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 117 毫秒
1.
Bailey BA  Larson RL 《Plant physiology》1989,90(3):1071-1076
Hydroxamic acids occur in several forms in maize (Zea mays L.) with 2,4-dihydroxy-7-methoxy-2H-1,4-benzoxazin-3(4H)-one (DIMBOA) being the predominant form and others including 2,4-dihydroxy-1,4-benzoxazin-3-one (DIBOA) being found at lower concentrations. Two enzymes capable of glucosylating hydroxamic acids were identified in maize protein extracts and partially purified and characterized. The total enzyme activity per seedling increased during the first 4 days of germination and was concurrent with the accumulation of DIMBOA. Purification of the enzymes by ammonium sulfate precipitation followed by Sephadex G-200 and Q-Sepharose gel chromatography resulted in a 13-fold increase in specific activity. The enzymes are initially separated into two peaks (peak 1 and peak 2) of activity by Q-Sepharose gel chromatography. The peak 1 glucosyltransferase had 3.6% of the DIMBOA glucosylating activity when DIBOA was used as substrate, whereas this percentage increased to 57% for the peak 2 enzyme. The enzyme in peak 2 has a Km of 174 micromolar for DIMBOA and a Km of 638 micromolar for DIBOA; the enzyme in peak 1 has a Km of 217 micromolar for DIMBOA and its activity on DIBOA was too low to determine a Km. The identification of two glucosyltransferases capable of glucosylating hydroxamic acids in vitro serves as an initial step in the characterization of the enzymes involved in production of hydroxamic acids in maize.  相似文献   

2.
The subcellular localization of hexose phosphorylating activity in extracts of pea stems has been studied by differential centrifugation and sucrose density gradient centrifugation. The hexokinase (EC 2.7.1.1) was associated with the mitochondria, whereas fructokinase (EC 2.7.1.4) was in the cytosolic fraction. Some properties of the mitochondrial hexokinase were studied. The enzyme had a high affinity for glucose (Km 76 micromolar) and mannose (Km 71 micromolar) and a relatively low affinity for fructose (Km 15.7 millimolar). The Km for MgATP was 180 micromolar. The addition of salts stimulated the activity of the hexokinase. Al3+ was a strong inhibitor at pH 7 but not at the optimum pH (8.2). The enzyme was not readily solubilized but, in experiments with intact mitochondria, was susceptible to proteolysis. A location on the outer mitochondrial membrane is suggested for the hexokinase of pea stems.  相似文献   

3.
A barley (Hordeum vulgare L.) mutant, nar1a (formerly Az12), deficient in NADH nitrate reductase activity is, nevertheless, capable of growth with nitrate as the sole nitrogen source. In an attempt to identify the mechanism(s) of nitrate reduction in the mutant, nitrate reductase from nar1a was characterized to determine whether the residual activity is due to a leaky mutation or to the presence of a second nitrate reductase. The results obtained indicate that the nitrate reductase in nar1a differs from the wild-type enzyme in several important aspects. The pH optima for both the NADH and the NADPH nitrate reductase activities from nar1a were approximately pH 7.7, which is slightly greater than the pH 7.5 optimum for the NADH activity and considerably greater than the pH 6.0 to 6.5 optimum for the NADPH activity of the wild-type enzyme. The nitrate reductase from nar1a exhibits greater NADPH than NADH activity and has apparent Km values for nitrate and NADH that are approximately 10 times greater than those of the wild-type enzyme. The nar1a nitrate reductase has apparent Km values of 170 micromolar for NADPH and 110 micromolar for NADH. NADPH, but not NADH, inhibited the enzyme at concentrations greater than 50 micromolar.  相似文献   

4.
Phosphoribulokinase (EC 2.7.1.19) was investigated in wild-type Chlamydomonas reinhardtii and in mutant strains deficient in this enzyme activity. Immunoblot analysis revealed substantial amounts of phosphoribulokinase in mutant 12-2B but none in mutant F-60. The pH optimum of the wild-type enzyme was 8.0 and that of the 12-2B enzyme was 6.5. The mutant kinase possessed a Km value for ribulose 5-phosphate of about 45 millimolar, nearly three orders of magnitude greater than the wild-type value of 56 micromolar. Km values for ATP in the range of 36 to 72 micromolar were observed with both wild-type and mutant enzymes. The Vmax of the wild-type enzyme was about 450 micromoles per minute per milligram of protein, and values for the mutant enzyme were 140 micromoles per minute per milligram at pH 6.5 and 36 micromoles per minute per milligram at pH 7.8. Thermal stabilities of the wild-type and mutant kinases were similar. Sequence analysis of the 12-2B phosphoribulokinase gene revealed a C to T transition that caused an arginine to cysteine change at position 64 of the enzyme. This arginine residue is conserved in phosphoribulokinases from vascular plants, algae, and photosynthetic bacteria and appears to function in binding ribulose 5-phosphate.  相似文献   

