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.
1. Exposure of smooth muscle cells (SMC) to glyceryl trinitrate (GTN, 75-600 microM) for 30 min led to a concentration-dependent increase in nitrite (NO2-), one of the breakdown products of nitric oxide (NO). This was not affected by 30 min pretreatment of the cells with 0.5 mM of sulphobromophthalein (SBP) an inhibitor of glutathione-S-transferase (GST), by metyrapone or SKF-525A inhibitors of cytochrome P450. These experiments were confirmed by organ bath studies using rabbit aortic strips denuded of endothelium and contracted with phenylephrine. Thus, a 30 min incubation of the strips with 0.5 mM SPB, metyrapone or SKF-525A did not affect the relaxations in response to GTN (10(-10)-10(-6) M). 2. Potentiation of the anti-platelet effect of GTN (44 microM) by endothelial cells (EC, 40 x 10(3) cells) was not affected by prior incubation of EC with SBP, metyrapone or SKF-525A (all at 0.5 mM). 3. Potentiation of the antiplatelet activity of GTN (11-352 microM) by small numbers of SMC (24 x 10(3) cells) or EC (40 x 10(3) cells) treated with indomethacin (10 microM) was attenuated when the SMC or EC were treated in culture with a high concentration of GTN (600 microM) for 18 h beforehand (referred to as 'tolerant' cells). In addition, tolerant SMC produced far less NO2- than non-tolerant SMC. 4. Exposure of non-tolerant SMC or EC (10(5) cells) to GTN (200 microM) for 3 min increased (3-4 fold) the levels of guanosine 3':5'-cyclic monophosphate (cyclic GMP).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The present studies yield that all 4 metabolites of glyceryl trinitrate (Nitro Mack, GTN) cause the same typical pharmacological effects as the parent substance. Continuous infusion of 4 mg/kg/min of glyceryl 1-nitrate (G-1-N) in the conscious dog results in a drop in blood pressure which is maintained for at least 20 min after stopping the infusion. Under comparable conditions the reduction in blood pressure caused by glyceryl 2-nitrate (G-2-N) (infused at 16 mg/kg/min) decreases by only 3 mmHg. The drop in blood pressure caused by a continuous infusion with GTN (8 micrograms/kg/min) disappeared 10 to 12 min after the end of the infusion. GTN is active in rat, dog and man for 15 to 30 min. Our experiments indicate that after oral administration to the rat or dog glyceryl 1,2-dinitrate (1,2-GDN) and glyceryl 1,3-dinitrate (1,2-GDN) are active for 3 h. Earlier experiments have shown that the administration of high oral doses of 1,2-GDN or 1,3-GDN (280 mg/kg) to the rat increases the time of action to 5-6 h. This is taken as indication that the dinitrates form pharmacologically active metabolites. 1,3-GDN (70 mg/kg p.o.) and 1,2-GDN (140 mg/kg p.o.) exhibited antianginal activity in the rat for 3 or 2 h, respectively. Orally administered GTN is only active for 15 min in the rat or dog. Its pharmacological activity after this time and up to 2-3 h on the rat (280 mg/kg p.o.) and on the dog (4 mg/kg p.o.) is brought about by the dinitrate metabolites (1,2-GDN, 1,3-GDN) and after this time probably only by G-1-N.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
The hepatic transformation of glyceryl trinitrate (GTN), commonly known as nitroglycerin, was studied in subcellular fractions prepared from rabbit livers. Both the cytosolic and microsomal fractions show activity toward GTN metabolism. Moreover, the formation of glyceryl dinitrates (GDNs) seems to be governed by different enzymatic processes in the two fractions. 1,2-GDN was preferentially formed in cytosolic fractions, whereas in microsomal fractions, 1,3-GDN was the predominant product. In cytosolic fractions, increasing starting concentrations of GTN led to a decrease in both the GTN degradation rate and the GDN ratio (1,2-GDN/1,3-GDN), which was mainly accounted for by saturation of the 1,2-GDN formation pathway. Various glutathione S-transferase (GST) inhibitors affected the rate of GDN formation differentially. In cytosolic fractions, 1-chloro-2,4-dinitrobenzene and iodomethane caused no change in the GDN ratio, while sulfobromophthalein, ethacrynic acid, and p-nitrobenzyl chloride decreased the GDN ratio, suggesting that different GST isozymes are inhibited by these agents. In microsomal fractions, no dose-dependent GTN metabolism and related change in the GDN ratios could be observed. With the exception of ethacrynic acid, addition of GST inhibitors did not decrease GDN metabolite production, and even in this case, no change in the GDN ratio was observed. The results suggest that different GTN metabolic pathways are present in the liver, most likely involving different GST isozymes.  相似文献   

