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.
We have examined the turnover of 5,6,7,8-tetrahydrobiopterin (BH4) and the effect of decreasing BH4 levels on in situ tyrosine hydroxylase (TH) activity and norepinephrine (NE) content in a homogeneous population of NE-containing neurons derived from the superior cervical ganglion (SCG) of the neonatal rat and maintained in tissue culture. Initial studies indicated that the level of BH4 within SCG cultures increased fourfold between 5 and 37 days in vitro (DIV). This increase in BH4 levels was determined to result from an increase in the rate of BH4 biosynthesis without a change in the rate of degradation. Regardless of culture age, the BH4 content of SCG neurons was observed to turn over with a half-life of approximately 2.5 h. BH4 synthesis by SCG neurons was found to be five times more sensitive to inhibition by 2,4-diamino-6-hydroxypyrimidine (DAHP) and 25 times less sensitive to inhibition by N-acetylserotonin than was previously reported for CNS neurons in culture. Under basal conditions, the rates of in situ TH activity and BH4 biosynthesis were similar. In response to inhibition of BH4 biosynthesis by DAHP and a 90-95% decrease in BH4 levels, in situ TH activity declined by 75%. NE levels declined by 30% following a 24-h period of inhibition of BH4 synthesis. After 2 days of BH4 synthesis inhibition, the level of NE was decreased by 47%. On treatment days 3 and 4, the decline in NE content plateaued at 24% of control levels.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
2,4-Diamino-6-hydroxypyrimidine (DAHP) is considered a specific inhibitor of BH(4) biosynthesis and is widely used in order to elucidate the possible biological function of BH(4) in various cells. In the present study, we found that both the synthesis of tetrahydrobiopterin (BH(4)) and expression of vascular cell adhesion molecule 1 (VCAM-1) were increased in human umbilical vein endothelial cells (HUVEC) treated with proinflammatory cytokines. Thus we examined the effects of DAHP to clarify whether BH(4) might be involved in the expression of VCAM-1 in HUVEC. DAHP reduced the levels of both BH(4) and VCAM-1 induced by TNF-alpha and IFN-gamma. However, the dose-response curves of DAHP for the suppression of the VCAM-1 level and that of BH(4) level were markedly different. Supplementation with sepiapterin failed to restore the depressed VCAM-1 level, although it completely restored the BH(4) level. Furthermore, DAHP significantly reduced the VCAM-1 level under the experimental conditions using TNF-alpha alone, which failed to induce BH(4) production. Taken together, these results indicate that DAHP inhibited the expression of VCAM-1 in a BH(4)-independent manner in HUVEC. In the present study, we also found that DAHP significantly suppressed the accumulation of cytokine-induced NF-kappaB (p65) in the nucleus as well as the mRNA levels of VCAM-1 and GTP cyclohydrolase I (GTPCH), the rate-limiting enzyme of BH(4) synthesis. The data obtained in this study suggest that DAHP reduced VCAM-1 and GTPCH protein synthesis at least partially via suppressing the NF-kappaB level in the nucleus of HUVEC.  相似文献   

3.
Interleukin (IL)-1 beta-induced inhibition of glucose-stimulated insulin secretion in rat islets of Langerhans is mediated in part by nitric oxide (NO). The NO synthase cofactor 5,6,7,8-tetrahydrobiopterin (BH(4)) supports NO synthesis in many cell types and IL-1 beta-induced NO generation and inhibition of insulin secretion have been previously correlated with intracellular BH(4 )levels in rat insulinoma cells. Using rat islets and the beta cell line BRIN-BD11, we have investigated whether synthesis of BH(4) limits IL-1beta-induced NO generation and inhibition of glucose-induced insulin secretion. IL-1 beta-induced NO generation by BRIN cells and islets was reduced by 2,4-diamino-6-hydroxypyrimidine (DAHP), an inhibitor of de novo BH(4) synthesis. Sepiapterin, the substrate for salvage pathway BH(4) synthesis, reversed this inhibitory effect of DAHP in islets but not BRIN cells. DAHP reversed IL-1 beta-induced inhibition of islet insulin secretion, an effect prevented by sepiapterin. We conclude that BH(4) generation is necessary for IL-1 beta-induced NO generation in rat islets and BRIN cells. While a contribution of non-NO mediators cannot be excluded, our results support the proposal that IL-1 beta-induced, NO-mediated inhibition of insulin secretion in rat islets is dependent on the NOS cofactor BH(4).  相似文献   