5.
Oligomeric structure and kinetic properties of NADP-malic enzyme, purified from sugarcane (Saccharam officinarum L.) leaves, were determined at either pH 7.0 and 8.0. Size exclusion chromatography showed the existence of an equilibrium between the dimeric and the tetrameric forms. At pH 7.0 the enzyme was found preferentially as a 125 kilodalton homodimer, whereas the tetramer was the major form found at pH 8.0. Although free forms of l-malate, NADP+, and Mg2+ were determined as the true substrates and cofactors for the enzyme at the two conditions, the kinetic properties of the malic enzyme were quite different depending on pH. Higher affinity for l-malate (Km = 58 micromolar), but also inhibition by high substrate (Ki = 4.95 millimolar) were observed at pH 7.0. l-Malate saturation isotherms at pH 8.0 followed hyperbolic kinetics (Km = 120 micromolar). At both pH conditions, activity response to NADP+ exhibited Michaelis-Menten behavior with Km values of 7.1 and 4.6 micromolar at pH 7.0 and 8.0, respectively. Negative cooperativity detected in the binding of Mg2+ suggested the presence of at least two Mg2+ - binding sites with different affinity. The Ka values for Mg2+ obtained at pH 7.0 (9 and 750 micromolar) were significantly higher than those calculated at pH 8.0 (1 and 84 micromolar). The results suggest that changes in pH and Mg2+ levels could be important for the physiological regulation of NADP-malic enzyme.  相似文献   

6.
Cytidine 5′-triphosphate (CTP):phosphatidate cytidyltransferase from the endoplasmic reticulum and mitochondria of Ricinus communis L. var Hale was characterized. The endoplasmic reticulum enzyme has a pH optimum of 6.5 and a divalent cation is required, Mn2+ being preferred and giving maximum activity at 2.5 millimolar. The estimated Km for CTP is 16.7 micromolar, but that for phosphatidate could not be determined accurately. The activity was inhibited by both deoxycholate and Triton X-100 at concentrations as low as 0.01% (w/w).

The mitochondrial enzyme has a pH optimum of 6.0 and a divalent cation requirement similar to that of the endoplasmic reticulum. Maximum stimulation of the reaction by substrates occurred with 1.5 millimolar phosphatidate (from egg phosphatidylcholine) and about 400 micromolar CTP. The apparent Km for phosphatidate could not be estimated accurately since activity was obtained in the absence of added lipid, apparently utilizing endogenous substrate. The Km estimated for CTP was altered by the presence of the detergent Triton X-100; in its absence the value was 33.3 micromolar, but in its presence the value was 66.7 micromolar. Inclusion of 0.6% (w/w) Triton X-100 in the assay mixture stimulated the activity about 2.5-fold.

  相似文献   

7.
Hexokinase II of Pea Seeds   总被引:4,自引:4,他引:0       下载免费PDF全文
A second hexokinase (EC 2.7.1.1) was obtained from pea seed (Pisum sativum L. var. Progress No. 9) extracts. The enzyme, termed hexokinase II, had a high affinity (Km, 48 micromolar) for glucose and a relatively low affinity (Km, 10 millimolar) for fructose. The Km for MgATP was 86 micromolar. Mg2+ was required for activity, but excess Mg2+ was inhibitory. MgADP inhibited hexokinase II. The addition of salts of monovalent cations increased hexokinase II activity. Al3+ was a strong inhibitor of the enzyme at pH 6.6 but not at the optimum pH (8.2). Citrate and 3-phosphoglycerate activated pea seed hexokinase II at pH 6.6, probably by coordinating with aluminum present as a contaminant in commercial ATP. The properties of hexokinase II are compared with those of the other three hexose kinases obtained from pea seed extracts. The possible role of these enzymes in plant carbohydrate metabolism is discussed.  相似文献   