4.
The pharmacokinetics and bioavailability of glyceryl trinitrate (GTN, CAS 55-63-0) and of its main metabolites, i.e. 1,2-glyceryl dinitrate (1,2-GDN, CAS 621-65-8) and 1,3-glyceryl dinitrate (1,3-GDN, CAS 623-87-0), were compared during a single 24-h application of a new GTN transdermal patch (Epinitril 10, hereinafter called EPI-10) or a reference patch (hereinafter called ND-10) releasing 10 mg GTN in 24 h. The study was an open, randomized balanced cross-over study on 24 healthy male volunteers to whom the patches were applied to the antero-lateral part of the thorax in two periods separated by a 3-day wash-out. Blood samples were collected before administration, during the 24-h patch application and at 0.5, 2 and 3 h after patch removal. Assayed in plasma were GTN, 1,2-GDN and 1,3-GDN using validated GC/MS methods with stable isotope-labeled internal standards (15N3-GTN, 15N2-1,2-GDN, and 15N2-1,3-GDN). The ratios of the AUCs of GTN, 1,2-GDN and 1,3-GDN measured during application of EPI-10 or of ND-10 were within the 0.85-1.25 limits required to assess equivalence of the extent of bioavailability. The ratios of the Cmax were within said limits for the signal metabolite 1,2-GDN and only slightly below (0.78-1.16) for the parent GTN. EPI-10 can therefore be considered equivalent to ND-10 also with regard to the rate of bioavailability. Under both patches GTN reached steady-state levels after 3-6 h of patch application and remained on sustained levels during the whole 24-h application. The plasma levels of 1,2-GDN were about 6 times higher than those of GTN. The plasma levels of 1,3-GDN were similar to those of GTN. Upon removal of the patches the concentrations of the three nitrates fell to negligible values within 3 h. Both patches were well tolerated at the application site. For its small size, thinness and transparency, EPI-10 is very patient friendly, a quality that improves compliance with the therapeutic regimen.  相似文献   

5.
The tissue distribution of glyceryl trinitrate (GTN) and its two dinitrate metabolites 1,2-glyceryl dinitrate (1,2-GDN) and 1,3-glyceryl trinitrate (1,3-GDN), was studied in GTN-tolerant and nontolerant male Sprague-Dawley rats. The concentrations of GTN, 1,2-GDN, and 1,3-GDN were measured in plasma, heart, brain, liver, aortic tissue, and adipose tissue at various time points after a subcutaneous dose of GTN (50 mg/kg). At the first time point (5 hr), concentrations of GTN, 1,2-GDN, and 1,3-GDN in plasma were equal for tolerant and nontolerant rats, but the elimination rate was altered for the tolerant rats as compared with nontolerant rats. In adipose tissue, the concentration of GTN was significantly higher as compared with concentrations of the dinitrate metabolites. In contrast, the other tissues studied showed significantly higher concentrations of the GDNs when compared with GTN. The 1,3-GDN/1,2-GDN ratio decreased with time for both tolerant and nontolerant rats. This study indicates that long-term GTN administration results not only in tolerance development, but also in altered pharmacokinetics of GTN, 1,2-GDN, and 1,3-GDN. The results also show that the 1,3-GDN/1,2-GDN ratio is dependent on the GTN concentration.  相似文献   

