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 15 毫秒
1.
Three isoforms of α-glucosidase (EC 3.2.1.20) have been extracted from pea (Pisum sativum L.) seedlings and separated by DEAE-cellulose and CM-Sepharose chromatography. Two α-glucosidase isoforms (αG1 and αG2) were most active under acid conditions, and appeared to be apoplastic. A neutral form (αG3) was most active near pH 7, and was identified as a chloroplastic enzyme. Together, the activity of αG1 and αG2 in apoplastic preparations accounted for 21% of the total acid α-glucosidase activity recovered from pea stems. The vast majority (86%) of the apoplastic acid α-glucosidase activity was due to αG1. The apparent Km values for maltose of αG1 and αG2 were 0.3 and 1.3 millimolar, respectively. The apparent Km for maltose of αG3 was 33 millimolar. The respective native molecular weights of αG1, αG2, and αG3 were 125,000, 150,000, and 110,000.  相似文献   

2.
Lysine-ketoglutarate reductase activity was detected and characterized in the developing endosperm of maize (Zea mays L.). The enzyme showed specificity for its substrates: lysine, α-ketoglutarate, and NADPH. Formation of the reaction product saccharopine was demonstrated. The pH optimum of the enzyme was close to 7, and the Km for lysine and α-ketoglutarate were 5.2 and 1.8 millimolar, respectively.  相似文献   

3.
Pyruvate decarboxylase (PDC) was purified from mature, dry maize kernels and from roots of anaerobically treated maize seedlings and partially characterized. PDC was purified to a specific activity of 96 units per milligram protein from kernels and to 41 units per milligram protein from root. The subunit molecular masses were estimated to be 61,000 and 60,000 for kernel PDC and 59,000 and 58,000 for root PDC. The pH optimum for each enzyme was 5.8. Since the pH optimum is nearly one pH unit below the value reported for the cytoplasm of anaerobically metabolizing maize roots (pH 6.7 ± 0.2), we investigated the effects of pH 5.8 and 6.6 on the cooperative kinetics observed for PDC from each source. The maximum Hill coefficients (nH) were much greater at each pH for the kernel PDC (pH 5.8, nH = 2.5 and pH 6.6, nH = 3.2) than for the root PDC (pH 5.8, nH = 1.4 and pH 6.6, nH = 1.8). The cooperative kinetics observed with respect to pyruvate were asymmetric. Potassium inhibited maize PDC and was competitive with pyruvate (root PDC Ki = 16 millimolar and kernel PDC Ki = 10 millimolar).  相似文献   

4.
The α-ketoglutarate carrier from corn shoot mitochondria (Zea mays L., B 73) was solubilized in Triton X-114 and partially purified by chromatography on hydroxyapatite and celite in the presence of cardiolipin. On SDS-gel electrophoresis, the hydroxyapatite/celite eluate showed various protein bands between 12 and 70 kilodaltons. When reconstituted into liposomes, the α-ketoglutarate transport protein catalyzed a phthalonate-sensitive α-ketoglutarate/α-ketoglutarate exchange. The protein was purified 60-fold with a recovery of 88% with respect to the mitochondrial extract. The protein yield was 0.6%. The properties of the reconstituted carrier, i.e. requirement for a counter-anion, substrate specificity, and inhibitor sensitivity, were similar to those of the α-ketoglutarate transport system as characterized in plant and animal mitochondria.  相似文献   

5.
After removal of myrosinase activity by concanavalin A-Sepharose 4B chromatography, cell-free extracts of light-grown cress (Lepidium sativum L.) seedlings, catalyzed the sulfation of desulfobenzylglucosinolate (Km, 0.23 millimolar) to benzylglucosinolate using PAPS (Km, 1 millimolar) as sulfur donor. Sulfotransferase activity, which was optimal at pH 9.0, was stimulated by MgCl2, MnCl2, β-mercaptoethanol, and dithiothreitol and was inhibited by ZnSO4 and SH-reagents. The enzyme also sulfated desulfoallyglucosinolate to allylglucosinolate (sinigrin) but was inactive towards all phenylpropanoids and flavonoids tested.  相似文献   