4.
Nitric oxide (NO) synthesis is induced in vascular smooth muscle cells by lipopolysaccharide (LPS) where it appears to mediate a variety of vascular dysfunctions. In some cell types tetrahydrobiopterin (BH4) synthesis has also been found to be induced by cytokines. Because BH4 is a cofactor for NO synthase, we investigated whether BH4 synthesis is required for LPS-induced NO production in rat aortic smooth muscle cells (RASMC). The total biopterin content (BH4 and more oxidized states) of untreated RASMC was below our limit of detection. However, treatment with LPS caused a significant rise in biopterin levels and an induction of NO synthesis; both effects of LPS were markedly potentiated by interferon-gamma. 2,4-Diamino-6-hydroxypyrimidine (DAHP), a selective inhibitor of GTP cyclohydrolase I, the rate-limiting enzyme for de novo BH4 synthesis, completely abolished the elevated biopterin levels induced by LPS. DAHP also caused a concentration-dependent inhibition of LPS-induced NO synthesis. Inhibition of NO synthesis by DAHP was reversed by sepiapterin, an agent which circumvents the inhibition of biopterin synthesis by DAHP by serving as a substrate for BH4 synthesis via the pterin salvage pathway. The reversal by sepiapterin was overcome by methotrexate, an inhibitor of the pterin salvage pathway. Sepiapterin, and to a lesser extent BH4, dose-dependently enhanced LPS-induced NO synthesis, indicating that BH4 concentration limits the rate of NO production by LPS-activated RASMC. Sepiapterin also caused LPS-induced NO synthesis to appear with an abbreviated lag period phase, suggesting that BH4 availability also limits the onset of NO synthesis. In contrast to the stimulation of LPS-induced NO synthesis, observed when sepiapterin was given alone, sepiapterin became a potent inhibitor of NO synthesis in the presence of methotrexate. This is attributable to a direct inhibitory action of sepiapterin on GTP cyclohydrolase I, an activity which is only revealed after blocking the metabolism of sepiapterin to BH4. Further studies with sepiapterin, methotrexate, and N-acetylserotonin (an inhibitor of the BH4 synthetic enzyme, sepiapterin reductase) indicated that the BH4 is synthesized in RASMC predominantly from GTP; however, a lesser amount may derive from pterin salvage. We demonstrate that BH4 synthesis is an absolute requirement for induction of NO synthesis by LPS in vascular smooth muscle. Our findings also suggest that pterin synthesis inhibitors may be useful for the therapy of endotoxin- and cytokine-induced shock.  相似文献   

5.
Inhibition of GTP cyclohydrolase I (GTPCH) has been used as a selective tool to assess the role of de novo synthesis of (6R)-5,6,7,8-tetrahydro-L-biopterin (BH4) in a biological system. Toward this end, 2,4-diamino-6-hydroxypyrimidine (DAHP) has been used as the prototypical GTPCH inhibitor. Using a novel real-time kinetic microplate assay for GTPCH activity and purified prokaryote-expressed recombinant proteins, we show that potent inhibition by DAHP is not the result of a direct interaction with GTPCH. Rather, inhibition by DAHP in phosphate buffer occurs via an indirect mechanism that requires the presence of GTPCH feedback regulatory protein (GFRP). Notably, GFRP was previously discovered as the essential factor that reconstitutes inhibition of pure recombinant GTPCH by the pathway end product BH4. Thus, DAHP inhibits GTPCH by engaging the endogenous feedback inhibitory system. We further demonstrate that L-Phe fully reverses the inhibition of GTPCH by DAHP/GFRP, which is also a feature in common with inhibition by BH4/GFRP. These findings suggest that DAHP is not an indiscriminate inhibitor of GTPCH in biological systems; instead, it is predicted to preferentially attenuate GTPCH activity in cells that most abundantly express GFRP and/or contain the lowest levels of L-Phe.  相似文献   

