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 812 毫秒
1.
The isoflavonoid phytoalexin pisatin is synthesized by Pisum sativum in response to microbial infection and certain other forms of stress. An enzyme which synthesizes pisatin by methylating the 3-hydroxyl of (+)6a-hydroxymaackiain (HMK) was extracted from CuCl2-stressed pea seedlings. The enzyme was enriched 370-fold by (NH4)2SO4 precipitation, DEAE chromatography, chromatofocusing, and hydrophobic interaction chromatography (HIC), to a specific activity of 8.2 microkatals per gram protein. Enzyme activity profiles from chromatofocusing and HIC columns suggested the presence of two isozymes, of pl 5.2 and 4.9. Nondenaturing gel filtration of the HIC-purified enzyme gave a single peak of activity at the same elution volume as BSA (66 kilodaltons); the active fractions showed two proteins upon SDS-PAGE, of Mr 66,000 and 43,000. The smaller protein was most abundant in chromatographic fractions containing peak enzyme activity throughout purification. In a partially purified preparation, this 43 kilodalton protein was the only one photoaffinity labelled by [3H]S-adenosyl-l-methionine. The purified enzyme preferred the (+) over the (−) stereoisomer of HMK and other pterocarpans; overall, (+)HMK was the best substrate. Km values were 2.3 micromolar for (+)HMK and 35 micromolar for S-adenosyl-l-methionine. The methyltransferase had a pH optimum of 7.9 and no apparent divalent cation requirement.  相似文献   

2.
Some isolates of the plant pathogen Nectria haematococca detoxify the isoflavonoid phytoalexin (−)maackiain by hydroxylation at carbon 6a. Precursor feeding studies strongly suggest that the penultimate step in (+)pisatin biosynthesis by Pisum sativum is 6a-hydroxylation of (+)maackiain. We have used 18O labeling to test the involvement of oxygenases in these two reactions. When fungal metabolism of maackiain took place under 18O2, the product was labeled with 99% efficiency; no label was incorporated by metabolism in H218O. Pisatin synthesized by pea pods in the presence of 18O2 or H218O was a mixture of molecules containing up to three labeled oxygen atoms. Primary mass spectra of such mixtures were complex but were greatly simplified by tandem MS. This analysis indicated that the 6a oxygen of pisatin was derived from H2O and not from O2. Labeling patterns for the other five oxygen atoms were consistent with the proposed pathway for biosynthesis of pisatin and related isoflavonoids. We conclude that the fungal hydroxylation of maackiain is catalyzed by an oxygenase, but the biosynthetic route to the 6a hydroxyl of pisatin is unknown.  相似文献   

3.
NADPH-dependent reduction of 2′,7-dihydroxy-4′,5′-methylenedioxyisoflavone to the isoflavanone sophorol, a proposed intermediate step in pisatin biosynthesis, was detected in extracts of Pisum sativum. This isoflavone reductase activity was inducible by treatment of pea seedlings with CuCl2. The timing of induction coincided with that of the 6a-hydroxymaackiain 3-O-methyltransferase, which catalyzes the terminal biosynthetic step. Neither enzyme was light inducible. Further NADPH-dependent metabolism of sophorol by extracts of Cucl2-treated seedlings was also observed; three products were radiolabeled when [3H]sophorol was the substrate, one of which is tentatively identified as maackiain.  相似文献   

4.
Pea (Pisum sativum) tissues, on treatment with aqueous CuCl2 synthesize the 6a-hydroxypterocarpan phytoalexin (+) - (6aR, 11aR) - pisatin. By supplying (?) - (6aR, 11aR) - maackiain during this induction process, sigruficant quantities of ( ? ) - (6aS, 11aS) - pisatin are produced, immature pods being most effective. Pisatin levels are considerably reduced when compared with the normal induction process, but may contain as much as 92% (?)-pisatin. This confirms that the 6a-hydroxylation of maackiain during the biosynthesis of pisatin must proceed with retention of configuration at C-6a.  相似文献   