6.
The novel sucrose derivative 1′-fluorosucrose (α-d-glucopyranosyl-β- d-1-deoxy-1-fluorofructofuranoside) was synthesized in order to help define mechanisms of sucrose entry into plant cells. Replacement of the 1′-hydroxyl by fluorine very greatly reduces invertase hydrolysis of the derivative (hydrolysis at 10 millimolar 1′-fluorosucrose is less than 2% that of sucrose) but does not reduce recognition, binding, or transport of 1′-fluorosucrose by a sucrose carrier. Transport characteristics of 1′-fluorosucrose were studied in three different tissues. The derivative is transported by the sucrose carrier in the plasmalemma of developing soybean cotyledon protoplasts with a higher affinity than sucrose (Km 1′-fluorosucrose 0.9 millimolar, Km sucrose 2.0 millimolar). 1′-Fluorosucrose is a competitive inhibitor of sucrose uptake with an apparent Ki also of 0.9 millimolar, while the Ki of sucrose competition of 1′-fluorosucrose uptake was 2.0 millimolar. Thus, both sugars are recognized at the same binding site in the plasmalemma. Both sucrose and 1′-fluorosucrose show very similar patterns of phloem translocation from an abraded leaf surface through the petiole indicating that recognition of 1′-fluorosucrose by sucrose carriers involved in phloem loading is likely as well.  相似文献   

7.
Mitochondria isolated from pea (Pisum sativum L.) leaves are able to transport the keto acid, oxaloacetate, from the reaction medium into he mitochondrial matrix at high rates. The rate of uptake by the mitochondria was measured as the rate of disappearance of oxaloacetate from the reaction medium as it was reduced by matrix malate dehydrogenase using NADH provided by glycine oxidation. The oxaloacetate transporter was identifed as being distinct from the dicarboxylate and the α-ketoglutarate transporters because of its inhibitor sensitivities and its inability to interact with other potential substrates. Phthalonate and phthalate were competitive inhibitors of oxaloacetate transport with Ki values of 60 micromolar and 2 millimolar, respectively. Butylmalonate, an inhibitor of the dicarboxylate and α-ketoglutarate transporters, did not alter the rate of oxaloacetate transport. In addition, a 1000-fold excess of malate, malonate, succinate, α-ketoglutarate, or phosphate had little effect on the rate of oxaloacetate transport. The Km for the oxaloacetate transporter was about 15 micromolar with a maximum velocity of over 500 nanomoles per milligram mitochondrial protein/min at 25°C. No requirement for a counter ion to move against oxaloacetate was detected and the highest rates of uptake occurred at alkaline pH values. An equivalent transporter has not been reported in animal mitochondria.  相似文献   

8.
Sucrose synthase of soybean nodules   总被引:6,自引:6,他引:0  
Sucrose synthase (UDPglucose: d-fructose 2-α-d-glucosyl transferase, EC 2.4.1.13) has been purified from the plant cytosolic fraction of soybean (Glycine max L. Merr cv Williams) nodules. The native enzyme had a molecular weight of 400,000. The subunit molecular weight was 90,000 and a tetrameric structure is proposed for soybean nodule sucrose synthase. Optimum activity in the sucrose cleavage and synthesis directions was at pH 6 and pH 9.5 respectively, and the enzyme displayed typical Michaelis-Menten kinetics. Soybean nodule sucrose synthase had a high affinity for UDP (Km, 5 micromolar) and a relatively low affinity for ADP (apparent Km, 0.13 millimolar) and CDP (apparent Km, 1.1 millimolar). The Km for sucrose was 31 millimolar. In the synthesis direction, UDPglucose (Km, 0.012 millimolar) was a more effective glucosyl donor than ADPglucose (Km, 1.6 millimolar) and the Km for fructose was 3.7 millimolar. Divalent cations stimulated activity in both the cleavage and synthesis directions and the enzyme was very sensitive to inhibition by heavy metals.  相似文献   

9.
Activated factor XII (FXIIa) is selectively inhibited by corn Hageman factor inhibitor (CHFI) among other plasma proteases. CHFI is considered a canonical serine protease inhibitor that interacts with FXIIa through its protease-binding loop. Here we examined whether the protease-binding loop alone is sufficient for the selective inhibition of serine proteases or whether other regions of a canonical inhibitor are involved. Six CHFI mutants lacking different N- and C-terminal portions were generated. CHFI-234, which lacks the first and fifth disulfide bonds and 11 and 19 amino acid residues at the N and C termini, respectively, exhibited no significant changes in FXIIa inhibition (Ki = 3.2 ± 0.4 nm). CHFI-123, which lacks 34 amino acid residues at the C terminus and the fourth and fifth disulfide bridges, inhibited FXIIa with a Ki of 116 ± 16 nm. To exclude interactions outside the FXIIa active site, a synthetic cyclic peptide was tested. The peptide contained residues 20–45 (Protein Data Bank code 1BEA), and a C29D substitution was included to avoid unwanted disulfide bond formation between unpaired cysteines. Surprisingly, the isolated protease-binding loop failed to inhibit FXIIa but retained partial inhibition of trypsin (Ki = 11.7 ± 1.2 μm) and activated factor XI (Ki = 94 ± 11 μm). Full-length CHFI inhibited trypsin with a Ki of 1.3 ± 0.2 nm and activated factor XI with a Ki of 5.4 ± 0.2 μm. Our results suggest that the protease-binding loop is not sufficient for the interaction between FXIIa and CHFI; other regions of the inhibitor also contribute to specific inhibition.  相似文献   