8.
Bromphenol blue, which was reduced with dithionite, was found to support nitrate reduction catalyzed by squash NADH:nitrate reductase at a rate about 5 times greater than NADH with freshly prepared enzyme and 10 times or more with enzyme having been frozen and thawed. Kinetic analysis of bromphenol blue as a substrate for squash nitrate reductase yielded apparent Km values of 60 micromolar for bromphenol blue at 10 millimolar nitrate and 500 micromolar for nitrate at 0.2 millimolar bromphenol blue. With the same preparation of enzyme the apparent Km values were 9 micromolar for NADH at 10 millimolar nitrate and 50 micromolar nitrate at 0.1 millimolar NADH. Bromphenol blue was found to be a noncompetitive inhibitor versus NADH with a Ki of 0.3 millimolar. When squash NADH:nitrate reductase activity was inactivated with p-hydroxymercuribenzoate or denatured by heating at 40°C, the bromphenol blue nitrate reductase activity was not lost. These results were taken to indicate that bromphenol blue and NADH donated electrons to nitrate reductase at different sites. When monoclonal antibodies prepared against corn and squash nitrate reductases were used to inhibit the nitrate reductase activities supported by NADH, bromphenol blue, and methyl viologen, differential inhibition was found which tended to indicate that the three electron donors were interacting with the enzyme at different sites. One monoclonal antibody prepared against squash nitrate reductase inhibited all three activities of both corn and squash nitrate reductase. It appears this antibody may bind to a highly conserved antigenic site in the nitrate binding region of the enzyme.  相似文献   

9.
Two enzymes capable of hydrolyzing fructose-1,6-bisphosphate (FBP) have been isolated from the foliose lichen Peltigera rufescens (Weis) Mudd. These enzymes can be separated using Sephadex G-100 and DEAE Sephacel chromatography. One enzyme has a pH optimum of 6.5, and a substrate affinity of 228 micromolar FBP. This enzyme does not require MgCl2 for activity, and is inhibited by AMP. The second enzyme has a pH optimum of 9.0, with no activity below pH 7.5. This enzyme responds sigmoidally to Mg2+, with half-saturation concentration of 2.0 millimolar MgCl2, and demonstrates hyperbolic kinetics for FBP (Km = 39 micromolar). This enzyme is activated by 20 millimolar dithiothreitol, is inhibited by AMP, but is not affected by fructose-2-6-bisphosphate. It is hypothesized that the latter enzyme is involved in the photosynthetic process, while the former enzyme is a nonspecific acid phosphatase.  相似文献   

10.
Microsomes from apical buds of pea (Pisum sativum L. var. Téléphone à rames) seedlings hydroxylate lauric acid at the ω-position. This oxidation is catalyzed by a cytochrome P-450 enzyme which differs from laurate hydroxylases previously described in microorganisms and mammals by its strict substrate specificity and the ability of low NADH concentrations to support unusually high oxidation rates. The apparent Km for lauric acid was 20 micromolar. NADPH- and NADH-dependent laurate hydroxylation followed non-Michaelian kinetics with apparent Km values ranging from 0.2 to 28 micromolar for NADPH, and 0.2 to 318 micromolar for NADH. When induced by the photomorphogenic photoreceptor phytochrome, the time course for the enhancement of laurate ω-hydroxylase was totally different from that of the cinnamic acid 4-hydroxylase, providing evidence for the existence of multiple cytochrome P-450 species in pea microsomes.  相似文献   

11.
S-adenosylmethionine decarboxylase of corn seedlings   总被引:2,自引:2,他引:0       下载免费PDF全文
Suzuki Y  Hirasawa E 《Plant physiology》1980,66(6):1091-1094
S-Adenosylmethionine decarboxylase (EC 4.1.1.50) has been purified 500-fold in 30% yield from the extract of etiolated corn seedlings (cv. Golden Crossbantam Bell). This preparation had a molecular weight of approximately 25,000. The Km value was 5 micromolar for S-adenosylmethionine. Methylglyoxal bis(guanylhydrazone), hydroxylamine, and sulfhydryl reagents (such as p-hydroxymercuriphenylsulfonate and N-ethylmaleimide) were effective inhibitors of this enzyme. Germination of corn seed was accompanied by a rapid increase in enzyme activity and maximum activity occurred in 5-day-old seedlings.  相似文献   