5.
Feeding experiments in cupric chloride-treated Pisum sativum pods and seedlings have demonstrated the preferential incorporation of (+)-(6aS,11aS)-[3H]maackiain over (?)-(6aR, 11aR)-[14C]maackiain into (+)-(6aR, 11aR)-pisatin, establishing that the 6a-hydroxylation of pterocarpans proceeds with retention of configuration. (+)- (6aR,11aR)-6a-hydroxymaackiain was similarly incorporated much better than (?)-(6aS,11aS)-6a- hydroxymaackiain. Where (?)-isomers were incorporated, optical activity measurements on the pisatin produced indicated significant synthesis of (?)-pisatin as well as the normal (+)-pisatin. 7,2′-Dihydroxy-4′,5′- methylenedioxyisoflav-3-ene and both enantiomers of 7,2′-dihydroxy-4′,5′-methylenedioxyisoflavan were poor precursors of pisatin.  相似文献   

6.
Frankia sp., the actinomycetous endophyte in nitrogen-fixing actinorhizal nodules, may differentiate two forms from its hyphae: vesicles and sporangia. In root nodules of Comptonia peregrina (L.) Coult. and Myrica gale L., sporangia may be either absent or present. Nitrogenase activity and symbiotic efficiency were contrasted in spore(+) and spore(−) nodules of these two host genera. Seedlings of C. peregrina nodulated with the spore(+) inoculum showed only 60% of the nitrogenase activity and 50% of the net size of their spore(−) counterparts after 12 weeks of culture. Measurements of acetylene reduction (i.e., nitrogenase activity) were coordinated with samplings of nodules for structural studies. Significant differences in acetylene reduction rates were discernible between spore(+) and spore(−) nodules commencing 4 weeks after nodulation, concomitant with the maturation of sporangia in the nodule. Spore(+) nodules ultimately reached less than half of the rate of nitrogenase activity of spore(−) nodules. Both types of nodules evolved only small amounts of molecular hydrogen, suggesting that both were equally efficient in recycling electrons lost to the reduction of hydrogen ions by nitrogenase. Respiratory cost of nitrogen fixation, expressed as the quotient of micromole CO2 to micromole ethylene evolved by excised nodules, was significantly greater in spore(+) than in spore(−) nodules. M. gale spore(−) nodules showed variable effectivity, though all had low CO2 to ethylene evolution ratios. M. gale spore(+) nodules resembled C. peregrina spore(+), with low effectivity and high respiratory cost for nitrogen fixation.  相似文献   

7.
Some isolates of the fungus Nectria haematococca Berk. and Br. can demethylate pisatin, a phytoalexin from pea (Pisum sativum L.). Pisatin demethylation appears to be necessary for tolerance to pisatin and virulence on pea, and is catalyzed by a microsomal cytochrome P-450. We now report solubilization of this enzyme from N. haematococca microsomes. Pisatin demethylase activity was obtained in the high speed supernatant of detergent treated microsomes, if detergent was removed before assay. The CO-binding spectrum of the soluble enzyme preparation indicated the presence of cytochrome P-450. Cholic acids were the most effective of the detergents tested for solubilizing enzyme activity. Loss of enzyme activity during solubilization was reduced by certain protease inhibitors, but not by substrate, reducing agents, antioxidants, or phospholipids. The most effective solubilization medium tested was 1% sodium cholate, 100 millimolar potassium phosphate, 500 millimolar sucrose, 1 millimolar phenylmethylsulfonyl fluoride, pH 7.5, which yielded approximately 30% of the pisatin demethylase and over 95% of the NADPH-cytochrome c reductase in the soluble fraction. Demethylase activity was lost when the reductase was removed by adsorption on 2′,5′-ADP-agarose. The demethylase activity of reductase-free fractions could be restored by adding a reductase preparation purified approximately 100-fold from microsomes of N. haematococca isolate 74-8-1, which does not demethylate pisatin. We conclude that pisatin demethylase requires NADPH-cytochrome c reductase for activity. The inability of some isolates to demethylate pisatin appears to be due to the absence of a suitable cytochrome P-450, rather than to a lack of functional reductase.  相似文献   

8.
Submitochondrial particles freshly prepared by sonication from pea cotyledon mitochondria showed low ATPase activity. Activity increased 20-fold on exposure to trypsin. The pea cotyledon submitochondrial particle ATPase was also activated by “aging” in vitro. At pH 7.0 addition of 1 millimolar ATP prevented the activation. ATPase of freshly prepared pea cotyledon submitochondrial particles had a substrate specificity similar to that of the soluble ATPase from pea cotyledon mitochondria, with GTPase > ATPase. “Aged” or trypsin-treated particles showed equal activity with the two substrates. NaCl and NaHCO3, which stimulate the ATPase but not the GTPase activity of the soluble pea enzyme, were stimulatory to both the ATPase and GTPase activities of freshly prepared submitochondrial particles. However, they were stimulatory only to the ATPase activity of trypsin-treated or “aged” submitochondrial particles. In contrast, the ATPase activity of rat liver submitochondrial particles was stimulated by HCO3, but inhibited by Cl, indicating that Cl stimulation is a distinguishing property of the pea mitochondrial ATPase complex.  相似文献   