6.
Nitroglycerin (GTN) is metabolized to 1,2-dinitroglycerin (1,2-GDN) and 1,3-dinitroglycerin (1,3-GDN) in vivo and in liver homogenates. 1,2-GDN and 1,3-GDN are converted to isomers of glyceryl mononitrate (GMN) in vivo. The denitration reactions yield inorganic nitrite (NO(-)(2)) which is oxidized to inorganic nitrate (NO(-)(3)). Denitration involves utilization of glutathione (GSH). In attempting to use the Bratton-Marshall assay for NO(-)(2) in studies of GTN metabolism in vitro, and in attempting to use Ellman's reagent for GSH in the same research, apparent concentrations of both NO(-)(2) and GSH were noticed lower than anticipated. Apparent mutual interference by NO(-)(2) and GSH in their respective assays was then found. Development of a specific liquid chromatographic method for measurement of NO(-)(2), NO(-)(3), GSH and oxidized glutathione (GSSG) permitted the study of the interaction of NO(-)(2) and GSH, which yielded NO(-)(3) and GSSG.  相似文献   

7.
In this study, we examined glyceryl trinitrate (GTN) biotransformation and cyclic GMP elevation in vascular smooth muscle before onset of GTN-induced relaxation. Isolated rabbit aortic strips (RAS) and strips of bovine pulmonary artery (BPA) and bovine pulmonary vein (BPV) were contracted submaximally and incubated with [3H]GTN. Before onset of GTN-induced vasodilation, the tissues were freeze-clamped and then analyzed for GTN, glyceryl-1,2-dinitrate (1,2-GDN), and glyceryl-1,3-dinitrate (1,3-GDN) and for cyclic GMP. Before onset of relaxation of RAS, BPA, and BPV, there was significant biotransformation of GTN to GDN and significant elevation of cyclic GMP. There was significantly greater biotransformation of GTN and elevation of cyclic GMP by BPV than by BPA incubated with the same concentration of GTN, which was temporally related with the more rapid onset of relaxation induced in BPV than in BPA. These results are consistent with the hypothesis that the magnitude of GTN biotransformation before vasodilation is the important determinant of subsequent tissue relaxation. In GTN biotransformation before vasodilation, there was preferential formation of 1,2-GDN. These data indicate that the mechanism of GTN biotransformation to 1,2-GDN is related to elevation of cyclic GMP and subsequent vasodilation.  相似文献   

8.
1,2- and 1,3-glyceryl dinitrates (1,2-GDN and 1,3-GDN) are the primary metabolites of glyceryl trinitrate, a commonly used anti-anginal agent. The goal of this study was to examine the pharmacokinetic properties of these metabolites in rats. Sprague-Dawley rats were infused intravenously with 0.25 or 2.0 micrograms min-1 of either 1,2- or 1,3-GDN for 70 min, during which steady state blood concentrations were achieved. Post-infusion blood samples were collected for 30 min. 1,2-GDN was found to possess slightly higher clearance (32.3 vs 20.8 ml min-1 kg-1) and volume of distribution (695 vs 454 ml kg-1) than 1,3-GDN; however, the two metabolites exhibited similar mean residence times (22.0 vs 21.8 min). Upon an 8-fold increase in the infusion rate, the pharmacokinetic parameters were not significantly altered for either 1,2- or 1,3-GDN. When each GDN was co-infused with an 8-fold higher dose of the other GDN, there were also no significant changes in the parameters.  相似文献   

9.
The pharmacokinetic characteristics and the bioavailability of glyceryl trinitrate (GTN, CAS 55-63-0) and of its main metabolities 1,2-glyceryl dinitrate (1,2-GDN, CAS 621-65-8) and 1,3-glyceryl dinitrate (1,3-GDN, CAS 623-87-0) during a single 24-h application of three strengths of a newly developed GTN transdermal patch (Epinitril) were investigated. The three strengths coded in this paper EPI-5, EPI-10 and EPI-15 have a nominal release rate of GTN of 5, 10 and 15 mg, respectively, in 24 h. The study was an open, randomized balanced cross-over study on 18 healthy male volunteers, to whom the patches were applied for 24 h to the antero-lateral part of the thorax in three periods, separated by an at least 3-day wash-out. Blood samples were collected before administration, during the 24-h patch application and 1, 3 and 6 h after patch removal. Assayed in plasma were GTN, 1,2-GDN and 1,3-GDN using validated GC/MS methods with stable isotope labeled internal standards (15N3-GTN, 15N2-1,2-GDN, and 15N2-1,3-GDN). GTN and its two metabolites reached the plateau already 3 h after application of the patches and remained on extended and fairly constant levels during the whole 24-h application. The plasma levels of the three nitrates were proportional to the strengths of the patches. The plasma levels of 1,2-GDN were about 6 times higher than those of GTN. The plasma levels of 1,3-GDN were similar to those of GTN. Upon removal of the patches the concentrations of the three nitrates fell to negligible values within 3 h, an important property when an intermittent GTN therapy is needed in order to avoid tolerance to the drug. The patches were well tolerated at the application site. For their good tolerability, small size and transparency, the new GTN patches are very patient friendly, a quality that improves compliance with the therapeutic regimen.  相似文献   

