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.
In an effort to asses the effect of Val311Met point mutation of Bacillus subtilis protoporphyrinogen oxidase on the resistance to diphenyl ether herbicides, a Val311Met point mutant of B. subtilis protoporphyrinogen oxidase was prepared, heterologously expressed in Escherichia coli, and the purified recombinant Val311Met mutant protoporphyrinogen oxidase was kinetically characterized. The mutant protoporphyrinogen oxidase showed very similar kinetic patterns to wild type protoporphyrinogen oxidase, with slightly decreased activity dependent on pH and the concentrations of NaCl, Tween 20, and imidazole. When oxyfluorfen was used as a competitive inhibitor, the Val311Met mutant protoporphyrinogen oxidase showed an increased inhibition constant about 1.5 times that of wild type protoporphyrinogen oxidase. The marginal increase of the inhibition constant indicates that the Val311Met point mutation in B. subtilis protoporphyrinogen oxidase may not be an important determinant in the mechanism that protects protoporphyrinogen oxidase against diphenyl ether herbicides.  相似文献   

2.
In barley (Hordeum vulgare L.) root cells, activity for oxidizing protoporphyrinogen to protoporphyrin (protoporphyrinogen oxidase), a step in chlorophyll and heme synthesis, was found both in the crude mitochondrial fraction and in a plasma membrane enriched fraction separated by a sucrose gradient technique utilized for preparing plasma membranes. The specific activity (expressed as nanomoles of protoporphyrin formed per hour per milligram protein) in the mitochondrial fraction was 8 and in the plasma membrane enriched fraction was 4 to 6. The plasma membrane enriched fraction exhibited minimal cytochrome oxidase activity and no carotenoid content, indicating little contamination with mitochondrial or plastid membranes. Etioplasts from etiolated barley leaves exhibited a protoporphyrinogen oxidase specific activity of 7 to 12. Protoporphyrinogen oxidase activity in the barley root mitochondrial fraction and etioplast extracts was more than 90% inhibited by assay in the presence of the diphenyl ether herbicide acifluorfen methyl, but the activity in the plasma membrane enriched fraction exhibited much less inhibition by this herbicide (12 to 38% inhibition) under the same assay conditions. Acifluorfen-methyl inhibition of the organellar (mitochondrial or plastid) enzyme was maximal upon preincubation of the enzyme with 4 mm dithiothreitol, although a lesser degree of inhibition was noted if the organellar enzyme was preincubated in the presence of other reductants such as glutathione or ascorbate. Acifluorfen-methyl caused only 20% inhibition if the enzyme was preincubated in buffer without reductants. Incubation of barley etioplast extracts with the earlier tetrapyrrole precursor coproporphyrinogen and acifluorfen-methyl resulted in the accumulation of protoporphyrinogen, which could be converted to protoporphyrin even in the presence of the herbicide by the addition of the plasma membrane enriched fraction from barley roots. These findings have implications for the toxicity of diphenyl ether herbicides, whose light induced tissue damage is apparently caused by accumulation of the photoreactive porphyrin intermediate, protoporphyrin, when the organellar protoporphyrinogen oxidase enzyme is inhibited by herbicides. Our results suggest that the protoporphyrinogen that accumulates as a result of herbicide inhibition of the organellar enzyme can be oxidized to protoporphyrin by a protoporphyrinogen oxidizing activity that is located at sites such as the plasma membrane, which is much less sensitive to inhibition by diphenylether herbicides.  相似文献   

3.
The specific binding of the herbicide acifluorfen 5-[2-chloro-4-(trifluoromethyl)phenoxy]-2-nitrobenzoic acid to corn etioplast membranes is competitively inhibited by protoporphyrinogen IX, the substrate of protoporphyrinogen oxidase. Three other peroxidizing molecules, oxadiazon [5-terbutyl-3-(2,4-dichloro-5-isopropoxyphenyl)-1,3,4-oxadiazol -2-one], LS 82556 [(S)3-N-(methylbenzyl)carbamoyl-5-propionyl-2,6-lutidine], and M&B 39279 [5-amino-4-cyano-1-(2,6-dichloro-4-trifluoromethylphenyl)pyrazol], also compete with acifluorfen for its binding site. The four herbicides thus bind to the same site, or to closely located sites, on the enzyme protoporphyrinogen oxidase.  相似文献   