6.
GTP cyclohydrolase I (GTPCH) catalyzes the first step in pteridine biosynthesis in Nocardia sp. strain NRRL 5646. This enzyme is important in the biosynthesis of tetrahydrobiopterin (BH4), a reducing cofactor required for nitric oxide synthase (NOS) and other enzyme systems in this organism. GTPCH was purified more than 5,000-fold to apparent homogeneity by a combination of ammonium sulfate fractionation, GTP-agarose, DEAE Sepharose, and Ultragel AcA 34 chromatography. The purified enzyme gave a single band for a protein estimated to be 32 kDa by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The molecular mass of the native enzyme was estimated to be 253 kDa by gel filtration, indicating that the active enzyme is a homo-octamer. The enzyme follows Michaelis-Menten kinetics, with a Km for GTP of 6.5 micromoles. Nocardia GTPCH possessed a unique N-terminal amino acid sequence. The pH and temperature optima for the enzyme were 7.8 and 56 degrees C, respectively. The enzyme was heat stable and slightly activated by potassium ion but was inhibited by calcium, copper, zinc, and mercury, but not magnesium. BH4 inhibited enzyme activity by 25% at a concentration of 100 micromoles. 2,4-Diamino-6-hydroxypyrimidine (DAHP) appeared to competitively inhibit the enzyme, with a Ki of 0.23 mM. With Nocardia cultures, DAHP decreased medium levels of NO2- plus NO3-. Results suggest that in Nocardia cells, NOS synthesis of nitric oxide is indirectly decreased by reducing the biosynthesis of an essential reducing cofactor, BH4.  相似文献   

7.
Tetrahydrobiopterin (BH4) is one of the cofactors of nitric oxide synthase (NOS), and the synthesis of BH4 is induced as well as inducible NOS (iNOS) by lipopolysaccharide (LPS) and/or cytokines. BH4 has a protective effect against the cytotoxicity induced by nitric oxide (NO) and/or reactive oxygen species in various types of cells. The purpose of this study was to examine whether or not an excess of BH4 is present during the production of NO by iNOS in LPS-treated de-endothelialized rat aorta. Addition of LPS (10 microg/ml) to the aorta bath solution caused L-arginine (L-Arg)-induced relaxation from 1.5 hr after the addition of LPS in de-endothelialized rat aorta pre-contracted with 30 mM KCl. The L-Arg-induced relaxation was prevented by NOS inhibitors. BH4 content also increased from 3 hr after the addition of LPS. mRNAs of iNOS and GTP cyclohydrolase I (GTPCH), a rate-limiting enzyme of BH4 synthesis, were increased from 1.5 hr after addition of LPS. Although the expression of iNOS and GTPCH mRNAs was observed in the media, the expression levels in the media were much lower than those in the adventitia. Ten millimolar 2,4-diamino-6-hydroxypyrimidine (DAHP), an inhibitor of GTPCH, strongly reduced L-Arg-induced relaxation, and decreased BH4 content to below the basal level in LPS-treated aorta, whereas 0.5 mM DAHP reduced the LPS-induced increase in BH4 content to the basal level but did not affect L-Arg-induced relaxation. The inhibition of L-Arg-induced relaxation by 10 mM DAHP was overcome by the addition of BH4 (10 microM). These results suggest that although BH4 is essential for NO production from iNOS, the increase in BH4 content above the basal level is not needed for eliciting L-Arg-induced relaxation by the treatment with LPS. Thus, an excess amount of BH4 may be synthesized during NO production by iNOS in LPS-treated rat aorta.  相似文献   