10.
The metabolism of glyceryl trinitrate (GTN) to nitric oxide (NO) was studied in the mouse macrophage cell line J774 and in the human monocytic cell line U937 in the absence or presence of Escherichia coli lipopolysaccharide (LPS). Two bioassay systems were used: inhibition of platelet aggregation and measurement of cGMP after stimulation by NO of guanylate cyclase in J774 cells. In addition, NO produced from GTN by cells or by cellular fractions was measured as nitrite (NO2-) one of its breakdown products. J774 cells (1.25 x 10(5) cells) treated with indomethacin (10 microM) enhanced the platelet inhibitory activity of GTN (22-352 microM) but not that of sodium nitroprusside (4 microM). This effect was abrogated by co-incubation with oxyhaemoglobin (oxyHb, 10 microM) indicating release of NO from GTN. U937 cells (up to 60 x 10(5)) did not metabolize GTN to NO. LPS (0.5 micrograms/mL for 18 hr) enhanced at least 2-fold the capacity of J774 cells but not that of U937 cells to form NO from GTN and this enhancement was attenuated when cycloheximide (10 micrograms/mL) was incubated together with LPS. In the absence of LPS stimulation, cycloheximide had no effect. Furthermore, when incubated with GTN (200 microM), J774 cells treated with LPS released more NO from GTN as indicated by a 3-fold greater increase in their level of cGMP which was prevented by oxyHb (10 microM). Incubation of J774 cells with GTN (75-600 microM) for 30 min led to a concentration-dependent increase in NO2- which was substantially reduced when the cells were boiled. The microsomal fraction was more potent than the cytosol in producing NO2- from GTN (1.2-2.4 mM). Release of NO2- from GTN by J774 cells was not affected by treating the cells with the NO synthase inhibitor, NG-monomethyl-L-arginine (MeArg, 300 microM). In J774 cells made tolerant to GTN, potentiation of the anti-platelet effects of GTN (11-352 microM) and release of NO2- from GTN was reduced. Thus, J774 cells but not U937 cells convert GTN to NO. This enzymic pathway (present mainly in the microsomal fraction of the J774 cells) is induced by LPS and is not regulated by endogenous NO released from L-Arg by the enzyme NO synthase. Furthermore, when compared to normal cells, tolerant J774 cells metabolize GTN to NO less effectively as assessed by a reduced capacity to potentiate the anti-platelet effect of GTN and to release NO2-.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

