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.
Limited tryptic proteolysis of spinach (Spinacia oleracea) ribulose bisphosphate carboxylase/oxygenase (ribulose-P2 carboxylase) resulted in the ordered release of two adjacent N-terminal peptides from the large subunit, and an irreversible, partial inactivation of catalysis. The two peptides were identified as the N-terminal tryptic peptide (acetylated Pro-3 to Lys-8) and the penultimate tryptic peptide (Ala-9 to Lys-14). Kinetic comparison of hydrolysis at Lys-8 and Lys-14, enzyme inactivation, and changes in the molecular weight of the large subunit, indicated that proteolysis at Lys-14 correlated with inactivation, while proteolysis at Lys-8 occurred much more rapidly. Thus, enzyme inactivation is primarily the result of proteolysis at Lys-14. Proteolysis of ribulose-P2 carboxylase under catalytic conditions (in the presence of CO2, Mg2+, and ribulose-P2) also resulted in ordered release of these tryptic peptides; however, the rate of proteolysis at lysyl residues 8 and 14 was reduced to approximately one-third of the rate of proteolysis of these lysyl residues under noncatalytic conditions (in the presence of CO2 and Mg2+ only). The protection of these lysyl residues from proteolysis under catalytic conditions could reflect conformational changes in the N-terminal domain of the large subunit which occur during the catalytic cycle.  相似文献   

2.
The tritium release assay of Hutton et al. (Anal. Biochem.16, 384, 1966) for prolyl hydroxylase and of Miller (Anal. Biochem.45, 202, 1972) for lysyl hydroxylase have been modified. The reaction is carried out on a microscale, and tritiated water is collected after passage of the trichloroacetic acid-soluble reaction products through a small Dowex-50 (H+) column instead of using a vacuum distillation apparatus as described in the original procedures. When measured by the modified procedures, both the prolyl and lysyl hydroxlyase reactions showed regions of linearity with respect to enzyme concentration, time and substrate concentration and were almost completely dependent on ascorbate and α-ketoglutarate. In addition, both reactions were completely inhibited by the iron chelator, α,α′-dipyridyl. The results indicate that these assay procedures are valid means of measuring prolyl and lysyl hydroxylase activities.  相似文献   

3.
《Insect Biochemistry》1987,17(8):1155-1161
Additional data are provided on the enzyme 2-deoxyecdysone C-2 hydroxylase which has been shown in a previous study (Kappler et al., 1986) to be a mitochondrial hydroxylase with some classical characteristics of a cytochrome P-450 monooxygenase but which appeared to be insensitive to CO. Using 18O2, we have now demonstrated that molecular oxygen is directly incorporated into ecdysone during the process of C-2 hydroxylation. Neither cumene hydroperoxide nor linoleyl hydroperoxide could support C-2 hydroxylation. When the reaction was sustained by α-ketoglutarate, addition of cofactors like Fe2+, ascorbate and catalase caused only a slight increase of the enzymatic activity whereas the α-ketoglutarate-dependent hydroxylation was largely decreased in the presence of malonate; these data eliminate the possible existence of a dioxygenase mechanism for C-2 hydroxylation.The paper also provides inhibition kinetics which indicate that 2-deoxy-20-hydroxyecdysone, 2,22-bisdeoxyecdysone and 2,22,25-trideoxyecdysone are competitive inhibitors of the C-2 hydroxylase whereas the 3-epi isomer of 2-deoxyecdysone is a non-competitive inhibitor.  相似文献   