12.
A novel acyltransferase from cotyledons of tomato (Lycopersicon esculentum Mill.), which catalyzes the transfer of caffeic acid from chlorogenic acid (5-O-caffeoylquinic acid) to glucaric and galactaric acids, was purified with a 2400-fold enrichment and a 4% recovery. The enzyme showed specific activities (theoretical Vmax per milligram of protein) of 625 nanokatals (caffeoylglucaric acid formation) and 310 nanokatals (caffeoylgalactaric acid formation). On sodium dodecyl sulfate-polyacrylamide gel electrophoresis it gave an apparent Mr of 40,000, identical to the value obtained by gel filtration column chromatography. Highest activity was found at pH 5.7, which was constant over a range of 20 to 120 millimolar K-phosphate. The isoelectric point of the enzyme was at pH 5.75. The reaction temperature optimum was at 38°C and the apparent energy of activation was calculated to be 57 kilojoules per mole. The apparent Km values were 0.4 millimolar for glucaric acid, 1.7 millimolar for galactaric acid, and with both acceptors as second substrates 20 millimolar for chlorogenic acid. The relative ratio of the Vmax/Km values for glucaric acid and galactaric acid was found to be 100:12. Substrate-competition experiments support the conclusion that one single enzyme is responsible for both the glucaric and galactaric acid ester formation with marked preference for glucaric acid. It is proposed that the enzyme be called chlorogenic acid:glucaric acid O-caffeoyltransferase (EC 2.3.1.-). The three caffeic acid-dependent enzyme activities involved in the formation of the glucaric and galactaric acid esters, the chlorogenic acid:glucaric acid caffeoyltransferase as the key activity as well as the caffeic acid:CoA ligase and the caffeoyl-CoA:quinic acid caffeoyltransferase as the preceding activities, were determined. The time course of changes in these activities were followed during development of the seedling in the cotyledons and growth of the young plant in the first and second leaf. The results from tomato seedlings suggest a sequential appearance of these enzymes.  相似文献   

13.
The levels of activity of 2-phosphoglycolate phosphatase in the green algae, Chlamydomonas reinhardtii and Chlorella vulgaris, were in the range of 37 to 60 micromoles per milligram chlorophyll per hour and in the blue-green algae, Anacystis nidulans and Anabaena variabilis were 204 to 310 micromoles per milligram chlorophyll per hour. The activity in each species was similar regardless of whether the algae were grown with air or 5% CO2 in air. The enzyme purified 530-fold from Chlamydomonas was stable, had a broad pH optimum between 6 and 8.5, and was specific for the hydrolysis of P-glycolate with a Km of 23 micromolar. The enzyme purified 18-fold from Anacystis was labile, had a sharp pH optimum at 6.3, and was also specific for P-glycolate with a Km of 94 micromolar. The molecular weight of the enzyme from Chlamydomonas was estimated to be 92,000 by gel filtration.

The phosphatase from both sources required a divalent cation for activity. The Chlamydomonas enzyme was most effectively activated by Co2+, but was also activated by Mg2+ (Ka = 30 micromolar), Mn2+, and Zn2+. The Anacystis enzyme was most effectively activated by Mg2+ (Ka = 140 micromolar), and was also activated by Co2+ and Mn2+, but not by Zn2+. Anions were also required for maximum activity of the enzyme from both sources. The Chlamydomonas enzyme was activated about 2- to 3-fold by chloride (Ka = 140 micromolar), bromide, nitrate, bicarbonate (Ka = 600 micromolar) and formate. The Anacystis enzyme was activated over 10-fold by chloride (Ka = 870 micromolar), bromide, iodide, and nitrate, but was not activated by bicarbonate or formate.

The properties of the algal enzymes were similar to those previously reported for higher plants. The levels and kinetic properties of the enzyme seemed sufficient to account for the flux through the glycolate pathway that occurs in these algae. The phosphatase was not associated with the ribulose 1,5-bisphosphate carboxylase/oxygenase responsible for P-glycolate formation in the carboxysomes of Anacystis.

