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.
We previously reported that a velvetleaf (Abutilon theophrasti Medic) biotype found in Maryland was resistant to atrazine because of an enhanced capacity to detoxify the herbicide via glutathione conjugation (JW Gronwald, Andersen RN, Yee C [1989] Pestic Biochem Physiol 34: 149-163). The biochemical basis for the enhanced atrazine conjugation capacity in this biotype was examined. Glutathione levels and glutathione S-transferase activity were determined in extracts from the atrazine-resistant biotype and an atrazine-susceptible or “wild-type” velvetleaf biotype. In both biotypes, the highest concentration of glutathione (approximately 500 nanomoles per gram fresh weight) was found in leaf tissue. However, no significant differences were found in glutathione levels in roots, stems, or leaves of either biotype. In both biotypes, the highest concentration of glutathione S-transferase activity measured with 1-chloro-2,4-dinitrobenzene or atrazine as substrate was in leaf tissue. Glutathione S-transferase measured with 1-chloro-2,4-dinitrobenzene as substrate was 40 and 25% greater in leaf and stem tissue, respectively, of the susceptible biotype compared to the resistant biotype. In contrast, glutathione S-transferase activity measured with atrazine as substrate was 4.4- and 3.6-fold greater in leaf and stem tissue, respectively, of the resistant biotype. Kinetic analyses of glutathione S-transferase activity in leaf extracts from the resistant and susceptible biotypes were performed with the substrates glutathione, 1-chloro-2,4-dinitrobenzene, and atrazine. There was little or no change in apparent Km values for glutathione, atrazine, or 1-chloro-2,4-dinitrobenzene. However, the Vmax for glutathione and atrazine were approximately 3-fold higher in the resistant biotype than in the susceptible biotype. In contrast, the Vmax for 1-chloro-2,4-dinitrobenzene was 30% lower in the resistant biotype. Leaf glutathione S-transferase isozymes that exhibit activity with atrazine and 1-chloro-2,4-dinitrobenzene were separated by fast protein liquid (anion-exchange) chromatography. The susceptible biotype had three peaks exhibiting activity with atrazine and the resistant biotype had two. The two peaks of glutathione S-transferase activity with atrazine from the resistant biotype coeluted with two of the peaks from the susceptible biotype, but peak height was three- to fourfold greater in the resistant biotype. In both biotypes, two of the peaks that exhibit glutathione S-transferase activity with atrazine also exhibited activity with 1-chloro-2,4-dinitrobenzene, with the peak height being greater in the susceptible biotype. The results indicate that atrazine resistance in the velvetleaf biotype from Maryland is due to enhanced glutathione S-transferase activity for atrazine in leaf and stem tissue which results in an enhanced capacity to detoxify the herbicide via glutathione conjugation.  相似文献   

2.
An unusual allantoinase from Dolichos biflorus has been purified 62-fold. The purified enzyme has an unusual pH activity profile with a shoulder at pH 4 and a peak at pH 7.5. This is due to a single enzyme which does not need metal ions for activation. In the fully reduced state the enzyme exhibits a single sharp peak at 7.5; when it is not in the sulfhydryl form (in the fully oxidized SS form?) the enzyme shows a single pH optimum at pH 4. Km values for (±)-allantoin were 5.5 mM at pH 4 and 1.43 mM at pH 7.5. Allantoinase activity has been demonstrated in the resting seed, and increased linearly with time during the first 5 days of seedling growth.  相似文献   

