accompanied by a release of histamine which was assessed by quantification of plasma levels of the histamine metabolite, N tau-methylhistamine. This suggests that the origin of the PGD2 release is not the mast cell. Only a modest increase (approximately 2-fold) in the urinary excretion of the prostacyclin metabolite, 2,3-dinor-6-keto-PGF1 alpha, occurred following ingestion of niacin and no increase in the excretion of the major urinary metabolite of PGE2 was found. These results indicate that the major vasodilatory PG released following ingestion of niacin is PGD2. The fact that markedly increased quantities of PGD2 are released suggests that PGD2 is the mediator of niacin-induced vasodilation in humans.  相似文献   

6.
The metabolic transformation of tritium-labeled prostaglandin D2 ([3H]PGD2) was investigated in the isolated Tyrode's-perfused rabbit liver. One major product was isolated and identified in the perfusate as a new prostanoid. The structure of this metabolite was further confirmed by gas chromatography-mass spectrometry and chemical methods to be 9 alpha,11 beta,15-L-trihydroxyprosta-5-cis, 13-trans-dienoic acid, namely (9 alpha,11 beta-PGF2). This new prostanoid was found to be an inhibitor of platelet aggregation and to cause constriction of canine coronary artery strips. These results suggested that on passage through the hepatic circulation exogenous PGD2 is converted to 9 alpha,11 beta-PGF2, the latter having a biological profile which differs from that of PGD2 and PGF2 alpha.  相似文献   

7.
This paper describes a new iodine-125 radioimmunoassay of 9alpha ,11beta-PGF2, and its use for the determination of urinary 9alpha,11beta-prostaglandin F2 after a selective one-step solid-phase extraction. The newly reported immunoassay is based on the use of 125I-tyrosyl methyl ester derivative of 9alpha,11beta-PGF2 and specific polyclonal antibody raised in rabbits.The assay detected as lowas 0.85 pg/tube 9alpha,11beta-PGF2, and the antibodyshowed lessthan 0.01 cross-reaction with PGF-ring metabolites (e.g., 8-iso-PGF2alpha, PGF2alpha 2,3-dinor-6-keto-PGF1alpha, and 5 more PGF-ring compounds). Both the intra-assay, and inter-assay CVs were lessthan 20% for internal controls containing low, medium and high concentrations of 9alpha,11beta-PGF2. Immuno-HPLC analysis showed a very low ratio of specific immunoreactivity in both non-extracted urine (6.5%), and in urine extracted on C18-silicacartridge (14.8%). By contrast, approximately 80% specific immunoreactivity could be achieved by using C2-silicaas the sorbent, acetonitrile: water (15:85, v/v) as wash solvent, and ethyl acetate as eluent of 9alpha,11beta-PGF2.This extraction procedure enabled a reasonably high extraction efficiency of 80.4 +/- 0.855 (mean +/- SEM, n=82), as determined by 3H-9alpha,11beta-PGF2. The new SPE/RIA method was applied for the determination of urinary 9alpha,11beta-PGF2 values in 50 healthy human volunteers. For the concentration and for the excretion rate 37.52 +/- 4.61 pg/ml (mean +/- SEM), and 3.50 + 0.35 ng/mmol creatinine (mean +/- SEM), respectively, was measured.The specificity of the SPE/RIA method was supported by the observed 69% decrease in 9alpha, 11beta-PGF2 excretion rate after acetylsalicylic acid treatment. The effect of nicotinic acid, a PGD2-stimulatory agent, was monitored by the urinary excretion of 9alpha ,11beta-PGF2 in 6 patients, by using the new SPE/RIA method. In patients responding with flushing symptoms nicotinic acid induced an increase of the urinary excretion of 9alpha,11beta-PGF2 in the range between 11% and 187%. In summary, the combination of the newly developed specific [125I] radioimmunoassay with solid-phase extraction on C2-silica cartridges enables the specific, sensitive, and reliable determination of 9alpha,11beta-PGF2 in human urine without the need for further laborious chromatographic purification before radioimmunoassay.  相似文献   