9.
Azole antifungal ketoconazole (KET) was demonstrated to activate aryl hydrocarbon receptor (AhR). Since clinically used KET is a racemic mixture of two cis-enantiomers (2R,4S)-(+)-KET and (2S,4R)-(−)-KET, we examined the effects of KET enantiomers on AhR signaling pathway. (+)-KET dose-dependently activated AhR in human gene reporter cell line AZ-AHR, and displayed 5–20× higher agonist activity (efficacy), as compared to (−)-KET; both enantiomers were AhR antagonists with equal potency (IC50). Consistently, (+)-KET strongly induced CYP1A1 mRNA and protein in human HepG2 cells, while (−)-KET exerted less than 10% of (+)-KET activity. In primary human hepatocytes, both enantiomers preferentially induced CYP1A2 over CYP1A1 mRNA and protein, and the potency of (+)-KET was slightly higher as compared to (−)-KET. Ligand binding assay with guinea pig liver cytosols revealed that both (+)-KET and (−)-KET are weak ligands of AhR that displaced [3H]-TCDD with comparable potency. Similarly, both enantiomers weakly transformed AhR to DNA-binding form with similar potency, as showed by EMSA, in guinea pig liver cytosolic extracts and nuclear extracts from mouse Hepa-1 cells. We also examined effects of KET on glucocorticoid receptor (GR), a regulator of AhR activity. Both KET enantiomers antagonized GR with similar potency, as revealed by gene reporter assay in AZ-GR cell line and down-regulation of tyrosine aminotransferase mRNA in human hepatocytes. Finally, we demonstrate enantiospecific antifungal activities of KET enantiomers in six Candida spp. strains. In conclusion, the significance of current study is providing the first evidence of enatiospecific effects of cis-enantiomers of ketoconazole on AhR-CYP1A pathway.  相似文献   

10.
Pisatin, a 6a-hydroxyl-pterocarpan phytoalexin from pea (Pisum sativum L.), is relatively unique among naturally occurring pterocarpans by virtue of the (+) stereochemistry of its 6a-11a C-C bond. However, pisatin synthesizing pea tissue has an isoflavone reductase, first identified in alfalfa, which acts on the (-) antipode. In order to establish the natural biosynthetic pathway to (+) pisatin, and to evaluate the possible involvement of intermediates with a (-) chirality in its biosynthesis, we administered chiral, tritium-labeled, isoflavanones and pterocarpans to pisatin-synthesizing pea cotyledons and compared the efficiency of their incorporation. Pea incorporated the isoflavanone, (-) sophorol, more efficiently than either its (+) antipode, or the pterocarpans (+) or (-) maackiain. (-) Sophorol was also metabolized by protein extracts from pisatin-synthesizing pea seedlings in a NADPH-dependent manner. Three products were produced. One was the isoflavene (7,2'-dihydroxy-4',5'-methylenedioxyisoflav-3-ene), and another had properties consistent with the isoflavanol (7,2'-dihydroxy-4',5'-methylenedioxyisoflavanol), the expected product for an isoflavanone reductase. A cDNA encoding sophorol reductase was also isolated from a cDNA library made from pisatin-synthesizing pea. The cloned recombinant sophorol reductase preferred (-) sophorol over (+) sophorol as a substrate and produced 7,2'-dihydroxy-4',5'-methylenedioxyisoflavanol. Although no other intermediates in (+) pisatin biosynthesis were identified, the results lend additional support to the involvement of intermediates of (-) chirality in (+) pisatin synthesis.  相似文献   