3.
Bacterial atrazine catabolism is initiated by the enzyme atrazine chlorohydrolase (AtzA) in Pseudomonas sp. strain ADP. Other triazine herbicides are metabolized by bacteria, but the enzymological basis of this is unclear. Here we begin to address this by investigating the catalytic activity of AtzA by using substrate analogs. Purified AtzA from Pseudomonas sp. strain ADP catalyzed the hydrolysis of an atrazine analog that was substituted at the chlorine substituent by fluorine. AtzA did not catalyze the hydrolysis of atrazine analogs containing the pseudohalide azido, methoxy, and cyano groups or thiomethyl and amino groups. Atrazine analogs with a chlorine substituent at carbon 2 and N-alkyl groups, ranging in size from methyl to t-butyl, all underwent dechlorination by AtzA. AtzA catalyzed hydrolytic dechlorination when one nitrogen substituent was alkylated and the other was a free amino group. However, when both amino groups were unalkylated, no reaction occurred. Cell extracts were prepared from five strains capable of atrazine dechlorination and known to contain atzA or closely homologous gene sequences: Pseudomonas sp. strain ADP, Rhizobium strain PATR, Alcaligenes strain SG1, Agrobacterium radiobacter J14a, and Ralstonia picketti D. All showed identical substrate specificity to purified AtzA from Pseudomonas sp. strain ADP. Cell extracts from Clavibacter michiganensis ATZ1, which also contains a gene homologous to atzA, were able to transform atrazine analogs containing pseudohalide and thiomethyl groups, in addition to the substrates used by AtzA from Pseudomonas sp. strain ADP. This suggests that either (i) another enzyme(s) is present which confers the broader substrate range or (ii) the AtzA itself has a broader substrate range.  相似文献   

4.
The in vivo rates of uptake and detoxification of alachlor and metolachlor were determined using Pioneer corn 3320 seedlings. Equal amounts of the radiolabeled herbicides were applied to etiolated coleoptiles and, at various intervals after treatment, the unabsorbed radioactivity was removed and quantified. Analysis of 80% methanol extracts by reverse phase liquid chromatography showed no significant differences in the rate of uptake of metolachlor and alachlor. However, the rate of glutathione conjugation of alachlor in vivo was two- to threefold greater than the rate for metolachlor at 2 and 4 hours after herbicide application. Since the initial step in detoxification is conjugation of the chloroacetanilide to glutathione, the activities of the enzymes responsible for conjugation, the glutathione-S-transferases (GST) were also analyzed in vitro, using crude extracts and the purified GST enzymes. The specific activities of the extracts were consistent with the results in vivo. Using alachlor as a substrate, the specific activity for glutathione conjugation was almost threefold higher than that for metolachlor. Kinetic analysis of purified GST III indicates that the enzyme has a higher affinity for alachlor (Kmapp = 1.69 millimolar) than for metolachlor (Kmapp = 8.9 millimolar).  相似文献   

5.
Pea (Pisum sativum L.) chloroplast D-enzyme (4-α-d-glucanotransferase, EC 2.4. 1.25) was purified greater than 750-fold and partially characterized. It is a dimer with a subunit Mr of ca. 50,000. Optimal activity is between pH 7.5 and 8.0 with maltotriose as substrate and the enzyme's Km for maltotriose is 3.3 millimolar. Chloroplast D-enzyme converts maltotriose to maltopentaose and glucose via the exchange of α-1,4-glycosidic linkages. Maltotriose acts either as a donor or acceptor of a maltosyl group. The enzyme has highest activity with maltotriose as substrate. As initial substrate degree of polymerization is increased to maltoheptaose, D-enzyme activity drops to zero at 10 millimolar substrate concentrations and by 70% at 1 millimolar concentrations. The enzyme cannot use maltose as a substrate. Glucose was found to be a suitable acceptor substrate for this D-enzyme. Addition of glucose to incubation mixtures, or production of glucose by D-enzyme, prevents the synthesis of maltodextrins larger than maltopentaose. Removal of glucose produced by D-enzyme activity with maltotriose as substrate resulted in the synthesis of maltopentaose and maltodextrins with sufficient degrees of polymerization to be suitable substrates for pea chloroplast starch phosphorylase. The possible role of D-enzyme in pea chloroplast starch metabolism is discussed.  相似文献   

6.
An improved assay method of a pseudomurein-degrading enzyme and its properties are described. The pseudomurein-degrading enzyme purified from Methanobacterium wolfei autolysate under an anoxic condition was assayed with the cell wall of Methanobacterium thermoautotrophicum as a substrate. By this improved method the enzyme activity was measured quantitatively and reproducibly. Moreover, the cell wall substrate can be stored in a freezer and used as needed, and the time required for an assay was as short as 1 h. The optimum pH and temperature of the enzyme was pH 6.8-7.4 and 75°C, respectively. Although the enzyme lost 50% of the activity upon heating at 75°C for 10 min in the absence of the cell wall substrate, it was more stable against heat inactivation in the presence of the substrate. Furthermore the inactivated enzyme recovered some of the activity by incubating with the substrate. Although the enzyme lost most of the activity under aerobic conditions, the activity was recovered under reducing conditions with Na2S·9H2O or DTT (dithiothreitol). The enzyme was also purified under aerobic conditions retaining the same specific activity as the anoxically purified enzyme. Using the partially purified enzyme the conditions preparing protoplasts of M. thermoautotrophicum was established.  相似文献   