8.
The full-length bovine lung prostaglandin(PG) F synthase cDNA was constructed from partial cDNA clones and ligated into bacterial expression vector pUC8 to develop expression plasmid pUCPF1. This plasmid permitted the synthesis of bovine lung PGF synthase in Escherichia coli. The recombinant bacteria overproduced a 36-KDa protein that was recognized by anti-PGF synthase antibody, and the expressed protein was purified to apparent homogeneity. The expressed protein reduced not only carbonyl compounds including PGD2 and phenanthrenequinone but also PGH2; and the Km values for phenanthrenequinone, PGD2, and PGH2 of the expressed protein were 0.1, 100, and 8 microM, respectively, which are the same as those of the bovine lung PGF synthase. The protein produced PGF2 alpha from PGH2, and 9 alpha, 11 beta-PGF2 from PGD2 at different active sites. Moreover, the structure of the purified protein from Escherichia coli was essentially identical to that of the native enzyme in terms of C-terminal sequence, sulfhydryl groups, and CD spectra except that the nine amino acids provided by the lac Z' gene of the vector were fused to the N-terminus. These results indicate that the expressed protein is essentially identical to bovine lung PGF synthase. We confirmed that PGF synthase is a dual function enzyme catalyzing the reduction of PGH2 and PGD2 on a single enzyme and that it has one binding site for NADPH.  相似文献   

9.
Prostaglandin D(2) (PGD(2)) is a cyclooxygenase (COX) product of arachidonic acid that activates D prostanoid receptors to modulate vascular, platelet, and leukocyte function in vitro. However, little is known about its enzymatic origin or its formation in vivo in cardiovascular or inflammatory disease. 11,15-dioxo-9alpha-hydroxy-2,3,4,5-tetranorprostan-1,20-dioic acid (tetranor PGDM) was identified by mass spectrometry as a metabolite of infused PGD(2) that is detectable in mouse and human urine. Using liquid chromatography-tandem mass spectrometry, tetranor PGDM was much more abundant than the PGD(2) metabolites, 11beta-PGF(2alpha) and 2,3-dinor-11beta-PGF(2alpha), in human urine and was the only endogenous metabolite detectable in mouse urine. Infusion of PGD(2) dose dependently increased urinary tetranor PGDM > 2,3-dinor-11beta-PGF(2alpha) > 11beta-PGF(2alpha) in mice. Deletion of either lipocalin-type or hemopoietic PGD synthase enzymes decreased urinary tetranor PGDM. Deletion or knockdown of COX-1, but not deletion of COX-2, decreased urinary tetranor PGDM in mice. Correspondingly, both PGDM and 2,3-dinor-11beta-PGF(2alpha) were suppressed by inhibition of COX-1 and COX-2, but not by selective inhibition of COX-2 in humans. PGD(2) has been implicated in both the development and resolution of inflammation. Administration of bacterial lipopolysaccharide coordinately elevated tetranor PGDM and 2,3-dinor-11beta-PGF(2alpha) in volunteers, coincident with a pyrexial and systemic inflammatory response, but both metabolites fell during the resolution phase. Niacin increased tetranor PGDM and 2,3-dinor-11beta-PGF(2alpha) in humans coincident with facial flushing. Tetranor PGDM is an abundant metabolite in urine that reflects modulated biosynthesis of PGD(2) in humans and mice.  相似文献   