11.
The effects of ultraviolet-B (UV-B) radiation on stomatal conductance (gs) in pea (Pisum sativum L.), commelina (Commelina communis L.), and oilseed rape (Brassica napus L.) plants were investigated. Plants were grown in a greenhouse either with three different high ratios of UV-B to photosynthetically active radiation or with no UV-B radiation. Pea plants grown in the highest UV-B radiation (0.63 W m−2) exhibited a substantial decrease of adaxial and abaxial gs (approximately 80% and 40%, respectively). With growth in 0.30 W m−2 of UV-B adaxial gs was decreased by 23%, with no effect on abaxial gs, and lower UV-B irradiance of 0.21 W m−2 had no effect on either surface. Although abaxial gs increased when leaves were turned over in control plants, it did not in plants grown with the highest UV-B. Adaxial gs in commelina and oilseed rape also decreased on exposure to high UV-B (0.63 W m−2). For previously unexposed pea plants the time course of the effect of UV-B on gs was slow, with a lag of approximately 4 h, and a time constant of approximately 3 h. We conclude that there is a direct effect of UV-B on stomata in addition to that caused by changes in mesophyll photosynthesis.  相似文献   

12.
Pisatin is the major phytoalexin produced by pea upon microbial infection. The enzyme that catalyzes the terminal step in the pisatin biosynthetic pathway is (+)6a-hydroxymaackiain 3-O-methyltransferase (HMM). We report here the isolation and characterization of two HMM cDNA clones (pHMM1 and pHMM2) made from RNA obtained from Nectria haematococca-infected pea tissue. The two clones were confirmed to encode HMM activity by heterologous expression in Escherichia coli/. The substrate specificity of the methyltransferases in E. coli was similar to the activity detected in CuCl2-treated pea tissue. Nucleotide sequence analysis of Hmm1 and Hmm2 revealed an open reading frame of 1080 bp and 360 amino acid residues which would encode 40.36 kda and 40.41 kDa polypeptides, respectively. The deduced amino acid sequence of HMM1 has 95.8% identity to HMM2, 40.6% identity to Zrp4, a putative O-methyltransferase (OMT) in maize root, and 39.1% to pBH72-F1, a putative OMT induced in barley by fungal pathogens or UV light. Comparison of the deduced amino acid sequences of the cDNA clones to OMTs from other higher plants identified the binding sites of S-adenosylmethionine (AdoMet). Southern blot analysis showed two closely linked genes with strong homology to Hmm in the pea genome.  相似文献   

13.
The uptake of (+)-S- and (−)-R-abscisic acid (ABA) by suspension culture cells of hopbush (Dodonaea viscosa L. Jacqu.) was followed over a range of temperatures, pH values, and time intervals. The natural (+)-S-ABA was taken up about five times faster than the unnatural (−)-R-ABA. Each 10°C rise in temperature from 1 to 31°C increased the rate of uptake (Q10) of (+)-S-ABA about 2.2-fold, whereas that of the (−)-R increased with a Q10 of 1.4. (+)-ABA was taken into the cells by a saturable carrier, but (−)-ABA and both enantiomers of 2-trans-ABA were not; they appeared to enter by passive diffusion. The uptake of (+)-ABA was linear over the first 8 hours but concentrations within the cells decreased after 2 hours to remain constant after 4 hours as rapid metabolism was induced. Electron microscopy of thin sections of the cells, combined with a stereological analysis of their shape, showed that the vacuoles comprised 80% of the cell volume and the cytoplasm plus nucleus comprised 20%. There were no photosynthetically active plastids in the cells. Concentrations of the endogenous ABA in the cytoplasm (pH 7.32) and vacuoles (pH 5.88) were calculated by applying the Henderson-Hasselbalch equation (ABA pKa 4.7) so that, provided no active metabolic redistribution occurred, the concentration in the cytoplasm was 7.9 micromolar and that in the vacuole was 0.3 micromolar. In vivo pH was measured by 31P nuclear magnetic resonance spectroscopy.  相似文献   

14.
Epigallocatechin-3-gallate (EGCG), a dietary polyphenol (flavanol) from green tea, possesses leishmanicidal and antitrypanosomal activity. Mitochondrial damage was observed in Leishmania treated with EGCG, and it contributed to the lethal effect. However, the molecular target has not been defined. In this study, EGCG, (+)-catechin and (−)-epicatechin were tested against recombinant arginase from Leishmania amazonensis (ARG-L) and rat liver arginase (ARG-1). The compounds inhibit ARG-L and ARG-1 but are more active against the parasite enzyme. Enzyme kinetics reveal that EGCG is a mixed inhibitor of the ARG-L while (+)-catechin and (−)-epicatechin are competitive inhibitors. The most potent arginase inhibitor is (+)-catechin (IC50 = 0.8 µM) followed by (−)-epicatechin (IC50 = 1.8 µM), gallic acid (IC50 = 2.2 µM) and EGCG (IC50 = 3.8 µM). Docking analyses showed different modes of interaction of the compounds with the active sites of ARG-L and ARG-1. Due to the low IC50 values obtained for ARG-L, flavanols can be used as a supplement for leishmaniasis treatment.  相似文献   