8.
Treatment with tetrahydrobiopterin (BH(4)) reduces blood pressure in spontaneously hypertensive rats (SHR). In the present study, we tested the hypothesis that chronic BH(4) reduces blood pressure in male SHR by reducing testosterone biosynthesis mediated by increasing nitric oxide (NO). Male SHR, aged 17-18 wk, intact or castrated, were treated for 1 wk with BH(4) (20 mg.kg(-1).day(-1) ip). After 1 wk, mean arterial pressure (MAP), serum testosterone, and nitrate/nitrite excretion (NO(x)) were measured. MAP was significantly higher in intact males than castrated males (179 +/- 2 vs. 155 +/- 4 mmHg, P < 0.001). In intact males, BH(4) caused a 17% reduction in MAP (148 +/- 2 mmHg), had no effect on NO(x), and reduced serum testosterone by 85% (24.09 +/- 2.37 vs. 3.72 +/- 0.73 ng/dl; P < 0.001). In castrated males, BH(4) had no effect on MAP (152 +/- 5 mmHg) but increased NO(x) by 38%. When castrated males were supplemented with testosterone, MAP increased to the same level as in intact males (180 +/- 7 mmHg), and BH(4) had no effect on MAP (182 +/- 7 mmHg) or NO(x). NO has been shown to decrease testosterone biosynthesis. Chronic sodium nitrite (70 mg.kg(-1).day(-1) x 1 wk) decreased MAP in intact males (150 +/- 4 mmHg) but had no effect on serum testosterone (21.46 +/- 3.08 ng/dl). The data suggest that BH(4) reduces testosterone synthesis and thereby reduces MAP in male SHR, an androgen-dependent model of hypertension. The mechanism(s) by which BH(4) reduces serum testosterone levels are not clear, but the data do not support a role for NO as a mediator.  相似文献   

9.
BACKGROUND: Nitric oxide (NO) has been shown to decrease myointimal hyperplasia in injured blood vessels. We hypothesize inducible No synthase (iNOS) gene transfer even at low efficiency will provide adequate local no production to achieve this goal. MATERIALS AND METHODS: A retroviral vector containing the human iNOS cDNA (DFGiNOS) was used to transfer the iNOS gene into vascular cells and isolated blood vessels to answer the following questions: can vascular endothelial and smooth muscle cells support iNOS activity and will low efficiency iNOS gene transfer suppress myointimal hyperplasia in injured porcine arteries? RESULTS: DFGiNOS-infected sheep pulmonary artery endothelial cells (SPAEC) expressed significant iNOS mRNA and protein, releasing nitrite levels of 155.0 +/- 10.7 nmol/mg protein/24 h vs. 5.5 +/- 1.1 by control cells. Transduced rat smooth muscle cells (RSMC) also expressed abundant iNOS mRNA and protein, but, in contrast to SPAEC, NO synthesis was dependent on exogenous tetrahydrobiopterin (BH4) (291.8 +/- 10.4 nmol nitrite/mg protein/24 hr with BH4, 37.7 +/- 2.6 without BH4). Only porcine arteries infected with DFGiNOS following balloon injury exhibited a 3-fold increase in total NO synthesis and a 15-fold increase in cGMP levels over control vessels in a BH4 dependent fashion, despite only a 1% gene transfer efficiency. Transfer of iNOS completely prevented the 53% increase in myointimal thickness induced by balloon catheter injury; the administration of a NOS inhibitor reversed this effect. CONCLUSIONS: These in vitro findings suggest that vascular iNOS gene transfer may be feasible. Furthermore, a low gene transfer efficiency may be sufficient to inhibit myointimal hyperplasia following arterial balloon injury, although a source of BH4 may be required.  相似文献   