4.
We previously have described a substance present in crude sonicates of L-929 cells which replaced ascorbate in vitro as a reductant for prolyl hydroxylase (B. Peterkofsky, D. Kalwinksy and R. Assad, 1980, Arch. Biochem. Biophys.199, 362–373). In the present study we found that almost 90% of the substance was particulate after differential centrifugation of stationary phase L-929 cell homogenates. The substance was not localized in nuclei or mitochondria and was found in the same fractions as microsomes, but these fractions also contained lysosomes and cell membranes. The reductant could not be solubilized from particles by Brij-35, indicating that it is an intrinsic component of a membrane rather than intracisternally located. The intramembranous cofactor, in the absence of ascorbate, participated in the in vitro hydroxylation of [4-3H]proline in radio-actively labeled, intracisternal unhydroxylated procollagen in isolated microsomes which also contained prolyl hydroxylase. Hydroxylation was determined by measuring tritiated water formed from release of the 4-trans tritium atom. Since it is unlikely that such participation could occur if the cofactor were located within the membrane of another subcellular organelle, we have concluded that it is in the same particle as prolyl hydroxylase and unhydroxylated procollagen, that is, the microsome. With the endogenous reductant the reaction was slower than with saturating ascorbate and was increased by NADH. Maximum hydroxylation with the endogenous reductant was close to that which could be achieved with ascorbate. These results provide strong evidence that the endogenous reductant alone can account for the phenomenon of ascorbate-independent proline hydroxylation in L-929 cells. As in the case of ascorbate, the microsomal reductant functioned only in the presence of α-ketoglutarate and Fe2+ and served as reductant for lysyl hydroxylase. It also was detected in the particulate fraction of virally transformed BALB 3T3 cells and in purified microsomes from bones of intact chick embryos. Since ascorbate could be taken up and concentrated in bone microsomes, it is unlikely that the endogenous reductant serves as an intermediary between ascorbate and intracisternal prolyl hydroxylase.  相似文献   

5.
Product inhibition studies with Rhodopseudomonas spheriodes NADP+ specific isocitrate dehydrogenase indicate that the enzyme mechanism involves the ordered addition of the substrates NADP+ and threo-ds-isocitrate and the ordered release of products CO2 (HCOs?), 2-ketoglutarate, and NADPH. In addition, the presence of a ternary complex consisting of enzyme, NADP+, and 2-ketoglutarate is indicated. Binding studies with radioactive substrates support the kinetically derived mechanism. The Rhodopseudomonas enzyme is dimeric and contains but a single active site. Different combinations of substrate were ineffective in causing gross changes in molecular structure as monitored by gel filtration techniques. A comparison of the amino acid composition of this enzyme with the bacterial enzyme from Azotobacter vinelandii indicate very significant differences in the amino acid compositions.  相似文献   

6.
The ATP.Mg-dependent type 1 protein phosphatase is inactive as isolated but can be activated in several different ways. In this report, we show that the phosphatase can also be activated by the Fe2+/ascorbate system. Activation of the phosphatase requires both Fe2+ ion and ascorbate and the level of activation is dependent on the concentrations of Fe2+ ion and ascorbate. In the presence of 20 mM ascorbate, the Fe2+ ion concentrations required for half-maximal and maximal activation are about 0.3 and 3mM, respectively. Several common divalent metal ions, including Co2+, Ni2+, Cu2+, Mg2+, and Ca2+ ions, cannot cooperate with ascorbate to activate the phosphatase, and SH-containing reducing agents such as 2-mercaptoethanol and dithiothreitol cannot cooperate with Fe2+ ion to activate the phosphatase, indicating that activation of the phosphatase by the Fe2+/ascorbate system is a specific process. Moreover, H2O2, a strong oxidizer, could significantly diminish the phosphatase activation by the Fe2+/ascorbate system, suggesting that reduction mechanism other than SH-SS interchange is a prerequisite for the Fe2+/ascorbate-mediated phosphatase activation. Taken together, the present study provides initial evidence for a new mode of type 1 protein phosphatase activation mechanism.  相似文献   

7.
The ATP.Mg-dependent type 1 protein phosphatase is inactive as isolated but can be activated in several different ways. In this report, we show that the phosphatase can also be activated by the Fe2+/ascorbate system. Activation of the phosphatase requires both Fe2+ ion and ascorbate and the level of activation is dependent on the concentrations of Fe2+ ion and ascorbate. In the presence of 20 mM ascorbate, the Fe2+ ion concentrations required for half-maximal and maximal activation are about 0.3 and 3mM, respectively. Several common divalent metal ions, including Co2+, Ni2+, Cu2+, Mg2+, and Ca2+ ions, cannot cooperate with ascorbate to activate the phosphatase, and SH-containing reducing agents such as 2-mercaptoethanol and dithiothreitol cannot cooperate with Fe2+ ion to activate the phosphatase, indicating that activation of the phosphatase by the Fe2+/ascorbate system is a specific process. Moreover, H2O2, a strong oxidizer, could significantly diminish the phosphatase activation by the Fe2+/ascorbate system, suggesting that reduction mechanism other than SH-SS interchange is a prerequisite for the Fe2+/ascorbate-mediated phosphatase activation. Taken together, the present study provides initial evidence for a new mode of type 1 protein phosphatase activation mechanism.Abbreviations MAPK mitogen-activated protein kinase - MCO metal ion-catalyzed oxidation - kinase FA the activating factor of ATP.Mg-dependent protein phosphatase - I2 inhibitor-2 - EDTA ethylenediaminetetraacetic acid - MBP myelin basic protein  相似文献   