7.
《Insect Biochemistry》1991,21(2):165-176
A lysosomal aspartic protease with cathepsin D activity, from the mosquito, Aedes aegypti, was purified and characterized. Its isolation involved ammonium sulfate (30–50%) and acid (pH 2.5) precipitations of protein extracts from whole previtellogenic mosquitoes followed by cation exchange chromatography. Purity of the enzyme was monitored by SDS-PAGE and silver staining of the gels. The native molecular weight of the purified enzyme as determined by polyacrylamide gel electrophoresis under nondenaturing conditions was 80,000. SDS-PAGE resolved the enzyme into a single polypeptide with Mr = 40,000 suggesting that it exists as a homodimer in its non-denatured state. The pI of the purified enzyme was 5.4 as determined by isoelectric focusing gel electrophoresis. The purified enzyme exhibits properties characteristic of cathepsin D. It utilizes hemoglobin as a substrate and its activity is completely inhibited by pepstatin-A and 6M urea but not by 10 mM KCN. Optimal activity of the purified mosquito aspartic protease was obtained at pH 3.0 and 45°C. With hemoglobin as a substrate the enzyme had an apparent Km of 4.2 μ M. Polyclonal antibodies to the purified enzyme were raised in rabbits. The specificity of the antibodies to the enzyme was verified by immunoblot analysis of crude mosquito extracts and the enzyme separated by both non-denaturing and SDS-PAGE. Density gradient centrifugation of organelles followed by enzymatic and immunoblot analyses demonstrated the lysosomal nature of the purified enzyme. The N-terminal amino acid sequence of the purified mosquito lysosomal protease (19 amino acids) has 74% identity with N-terminal amino acid sequence of porcine and human cathepsins D.  相似文献   

8.
Purified hepatic soluble guanylate cyclase (EC 4.6.1.2) had maximal specific activities in the unactivated state of 0.4 and 1 μmol cyclic GMP min?1 mg protein?1, when MgGTP and MnGTP, respectively, were used as substrates. The apparent Km for GTP was 85 or 10 μm in the presence of excess Mg2+ or Mn2+, respectively. Guanylate cyclase purified as described was deficient in heme but could be readily reconstituted with heme by reacting enzyme with hematin and excess dithiothreitol at 4 °C and pH 7.8. Unpurified guanylate cyclase was activated 20- to 84-fold by NO, nitroso compounds, NO-heme, and protoporphyrin IX. The purified enzyme was only slightly (2- to 3-fold) activated by NO and nitroso compounds but was markedly (50-fold) activated by NO-heme and protoporphyrin IX, achieving maximal specific activities of 10 μmol cyclic GMP min?1 mg protein?1. Enzyme activation by NO and nitroso compounds was restored by addition of hematin or by reconstitution of guanylate cyclase with heme. Excess hematin, however, inhibited enzyme activity. A partially purified heat-stable factor (activation-enhancing factor) was found to enhance (2- to 35-fold) enzyme activation without directly stimulating guanylate cyclase. In the presence of optimal concentrations of hematin, enzyme activation was still increased (2-fold) by the activation-enhancing factor but not by bovine serum albumin. Guanylate cyclase was markedly inhibited by SH reactive agents such as cystine, o-iodosobenzoic acid, periodate, and 5,5′-dithiobis (2-nitrobenzoic acid). In addition, CN? and FMN inhibited enzyme activation by NO-heme, but not by protoporphyrin IX, and did not affect basal enzymatic activity. Hepatic soluble guanylate cyclase appears to possess SH groups required for catalysis and to require heme and/or other unknown factors for the full expression of enzyme activation by NO and nitroso compounds.  相似文献   