10.
We studied the synthesis and release of nerve growth factor (NGF) in cultured rat iris with a two-site enzyme immunoassay by measuring the time course of NGF levels remaining in the iris and relased into the medium up to 72 h. For up to 3 h, the NGF levels in the iris did not change significantly. After that, they increased to a maximal level of 350 +/- 30 pg NGF/iris at 19 h, which is 200 times higher than the in vivo content. Between 20 and 72 h in culture, the NGF level decreased to 130 +/- 10 pg NGF/iris, whereas general protein synthesis did not change during that time period. Maximal rate of NGF production (203 pg NGF/h/iris) was seen between 9 and 12 h in culture. In the medium, NGF levels were first detectable after 6 h. Levels then increased with a time course similar to that seen within the iris, reaching a maximal level of 1,180 +/- 180 pg after 19 h in vitro, and then did not significantly change for up to 48 h. The NGF production of the densely sympathetically innervated dilator was three times higher than that of the predominantly cholinergically innervated sphincter. The NGF production was blocked by inhibitors of messenger RNA synthesis (actinomycin D) and of polyadenylation (9-beta-D-arabinofuranosyladenine) as well as by inhibitors of translation (cycloheximide). Monensin, which interferes with the transport of proteins through the Golgi apparatus, decreased NGF levels to 8-12% of controls in the medium, suggesting that the Golgi apparatus is involved in the intracellular processing of NGF.  相似文献   

11.
Cell culture lines were established from the transplantable mouse hepatomas H6 and H129. Both cell lines had a doubling time about 30 h when maintained in medium containing 5% foetal bovine serum. H6 cells contained about 3-4 times more DNA-dependent RNA polymerase I (Pol I; ribonucleoside triphosphate--RNA nucleotidyltransferase, EC 2.7.7.6) than did H129 cells. Moreover, the H6-cell enzyme was more heat-labile than that from H129 cells. Steady-state contents of 28S rRNA were measured in both cell lines. Exponentially growing cultures of H6 cells contained about 6.5pg of 28S rRNA/cell, and similar cultures of H129 cells contained about 5.8pg/cell. Stationary cultures of both cell lines contained about 2pg of 28S rRNA/cell. By two different techniques, the half-time for turnover of 28S rRNA was estimated to be 16-17h for both H6 and H129 cells. Knowing the turnover rate and the steady-state concentration, one may calculate that both H6 and H129 cells synthesize 28S rRNA at a rate of about 0.25 pg/h per cell. The amount of template-bound Pol I activity was similar in nuclei isolated from H6 and H129 cell cultures. These data indicate that, although H6 cells contained 3-4 times more Pol I than did H129 cells, both cell lines synthesized rRNA at about the same rate.  相似文献   

12.
The developmental expression of the alpha-subunit of Go was examined in neuronal cultures derived from rat mesencephalon (MES) and hypothalamus (HYP). These cultures were essentially free of contaminating glia and were maintained as a stable population for periods up to 3 weeks. Immunoblotting utilizing specific antisera against Go indicated that in neurons from both brain regions, membrane concentrations of Go increased dramatically during the first 2 weeks in vitro. Thereafter, increases in the amount of Go per neuron kept pace with increasing process (axons and dendrites) formation. Multiple forms of immunoreactive Go were detected in MES and HYP neurons, and the proportions of these forms changed between 4 and 14 days in culture. Finally, increasing neuron density significantly increased membrane levels of Go in MES but not HYP cultures.  相似文献   

13.
The objective was to determine the effects of an inhibin alpha (1-32) fragment gene on proliferation, apoptosis, and steroidogenesis of bovine granulosa cells (GC) isolated from medium and small follicles (diameter >4-8 and 1-4mm, respectively), and the effect of GC, previously transfected with pEGISI, on oocyte maturation and in vitro embryo development. To enhance expression of the inhibin alpha (1-32) fragment, GC were transfected with pEGISI. Transfection inhibited (P<0.05) GC proliferation (88.8+/-2.1%; mean+/-S.E.M.) compared to the control and EGFP groups (100% and 97.5+/-2.1%) from medium follicles, with no significant effect on GC from small follicles. Apoptosis was higher (P<0.01) in transfected GC than in controls. Transfection increased (P<0.05) estradiol synthesis from both medium and small follicles (0.57+/-0.13 and 0.86+/-0.13 pg/mL vs. 0.19+/-0.05 and 0.35+/-0.09 pg/mL in controls) after culturing for 48 h, with suppression (P<0.05) in transfected GC after 96 h. Transfection reduced (P<0.05) progesterone synthesis in GC from both medium and small follicles (24.5+/-3.4 and 75.4+/-4.6 ng/mL vs. 45.42+/-5.33 and 117.32+/-11.99 ng/mL in controls) after culture for 48 h, with no significant difference after 96 h. Maturation rate of oocytes co-cultured with transfected GC from medium follicles was decreased relative to control (61.5+/-6.8% vs. 71.2+/-5.7%, P<0.05), with no significant effect on embryo development. In conclusion, overexpression of inhibin alpha (1-32) fragment regulated GC development; effects on subsequent oocyte maturation were both time- and stage-dependent.  相似文献   