10.
Prostaglandin H(2) (PGH(2)) formed from arachidonic acid is an unstable intermediate and is efficiently converted into more stable arachidonate metabolites by the action of enzymes. Prostaglandin F synthase (PGFS) has dual catalytic activities: formation of PGF(2)(alpha) from PGH(2) by the PGH(2) 9,11-endoperoxide reductase activity and 9alpha,11beta-PGF(2) (PGF(2)(alphabeta)) from PGD(2) by the PGD(2) 11-ketoreductase activity in the presence of NADPH. Bimatoprost (BMP), which is a highly effective ocular hypotensive agent, is a PGF(2)(alpha) analogue that inhibits both the PGD(2) 11-ketoreductase and PGH(2) 9,11-endoperoxide reductase activities of PGFS. To examine the catalytic mechanism of PGH(2) 9,11-endoperoxide reductase, a crystal structure of PGFS[NADPH + BMP] has been determined at 2.0 A resolution. BMP binds near the PGD(2) binding site, but the alpha- and omega-chains of BMP are locate on the omega- and alpha-chains of PGD(2), respectively. Consequently, the bound BMP and PGD(2) direct their opposite faces of the cyclopentane moieties toward the nicotinamide ring of the bound NADP. The alpha- and omega-chains of BMP are involved in H-bonding with protein residues, while the cyclopentane moiety is surrounded by water molecules and is not directly attached to either the protein or the bound NADPH, indicating that the cyclopentane moiety is movable in the active site. From the complex structure, two model structures of PGFS containing PGF(2)(alpha) and PGH(2) were built. On the basis of the model structures and inhibition data, a putative catalytic mechanism of PGH(2) 9,11-endoperoxide reductase of PGFS is proposed. Formation of PGF(2)(alpha) from PGH(2) most likely involves a direct hydride transfer from the bound NADPH to the endoperoxide of PGH(2) without the participation of specific amino acid residues.  相似文献   

11.
The accumulation of inositol phosphates (IPs) in response to prostaglandins (PGs) was studied in NG108-15 cells preincubated with myo-[3H]inositol. As a positive control, bradykinin caused accumulation of IPs transiently at an early phase (within 1 min) and continuously during a late phase (15-60 min) of incubation in the cells. PGD2 and PGF2 alpha did not significantly cause the accumulation of IPs at an early phase but significantly stimulated inositol bisphosphate (IP2) and inositol monophosphate (IP) formation at late phase of incubation. The maximum stimulation was obtained at greater than 10(-7) M concentrations of these PGs, the levels being three-and twofold for IP2 and IP1, respectively. 9 alpha, 11 beta-PGF2 has a slight effect but PGE2 and the metabolites of PGD2 and PGF2 alpha have no effect up to 10(-6)M. The effects of PGD2 and PGF2 alpha were not additive, but the effect of each PG was additive to that of bradykinin at a late phase of incubation. Inositol 1-monophosphate was mainly identified in the stimulation by 10(-5) M PGD2 and 10(-5) M PGF2 alpha, whereas both inositol 1-monophosphate and inositol 4-monophosphate were produced in the stimulation by 10(5) M bradykinin. Depletion of extracellular Ca2+ diminished the stimulatory effect of PGD2 and PGF2 alpha and late-phase effect of bradykinin, but simple Ca2+ influx into the cells by high K+, ionomycin, or A23187 failed to cause such late-phase effects. These results suggest that PGD2 and PGF2 alpha specifically stimulate hydrolysis of inositol phospholipids.  相似文献   

12.
The major eicosanoid produced within the rat liver, prostaglandin (PG) D2, wa studied for its ability to interact with the various liver cell types. It appeared that PGD2 bound specifically to parenchymal liver cells, whereas the binding of PGD2 to Kupffer and endothelial liver cells was quantitatively unimportant. Maximally 700 pg of PGD2/mg of parenchymal-cell protein could be bound by a high-affinity site (1 x 10(6) PGD2-binding sites/cell). The recognition site for PGD2 is probably a protein because trypsin treatment of the cells virtually abolished the high-affinity binding. High-affinity binding of PGD2 was a prerequisite for the induction of a metabolic effect in isolated parenchymal liver cells, i.e. the induction of glycogenolysis. High-affinity binding of PGD2 by parenchymal cells was coupled to the conversion of PGD2 into three metabolites, whereas no conversion of PGD2 by Kupffer and endothelial liver cells was noticed. The temperature-sensitivity of the conversion of PGD2 was consistent with a conversion of PGD2 on or in the vicinity of the cell membrane. One of the PGD2 metabolites could be identified as 9 alpha, 11 beta-PGF2. It can be calculated that the conversion rate of PGD2 by parenchymal liver cells exceeds the production rate of PGD2 by Kupffer plus endothelial liver cells, indicating that PGD2 is meant to exert its activity within the liver. The present finding that PGD2 formed by the non-parenchymal liver cells is recognized by a specific receptor on parenchymal liver cells and that binding, conversion and metabolic effect of PGD2 are interlinked by this receptor provides further support for the specific role of PGD2 in the intercellular communication in the liver.  相似文献   