9.
An oxidase activity utilizing reduced nicotinamide adenine dinucleotide phosphate (NADPH) and producing H2O2 was observed in intact adipocytes of rat, as well as in the isolated plasma membranes of these cells. A stoichiometry of 1 mol of H2O2 production per mole of NADPH disappearance was found with isolated plasma membranes. Activation of this enzyme (R) was produced by pretreatment of cells with insulin, dithiothreitol, or sulfhydryl inhibitors, e.g., p-chloromercuribenzoate or tosyl-l-lysine chloromethyl ketone. All of these agents also stimulated glucose oxidation via the hexose monophosphate shunt. Activation of R was also observed with biologically active derivatives of insulin, e.g., proinsulin or desalanine insulin, but not with an inactive derivative, desoctapeptide insulin. The enzyme could not be activated by exposing the cells to membrane perturbants, e.g., hypotonic conditions or Triton X-100 (0.01–0.1%). The enzyme activity in the plasma membrane had a pH optimum at 6.0 and, from the Lineweaver-Burke plot, V was determined at 230 nmol and Km for NADPH was at 5.8 × 10?5, m. The activity remained unaltered in the presence of sodium azide or cyanide. Preincubation of adipocytes with insulin or SH reagents or direct addition of oxidants, e.g., H2O2, potassium ferricyanide, or phenazine methosulfate, to the membranes also caused inhibition of adenylate cyclase (AC). This enzyme activity could be restored in these preparations by adding thiols. It is suggested that inhibition of AC in whole cells in response to insulin may be caused by oxidation of its SH groups by the H2O2 generated from the activated NADPH oxidase. Reversal of this inhibition may involve cellular reducing equivalents. The evidence suggests that the plasma membrane enzymes, i.e., NADPH oxidase and adenylate cyclase, are controlled, in part, by the intracellular redox potential.  相似文献   

10.
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   

11.
Attraction of the German cockroach, Blattella germanica, to esters was tested in a heavily infested room. Sticky traps of cardboard baited with hexyl hexanoate and pentanoate captured higher numbers than an empty control. Recluction or extension of the alcohol moiety even by a C1 unit resulted in a decrease of activity. The suggested importance of the C6 moiety which is known to make a ring form with the least effort led to examination of the activity of the corresponding cyclohexanol esters; the result was some tenfold increase in activity. A further tenfold increase in attractiveness was attained by inverting the O and C adjacent to the ester carbonyl function, i.e. by modifying cyclohexyl pentanoate into propyl cyclohexaneacetate.  相似文献   

12.
Two major endoproteinases were purified from senescing primary barley leaves. The major enzyme (EP1) appeared to be a thiol proteinase and accounted for about 85% of the total proteolytic activity measured in vitro. This proteinase was purified 5,800-fold and had a molecular weight of 28,300. It was highly unstable in the absence of dithiothreitol or at a pH greater than 7.5. Leupeptin, at a concentration of 10 micromolar, inhibited this enzyme 100%. A second proteinase (EP2) was purified approximately 50-fold and had a molecular weight of 67,000. It was inhibited 20% by 1 millimolar dithiothreitol and 50% by 1 millimolar phenylmethyl sulfonylfluoride. EP2 contributed about 15% of the total proteolytic activity measured in vitro. Both proteinases hydrolyzed a variety of artificial and protein substrates, and both had pH optima of 5.5 to 5.7 when either azocasein or [14C]ribulose-1,5-bisphosphate carboxylase ([14C]RuBPCase) was the substrate. The thiol endoproteinase hydrolyzed azocasein linearly but hydrolyzed [14C]RuBPCase biphasically. A third endoproteinase (EP3), not detected by standard proteolytic assays, was observed when [14C]RuBPCase was the substrate.  相似文献   