14.
Escherichia coli phenylalanine-sensitive 3-deoxy-arabino-heptulosonate 7-phosphate synthase (DAHP synthase) catalyzes the net aldol condensation of phosphoenolpyruvate and erythrose 4-phosphate to form 3-deoxy-D-arabino-heptulosonate 7-phosphate and inorganic phosphate. For the first time, the presteady-state kinetic analysis of the Phe-sensitive DAHP synthase from E. coli is reported. The steady-state and presteady-state kinetic parameters of the DAHP synthase reconstituted with Mn(II), Cu(II), and Zn(II) were compared. These studies showed the following: 1) product release is rate-limiting for all of the three metal ions studied under physiologically relevant conditions; 2) concentration of the active sites of the metal-containing DAHP synthase is increasing from Mn- (30%) to Zn- (52%) and to Cu-DAHP synthase (88%); 3) rate constant for product formation is higher in Mn- (130-200 s(-1)) than Cu- (55 s(-1)) and Zn-DAHP synthase (6.8 s(-1)); and 4) steady-state rate (rate constant for product release) is higher for the Mn- (70 s(-1)) than for Cu- (5.6 s(-1)) and Zn-DAHP synthase (1.8 s(-1)). In addition, an examination of the reaction kinetics at lower pH reveals that for Cu-DAHP synthase, product release is no longer rate-limiting, whereas the Mn- and Zn-DAHP synthase show a slower rate of product formation, suggesting that the intermediate formation becomes rate-limiting in product formation. Also, a deuterium-isotope effect on the burst rate constant of product formation for Mn-DAHP synthase was observed at pH 6.0. This supports the hypothesis that the role of metal ion in E. coli DAHP synthase is to position the amino acids with the appropriate geometry required to coordinate and activate the water molecule.  相似文献   

15.
The rate of (13)C-label incorporation into both aspartyl (NAA C3) and acetyl (NAA C6) groups of N-acetyl aspartate (NAA) was simultaneously measured in the rat brain in vivo for up to 19 h of [1-(13)C]glucose infusion (n = 8). Label incorporation was detected in NAA C6 approximately 1.5 h earlier than in NAA C3 because of the delayed labeling of the precursor of NAA C3, aspartate, compared to that of NAA C6, glucose. The time courses of NAA were fitted using a mathematical model assuming synthesis of NAA in one kinetic compartment with the respective precursor pools of aspartate and acetyl coenzyme A (acetyl-CoA). The turnover rates of NAA C6 and C3 were 0.7 +/- 0.1 and 0.6 +/- 0.1 micromol/(g h) with the time constants 14 +/- 2 and 13 +/- 2 h, respectively, with an estimated pool size of 8 micromol/g. The results suggest that complete label turnover of NAA from glucose occurs in approximately 70 h. Several hours after starting the glucose infusion, label incorporation into glutathione (GSH) was also detected. The turnover rate of GSH was 0.06 +/- 0.02 micromol/(g h) with a time constant of 13 +/- 2 h. The estimated pool size of GSH was 0.8 micromol/g, comparable to the cortical glutathione concentration. We conclude that NAA and GSH are completely turned over and that the metabolism is extremely slow (< 0.05% of the glucose metabolic rate).  相似文献   