10.
Aspartate or glutamate stimulated the rate of light-dependent malate decarboxylation by isolated Zea mays bundle sheath chloroplasts. Stimulation involved a decrease in the apparent Km (malate) and an increased maximum velocity of decarboxylation. In the presence of glutamate other dicarboxylates (succinate, fumarate) competitively inhibited malate decarboxylation by intact chloroplasts with respect to malate with an apparent Ki of about 6 millimolar. For comparison the Ki for inhibition of nicotinamide adenine dinucleotide phosphate-malic enzyme from freshly lysed chloroplasts by these dicarboxylates was 15 millimolar. A range of compounds structurally related to aspartate stimulated malate decarboxylation by intact chloroplasts. Ka values for stimulation at 5 millimolar malate were 1.7, 5, and 10 millimolar for l-glutamate, l-aspartate, and β-methyl-dl-aspartate, respectively. Certain compounds, notably cysteic acid, which stimulated malate decarboxylation by intact chloroplasts inhibited malate decarboxylation by nicotinamide adenine dinucleotide phosphate-malic enzyme obtained from lysed chloroplasts and assayed under comparable conditions. It was concluded that aspartate, glutamate, and related compounds affect the transport of malate into the intact chloroplasts and that malate translocation does not take place on the general dicarboxylate translocator previously reported for higher plant chloroplasts.  相似文献   

11.
Properties of an Aminotransferase of Pea (Pisum sativum L.)   总被引:2,自引:2,他引:0  
A transaminase (aminotransferase, EC 2.6.1) fraction was partially purified from shoot tips of pea (Pisum sativum L. cv. Alaska) seedlings. With α-ketoglutarate as co-substrate, the enzyme transaminated the following aromatic amino acids: d,l-tryptophan, d,l-tyrosine, and d,l-phenylalanine, as well as the following aliphatic amino acids: d,l-alanine, d,l-methionine, and d,l-leucine. Of other α-keto acids tested, pyruvate and oxalacetate were more active than α-ketoglutarate with d,l-tryptophan. Stoichiometric yields of indolepyruvate and glutamate were obtained with d,l-tryptophan and α-ketoglutarate as co-substrates. The specific activity was three times higher with d-tryptophan than with l-tryptophan.  相似文献   

12.
Previous results (TJ Buckhout, Planta [1989] 178: 393-399) indicated that the structural specificity of the H+-sucrose symporter on the plasma membrane from sugar beet leaves (Beta vulgaris L.) was specific for the sucrose molecule. To better understand the structural features of the sucrose molecule involved in its recognition by the symport carrier, the inhibitory activity of a variety of phenylhexopyranosides on sucrose uptake was tested. Three competitive inhibitors of sucrose uptake were found, phenyl-α-d-glucopyranoside, phenyl-α-d-thioglucopyranoside, and phenyl-α-d-4-deoxythioglucopyranoside (PDTGP; Ki = 67, 180, and 327 micromolar, respectively). The Km for sucrose uptake was approximately 500 micromolar. Like sucrose, phenyl-α-d-thioglucopyranoside and to a lesser extent, PDTGP induced alkalization of the external medium, which indicated that these derivatives bound to and were transported by the sucrose symporter. Phenyl-α-d-3-deoxy-3-fluorothioglucopyranoside, phenyl-α-d-4-deoxy-4-fluorothioglucopyranoside, and phenyl-α-d-thioallopyranoside only weakly but competively inhibited sucrose uptake with Ki values ranging from 600 to 800 micromolar, and phenyl-α-d-thiomannopyranoside, phenyl-β-d-glucopyranoside, and phenylethyl-β-d-thiogalactopyranoside did not inhibit sucrose uptake. Thus, the hydroxyl groups of the fructose portion of sucrose were not involved in a specific interaction with the carrier protein because phenyl and thiophenyl derivatives of glucose inhibited sucrose uptake and, in the case of phenyl-α-d-thioglucopyranoside and PDTGP, were transported.  相似文献   