4.
Several diphenyl ether herbicides, such as acifluorfen methyl, have been previously shown to cause large accumulations of the heme and chlorophyll precursor, protoporphyrin, in plants. Lightinduced herbicidal damage is mediated by the photoactive porphyrin. Here we investigate whether diphenyl ether herbicides can affect porphyrin synthesis in rat and chick hepatocytes. In rat hepatocyte cultures, protoporphyrin, as well as coproporphyrin, accumulated after treatment with acifluorfen or acifluorfen methyl. Combination of acifluorfen methyl with an esterase inhibitor to prevent the conversion of acifluorfen methyl to acifluorfen resulted in a greater accumulation of porphyrins than caused by acifluorfen methyl or acifluorfen alone. In vitro enzyme studies of hepatic mitochondria isolated from rat and chick embryos demonstrated that protopor-phyrinogen oxidase, the penultimate enzyme of heme biosynthesis, was inhibited by low concentrations of acifluorfen, nitrofen, or acifluorfen methyl with the latter being the most potent inhibitor. These findings indicate that diphenyl ether treatment can cause protoporphyrin accumulation in rat hepatocyte cultures and suggest that this accumulation was associated with the inhibition of protoporphyrinogen oxidase. In cultured chick embryo hepatocytes, treatment with acifluorfen methyl plus an esterase inhibitor caused massive accumulation of uroporphyrin rather than protoporphyrin or coproporphyrin. Specific isozymes of cytochrome P450 were also induced in chick embryo hepatocytes. These effects were not observed in the absence of an esterase inhibitor. These results suggest that diphenyl ether herbicides can cause uroporphyrin accumulation similar to that induced by other cytochrome P450-inducing chemicals such as polyhalogenated aromatic hydrocarbons in the chick hepatocyte system.  相似文献   

5.
The use of herbicides to control undesirable vegetation has become a universal practice. For the broad application of herbicides the risk of damage to crop plants has to be limited. We introduced a gene into the genome of tobacco (Nicotiana tabacum) plants encoding the plastid-located protoporphyrinogen oxidase of Arabidopsis, the last enzyme of the common tetrapyrrole biosynthetic pathway, under the control of the cauliflower mosaic virus 35S promoter. The transformants were screened for low protoporphyrin IX accumulation upon treatment with the diphenyl ether-type herbicide acifluorfen. Leaf disc incubation and foliar spraying with acifluorfen indicated the lower susceptibility of the transformants against the herbicide. The resistance to acifluorfen is conferred by overexpression of the plastidic isoform of protoporphyrinogen oxidase. The in vitro activity of this enzyme extracted from plastids of selected transgenic lines was at least five times higher than the control activity. Herbicide treatment that is normally inhibitory to protoporphyrinogen IX oxidase did not significantly impair the catalytic reaction in transgenic plants and, therefore, did not cause photodynamic damage in leaves. Therefore, overproduction of protoporphyrinogen oxidase neutralizes the herbicidal action, prevents the accumulation of the substrate protoporphyrinogen IX, and consequently abolishes the light-dependent phytotoxicity of acifluorfen.  相似文献   