15.
In peroxisomes isolated from pea leaves (Pisum sativum L.) the production of superoxide free radicals (O2) by xanthine and NADH was investigated. In peroxisomal membranes, 100 micromolar NADH induced the production of O2 radicals. In the soluble fractions of peroxisomes, no generation of O2 radicals was observed by incubation with either NADH or xanthine, although xanthine oxidase was found located predominantly in the matrix of peroxisomes. The failure of xanthine to induce superoxide generation was probably due to the inability to fully suppress the endogenous Mn-superoxide dismutase activity by inhibitors which were inactive against xanthine oxidase. The generation of superoxide radicals in leaf peroxisomes together with the recently described production of these oxygen radicals in glyoxysomes (LM Sandalio, VM Fernández, FL Rupérez, LA del Río [1988] Plant Physiol 87: 1-4) suggests that O2 generation could be a common metabolic property of peroxisomes and further supports the existence of active oxygen-related rôles for peroxisomes in cellular metabolism.  相似文献   

16.
The nitrodiphenyl ether herbicide 5-[2-chloro-4-(trifluoromethyl)phenoxy]-2-nitroacetophenone oxime-O-(acetic acid, methyl ester) (DPEI) induced an abnormal accumulation of protoporphyrin IX in darkness in the green alga Chlamydomonas reinhardtii, as determined by high-performance liquid chromatography and spectrofluorimetry. It also inhibited the increase in cell density of the alga in light-grown cultures with an I50 (concentration required to decrease cell density increase to 50% of the noninhibited control value) of 0.16 μm. The relative ability of four peroxidizing diphenyl ether herbicides to cause tetrapyrrole accumulation in C. reinhardtii correlated qualitatively with their ability to inhibit the increase in cell density in light-grown cultures. The purified S(−) enantiomer of the optically active phthalide DPE 5-[2-chloro-4-(trifluoromethyl)phenoxy]-3-methylphthalide (DPEIII), which has greater herbicidal activity than the R(+) isomer, induces a 4- to 5-fold greater tetrapyrrole accumulation than the R(+) isomer. The I50 for inhibition of increase in cell density in light-grown cultures of C. reinhardtii by the S(−) isomer (0.019 μm) is less than 25% of that for the R(+) isomer. DPEIII inhibits protoporphyrinogen IX oxidase activity in pea (Pisum sativum) etioplast lysates, with the S(−) enantiomer showing considerably greater potency than the R(+) isomer and the racemic mixture showing a potency intermediate between the two. The results indicate that the site at which DPEs inhibit protoporphyrinogen IX oxidase shows chiral discrimination and provide further evidence for the link between inhibition of this enzyme, protoporphyrin IX accumulation, and the phytotoxicity of DPE herbicides.  相似文献   

17.
Antifungal drug ketoconazole causes severe drug-drug interactions by influencing gene expression and catalytic activity of major drug-metabolizing enzyme cytochrome P450 CYP3A4. Ketoconazole is administered in the form of racemic mixture of two cis-enantiomers, i.e. (+)-ketoconazole and (−)-ketoconazole. Many enantiopure drugs were introduced to human pharmacotherapy in last two decades. In the current paper, we have examined the effects of ketoconazole cis-enantiomers on the expression of CYP3A4 in human hepatocytes and HepG2 cells and on catalytic activity of CYP3A4 in human liver microsomes. We show that both ketoconazole enantiomers induce CYP3A4 mRNA and protein in human hepatocytes and HepG2 cells. Gene reporter assays revealed partial agonist activity of ketoconazole enantiomers towards pregnane X receptor PXR. Catalytic activity of CYP3A4/5 towards two prototypic substrates of CYP3A enzymes, testosterone and midazolam, was determined in presence of both (+)-ketoconazole and (−)-ketoconazole in human liver microsomes. Overall, both ketoconazole cis-enantiomers induced CYP3A4 in human cells and inhibited CYP3A4 in human liver microsomes. While interaction of ketoconazole with PXR and induction of CYP3A4 did not display enantiospecific pattern, inhibition of CYP3A4 catalytic activity by ketoconazole differed for ketoconazole cis-enantiomers ((+)-ketoconazole IC50 1.69 µM, Ki 0.92 µM for testosterone, IC50 1.46 µM, Ki 2.52 µM for midazolam; (−)-ketoconazole IC50 0.90 µM, Ki 0.17 µM for testosterone, IC50 1.04 µM, Ki 1.51 µM for midazolam).  相似文献   