8.
A gibberellin 2β-hydroxylase has been purified from mature seeds ofPhaseolus vulgaris. The enzyme is of molecular weight 36,000 and has the characteristics of a dioxygenase; the cofactors areα-ketoglu-tarate, Fe2+ and ascorbate, and activity is stimulated by catalase. The Vmax of the enzyme is 6.86 nmole h?1 mg?1, and the Km values for [1,2-3H2]GA1 andα-ketoglutarate are 0.085 μM and 21 μM, respectively. The purified enzyme preparation catalyzes hydroxylation of GA1, GA4, GA9, and GA20 but exhibits a marked preference for the 3-hydroxylated gibberellins as substrate.  相似文献   

9.
The kinetics of the lysyl hydroxylase (peptidyllysine, 2-oxoglutarate:oxygen 5-oxidoreductase, EC 1.14.11.4) reaction were studied using enzyme from chick embryos by varying the concentration of one substrate in the presence of different fixed concentrations of the second substrate, while the concentrations of the other substrates were held constant. Intersecting lines were obtained in double-reciprocal plots for all possible pairs involving Fe2+, alpha-ketoglutarate, O2 and the peptide substrate, whereas parallel lines were obtained for pairs comprising ascorbate and each of the other substrates. The pair composed of Fe2+ and alpha-ketoglutarate gave an asymmetrical initial veolcity pattern, indicating binding of these two reactants in this order, that of Fe2+ being at thermodynamic equilibrium. The initial velocity patterns are identical with those reported for prolyl 4-hydroxylase, and the apparent Km and Kd values calculated from these data are also very similar. The largest difference was fo-nd in Km and Kd for alpha-ketoglutarate, which were about 4 times the corresponding values for prolyl 4-hydroxylase. Ascorbate was found to be a quite specific requirement for lysyl hydroxylase, but the enzyme catalyzed its reaction for a short time at a high rate in the complete absence of this vitamin, suggesting that the reaction with ascorbate does not occur during each catalytic cycle. Lysyl hydroxylase catalyzed an uncoupled decarboxylation of alpha-ketoglutarate in the absence of the peptide substrate, the rate being about 4% of that observed in the presence of a saturating concentration of the peptide substrate. This uncoupled decarboxylation required the same cosubstrates as the complete reaction.  相似文献   

10.
Hydroxylation of 6-N-trimethyl-l-lysine(lys(Me3)) to 3-hydroxy-6-N-trimethyl-l-lysine(3-HO-lys(Me3)) by several rat tissues has been examined and compared. The kidney enzyme, which previously was shown to require molecular oxygen and α-ketoglutarate as cosubstrates, ferrous iron and ascorbate as cofactors, and to be stimulated by catalase, has a broad pH optimum ranging between 6.5 to 7.5 at 37 °C. As determined with crude tissue extracts from kidney, liver, heart, and skeletal muscle, similar apparent Km values were obtained for substrate, cosubstrates, and cofactors. In view of similar kinetic parameters among the several lys(Me3) hydroxylases examined in rat tissues, and the fact that the level of skeletal muscle lys(Me3) hydroxylase activity is comparable to that of heart, liver, and kidney, because of its large total mass, skeletal muscle may contribute significantly to the biosynthesis of l-carnitine from lys(Me3). The most effective inhibitors found, competitive with lys(Me3), were 2-N-acetyl-6-N-trimethyl-l-lysine, 6-N-monomethyl-l-lysine, and 6-N-dimethyl-l-lysine. l-2-Amino-6-N-trimethylammonium-4-hexynoate, d-2-amino-6-N-trimethylammonium-4-hexynoate, and dl2-amino-6-N-trimethylammonium-cis-4-hexenoate, also inhibited hydroxylase activity but by a yet undetermined mechanism. Oxalacetate, succinate, and citrate inhibited the hydroxylation reaction by competing with α-ketoglutarate. The binding of ferrous iron to the enzyme was competitively inhibited by ions of “soft metals” (e.g., Cd2+, Zn2+) but not by those of “hard metals” (e.g., Ca2+, Mg2+). Preincubation of the crude kidney enzyme for 15 min at 37 °C with mercuriphenylsulfonate, N-ethylmaleimide, iodoacetate, or iodoacetamide resulted in considerable inhibition of 3-HO-lys(Me3) formation. The degree of inhibition by N-ethylmaleimide could be reduced by including Zn (II) during preincubation of the enzyme. The effects of “soft” metals and sulfhydryl reagents on the enzyme suggest that sulfhydryl groups are required for ferrous iron binding in the active site.  相似文献   