13.
Digestion of the native pig kidney fructose 1,6-bisphosphatase tetramer with subtilisin cleaves each of the 35,000-molecular-weight subunits to yield two major fragments: the S-subunit (Mr ca. 29,000), and the S-peptide (Mr 6,500). The following amino acid sequence has been determined for the S peptide: AcThrAspGlnAlaAlaPheAspThrAsnIle Val ThrLeuThrArgPheValMetGluGlnGlyArgLysAla ArgGlyThrGlyGlu MetThrGlnLeuLeuAsnSerLeuCysThrAlaValLys AlaIleSerThrAla z.sbnd;ValArgLysAlaGlyIleAlaHisLeuTyrGlyIleAla. Comparison of this sequence with that of the NH2-terminal 60 residues of the enzyme from rabbit liver (El-Dorry et al., 1977, Arch. Biochem. Biophys.182, 763) reveals strong homology with 52 identical positions and absolute identity in sequence from residues 26 to 60.Although subtilisin cleavage of fructose 1,6-bisphosphatase results in diminished sensitivity of the enzyme to AMP inhibition, we have found no AMP inhibition-related amino acid residues in the sequenced S-peptide. The loss of AMP sensitivity that occurs upon pyridoxal-P modification of the enzyme does not result in the modification of lysyl residues in the S-peptide. Neither photoaffinity labeling of fructose 1,6-bisphosphatase with 8-azido-AMP nor modification of the cysteinyl residue proximal to the AMP allosteric site resulted in the modification of residues located in the NH2-terminal 60-amino acid peptide.  相似文献   

14.
15.
Studies on a gram-positive hydrogen bacterium,Nocardia opaca 1 b   总被引:7,自引:0,他引:7  
Nocardia opaca strain 1 b has a NAD-dependent hydrogenase (hydrogen dehydrogenase). The enzyme has been purified from autotrophically grown cells and tested for optimal assay conditions and stability. The purification procedure involved protamine sulfate treatment, ammonium sulfate precipitation, and separation by DEAE-cellulose and Sephadex G-200 chromatography and resulted in a 63-fold increase of specific activity at a 11.7% enzyme recovery. The final specific activity was 103 μmoles H2/min·mg protein. The purified enzyme was dependent on nickel and magnesium ions at 0.5 and 5.0 mM concentrations, respectively, as well as flavin mononucleotide at a 5–10 μM concentration. Straight enzyme kinetics were achieved by preincubating the enzyme in the presence of NADH2. A high stability of the enzyme was observed in 0.1 M potassium phosphate buffer, pH 6.5, in the presence of 0.5 mM nickel and 5 mM magnesium ions under hydrogen atmosphere. Even under air the enzyme was remarkably stable, although less than under hydrogen. From double reciprocal plots of substrate saturation curves the Michaelis-Menten constants were calculated: For saturating NAD-concentration the K m was 0.063 mM H2 and for saturating hydrogen concentration the K m was 0.123 mM NAD.  相似文献   

16.
An acid phosphatase from Trichoderma harzianum was purified in a single step using a phenyl-Sepharose chromatography column. A typical procedure showed 22-fold purification with 56% yield. The purified enzyme showed as a single band on SDS-PAGE with an apparent molecular weight of 57.8 kDa. The pH optimum was 4.8 and maximum activity was obtained at 55°C. The enzyme retained 60% of its activity after incubation at 55°C for 60 min. The K m and V max values for p-nitrophenyl phosphate (p-NPP) as a substrate were 165 nM and 237 nM min?1, respectively. The enzyme was partially inhibited by inorganic phosphate and strongly inhibited by tungstate. Broad substrate specificity was observed with significant activities for p-NPP, ATP, ADP, AMP, fructose 6-phosphate, glucose 1-phosphate and phenyl phosphate.  相似文献   

17.
Aspergillus oryzae aminohydrolase free acid phosphodiesterase catalyzes nicotinamide adenine dinucleotide to deamino-NAD and ammonia. The enzyme was purified to homogeneity by a combination of acetone precipitation, anion exchange chromatography and gel filtration chromatography. The enzyme was purified 230.5 fold. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of the purified enzyme showed a single protein band of MW 94 kDa. The enzyme displayed maximum activity at pH 5 and 40 °C with NAD as substrate. The enzyme activity appeared to be stable up to 40 °C. The enzyme activity was enhanced slightly by addition of Na+ and K+, whereas inhibited strongly by addition of Ag+, Mn2+, Hg2+ and Cu2+ to the reaction mixtures. The enzyme hydrolyzes several substrates, suggesting a probable non-specific nature. The enzyme catalyzes the hydrolytic cleavage of amino group of NAD, adenosine, AMP, CMP, GMP, adenosine, cytidine and cytosine to the corresponding nucleotides, nucleosides or bases and ammonia. The substrate concentration–activity relationship is the hyperbolic type and the apparent Km and Kcat for the tested substrates were calculated.  相似文献   