16.
GTP cyclohydrolase 1 is the rate-limiting enzyme in production of tetrahydrobiopterin, a necessary cofactor for endothelial nitric oxide synthase. We tested the hypothesis that inhibition of tetrahydrobiopterin synthesis impairs endothelium-dependent relaxation and increase blood pressure in rats. 2,4-Diamino-6-hydroxypyrimidine (DAHP), a GTP cyclohydrolase 1 inhibitor, was given in drinking water (approximately 120 mg.kg(-1).day(-1)) to male Sprague-Dawley rats for 3 days. Systolic blood pressures were measured (tail-cuff procedure) for 3 days before and each day during DAHP treatment. Blood pressure was significantly increased after DAHP treatment (122 +/- 2 vs. 154 +/- 3 mmHg before and after DAHP, respectively; P < 0.05). Endothelium-intact aortic segments from pentobarbital sodium-anesthetized rats were isolated and hung in organ chambers for measurement of isometric force generation. Aortas from DAHP-treated rats exhibited a decreased maximal relaxation to ACh compared with controls [% relaxation from phenylephrine (10-7 M)-induced contraction: DAHP 57 +/- 6% vs. control 79 +/- 4%; P < 0.05]. Relaxation responses to A-23187 were also decreased in aortas from DAHP-treated rats compared with controls. Incubation with sepiapterin (10-4 M, 1 h), which produces tetrahydrobiopterin via a salvage pathway, restored relaxation to ACh in aortas from DAHP-treated rats. Superoxide dismutase significantly increased ACh-induced relaxation in aortas from DAHP-treated rats, whereas catalase had no effect. Endothelium-independent relaxation to sodium nitroprusside in aortas from DAHP-treated rats was not different from control rats; however, nitric oxide synthase inhibition increased sensitivity to sodium nitroprusside in aortas from DAHP-treated rats. These results support the hypothesis that GTP cyclohydrolase 1 inhibition decreases relaxation and increases blood pressure in rats.  相似文献   

17.
Using a specific 13C NMR localization method, 13C label incorporation into the glycogen C1 resonance was measured while infusing [1-(13)C]glucose in intact rats. The maximal concentration of [1-(13)C]glycogen was 5.1 +/- 0.6 micromol g(-1) (mean +/- SE, n = 8). During the first 60 min of acute hyperglycemia, the rate of 13C label incorporation (synthase flux) was 2.3 +/- 0.7 micromol g(-1) h(-1) (mean +/- SE, n = 9 rats), which was higher (p < 0.01) than the rate of 0.49 +/- 0.14 micromol g(-1) h(-1) measured > or = 2 h later. To assess whether the incorporation of 13C label was due to turnover or net synthesis, the infusion was continued in seven rats with unlabeled glucose. The rate of 13C label decline (phosphorylase flux) was lower (0.33 +/- 0.10 micromol g(-1) h(-1)) than the initial rate of label incorporation (p < 0.01) and appeared to be independent of the duration of the preceding infusion of [1-(13)C]glucose (p > 0.05 for correlation). The results implied that net glycogen synthesis of approximately 3 micromol g(-1) had occurred, similar to previous reports. When infusing unlabeled glucose before [1-(13)C]glucose in three studies, the rate of glycogen C1 accumulation was 0.46 +/- 0.08 micromol g(-1) h(-1). The results suggest that steady-state glycogen turnover rates during hyperglycemia are approximately 1% of glucose consumption.  相似文献   

18.
When intact human erythrocytes are incubated at metabolic steady state in a chloride-free medium containing [32P]Pi, there is rapid labelling of the gamma-phosphate of ATP, followed by a slower labelling of the monoester phosphate groups of phosphatidylinositol 4-phosphate (PtdIns4P) and phosphatidylinositol 4,5-bisphosphate [PtdIns(4,5)P2] [King, Stephens, Hawkins, Guy & Michell (1987) Biochem. J. 244, 209-217]. We have analysed the early kinetics of the labelling of these phosphate groups, in order to determine: (a) the steady-state rates of the interconversions of phosphatidylinositol, PtdIns4P and PtdIns(4,5)P2; and (b) the fractions of the total cellular complement of PtdIns4P and PtdIns(4,5)P2 that participate in this steady-state turnover. The experimental data most closely fit a pattern of PtdIns4P and PtdIns(4,5)P2 turnover in which one-quarter of the total cellular complement of each lipid is in the metabolic pool that participates in rapid metabolic turnover, with rate constants of 0.028 min-1 for the interconversion of PtdIns and PtdIns4P, and of 0.010 min-1 for the PtdIns4P/PtdIns(4,5)P2 cycle. These rate constants represent metabolic fluxes of approx. 2.1 nmol of lipid/h per ml of packed erythrocytes between PtdIns and PtdIns4P and of approx. 5.7 nmol/h per ml of cells between PtdIns4P and PtdIns(4,5)P2.  相似文献   