  相似文献   

14.
Acetyl-CoA carboxylase from the diatom Cyclotella cryptica has been purified to near homogeneity by the use of ammonium sulfate fractionation, gel filtration chromatography, and affinity chromatography with monomeric avidin-agarose. The specific activity of the final preparation was as high as 14.6 micromoles malonyl-CoA formed per milligram protein per minute, indicating a 600-fold purification. Native acetyl-CoA carboxylase has a molecular weight of approximately 740 kilodaltons and appears to be composed of four identical biotin-containing subunits. The enzyme has maximal activity at pH 8.2, but enzyme stability is greater at pH 6.5. Km values for MgATP, acetyl-CoA, and HCO3- were determined to be 65, 233, and 750 micromolar, respectively. The purified enzyme is strongly inhibited by palmitoyl-CoA, and is inhibited to a lesser extent by malonyl-CoA, ADP, and phosphate. Pyruvate stimulates enzymatic activity to a slight extent. Acetyl-CoA carboxylase from Cyclotella cryptica is not inhibited by cyclohexanedione or aryloxyphenoxypropionic acid herbicides as strongly as monocot acetyl-CoA carboxylases; 50% and 0% inhibition was observed in the presence of 23 micromolar clethodim and 100 micromolar haloxyfop, respectively.  相似文献   

15.
A 3′ -phosphoadenosine 5′ -phosphosulfate (PAPS):desulfoglucosinolate sulfotransferase (EC 2.8.2-) was extensively purified from light-grown cress (Lepidium sativum L.) seedlings by gel filtration and concanavalin A-Sepharose 4B, Matrex Gel Green A, and Mono Q fast protein liquid chromatography. The purified enzyme, which required bovine serum albumin for stabilization, had a native molecular weight of 31,000 ± 5,000 and an apparent isoelectric point of 5.2. Using PAPS (Km 60 micromolar) as sulfur donor, it catalyzed the sulfation of desulfobenzylglucosinolate (Km 82 micromolar), desulfo-p-hydroxybenzylglucosinolate (Km 670 micromolar), and desulfoallylglucosinolate (Km 6.5 millimolar) at an optimal pH of 9.0. All other potential substrates tested, including flavonoids, flavonoid glycosides, cinnamic acids, and phenylacetaldoxime, were not sulfated. Sulfotransferase activity was stimulated by MgCl2, MnCl2 and reducing agents and inhibited by ZnCl2, PbNO3 NiCl2 and the reaction product PAP. The thiol reagents N-ethylmaleimide, p-chloromercuriphenylsulfonic acid, and 5,5′ -dithio-bis-(2-nitrobenzoic acid) were also potent inhibitors, but the enzyme was protected from covalent modification by β-mercaptoethanol. The kinetics of desulfobenzylglucosinolate sulfation were consistent with a rapid equilibrium ordered mechanism with desulfobenzylglucosinolate binding first and PAPS second.  相似文献   

16.
A salicylic acid (SA)-inducible uridine 5′-diphosphate (UDP)-glucose:SA 3-O-glucosyltransferase was extracted from oat (Avena sativa L. cv Dal) roots. Reverse phase high-performance liquid chromatography or anion exchange chromatography was used to separate SA from the product, β-O-d-glucosylsalicylic acid. The soluble enzyme was purified 176-fold with 5% recovery using a combination of pH fractionation, anion exchange, gel filtration, and chromatofocusing chromatography. The partially purified protein had a native molecular weight of about 50,000, an apparent isoelectric point at pH 5.0, and maximum activity at pH 5.5. The enzyme had a Km of 0.28 mm for UDP-glucose and was highly specific for this sugar donor. More than 20 hydroxybenzoic and hydroxycinnamic acid derivatives were assayed as potential glucose acceptors. UDP-glucose:SA 3-O-glucosyltransferase activity was highly specific toward SA (Km = 0.16 mm). The enzyme was inhibited by UDP and uridine 5′-triphosphate but not by up to 7.5 mm uridine 5′-monophosphate.  相似文献   

17.
Doehlert DC 《Plant physiology》1989,89(4):1042-1048
Four forms of hexose kinase activity from developing maize (Zea mays L.) kernels have been separated by ammonium sulfate precipitation, gel filtration chromatography, blue-agarose chromatography, and ion exchange chromatography. Two of these hexose kinases utilized d-glucose most effectively and are classified as glucokinases (EC 2.7.1.2). The other two hexose kinases utilized only d-fructose and are classified as fructokinases (EC 2.7.1.4). All hexose kinases analyzed had broad pH optima between 7.5 and 9.5 with optimal activity at pH 8.5. The two glucokinases differed in substrate affinities. One form had low Km values [Km(glucose) = 117 micromolar, Km(ATP) = 66 micromolar] whereas the other form had much higher Km values [Km(glucose) = 750 micromolar, Km(ATP) = 182 micromolar]. Both fructokinases had similar substrate saturation responses. The Km(fructose) was about 130 micromolar and the Km(ATP) was about 700 micromolar. Both exhibited uncompetitive substrate inhibition by fructose [Ki(fructose) = 1.40 to 2.00 millimolar]. ADP inhibited all four hexose kinase activities, whereas sugar phosphates had little effect on their activities. The data suggest that substrate concentrations are an important factor controlling hexose kinase activity in situ.  相似文献   