18.
The Mg-nucleoside triphosphatase activity associated with the inner envelope membrane of the pea chloroplast is comprised of at least two components, a major activity that is sensitive to vanadate and sodium fluoride and a minor insensitive activity. The vanadate/fluoride sensitive activity has been partially purified (about 35-fold) from Triton X-100 solubilized membranes by DEAE-Sephadex chromatography and sucrose density gradient centrifugation. The partially purified enzyme resembles the membrane-bound activity in requiring either Mg2+ or Mn2+, having a broad specificity for nucleoside triphosphates, having a Km for ATP of 0.18 millimolar, and being inhibited by N-ethylmaleimide, but insensitive to sodium azide and dicyclohexylcarbodiimide. The partially purified enzyme obtained after sucrose gradient centrifugation has a markedly increased sensitivity to inhibition by inorganic pyrophosphate compared with the less pure enzyme. Pyrophosphate is not a substrate of either the membrane-bound or partially purified enzyme.  相似文献   

19.
Atrazine chlorohydrolase, TrzN (triazine hydrolase or atrazine chlorohydrolase 2), initiates bacterial metabolism of the herbicide atrazine by hydrolytic displacement of a chlorine substituent from the s-triazine ring. The present study describes crystal structures and reactivity of wild-type and active site mutant TrzN enzymes. The homodimer native enzyme structure, solved to 1.40 Å resolution, is a (βα)8 barrel, characteristic of members of the amidohydrolase superfamily. TrzN uniquely positions threonine 325 in place of a conserved aspartate that ligates the metal in most mononuclear amidohydrolases superfamily members. The threonine side chain oxygen atom is 3.3 Å from the zinc atom and 2.6 Å from the oxygen atom of zinc-coordinated water. Mutation of the threonine to a serine resulted in a 12-fold decrease in kcat/Km, largely due to kcat, whereas the T325D and T325E mutants had immeasurable activity. The structure and kinetics of TrzN are reminiscent of carbonic anhydrase, which uses a threonine to assist in positioning water for reaction with carbon dioxide. An isosteric substitution in the active site glutamate, E241Q, showed a large diminution in activity with ametryn, no detectable activity with atratone, and a 10-fold decrease with atrazine, when compared with wild-type TrzN. Activity with the E241Q mutant was nearly constant from pH 6.0 to 10.0, consistent with the loss of a proton-donating group. Structures for TrzN-E241Q were solved with bound ametryn and atratone to 1.93 and 1.64 Å resolution, respectively. Both structure and kinetic determinations suggest that the Glu241 side chain provides a proton to N-1 of the s-triazine substrate to facilitate nucleophilic displacement at the adjacent C-2.  相似文献   

20.
An enzyme has been isolated from blue-green algae Oscillatoria sp. which utilizes the product, 13-hydroperoxy-9, 11-octadecadienoic acid (13-HPOD), of lipoxygenase for its substrate. This enzyme, termed hydroperoxide lyase, converts the conjugated diene 13-hydroperoxide of linoleic acid to 13-oxotrideca-9, 11-dienoic acid. The structure of the latter has been determined by ultraviolet spectroscopy and mass spectrometry. 9-HPOD is not a substrate for this enzyme. The hydroperoxide lyase from Oscillatoria sp. has a maximum of activity at pH 6.4 and 30°C. The molecular weight of the enzyme was estimated at 56,000. The enzyme was not inhibited by BW 755C, but was inhibited by molecules containing more than one hydroxyl group. Quercetin was found to be the best inhibitor of the enzyme activity. The purified hydroperoxide lyase from Oscillatoria sp. showed an apparent Km of 7.4 micromolar and a Vmax of 35 nanomoles per minute per milligram of protein for 13-HPOD. An enzymatic pathway for the biogenesis of oxodienoic acid from linoleic acid is proposed. This involves the sequential activity of lipoxygenase and hydroperoxide lyase enzymes.  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