6.
Jacobs JM  Jacobs NJ 《Plant physiology》1993,101(4):1181-1187
We have investigated the formation of porphyrin intermediates by isolated barley (Hordeum vulgare) plastids incubated for 40 min with the porphyrin precursor 5-aminolevulinate and in the presence and absence of a diphenylether herbicide that blocks protoporphyrinogen oxidase, the enzyme in chlorophyll and heme synthesis that oxidizes protoporphyrinogen IX to protoporphyrin IX. In the absence of herbicide, about 50% of the protoporphyrin IX formed was found in the extraplastidic medium, which was separated from intact plastids by centrifugation at the end of the incubation period. In contrast, uroporphyrinogen, an earlier intermediate, and magnesium protoporphyrin IX, a later intermediate, were located mainly within the plastid. When the incubation was carried out in the presence of a herbicide that inhibits protoporphyrinogen oxidase, protoporphyrin IX formation by the plastids was completely abolished, but large amounts of protoporphyrinogen accumulated in the extraplastidic medium. To detect extraplastidic protoporphyrinogen, it was necessary to first oxidize it to protoporphyrin IX with the use of a herbicide-resistant protoporphyrinogen oxidase enzyme present in Escherichia coli membranes. Protoporphyrinogen is not detected by some commonly used methods for porphyrin analysis unless it is first oxidized to protoporphyrin IX. Protoporphyrin IX and protoporphyrinogen found outside the plastid did not arise from plastid lysis, because the percentage of plastid lysis, measured with a stromal marker enzyme, was far less than the percentage of these porphyrins in the extraplastidic fraction. These findings suggest that of the tetrapyrrolic intermediates synthesized by the plastids, protoporphyrinogen and protoporphyrin IX, are the most likely to be exported from the plastid to the cytoplasm. These results help explain the extraplastidic accumulation of protoporphyrin IX in plants treated with photobleaching herbicides. In addition, these findings suggest that plastids may export protoporphyrinogen or protoporphyrin IX for mitochondrial heme synthesis.  相似文献   

7.
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   

8.
It is now generally accepted that protoporphyrinogen oxidase is the target-enzyme for diphenyl-ether-type herbicides. Recent studies [Camadro, J-M., Matringe M., Scalla, R. & Labbe, P. (1991) Biochem. J. 277, 17-21] have revealed that in maize, diphenyl ethers competitively inhibit protoporphyrinogen oxidase with respect to its substrate, protoporphyrinogen IX. In this study, we show that, in purified pea etioplast, [3H]acifluorfen specifically binds to a single class of high-affinity binding sites with an apparent dissociation constant of 6.2 +/- 1.3 nM and a maximum density of 29 +/- 5 nmol/g protein. [3H]Acifluorfen binding reaches equilibrium in about 1 min at 30 degrees C. Half dissociation occurs in less than 30 s, indicating that the binding is fully reversible. The specificity of [3H]acifluorfen binding to protoporphyrinogen oxidase is examined. [3H]Acifluorfen binding is inhibited by all the peroxidizing molecules tested. The phthalimide derivative, N-(4-chloro-2-fluoro-5-isopropoxy)phenyl-3,4,5,6-tetra hydrophthalimide, exerts a mixed-competitive inhibition on this binding. The effects of all these molecules on the binding of [3H]acifluorfen are tightly linked to their capacity to inhibit pea etioplast protoporphyrinogen oxidase activity. Furthermore, protoporphyrinogen IX, the substrate of the reaction catalyzed by protoporphyrinogen oxidase, was able to competitively inhibit the binding of [3H]acifluorfen. In contrast, protoporphyrin IX, the product of the reaction, did not inhibit this binding. All these results provide clear evidence that in pea etioplasts, [3H]acifluorfen exclusively binds to protoporphyrinogen oxidase, that the protoporphyrinogen oxidase inhibitors tested so far bind to the same region of the enzyme and that this region overlaps the catalytic site of the enzyme.  相似文献   