13.
Previous studies have shown that the natural prostanoids, PGE2, PGE1 and PGF2 alpha are potent stimulators of bone resorption. In this study, we have examined the effects of alterations in the cyclopentane ring of these prostanoids for their effect on the resorptive response of cultured long bones from 19-day fetal rats as measured by the release of previously incorporated 45Ca. Indomethacin (10(-6)M) was added to minimize endogenous prostaglandin production. In this system PGE2 and PGE1, the 9 keto, 11 alpha hydroxy compounds, were approximately equally effective at concentrations of 10(-8) to 10(-6) M. The 9 alpha hydroxy, 11 alpha hydroxy compound, PGF2 alpha, was active at 10(-7) to 10(-5) M. In contrast, the 9 alpha hydroxy, 11-keto compound, PGD2, showed only a minimal stimulation of bone resorption at 10(-5) M. While these data suggested that the 11 alpha hydroxy group was important for bone resorbing activity, 11 beta PGE2 and 11-deoxy PGE1 were only slightly less potent than their physiologic counterparts. Both 9 beta, 11 alpha PGF2 and 9 alpha, 11 beta PGF2 were less potent than PGF2 alpha but did cause substantial stimulation of bone resorption and were equally effective at 10(-6) to 10(-5) M. 9 alpha, 11 beta PGF2 alpha is of particular interest since it is major metabolite of PGD2. These results suggest that the binding of prostanoids to the receptor which mediates bone resorption is affected by changes at the 9 and 11 positions of the pentane ring but do not support the hypothesis that the 11 alpha OH function is essential for this biological activity.  相似文献   

14.
Prostaglandin H(2) (PGH(2)) formed from arachidonic acid is an unstable intermediate and is efficiently converted into more stable arachidonate metabolites (PGD(2), PGE(2), and PGF(2)) by the action of three groups of enzymes. Prostaglandin F synthase (PGFS) was first purified from bovine lung and catalyzes the formation of 9 alpha,11 beta-PGF(2) from PGD(2) and PGF(2)(alpha) from PGH(2) in the presence of NADPH. Human PGFS is 3 alpha-hydroxysteroid dehydrogenase (3 alpha-HSD) type II and has PGFS activity and 3 alpha-HSD activity. Human lung PGFS has been crystallized with the cofactor NADP(+) and the substrate PGD(2), and with the cofactor NADPH and the inhibitor rutin. These complex structures have been determined at 1.69 A resolution. PGFS has an (alpha/beta)(8) barrel structure. The cofactor and substrate or inhibitor bind in a cavity at the C-terminal end of the barrel. The cofactor binds deeply in the cavity and has extensive interactions with PGFS through hydrogen bonds, whereas the substrate (PGD(2)) is located above the bound cofactor and has little interaction with PGFS. Despite being largely structurally different from PGD(2), rutin is located at the same site of PGD(2), and its catechol and rhamnose moieties are involved in hydrogen bonds with PGFS. The catalytic site of PGFS contains the conserved Y55 and H117 residues. The carbonyl O(11) of PGD(2) and the hydroxyl O(13) of rutin are involved in hydrogen bonds with Y55 and H117. The cyclopentane ring of PGD(2) and the phenyl ring of rutin face the re-side of the nicotinamide ring of the cofactor. On the basis of the catalytic geometry, a direct hydride transfer from NADPH to PGD(2) would be a reasonable catalytic mechanism. The hydride transfer is facilitated by protonation of carbonyl O(11) of PGD(2) from either H117 (at low pH) or Y55 (at high pH). Since the substrate binding cavity of PGFS is relatively large in comparison with those of AKR1C1 and AKR1C2, PGFS (AKR1C3) could catalyze the reduction and/or oxidation reactions of various compounds over a relatively wide pH range.  相似文献   