18.
An acid carboxypeptidase (EC 3.4.16.1) has been isolated from the culture filtrate of a wood-degrading Basidiomycete, Pycnoporus sanguineus and the molecular and enzymatic properties of the enzyme were determined. The extracellular acid carboxypeptidase was homogeneous on polyacrylamide gel electrophoresis at pH 9.4 and SDS-disc gel electrophoresis. The MWs as determined by gel filtration and SDS-gel electrophoresis were 50 000 and 54 000, respectively. The isoelectric point was pH 4.78 using electrofocusing. The purified enzyme had a pH optimum of 3.4, a Km of 0.74 mM and a kcat of 16/sec with benzyloxycarbonyl-l-glutamyl-l-tyrosine. The Km and kcat values for bradykinin at pH 3.4 and 30° were 2.0 mM and 25/sec. Values for angiotensin at pH 3.4 and 30° were 0.76 mM and 2.4/sec, respectively.  相似文献   

19.
UDP-glucose:dolichylphosphate glucosyltransferase has been purified 734-fold from Triton X-100 solubilized mung bean (Phaseolus aureus) microsomes. The partially purified enzyme has broad pH optima of activity from 6.0 to 7.0 and is maximally stimulated with 10 millimolar MgCl2. The Km for UDP-glucose was determined as 27 micromolar, and the Km for dolichol-P was 2 micromolar. Using the UDP-glucose photoaffinity analog, 5-azido-UDP-glucose, a polypeptide of 39 kilodaltons on sodium dodecyl sulfate-polyacrylamide gels was identified as the catalytic subunit of the enzyme. Photoinsertion into this 39-kilodalton polypeptide with [32P]5-azido-UDP-glucose was saturable, and was maximally protected with the native substrate UDP-glucose. 5-Azido-UDP-glucose behaves competitively with UDP-glucose in enzyme assays, and upon photolysis inhibits activity in proportion to its concentration. This study represents the first subunit identification of a plant glycosyltransferase involved in the biosynthesis of the lipid-linked oligosaccharides that are precursors of N-linked glycoproteins.  相似文献   

20.
Vicente JA  Vale MG 《Plant physiology》1991,96(4):1345-1353
The phosphohydrolase activity of a light microsomal fraction isolated from corn roots (Zea mays L. cv LG 55) was investigated. The fraction, which appears to be enriched in endoplasmic reticulum and Golgi membranes, has ATPase and pyrophosphatase activities that hydrolyze ATP and pyrophosphate at an optimum pH of 7.0, with Km values of about 160 and 240 micromolar and with Vmax values of about 200 and 50 nanomoles substrate hydrolyzed per milligram protein per minute, respectively. These enzymes differ in their sensitivity to anions and inhibitors. The ATPase is stimulated by sulfate anions, whereas pyrophosphatase is inhibited by molybdate. Furthermore, the simultaneous addition of ATP and pyrophosphate to the reaction medium increases phosphohydrolysis, suggesting that separate enzymes are operating in the membranes. We also observed that pyrophosphate competitively inhibits the ATPase, whereas ATP has no significant effect on the pyrophosphatase. On the other hand, we observed a detergent-stimulated, molybdate-insensitive inosine diphosphatase activity which, in the native state, hydrolyzes inosine diphosphate with a Km of about 700 micromolar and a Vmax of about 450 nanomoles inosine diphosphate hydrolyzed per milligram protein per minute. In the solubilized form, the enzyme appears to be fully active, exhibiting lower Km values to hydrolyze inosine diphosphate. Furthermore, we found that native inosine diphosphatase is inhibited either by ATP or pyrophosphate, whereas inosine diphosphate inhibits the ATPase, but has no significant effect on the pyrophosphatase. It appears that inosine diphosphate is a positive modulator of the inosine diphosphatase, whereas ATP and pyrophosphate act as negative modulators of this 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