13.
Myrosinase (β-thioglucoside glucohydrolase, EC 3.2.3.1) was purified to apparent homogeneity from light-grown cress (Lepidium sativum L.) seedlings. This enzyme, which catalyzes hydrolysis of the glucosinolate sinigrin (Km, 115 micromolar) at an optimum pH of 5.5 in sodium citrate buffer, had a native molecular weight of 130 ± 5 kilodaltons and an isoelectric point of 4.7 to 4.9. SDS-PAGE revealed two polypeptides with molecular weights of 62 and 65 kilodaltons. Both subunits contained carbohydrate as shown by periodic acid-Schiff staining. The purified enzyme hydrolyzed p-nitrophenyl-β-d-glucoside (Km, 2.0 millimolar) at an optimum pH of 6.5 in phosphate buffer. The indolizidine alkaloid castanospermine, a known inhibitor of O-glycosidases, competitively inhibited the hydrolyses of sinigrin (thioglucosidase activity) and p-nitrophenyl-β-d-glucoside (O-glucosidase activity) with Ki values of 5 and 6 micromolar, respectively. In contrast, the related polyhydroxyalkaloids swainsonine and deoxynojirimycin were without effect upon these hydrolyses.  相似文献   

14.
Two ammonium-inducible, chloroplast-localized NADP-specific glutamate dehydrogenase isoenzymes were purified to homogeneity from Chlorella sorokiniana. These isoenzymes were homopolymers of either α- or β-subunits with molecular weights of 55,500 or 53,000, respectively. The α-isoenzyme was preferentially induced at low ammonium concentrations (2 millimolar or lower), whereas only the β-isoenzyme accumulated after cells were fully induced (120 minutes) at high ammonium concentrations (29 millimolar). Purification of isoenzymes was achieved by (NH4)2SO4 fractionation, gel-filtration, anion-exchange fast protein liquid chromatography, and affinity chromatography. The α- and β-isoenzymes were separated by their differential binding to Type 4 nicotinamide adenine dinucleotide phosphate-Sepharose. Both isoenzymes bound to an antibody affinity column to which purified antibody (prepared against β-isoenzyme) was covalently attached. Peptide mapping of the subunits showed them to have a high degree of sequence homology. Both subunits were synthesized in vitro from precursor protein(s) with a molecular weight of 58,500. Although the subunits have similar chemical, physical, and antigenic properties, their holoenzymes have strikingly different ammonium Km values. The ammonium Km of the β-isoenzyme remained constant at approximately 75 millimolar, whereas this Km of the α-isoenzyme ranged from 0.02 to 3.5 millimolar, depending upon nicotinamide adenine dinucleotide phosphate concentration.  相似文献   

15.
Acetohydroxamate (AHA) and aminooxyacetate (AOA) were found to be potent inhibitors of purified NADPH(NADH)-dependent glyoxylate reductase from spinach (Spinacia oleracea) leaves. AHA was a noncompetitive (ro mixed) inhibitor of the NADPH-dependent activity of the reductase with a Ki of 0.33 millimolar. With NADH serving as a cofactor, AHA preferentially bound to the same form of the enzyme as glyoxylate, exhibiting a Ki of 0.31 millimolar. Glycine hydroxamate and l-glutamic acid-γ-hydroxamate were also inhibitory, but to a lesser extent than AHA. Inhibition by AOA (Ki of 1.8 millimolar) was enhanced by increased concentrations of glyoxylate, indicating that the inhibitor preferentially reacted with the glyoxylate-bound form of the enzyme. Glycidate, an effector of glycolate metabolism in leaves, was found to be a much weaker inhibitor of the enzyme with a Ki of 21 millimolar. While the inhibition by both AHA and AOA was fully reversible, glycidate acted as a tight-binding inhibitor. These findings are discussed with respect to the use of AHA, AOA, and glycidate as inhibitors of photorespiratory carbon metabolism in leaves. Caution is recommended in the use of these inhibitors with intact tissue experiments due to their lack of specificity.  相似文献   