18.
This work was designed to investigate the structural features required for a branched xyloglucan nonasaccharide (XG9; composition: glucose4xylose3galactose1fucose1) to exhibit anti-auxin activity in the pea (Pisum sativum L.) stem segment straight-growth bioassay. Oligosaccharides were prepared by cellulase-catalyzed hydrolysis of Rosa xyloglucan, and tested for auxin antagonism. The quantitatively major hepta-, octa-, and decasaccharides (XG7, XG8, and XG10) showed no antiauxin activity at the concentrations tested and did not interfere with the antiauxin effect of 10−9 molar XG9 when coincubated at equimolar concentrations. The results indicate that the XG9-recognition system in pea stem segments is highly discriminating. A terminal α-l-fucose residue is essential for the antiauxin activity of XG9 and a neighboring terminal β-d-galactose residue can abolish the activity; possible reasons for the effect of the galactose residue are discussed. A sample of XG9 extensively purified by gel-permeation chromatography followed by paper chromatography in two solvent systems still exhibited antiauxin activity with a concentration optimum around 10−9 molar. This diminishes the likelihood that the antiauxin activity reported for previous nonsaccharide preparations was due to a compound other than XG9.  相似文献   

19.
Naturally occurring (+)-trans-isoalliin, (RCRS)-(+)-trans-S-1-propenyl-L-cysteine sulfoxide, is a major cysteine sulfoxide in onion. The importance of producing it synthetically to support further research is very well recognized. The (+)-trans-isoalliin is prepared by chemical synthesis and reversed-phase (RP)-HPLC. First, S-2-propenyl-L-cysteine (deoxyalliin) is formed from L-cysteine and allyl bromide, which is then isomerized to S-1-propenyl-L-cysteine (deoxyisoalliin) by a base-catalyzed reaction. A mixture of cis and trans forms of deoxyisoalliin is formed and separated by RP-HPLC. Oxidation of the trans form of deoxyisoalliin by H2O2 produces a mixture of (−)- and (+)-trans-isoalliin. Finally, RP-HPLC is used successfully in separating (−)- and (+)-trans-isoalliin, and hence, (+)-trans-isoalliin is synthesized for the first time in this study. In addition, the (±) diastereomers of cis-isoalliin are also separated and purified by RP-HPLC.  相似文献   

20.
Group V (GV) phospholipase A2 (PLA2) is a member of the family of secreted PLA2 (sPLA2) enzymes. This enzyme has been identified in several organs, including the kidney. However, the physiologic role of GV sPLA2 in the maintenance of renal function remains unclear. We used mice lacking the gene encoding GV sPLA2 (Pla2g5−/−) and wild-type breeding pairs in the experiments. Mice were individually housed in metabolic cages and 48-h urine was collected for biochemical assays. Kidney samples were evaluated for glomerular morphology, renal fibrosis, and expression/activity of the (Na+ + K+)-ATPase α1 subunit. We observed that plasma creatinine levels were increased in Pla2g5−/− mice following by a decrease in creatinine clearance. The levels of urinary protein were higher in Pla2g5−/− mice than in the control group. Markers of tubular integrity and function such as γ-glutamyl transpeptidase, lactate dehydrogenase, and sodium excretion fraction (FENa+) were also increased in Pla2g5−/− mice. The increased FENa+ observed in Pla2g5−/− mice was correlated to alterations in cortical (Na+ + K+) ATPase activity/ expression. In addition, the kidney from Pla2g5−/− mice showed accumulation of matrix in corticomedullary glomeruli and tubulointerstitial fibrosis. These data suggest GV sPLA2 is involved in the maintenance of tubular cell function and integrity, promoting sodium retention through increased cortical (Na+ + K+)-ATPase expression and activity.  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