9.
Plant protoporphyrinogen oxidase is of particular interest since it is the last enzyme of the common branch for chlorophyll and heme biosynthetic pathways. In addition, it is the target enzyme for diphenyl ether-type herbicides, such as acifluorfen. Two distinct methods were used to investigate the localization of this enzyme within Percoll-purified spinach chloroplasts. We first assayed the enzymatic activity by spectrofluorimetry and we analyzed the specific binding of the herbicide acifluorfen, using highly purified chloroplast fractions. The results obtained give clear evidence that chloroplast protoporphyrinogen oxidase activity is membrane-bound and is associated with both chloroplast membranes, i.e. envelope and thylakoids. Protoporphyrinogen oxidase specific activity was 7-8 times higher in envelope membranes than in thylakoids, in good agreement with the number of [3H]acifluorfen binding sites in each membrane system: 21 and 3 pmol/mg protein, respectively, in envelope membranes and thylakoids. On a total activity basis, 25% of protoporphyrinogen oxidase activity were associated with envelope membranes. The presence of protoporphyrinogen oxidase in chloroplast envelope membranes provides further evidence for a role of this membrane system in chlorophyll biosynthesis. In contrast, the physiological significance of the enzyme associated with thylakoids is still unknown, but it is possible that thylakoid protoporphyrinogen oxidase could be involved in heme biosynthesis.  相似文献   

10.
首次利用体外α-葡萄糖苷酶抑制模型对内蒙古产3种蒺藜科植物的9个提取物进行活性评价,并与阳性对照Acarbose比较,发现3种植物均有抑制α-葡萄糖苷酶活性。其中白刺石油醚提取物对α-葡萄糖苷酶的抑制活性(IC50=81.80 mg/L)最高,其余依次为小果白刺乙酸乙酯提取物(IC50=610.29 mg/L),霸王石油醚(IC50=627.22 mg/L)和乙酸乙酯提取物(IC50=838.40 mg/L),它们的抑制活性远大于阳性对照Acarbose(IC50=1103.01 mg/L)。结果发现,不同植物不同溶剂提取物的α-葡萄糖苷酶抑制活性不同。同一植物不同溶剂提取物相比较,甲醇提取物的α-葡萄糖苷酶抑制活性不及乙酸乙酯和石油醚提取物。  相似文献   

11.
Seedlings of Triticum aestivum L. and Secale cereale L. were grown in the presence of six different (five having different chemical structures) chlorosis-inducing herbicides: aminotriazole and its derivative SDR 5175, haloxidine, Sandoz 6706, fluometuron, and EMD-IT 5914. Concentrations were applied which allowed the leaves to grow normally and to reach normal total amino nitrogen contents but evoked a complete chlorosis (less than 6% chlorophyll). The effects of the herbicides on the accumulation of several chloroplast constituents and on peroxisomal and mitochondrial marker enzyme activities were compared. Wheat and rye, in general, gave very similar results, wheat being more sensitive to unspecific inhibitory effects.

In dark-grown plants, the herbicides had no or only minor effects on the rRNA pattern and on enzyme activities of the leaves. In the light, all herbicides applied prevented the accumulation of carotenoids and of chloroplastic rRNA. Consequently, ribulose-1,5-bisphosphate carboxylase activity was virtually absent. After all herbicide treatments in light, the leaves contained only rather low catalase activity. In the presence of aminotriazole and haloxidine, the chloroplast-specific NADP-glyceraldehyde-3-phosphate dehydrogenase and the peroxisomal enzymes glycolate oxidase and hydroxypyruvate reductase had high or even normal activities, as in untreated leaves. In leaves treated with Sandoz 6706, fluometuron, or EMDIT 5914, the activities of the latter three enzymes were, in parallel, only very low. Some herbicides interfered with enzyme activities in vitro, particularly with those of catalase and of glycolate oxidase. Among mitochondrial enzymes, cytochrome c oxidase activity was either unaffected or lower, while fumarase had considerably higher activities in the herbicide-treated, as compared to untreated leaves. The specific effects on peroxisomal enzymes cannot be explained by the hypothesis of herbicide-induced photodestructions in carotene-deficient plastids. Alternative explanations for the genesis of the chlorosis are discussed.