11.
The GA20 3β-hydroxylase present in immature seeds of Phaseolus vulgaris has been partially purified and characterized. The physical characteristics of the enzyme are similar to those of the GA 2β-hydroxylases present in mature and immature seeds of Pisum sativum. It is acid-labile, hydrophobic, and of Mr 45,000. The enzyme catalyzes the synthesis of GA1, GA5, and GA29 from GA20. Activity is dependent upon the presence of Fe2+, ascorbate, 2-oxoglutarate, and oxygen. 2-Oxoglutarate does not function as a cosubstrate; in the presence of the enzyme, succinate is not a reaction product.  相似文献   

12.
Abstract: Tyrosine hydroxylase activity is reversibly modulated by the actions of a number of protein kinases and phosphoprotein phosphatases. A previous report from this laboratory showed that low-molecular-weight substances present in striatal extracts lead to an irreversible loss of tyrosine hydroxylase activity under cyclic AMP-dependent phosphorylation conditions. We report here that ascorbate is one agent that inactivates striatal tyrosine hydroxylase activity with an EC50 of 5.9 μM under phosphorylating conditions. Much higher concentrations (100 mM) fail to inactivate the enzyme under nonphosphorylating conditions. Isoascorbate (EC50, 11 μM) and dehydroascorbate (EC50, 970 μM) also inactivated tyrosine hydroxylase under phosphorylating but not under nonphosphorylating conditions. In contrast, ascorbate sulfate was inactive under phosphorylating conditions at concentrations up to 100 mM. Since the reduced compounds generate several reactive species in the presence of oxygen, the possible protecting effects of catalase, peroxidase, and superoxide dismutase were examined. None of these three enzymes, however, afforded any protection against inactivation. We also examined the effects of ascorbate and its congeners on the activity of tyrosine hydroxylase purified to near homogeneity from a rat pheochromocytoma. This purified enzyme was also inactivated by the same agents that inactivated the impure corpus striatal enzyme. Under conditions in which ascorbate almost completely abolished enzyme activity, we found no indication for significant prote-olysis of the purified enzyme as determined by sodium do-decyl sulfate-polyacrylamide gel electrophoresis. We also found that pretreatment of PC12 cells in culture for 4 h with 1 mM ascorbate, dehydroascorbate, or isoascorbate (but not ascorbate sulfate) also decreased tyrosine hydroxylase activity 25–50%. The inactivation seen under in vitro conditions appears to have a counterpart under more physiological conditions.  相似文献   

13.
Crude preparations of lysyl hydroxylase were extracted from chick-embryo tendons synthesizing exclusively type I collagen, chick-embryo sterna synthesizing exclusively type II collagen and HT-1080 sarcoma cells synthesizing exclusively type IV collagen. No differences were found in the Km values for Fe2+, 2-oxoglutarate and ascorbate between these three enzymes preparations. Similarly no differences were found in the Km values for type I and type II protocollagens and the rate at which type IV protocollagen is hydroxylated between these enzyme preparations. The extent to which type I protocollagen could be hydroxylated by the three enzymes was likewise identical. These data strongly argue against the existence of collagen-type-specific lysyl hydroxylase isoenzymes.  相似文献   