19.
Bovine embryos recovered from superovulated donors on Days 8-18 postestrus were cultured in vitro in a tissue perifusion system to quantify hormone secretion. Embryos were cultured for 24 h at 37 degrees C in Ham's F-10 medium supplemented 5% v/v with heat-treated, charcoal-stripped calf serum; 100 IU/ml penicillin; and 100 micrograms/ml streptomycin. The medium was saturated with 5% CO2 in air and perifused at 50 microliters/min (3 ml/h). Estrone (E1) estradiol (E2), progesterone (P4), prostaglandin E2 (PGE2), and prostacyclin (PGI2) were quantified by RIA in 6-h pools of perifusate fractions. Estrone was measurable (pg/h/embryo; mean +/- SE) on Days 13 (10.80 +/- 4.56) and 15 (34.80 +/- 9.80); E2 on Days 11 (36.80), 12 (81.28 +/- 29.80), 13 (11.75 +/- 4.09), 15 (157.20 +/- 112.60), and 16 (30.26 +/- 8.76); and P4 (ng/h/embryo) on Days 13 (0.5-1.0) and 17 (approximately 1.5). PGE2 was secreted by Day 10 bovine embryos during the last 6 h of culture (19-24 h) and throughout culture for Day 11-18 embryos. The rate of PGE2 secretion increased (p less than 0.05) over the previous days(s) at Days 13 and 17. The mean (+/- SE) secretion rates (pg/h/embryo) for the 24-h culture by embryonic ages were as follows: Day 11 (63.39 +/- 14.61), 12 (172.10 +/- 30.90), 13 (3094.08 +/- 283.35), 14 (1633.89 +/- 49.98), 15 (3739.23 +/- 1082.79), 16 (4955.37 +/- 1381.83), 17 (11893.23 +/- 1188.48), and 18 (13827.99 +/- 3587.88).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Tetrahydrobiopterin (BH4) is an essential co-factor for nitric oxide synthases (NOS). The aim of the present work was to study whether BH4 deficiency affects the vulnerability of neurones in primary culture to hypoxia. Intracellular BH4 levels were depleted by pre-incubating neurones with 5 mm 2,4-diamino-6-hydroxypyrimidine (DAHP) for 18 h, after which cells were exposed for 1 h to normoxic or hypoxic conditions. Our results showed that whereas neurones were resistant to hypoxia-induced cellular damage, BH4 deficiency in neurones led to oxidative stress, mitochondrial depolarization, ATP depletion and necrosis after 1 h of hypoxia. Indeed, hypoxia specifically inhibited mitochondrial complex IV activity in BH4-deficient neurones. All these effects were counteracted when neuronal BH4 levels were restored by incubating cells with exogenous BH4 during the hypoxic period. Moreover, hypoxia-induced damage in BH4-deficient neurones was prevented when Nomega-nitro-l-arginine monomethyl ester (NAME), haemoglobin or superoxide dismutase plus catalase were present during the hypoxic period, suggesting that peroxynitrite might be involved in the process. In fact, BH4 deficiency elicited neuronal NO dysfunction, resulting in an increase in peroxynitrite generation by cells, as shown by the enhancement in tyrosine nitration; this was prevented by supplements of BH4, NAME, haemoglobin or superoxide dismutase plus catalase during hypoxia. Our results suggest that BH4 deficiency converts neuronal NOS into an efficient peroxynitrite synthase, which is responsible for the increase in neuronal vulnerability to hypoxia-induced mitochondrial damage and necrosis.  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