  相似文献   

12.
Protoporphyrinogen oxidase has been solubilized from plasma membranes of Desulfovibrio gigas. The enzyme was purified to apparent homogeneity with single silver-stained protein bands on isoelectric focusing and sodium dodecyl sulfate-polyacrylamide gels. This protoporphyrinogen oxidase has a molecular weight (Mr) of 148,000 and is composed of three dissimilar subunits of Mrs 12,000, 18,500, and 57,000, which are held together by sulfhydryl bonds. Unlike other protoporphyrinogen oxidases, which use molecular oxygen as an electron acceptor, this enzyme does not couple to oxygen. The protoporphyrinogen oxidase donates electrons to 2,6-dichlorophenol-indophenol but not to NAD+, NADP+, flavin adenine dinucleotide, or flavin mononucleotide. The natural physiological electron acceptor of the protoporphyrinogen oxidase from D. gigas is unknown. By using 2,6-dichlorophenol-indophenol as the electron acceptor, the Km and Vmax values for oxidation of protoporphyrinogen were determined to be 21 microM and 8.38 nmol/min per 70 micrograms of protein, respectively. The catalytic rate constant, Kcat, was calculated to be 17.7 mol of protoporphyrin formed per mole of enzyme per min of incubation, and the Kcat/Km was 0.84. Energies of activation were calculated from Arrhenius plots with 7,429 cal (ca. 31,080 J)/mol per degree below 10 degrees C and 1,455 cal (ca. 6,088, J)/mol per degree above 10 degrees C. Optimum enzyme activity was at 23 degrees C, and inhibition was observed with both N-ethylmaleimide and iodoacetamide.  相似文献   

13.
采用96微孔板法测定草麻黄抑制α-葡萄糖苷酶活性.草麻黄正丁醇(IC50=6.86 μg/mL)、乙酸乙酯(IC50 =77.28 μg/mL)和石油醚部位(IC50=190.20 μg/mL)抑制活性远高于阳性对照阿卡波糖(IC50=1081.27μg/mL).研究表明,草麻黄各部位均具有很好的α-筒萄糖苷酶抑制活性,可进行活性追踪分离活性成分.  相似文献   

14.
The mitochondrial location of protoporphyrinogen oxidase   总被引:4,自引:0,他引:4  
Using the digitonin method and subsequent fractionation of rat liver mitochondria, protoporphyrinogen oxidase (penultimate enzyme in the heme biosynthesis pathway) was found to be closely associated with the mitochondrial inner membrane fraction. Chemical treatment with non-specific probes (trypsin and diazobenzene sulfonate) of either intact or inverted mitoplasts, indicated that protoporphyrinogen oxidase was anchored within the lipid bilayer of the inner membrane. Protoporphyrinogen had an equal access to the active site of the enzyme from both sides of the inner membrane and its transformation to protoporphyrin did not appear to be energy-dependent. Studies of protoporphyrinogen synthesis from exogenously added coproporphyrinogen in either intact or hypoosmotically treated mitochondria underlined the importance of the peculiar submitochondrial location of coproporphyrinogen oxidase and protoporphyrinogen oxidase for the transfer of substrates to the inner membrane.  相似文献   

15.
A new spectrophotometric assay for protoporphyrinogen oxidase activity has been developed, involving enzymatic generation of protoporphyrinogen in the incubation medium. This assay, more sensitive and reliable than those previously described, can be used to measure this activity in yeast mitochondrial membranes, rat liver mitochondria and E. coli membranes. By measuring protoporphyrinogen oxidase activity in different wild type and heme-mutant yeast strains, it was shown that 1) one heme-mutant was totally lacking this activity, 2) different factors might control its level in yeast.  相似文献   

16.
Protoporphyrinogen oxidase (EC 1.3.3.4, PPOX) is the last enzyme in the branched tetrapyrrole biosynthetic pathway, before its substrate protoporphyrin is directed to the Mg and Fe branches for chlorophyll and haem biosynthesis, respectively. The enzyme exists in many plants in two similar isoforms, which are either exclusively located in plastids (PPOX I) or in mitochondria and plastids (PPOX II). Antisense RNA expression inhibited the formation of PPOX I in transgenic tobacco plants, which showed reduced growth rate and necrotic leaf damage. The cytotoxic effect is attributed to accumulation of photodynamically acting protoporphyrin. The expression levels of PPOX I mRNA and protein and the cellular enzyme activities were reduced to similar extents in transgenic plants grown under low- or high-light conditions (70 and 530 mumol photons m(-2) sec(-1)). More necrotic leaf lesions were surprisingly generated under low- than under high-light exposure. Several reasons were explored to explain this paradox and the intriguing necrotic phenotype of PPOX-deficient plants under both light intensity growth conditions. The same reduction of PPOX expression and activity under both light conditions led to similar initial protoporphyrin, but to faster decrease in protoporphyrin content during high light. It is likely that a light intensity-dependent degradation of reduced and oxidized porphyrins prevents severe photodynamic leaf damage. Moreover, under high-light conditions, elevated contents of reduced and total low-molecular-weight antioxidants contribute to the protection against photosensitizing porphyrins. These reducing conditions stabilize protoporphyrinogen in plastids and allow their redirection into the metabolic pathway.  相似文献   