16.
The nicotinamide adenine dinucleotide-specific glutamate dehydrogenase (l-glutamate:NAD+ oxidoreductase, EC 1.4.1.2) of Chlorella sorokiniana was purified 1,000-fold to electrophoretic homogeneity. The native enzyme was shown to have a molecular weight of 180,000 and to be composed of four identical subunits with a molecular weight of 45,000. The N-terminal amino acid was determined to be lysine. The pH optima for the aminating and deaminating reactions were approximately 8 and 9, respectively. The Km values for α-ketoglutarate, NADH, NH4+, NAD+, and l-glutamate were 2 mm, 0.15 mm, 40 mm, 0.15 mm, and 60 mm, respectively. Whereas the Km for α-ketoglutarate and l-glutamate increased 10-fold, 1 pH unit above or below the pH optima for the aminating or deaminating reactions, respectively, the Km values for NADH and NAD+ were independent of change in pH from 7 to 9.6. By initial velocity, product inhibition, and equilibrium substrate exchange studies, the kinetic mechanism of enzyme was shown to be consistent with a bi uni uni uni ping-pong addition sequence. Although this kinetic mechanism differs from that reported for any other glutamate dehydrogenase, the chemical mechanism still appears to involve the formation of a Schiff base between α-ketoglutarate and an ε-amino group of a lysine residue in the enzyme. The physical, chemical, and kinetic properties of this enzyme differ greatly from those reported for the NH4+-inducible glutamate dehydrogenase in this organism.  相似文献   

17.
Experiments were performed to determine the effect of plasmalemma ATPase inhibitors on cell potentials (Ψ) and K+ (86Rb) influx of corn root tissue over a wide range of K+ activity. N,N′Dicyclohexylcarbodiimide (DCCD), oligomycin, and diethylstilbestrol (DES) pretreatment greatly reduced active K+ influx and depolarized Ψ at low, but not at high, K+ activity (K°). More comprehensive studies with DCCD and anoxia showed nearly complete inhibition of the active component of K+ influx over a wide range of K°, with no effect on the apparent permeability constant. DCCD had no effect on the electrogenic component of the cell potential (Ψp) above 0.2 millimolar K°. Net proton efflux was rapidly reduced 80 to 90% by DCCD. Since tissue ATP content and respiration were only slightly affected by the DCCD-pretreatment, the inhibitions of active K+ influx and Ψp at low K° can be attributed to inhibition of the plasmalemma ATPase.  相似文献   

18.
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   

19.
Chloroplasts from CAM-Mesembryanthemum crystallinum can transport phosphoenolpyruvate (PEP) across the envelope. The initial velocities of PEP uptake in the dark at 4°C exhibited saturation kinetics with increasing external PEP concentration. PEP uptake had a Vmax of 6.46 (±0.05) micromoles per milligram chlorophyll per hour and an apparent Kmpep of 0.148 (±0.004) millimolar. The uptake was competitively inhibited by Pi (apparent Ki = 0.19 millimolar), by glycerate 3-phosphate (apparent Ki = 0.13 millimolar), and by dihydroxyacetone phosphate, but malate and pyruvate were without effect. The chloroplasts were able to synthesize PEP when presented with pyruvate. PEP synthesis was light dependent. The prolonged synthesis and export of PEP from the chloroplasts required the presence of Pi or glycerate 3-phosphate in the external medium. It is suggested that the transport of pyruvate and PEP across the chloroplasts envelope is required during the gluconeogenic conversion of carbon from malate to storage carbohydrate in the light.  相似文献   

20.
An α-l-arabinofuranosidase has been purified 1043-fold from radish (Raphanus sativus L.) seeds. The purified enzyme was a homogeneous glycoprotein consisting of a single polypeptide with an apparent molecular weight of 64,000 and an isoelectric point value of 4.7, as evidenced by denaturing gel electrophoresis and reversed-phase or size-exclusion high-performance liquid chromatography and isoelectric focusing. The enzyme characteristically catalyzes the hydrolysis of p-nitrophenyl α-l-arabinofuranoside and p-nitrophenyl β-d-xylopyranoside in a constant ratio (3:1) of the initial velocities at pH 4.5, whereas the corresponding α-l-arabinopyranoside and β-d-xylofuranoside are unsusceptible. The following evidence was provided to support that a single enzyme with one catalytic site was responsible for the specificity: (a) high purity of the enzyme preparation, (b) an invariable ratio of the activities toward the two substrates throughout the purification steps, (c) a parallelism of the activities in activation with bovine serum albumin and in heat inactivation of the enzyme as well as in the inhibition with heavy metal ions and sugars such as Hg2+, Ag+, l-arabino-(1→4)-lactone, and d-xylose, and (d) results of the mixed substrate kinetic analysis using the two substrates. The enzyme was shown to split off α-l-arabinofuranosyl residues in sugar beet arabinan, soybean arabinan-4-galactan, and radish seed and leaf arabinogalactan proteins. Arabinose and xylose were released by the action of the enzyme on oat-spelt xylan. Synergistic action of α-l-arabinofuranosidase and β-d-galactosidase on radish seed arabinogalactan protein resulted in the extensive degradation of the carbohydrate moiety.  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