11.
This review discusses the pharmacokinetics and pharmacodynamics of glyceryl trinitrate (nitroglycerin; GTN) pertinent to clinical medicine. The pharmacokinetics of GTN associated with various dose regimens are characterised by prominent intra- and inter-individual variability. It is, nevertheless, important to clearly understand the pharmacokinetics and characteristics of GTN to optimise its use in clinical practice and, in particular, to obviate the development of tolerance. Measurements of plasma concentrations of GTN and of 1,2-glyceryl dinitrate (1,2-GDN), 1,3-glyceryl dinitrate (1,3-GDN), 1-glyceryl mononitrate (1-GMN), and 2-glyceryl mononitrate (2-GMN), its four main metabolites, remain difficult and require meticulous techniques to obtain reliable results. Since GDNs have an effect on haemodynamic function, pharmacokinetic analyses that include the parent drug as well as the metabolites are important. Although the precise mechanisms of GTN metabolism have not been elucidated, two main pathways have been proposed for its biotransformation. The first is a mechanism-based biotransformation pathway that produces nitric oxide (NO) and contributes directly to vasodilation. The second is a clearance-based biotransformation or detoxification pathway that produces inorganic nitrite anions (NO(2) -). NO(2) - has no apparent cardiovascular effect and is not converted to NO in pharmacologically relevant concentrations in vivo. In addition, several non-enzymatic and enzymatic systems are capable of metabolising GTN. This complex metabolism complicates considerably the evaluation of the pharmacokinetics and pharmacodynamics of GTN. Regardless of the route of administration, concentrations of the metabolites exceed those of the parent compound by several orders of magnitude. During continuous steady-state delivery of GTN, for instance by a patch, concentrations of 1,2-GDN are consistently 2-7 times higher than those of 1,3-GDN, and concentrations of 2-GMN are 4-8 times higher than those of 1-GMN. Concentrations of GDNs are approximately 10 times higher, and of GMNs approximately 100 times higher, than those of GTN during sustained administration. The development of tolerance is closely related to the metabolism of GTN, and can be broadly categorised as haemodynamic tolerance versus vascular tolerance. Efforts are warranted to circumvent the development of tolerance and facilitate the use of GTN in clinical practice. Although this remains to be accomplished, it is likely that, in the near future, regimens will be developed based on a full understanding of the pharmacokinetics and pharmacodynamics of GTN and its metabolites.  相似文献   

12.
The aim of the present study was to determine the concentrations of nitroglycerin (glyceryl trinitrate, GTN, CAS 55-63-0) and its two main stable metabolites; 1,2-dinitroglycerin (1,2-glyceryl dinitrate, GDN, CAS 621-65-8) and 1,3-dinitroglycerin (1,3-GDN, CAS 623-87-0) in human plasma using a capillary gas chromatography method with an electron capture detection. Using the GC conditions, linear calibrations were obtained for 1,3-GDN from 0.14 to 3 ng/mL, for 1,2-GDN from 0.06 to 6 ng/mL, and for GTN from 0.01 to 0.3 ng/mL in plasma samples by the following calibration curve equations: [y = 0.1924x - 0.0088 (r = 0.999)], [y = 0.2273x + 0.0164 (r = 0.995)], [y = 17.434x - 0.0751] for 1,3-GDN, 1,2-GDN, and GTN respectively. The calculated limits of quantification values for GTN, 1,2-GDN, and 1,3-GDN were 0.03 ng/mL, 0.2 ng/mL, and 0.15 ng/mL respectively. This method was verified with a bioequivalence study of an Iranian brand of oral sustained release nitroglycerin with an innovator formulation.  相似文献   

13.
The possible role of glutathione S-transferases (GTSs) in vascular glyceryl trinitrate (GTN) metabolism was investigated. GTN degradation to form its dinitrate metabolites (GDNs) in the 9000g (9k) supernatant fraction of bovine coronary arteries (BCA) was examined. BCAs were homogenized with a 3x volume of phosphate buffer, and the 9k fraction was obtained by centrifugation. GTN (40 ng/ml; 1.76 x 10–7 M) was incubated for 2 hr in the 9k fraction of BCA in the presence of reduced glutathione (2 x 10–3 M). Samples were taken at 10, 20, 40, 60, and 120 min. GTN was observed to degrade readily, exhibiting a half-life of 26 min in the incubate. While both 1,2- and 1,3-GDNs were generated from GTN, formation of 1,3-GDN was predominant (GDN ratio, as 1,2/1,3-GDN, = 0.7–0.8). Coincubation with 2 x 10–5 Mconcentrations of two GST inhibitors, sulfobromophthalein (SBP) and ethacrynic acid (ECA), decreased the rate of GTN loss. The GTN half-lives in SBP- and ECA-treated incubations were 66 and 84 min, respectively. In addition, the pattern of GDN formation was also altered. The resultant GDN ratios exceeded unity in the presence of these inhibitors, indicating that 1,3-GDN formation was attenuated to a greater extent than that of 1,2-GDN. These data suggest that vascular GTN metabolism in BCA is carried out by cytosolic GST isozymes which possess a preference for C-2 denitration of GTN.  相似文献   