17.
We describe fluorometric assays for two enzymes of the heme pathway, coproporphyrinogen oxidase and protoporphyrinogen oxidase. Both assays are based on measurement of protoporphyrin IX fluorescence generated from coproporphyrinogen III by the two consecutive reactions catalyzed by coproporphyrinogen oxidase and protoporphyrinogen oxidase. Both enzymatic activities are measured by recording protoporphyrin IX fluorescence increase in air-saturated buffer in the presence of EDTA (to inhibit ferrochelatase that can further metabolize protoporphyrin IX) and in the presence of dithiothreitol (that prevents nonenzymatic oxidation of porphyrinogens to porphyrins). Coproporphyrinogen oxidase (limiting) activity is measured in the presence of a large excess of protoporphyrinogen oxidase provided by yeast mitochondrial membranes isolated from commercial baker's yeast. These membranes are easy to prepare and are stable for at least 1 year when kept at -80 degrees C. Moreover they ensure maximum fluorescence of the generated protoporphyrin (solubilization effect), avoiding use of a detergent in the incubation medium. The fluorometric protoporphyrinogen oxidase two-step assay is closely related to that already described (J.-M. Camadro, D. Urban-Grimal, and P. Labbe, 1982, Biochem. Biophys. Res. Commun. 106, 724-730). Protoporphyrinogen is enzymatically generated from coproporphyrinogen by partially purified yeast coproporphyrinogen oxidase. The protoporphyrinogen oxidase reaction is then initiated by addition of the membrane fraction to be tested. However, when very low amounts of membrane are used, low amounts of Tween 80 (less than 1 mg/ml) have to be added to the incubation mixture to solubilize protoporphyrin IX in order to ensure optimal fluorescence intensity. This detergent has no effect on the rate of the enzymatic reaction when used at concentrations less than 2 mg/ml. Activities ranging from 0.1 to 4-5 nmol protoporphyrin formed per hour per assay are easily and reproducibly measured in less than 30 min.  相似文献   

18.
Blebbistatin is a myosin II-specific inhibitor. However, the mechanism and tissue specificity of the drug are not well understood. Blebbistatin blocked the chemotaxis of vascular smooth muscle cells (VSMCs) toward sphingosylphosphorylcholine (IC(50) = 26.1 +/- 0.2 and 27.5 +/- 0.5 microM for GbaSM-4 and A7r5 cells, respectively) and platelet-derived growth factor BB (IC(50) = 32.3 +/- 0.9 and 31.6 +/- 1.3 muM for GbaSM-4 and A7r5 cells, respectively) at similar concentrations. Immunofluorescence and fluorescent resonance energy transfer analysis indicated a blebbistatin-induced disruption of the actin-myosin interaction in VSMCs. Subsequent experiments indicated that blebbistatin inhibited the Mg(2+)-ATPase activity of the unphosphorylated (IC(50) = 12.6 +/- 1.6 and 4.3 +/- 0.5 microM for gizzard and bovine stomach, respectively) and phosphorylated (IC(50) = 15.0 +/- 0.6 microM for gizzard) forms of purified smooth muscle myosin II, suggesting a direct effect on myosin II motor activity. It was further observed that the Mg(2+)-ATPase activities of gizzard myosin II fragments, heavy meromyosin (IC(50) = 14.4 +/- 1.6 microM) and subfragment 1 (IC(50) = 5.5 +/- 0.4 microM), were also inhibited by blebbistatin. Assay by in vitro motility indicated that the inhibitory effect of blebbistatin was reversible. Electron-microscopic evaluation showed that blebbistatin induced a distinct conformational change (i.e., swelling) of the myosin II head. The results suggest that the site of blebbistatin action is within the S1 portion of smooth muscle myosin II.  相似文献   