14.
The kinetics of a Mn2+-requiring, NADP+-specific isocitrate dehydrogenase from Salmonella typhimurium have been examined by the measurement of initial velocity rates in the presence and absence of the reaction products. The binding of each of the cosubstrates, isocitrate, and NADP+, is not independent of the other, and the isocitrate-Mn2+ complex is the kinetically important substrate species. All of the reaction products, α-ketoglutarate, CO2, and NADPH are competitive with both cosubstrates and the mechanism appears to be of the rapid equilibrium random type. The enzyme has been purified to homogeneity and has an isoelectric point at pH 4.0–4.2, and an apparent molecular weight of 102,000.  相似文献   

15.
The ring expansion enzyme (desacetoxycephalosporin C synthetase) of cephalosporin C biosynthesis in Cephalosporium acremonium has been purified 40-fold, using gel filtration on LKB Ultrogel AcA54. Purity was about twice that previously attained. The purified enzyme showed an almost absolute requirement for α-ketoglutarate, and most other acids were inactive or only weakly active. The order of addition of the cofactors and reactants was found to exert a major effect on enzyme activity. The major negative effect was caused by α-ketoglutarate or penicillin N being added to the enzyme first. The most effective order of addition to the enzyme was found to be the simultaneous addition of Fe2+, ascorbate and ATP, followed by α-ketoglutarate and then penicillin N. The negative effect of the required reactant, α-ketoglutarate, when added too early, coincides with observations made about other α-ketoglutarate-dependent dioxygenases even though, in the case of the synthetase, one atom of oxygen does not end up in the product, desacetoxycephalosporin C.  相似文献   

16.
Root cultures of various solanaceous plants grow well in vitro and produce large amounts of tropane alkaloids. Enzyme activity that converts hyoscyamine to 6β-hydroxyhyoscyamine is present in cell-free extracts from cultured roots of Hyoscyamus niger L. The enzyme hyoscyamine 6β-hydroxylase was purified 3.3-fold and characterized. The hydroxylation reaction has absolute requirements for hyoscyamine, 2-oxoglutarate, Fe2+ ions and molecular oxygen, and ascorbate stimulates this reaction. Only the l-isomer of hyoscyamine serves as a substrate; d-hyoscyamine is nearly inactive. Comparisons were made with a number of root, shoot, and callus cultures of the Atropa, Datura, Duboisia, Hyoscyamus, and Nicotiana species for the presence of the hydroxylase activity. Decarboxylation of 2-oxoglutarate during the conversion reaction was studied using [1-14C]-2-oxoglutarate. A 1:1 stoichiometry was shown between the hyoscyamine-dependent formation of CO2 from 2-oxoglutarate and the hydroxylation of hyoscyamine. Therefore, the enzyme can be classified as a 2-oxoglutarate-dependent dioxygenase (EC 1.14.11.-). Both the supply of hyoscyamine and the hydroxylase activity determine the amounts of 6β-hydroxyhyoscyamine and scopolamine produced in alkaloid-producing cultures.  相似文献   

17.
In permanent blastulae of the sea urchin, which were obtained by culture in SO2?4-free artificial seawater from the time of fertilization, ascorbate and α-ketoglutarate, activators of protocollagen proline hydroxylase, induced the formation of archenteron. By adding either ascorbate or α-ketoglutarate to the SO2?4-free culture at 12 hr of fertilization, spherical embryos with archenteron were obtained by successive 12 hr cultures at 20°C. The embryos thus obtained did not develop to plutei. Archenteron formation induced by these compounds in SO2?4-free-cultured embryos, as well as in the normal embryos, was inhibited by α,α′-dipyridyl, an inhibitor of protocollagen proline hydroxylase. Glutamate, malate, citrate, and fumarate did not stimulate archenteron formation in SO2?4-free cultured embryos. In the SO2?4-free-cultured embryos exposed to [14C]proline, considerable radioactivity was found in hot trichloroacetic acid-extractable proteins but the radioactivity of [14C]hydroxyproline residue, produced by hydroxylation of proline residue of protocollagen, was markedly lower than that in normal embryos. In the presence of ascorbate and α-ketoglutarate, the radioactivity of [14C]hydroxyproline residue became high and was lowered by α,α′-dipyridyl. Archenteron formation induced by ascorbate and α-ketoglutarate in the embryos kept in SO2?4-free artificial seawater probably results from the stimulated protocollagen hydroxylation.  相似文献   