14.
The metabolism of nitroglycerin (GTN) to 1,2- and 1,3-glyceryl dinitrate (GDN) by hairless mouse skin in vitro has been measured. In the first set of experiments, GTN was incubated with the 9000g supernatant of fresh, homogenized tissue in the presence and absence of glutathione (GSH), a cofactor for glutathione-S-transferase. After 2 hr of incubation with GSH, 30% of the initially present GTN had been converted to 1,2- and 1,3-GDN; without GSH, less than 5% of the GTN was metabolized. The ratio of 1,2-GDN to 1,3-GDN produced by the homogenate was 1.8– 2.1. In the second series of studies, GTN was administered topically to freshly excised, intact hairless mouse skin in conventional in vitro diffusion cells. The concurrent transport and metabolism of GTN was then monitored by sequential analysis of the receptor phase perfusing the dermal side of the tissue. Three topical formulations were used: a low concentration (1 mg/ml) aqueous solution, a 2% ointment, and a transder-mal delivery system. Delivery of total nitrates (GTN + 1,2-GDN + 1,3-GDN) into the receptor phase was similar for ointment and patch formulations and much greater than that from the solution. The percentage metabolites formed, however, was greatest for the solution (61% and 2 hr, compared to 49% for the patch and 35% for the ointment). As has been noted before, therefore, the relative level of skin metabolism is likely to be greatest when the transepidermal flux is small. Distinct from the homogenate experiments, the 1,2/1,3-GDN ratios in the penetration studies were in the range 0.7– 0.9. It would appear that homogenization of the skin permits GTN to be exposed to a different distribution of enzymes than that encountered during passive skin permeation.  相似文献   

15.
Intravenous infusions of nitroglycerin (GTN), 1,2-glyceryl dinitrate (1,2-GDN), and 1,3-glyceryl dinitrate (1,3-GDN) were given to four conscious dogs at 10 g/min, 30 g/min, 50 g/min, and 70 g/min of GTN and 20 g/min and 100 g/min of GDNs. The steady state plasma concentrations (Css)of GTN were reached after about 60 min whereas for 1,2-GDN and 1,3-GDN the Csswere reached at about 150 min after the infusion began. Except for one dog, the Cssof GTN were not proportional to infusion rate, however, all dogs together showed a good linear relationship between Cssof GTN and infusion rates with an average correlation coefficient of 0.917±0.102. Large variability in GTN clearance after various infusion rates was observed in all dogs. The Cssratios of 1,2-GDN/GTN and 1,3-GDN/GTN yield overall averages of 31.5 ±17.2 and 5.47 ±3.19,respectively. Average Cssratios of metabolites 1,2-GDN/1,3-GDN were 5.78±1.23. This ratio is different from those obtained after iv bolus and oral dosing indicating that the biotransformation of GTN to 1,2-GDN and 1,3-GDN differs for each dosing route. The clearances for 1,2-GDN and 1,3-GDN were not changed over the dose range of 20 g/min to 100 g/min. Terminal half-lives of 1,2-GDN and 1,3-GDN postinfusion were similar to those values obtained after a single bolus dose (45 min). It appears that all the GTN dose at steady state can be accounted for by the formation of measurable 1,2-GDN and 1,3-GDN. Large intra- and interdog variations in systolic blood pressure decrease (SPD) following infusions of GTN were observed, however, all dogs showed a clear systolic blood pressure decrease when the highest infusion rate (70 g/min) was given. No significant systolic blood pressure drop was detected following 20 g/min infusions of 1,2-GDN or 1,3-GDN. It was clear that systolic blood pressure in all dogs decreased following 100 g/min infusions of 1,2-GDN or 1,3-GDN. When SPD values were plotted vs. log GTN concentrations following the infusion of 70 g/min of GTN in all four dogs, a counterclockwise hysteresis was observed indicating the significant contribution of the active dinitrate metabolites to GTN pharmacodynamics.This work was supported in part by NIH grant HL32243.  相似文献   