19.
首次采用96微孔板法检测贵州和河南产凹叶厚朴抑制α-葡萄糖苷酶活性;并采用DPPH、ABTS和FRAP三种方法测定其抗氧化活性.贵州产凹叶厚朴乙酸乙酯(IC50 =7.22 μg,/mL)和正丁醇提取部位(IC50=36.59 μg/mL),河南产凹叶厚朴石油醚(IC50=107.04 μg/mL)和乙酸乙酯提取部位(IC50=17.17μg/mL),它们的活性都远高于于阳性对照Acarhose( IC50=1081.27 μg/mL).贵州产凹叶厚朴乙酸乙酯提取部位清除ABTS自由基的能力最强(IC50=8.81 μg/mL),强于阳性对照BHT(IC50=11.94 μg/mL);其次为河南产凹叶厚朴乙酸乙酯提取部位(IC50=12.73 μg/mL).研究结果表明,贵州产凹叶厚朴乙酸乙酯提取部位抑制α-葡萄糖苷酶和抗氧化活性最好.  相似文献   

20.
An intracellular (ATP + Mg2+)-dependent Ca2+ pumping mechanism has been identified and characterized within the cultured clonal neuroblastoma cell line N1E-115. Using cell suspensions treated with 0.005% saponin which selectively permeabilizes the plasma membrane in 95-98% of the cells, it was possible to show clearly that the intracellular Ca2+ pump mechanism is of non-plasma membrane origin and therefore can be compared directly with the Ca2+ pump characterized in detail in synaptosomal membrane vesicles (Gill, D. L., Grollman, E. F., and Kohn, L. D. (1981) J. Biol. Chem. 256, 184-192; Gill, D. L., Chueh, S. H., and Whitlow, C. L. (1984) J. Biol. Chem. 259, 10807-10813) which was proven by flux reversal studies to be derived from the neural plasma membrane (Gill, D. L. (1982) J. Biol. Chem. 257, 10986-10990). The intracellular Ca2+ pump in N1E-115 cells is distinct from mitochondrial Ca2+ accumulation and is increased up to 8-fold higher as cells reach confluency. In similarity to the neural plasma membrane pump, the intracellular Ca2+ pump within N1E-115 cells has high affinity for Ca2+ (Km = 0.28 microM), is dependent on both ATP (Km = 26 microM) and either Mg2+ or Mn2+ which half-maximally activate Ca2+ pumping at 0.35 mM and 0.32 mM, respectively, and shows similar specificity for Sr2+ and Ba2+ which half-maximally inhibit Ca2+ transport at 50 microM and 1.5 mM, respectively. In contrast to the neural plasma membrane pump, the intracellular Ca2+ pump displays approximately 40-fold higher sensitivity to La3+ (IC50 = 5 microM) and an apparent 400-fold lower sensitivity to VO4(3-) (IC50 = 185 microM), although the inhibitory effectiveness of VO4(3-) is increased 37-fold by a 15-min preincubation of the permeabilized cells with VO4(3-) in the absence of ATP (apparent IC50 = 5 microM). In further contrast to the neural plasma membrane Ca2+ pump, the intracellular pump within N1E-115 cells is stimulated more than 20-fold by oxalate (giving prolonged linear Ca2+ accumulation), is resistant to low saponin concentrations, and is not modified by calmodulin even after extensive treatment with ethylene glycol bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid and/or calmodulin antagonist drugs. However, calmidazolium is effective in inhibiting the intracellular Ca2+ pump with an IC50 of approximately 2 microM.  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