15.
Methods for the profiling of prostaglandin D2 (PGD2), prostaglandin E2 (PGE2), prostaglandin F2 alpha (PGF2 alpha), 15(S),9 alpha,11 beta-trihydroxyprosta-5Z,13E-dien-1-oic acid (9 alpha,11 beta-PGF2), 6-keto-prostaglandin F1 alpha (6kPGF1 alpha), and thromboxane B2 (TxB2) in bronchoalveolar lavage (BAL) fluids from human subjects by combined capillary gas chromatography-mass spectrometry are described. Aliquots (5 ml) of BAL fluid obtained using a standardized lavage protocol were extracted on octadecylsilyl silica cartridges after addition of 0.8 to 2.0 nanograms of tetradeuterated analogs of PGE2, PGF2 alpha, and 6kPGF1 alpha as internal standards. Eluted analytes and internal standards were prepared for vapor phase analysis by sequential reactions resulting in the formation of methyloxime-pentafluorobenzyl ester-trimethylsilyl ether derivatives. The derivatized analytes were detected by simultaneous monitoring of ions at six different masses characteristic for each of the derivatized prostanoids. The samples were of adequate purity for identification and quantitation of each of the prostanoids with detection limits of 0.1 to 0.2 picograms of each analyte per milliliter of BAL fluid. The time required for analysis of each sample was approximately 30 minutes. Standard curves of unlabeled species of the six prostanoids extracted after addition to BAL fluid were linear over a range from subpicogram to nanogram quantities. The differences between the amounts of prostanoid added and the amounts of prostanoid measured were typically less than 19%, and the intra-assay coefficients of variation for repeated measurements of a single sample were less than 20%. PGE2, PGD2, PGF2 alpha, and TxB2 were detectable in BAL fluids from normal subjects with levels of each of these compounds being less than 2.6 picograms/ml. BAL fluids from patients with lung disease presented qualitative and quantitative profiles of prostanoids markedly different than those from normal subjects. These analytical methods provide a basis for in vivo comparisons of prostanoid profiles in the lower respiratory tract of man and should be readily adaptable for use in a variety of clinical studies.  相似文献   

16.
The present study investigates the contribution of gastric mast cells on PGD2 generation in rat gastric mucosa. Cold-restraint induced stress or i.v. carbachol injection methods were used for gastric mast cell degranulation. In 19 stressed, 15 carbachol-infused and 14 control rats, gastric mast cell counts and gastric mucosa PGD2 assay were performed. Gastric mucosal content of PGF2 alpha was also determined in carbachol infused and control rats. The mean number of gastric mast cells was significantly lower in stressed and carbachol infused than in control rats. Despite these differences in gastric mast cell counts, neither PGD2 or PGF2 alpha contents in the gastric mucosa were significantly different in mast cells degranulated rats than in control animals. These results suggest another source of PGD2 in the rat gastric mucosa other than mast cells.  相似文献   

17.
Anhydrolevuglandin D2 (AnLGD2), which is produced from PGH2 by a water-induced rearrangement and subsequent dehydration, is uterotonic. However, increasing concentrations caused decreased responses of the uterine horns. AnLGD2 inhibited responses of uteri to stimulation by specific prostaglandins. PGF2 alpha was inhibited at an AnLGD2:PGF2 alpha ratio of 0.05:1 with 5 to 25 pg/ml concentrations of PGF2 alpha. The response to PGD2 was inhibited at an AnLGD2:PGD2 ratio of 0.05:1 with PGD2 concentrations of 5 to 75 pg/ml. In contrast, the uterotonic effects of PGE2 were not inhibited by AnLGD2. When AnLGD2 was added to baths with contracting uteri it inhibited contractions less if the exposure period was 5 min than if it was 10 min. The longer exposure times produced prolonged inhibition of contractile activity with bath concentrations of AnLGD2 as little as 2.5 pg/ml.  相似文献   