16.
Bioactivation of nitroglycerin (GTN) into an activator of soluble guanylate cyclase (sGC) is essential for the vasorelaxant effect of the drug. Besides several enzymes that catalyze GTN bioactivation, the reaction with cysteine is the sole nonenzymatic mechanism known so far. Here we show that a reaction with ascorbate results in GTN bioactivation. In the absence of ascorbate, GTN did not affect the activity of purified sGC. However, the enzyme was activated to approximately 20% of maximal NO-stimulated activity by GTN in the presence of 10 mM ascorbate with an EC(50) value of 27.3 +/- 4.9 microM GTN. The EC(50) value of ascorbate was 0.11 +/- 0.011 mM. Activation of sGC was sensitive to oxyhemoglobin, superoxide, and a heme-site enzyme inhibitor. GTN had no effect when ascorbate was replaced by 1000 U of superoxide dismutase per milliliter. Ascorbate is known to reduce inorganic nitrite to NO. In the presence of 10 mM ascorbate, approximately 30 microM nitrite caused the same increase in sGC activity as 0.3 mM GTN. Determination of ascorbate-driven 1,2- and 1,3-glycerol dinitrate formation indicated that a 140 nM concentration of products was generated from 0.3 mM GTN within 10 min, excluding nitrite as a relevant intermediate. Our results suggest that a reaction between GTN and ascorbate or an ascorbate-derived species yields an activator of sGC with NO-like chemical properties. This reaction may contribute to GTN bioactivation in blood vessels under conditions of GTN tolerance and ascorbate supplementation.  相似文献   

17.
The effects of skin storage, skin preparation, skin pretreatment with a penetration enhancer, and skin barrier removal by adhesive tape-stripping on the concurrent cutaneous transport and metabolism of nitroglycerin (GTN) have been studied in vitro using hairless mouse skin. Storing the skin for 10 days at 4°C did not alter barrier function to total nitrate flux [GTN + 1,2-glyceryl dinitrate (1,2-GDN) + 1,3-glyceryl dinitrate (1,3-GDN)]. However, metabolic function was significantly impaired and suggested at least fivefold loss of enzyme activity. Heating skin to 100°C for 5 min appreciably damaged hairless mouse skin barrier function. The ability to hydrolyze GTN was still present, however, and remained constant over the 10-hr experimental period, in contrast to the control, which showed progressively decreasing enzymatic function with time. Pretreatment of hairless mouse skin in vivo (prior to animal sacrifice, tissue excision, and in vitro transport/metabolism studies) with 1-dodecylazacyclo-heptan-2-one (Azone), a putative penetration enhancer, significantly lowered the skin barrier to nitrate flux (relative to the appropriate control). Again, barrier perturbation resulted in essentially constant metabolic activity over the observation period. The ratio of metabolites formed (1,2-GDN/1,3-GDN) was increased from less than unity to slightly above 1 by the Azone treatment. Adhesive tape-stripping gradually destroyed skin barrier function by removal of the stratum corneum. The effects of 15 tape-strips were identical to those of Azone pretreatment: a greatly enhanced flux, a constant percentage formation of metabolites over 10 hr (once again), and an increase in the 1,2-GDN/1,3 GDN ratio. Overall, the experiments caution that, for transdermal drug delivery candidates susceptible to skin metabolism, the status of barrier function (enhancer pretreated, skin damage or disease, etc.) may significantly affect systemic availability.  相似文献   