18.
Dynamic equilibria in iron uptake and release by ferritin   总被引:7,自引:0,他引:7  
The function of ferritins is to store and release ferrous iron. During oxidative iron uptake, ferritin tends to lower Fe2+ concentration, thus competing with Fenton reactions and limiting hydroxy radical generation. When ferritin functions as a releasing iron agent, the oxidative damage is stimulated. The antioxidant versus pro-oxidant functions of ferritin are studied here in the presence of Fe2+, oxygen and reducing agents. The Fe2+-dependent radical damage is measured using supercoiled DNA as a target molecule. The relaxation of supercoiled DNA is quantitatively correlated to the concentration of exogenous Fe2+, providing an indirect assay for free Fe2+. After addition of ferrous iron to ferritin, Fe2+ is actively taken up and asymptotically reaches a stable concentration of 1–5 m. Comparable equilibrium concentrations are found with plant or horse spleen ferritins, or their apoferritins. After addition of ascorbate, iron release is observed using ferrozine as an iron scavenger. Rates of iron release are dependent on ascorbate concentration. They are about 10 times larger with pea ferritin than with horse ferritin. In the absence of ferrozine, the reaction of ascorbate with ferritins produces a wave of radical damage; its amplitude increases with increased ascorbate concentrations with plant ferritin; the damage is weaker with horse ferritin and less dependent on ascorbate concentrations.  相似文献   

19.
The effect of lactic acid (lactate) on Fenton based hydroxyl radical (·OH) production was studied by spin trapping, ESR, and fluorescence methods using DMPO and coumarin-3-carboxylic acid (3-CCA) as the ·OH traps respectively. The ·OH adduct formation was inhibited by lactate up to 0.4mM (lactate/iron stoichiometry = 2) in both experiments, but markedly enhanced with increasing concentrations of lactate above this critical concentration. When the H2O2 dependence was examined, the DMPO-OH signal was increased linearly with H2O2 concentration up to 1 mM and then saturated in the absence of lactate. In the presence of lactate, however, the DMPO-OH signal was increased further with higher H2O2 concentration than 1 mM, and the saturation level was also increased dependent on lactate concentration. Spectroscopic studies revealed that lactate forms a stable colored complex with Fe3+ at lactate/Fe3+ stoichiometry of 2, and the complex formation was strictly related to the DMPO-OH formation. The complex formation did not promote the H2O2 mediated Fe3+ reduction. When the Fe3+-lactate (1:2) complex was reacted with H2O2, the initial rate of hydroxylated 3-CCA formation was linearly increased with H2O2 concentrations. All the data obtained in the present experiments suggested that the Fe3+-lactate (1:2) complex formed in the Fenton reaction system reacts directly with H2O2 to produce additional ·OH in the Fenton reaction by other mechanisms than lactate or lactate/Fe3+ mediated promotion of Fe3+/Fe2+ redox cycling.  相似文献   

20.
Abstract– Detergent-solubilized tyrosine hydroxylase from the caudate nucleus of the sheep was purified 3-fold by affinity chromatography on 3-iodotyrosine modified agarose. Supplementation of the standard assay with 1 mM Fe2+ resulted in only slight stimulation of the enzymic activity. The enzyme was, however, markedly inhibited by certain complexing agents specific for either Fe2+ or Fe3+. Double reciprocal plots of the kinetic data for a representative complexing agent, bathophenanthroline, showed the inhibition to be competitive with O2 (apparent Km 0.15 mM) and noncompetitive with either l -tyrosine or the synthetic tetrahydropterin cofactor DMPH4 (apparent Km's 0.12 and 0.29 mM, respectively). The combined inhibition and kinetics studies strongly suggest that brain tyrosine hydroxylase is an iron enzyme. Furthermore, the prosthetic iron very likely participates directly in catalytic function, presumably by binding molecular oxygen. The activity of ammonium sulphate-precipitated enzyme was found to be stimulated 2-fold by Fe2+, catalase or peroxidase, while untreated enzyme was markedly less affected by these agents. Since the only ostensible difference between the two preparations was the extensive aggregation present in the former case, the change in physical state evoked by ammonium sulphate precipitation appeared to be somehow related to this peculiar property of the enzyme.  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