18.
The influence of daltroban (BM13.505; SK&F 96148), a thromboxane (Tx) A2-receptor-blocking agent, on responses to the TxA2 mimics U-46619 and U-44069 was investigated in the pulmonary vascular bed of the intact-chest cat under constant-flow conditions. Daltroban (5 mg/kg iv) had no significant effect on mean baseline vascular pressures but significantly decreased responses to the TxA2 mimics without altering responses to prostaglandin (PG) F2 alpha or PGD2 or the PGD2 metabolite 9 alpha, 11 beta-PGF2. Dose-response curves for U-46619 and U-44069 were shifted to the right in a parallel manner, and daltroban had no significant effect on responses to norepinephrine, serotonin, angiotensin II, BAY K 8644, endothelin-(ET) 1, ET-2, or platelet-activating factor (PAF). After administration of daltroban, responses to U-46619 returned to 50% of control in 90 min and responses to the PG and TxA2 precursor arachidonic acid were decreased significantly. These results suggest that daltroban selectively antagonizes TxA2-receptor-mediated responses in a competitive and reversible manner. These data provide support for the hypothesis that discrete TxA2 receptors unrelated to receptors stimulated by PGF2 alpha, PGD2, or 9 alpha, 11 beta-PGF2 are present in the pulmonary vascular bed of the cat. The present data suggest that pulmonary vasoconstrictor responses to PAF and ET peptides are not dependent on activation of TxA2 receptors in the cat.  相似文献   

19.
BACKGROUND: Prostaglandin D2 (PGD2) is released from mast cells during the allergic response. OBJECTIVE: Since PGD2 has been shown to induce nasal congestion in humans, we investigated the distribution of hematopoietic prostaglandin D synthase (PGDS) and the two PGD2 receptors, DP and CRTH2 in human nasal mucosa from healthy subjects and subjects suffering from polyposis, a severe form of chronic rhinosinusitis. METHODS: DP mRNA expression was detected by in situ hybridization while PGDS, CRTH2 and various leukocyte markers expression were revealed by immunohistochemistry. RESULTS: In the normal mucosa, PGDS was only detected in few resident mast cells while CRTH2 was undetectable. In contrast, DP receptor mRNA was detected in epithelial goblet cells, serous glands and in the vasculature. In the nasal mucosa of subjects suffering from polyposis: (1) PGDS was detected in mast cells and other large infiltrating inflammatory cells, (2) both DP mRNA and CRTH2 were detected in eosinophils and (3) CRTH2 was detected on a subset of infiltrating T cells. Although DP mRNA could not be detected in the T cells invading the nasal mucosa, it was found to be expressed in the T cells present in the lymph node and the thymus from normal individuals. CONCLUSION: This study indicates that cells capable of producing PGD2 are present in the nasal mucosa and that both PGD2 receptors, DP and CRTH2, might play a role in inflammatory disease of the upper airways.  相似文献   

20.
PGD(2) is a key mediator of allergic inflammatory diseases that is mainly synthesized by mast cells, which constitutively express high levels of the terminal enzyme involved in PGD(2) synthesis, the hematopoietic PGD synthase (H-PGDS). In this study, we investigated whether eosinophils are also able to synthesize, and therefore, supply biologically active PGD(2). PGD(2) synthesis was evaluated within human blood eosinophils, in vitro differentiated mouse eosinophils, and eosinophils infiltrating inflammatory site of mouse allergic reaction. Biological function of eosinophil-derived PGD(2) was studied by employing inhibitors of synthesis and activity. Constitutive expression of H-PGDS was found within nonstimulated human circulating eosinophils. Acute stimulation of human eosinophils with A23187 (0.1-5 μM) evoked PGD(2) synthesis, which was located at the nuclear envelope and was inhibited by pretreatment with HQL-79 (10 μM), a specific H-PGDS inhibitor. Prestimulation of human eosinophils with arachidonic acid (10 μM) or human eotaxin (6 nM) also enhanced HQL-79-sensitive PGD(2) synthesis, which, by acting on membrane-expressed specific receptors (D prostanoid receptors 1 and 2), displayed an autocrine/paracrine ability to trigger leukotriene C(4) synthesis and lipid body biogenesis, hallmark events of eosinophil activation. In vitro differentiated mouse eosinophils also synthesized paracrine/autocrine active PGD(2) in response to arachidonic acid stimulation. In vivo, at late time point of the allergic reaction, infiltrating eosinophils found at the inflammatory site appeared as an auxiliary PGD(2)-synthesizing cell population. Our findings reveal that eosinophils are indeed able to synthesize and secrete PGD(2), hence representing during allergic inflammation an extra cell source of PGD(2), which functions as an autocrine signal for eosinophil activation.  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