18.
The percutaneous penetration kinetics of the antianginal, nitroglycerin (GTN), and its primary metabolites, 1,2- and 1,3-glyceryl dinitrate (1,2- and 1,3-GDN), were evaluated in vitro, using full-thickness hairless mouse skin. GTN and the 1,2- and 1,3-GDNs were applied (a) in aqueous solution as pH 7.4 phosphate-buffered saline (PBS) and (b) incorporated into lipophilic ointment formulations. The cutaneous transformation of GTN to its dinitrate metabolites was detected, but no interconversion between 1,2-GDN and 1,3-GDN was observed. Following application of the nitrates in PBS solution, all three compounds exhibited steady-state transport kinetics. The steady-state flux of GTN (8.9 ± 1.5 nmol cm–2 hr–1) was significantly greater (P < 0.05) than those of 1,2-GDN (0.81 ± 0.54 nmol cm–2 hr–1) and 1,3-GDN (0.72 ± 0.20 nmol cm–2 hr–1). The corresponding permeability coefficient () for GTN (20 ± 3 × 10–3 cm hr–1) was significantly larger than the corresponding values for 1,2-GDN (1.4 ± 0.9 × 10–3 cm hr–1) and 1,3-GDN (1.2 ± 0.4 × 10–3 cm hr–1), which were statistically indistinguishable (P > 0.05). Further analysis of the transport data showed that the differences between GTN and the GDNs could be explained by the relative stratum corneum/water partition coefficient (K s) values of the compounds. The apparent partition parameters, defined as = K s · h [where h is the diffusion path length through stratum corneum (SC)] were 19.8 ± 2.5 × 10–2 cm for GTN and 1.91 ± 1.07 × 10–2 and 1.81 ± 0.91 × 10–2 cm for 1,2- and 1,3-GDN, respectively. However, when the nitrates were administered in an ointment base, the apparent partition parameter (') and permeability coefficient (') of GTN markedly decreased, to 2.51 ± 0.75 × 10–2 cm and 1.6 ± 0.3 × 10–3 cm hr–1, respectively. In contrast, the ' and ' results for 1,2- and 1,3-GDN were not significantly different (P > 0.05) from the corresponding and values, which were measured following dosing as aqueous solutions. As a result, the steady-state fluxes of all three nitrates from the ointment formulation were comparable (GTN, 154 ± 28 nmol cm–2 hr–1; 1,2-GDN, 162 ± 22 nmol cm–2 hr–1; 1,3-GDN, 162 ± 34 nmol cm–2 hr–1). It follows that the dinitrates can be as efficiently delivered across the skin as GTN when a suitable formulation is employed. This finding may support transdermal therapy using 1,2- or 1,3-GDN if, indeed, they are found to be pharmacologically effective.  相似文献   

19.
1. The inhibitory activity of astrocytoma cells (0.25-3 x 10(5)) treated with indomethacin (10 microM) on platelet aggregation was enhanced by incubating the cells with E. coli lipopolysaccharide (LPS, 0.5 micrograms ml-1) for 18 h. This effect was attenuated when cycloheximide (10 micrograms ml-1) was incubated together with LPS. The inhibition of platelet aggregation by cells treated with LPS was potentiated by superoxide dismutase (60 u ml-1) and ablated by oxyhaemoglobin (oxyHb, 10 microM) or NG-monomethyl-L-arginine (L-NMMA, 30-300 microM). The effects of L-NMMA were reversed by co-incubation with L-arginine (L-Arg, 100 microM) but not D-arginine (D-Arg, 100 microM). LPS also increased the levels of nitrite in the culture media and this increase was ablated by co-incubation with L-NMMA (300 microM) or cycloheximide (10 micrograms ml-1). 2. Astrocytoma cells (0.5 x 10(5)) treated with indomethacin (10 microM) enhanced the platelet inhibitory activity of glyceryl trinitrate (GTN, 11-352 microM) but not that of sodium nitroprusside (4 microM). Furthermore, when incubated with GTN (200 microM) a 4 fold increase in the levels of guanosine 3':5'-cyclic monophosphate (cyclic GMP) was observed. These effects were abrogated by co-incubation with oxyHb (10 microM) but not with L-NMMA (300 microM). Treatment of the cells with LPS (0.5 micrograms ml-1) for 18 h did not enhance their capacity to form NO from GTN. 3. Thus, in cultured astrocytoma cells, LPS enhances the formation of nitric oxide from endogenous L-arginine.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
The availability of glyceryl trinitrate (GTN) and the differential formation of dinitrate metabolites (GDNs) in various organs as a function of routes of administration were investigated in the rat. GTN was infused at 2.0 µg/min via the left femoral vein (LFV), left external jugular vein (LJV), left femoral artery (LFA), and hepatic portal vein (HPV). Blood concentrations of GTN and GDNs were measured in femoral arterial samples. Different infusions yielded GTN steady-state concentrations in the following rank order: LJV LFV > LFA HPV. Furthermore, the GDN formation ratios (1,2-GDN/1,3-GDN) are different: LFV LJV > LFA > HPV. The availabilities of GTN through the leg, vein, and liver were derived. GTN is significantly extracted and metabolized in these organs, and the leg and the vein prefer 1,2-GDN formation, while the liver forms 1,3-GDN predominantly.  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