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.
The shikimate pathway leads to the biosynthesis of aromatic amino acids essential for protein biosynthesis and the production of a wide array of plant secondary metabolites. Among them, quinate is an astringent feeding deterrent that can be formed in a single step reaction from 3-dehydroquinate catalyzed by quinate dehydrogenase (QDH). 3-Dehydroquinate is also the substrate for shikimate biosynthesis through the sequential actions of dehydroquinate dehydratase (DQD) and shikimate dehydrogenase (SDH) contained in a single protein in plants. The reaction mechanism of QDH resembles that of SDH. The poplar genome encodes five DQD/SDH-like genes (Poptr1 to Poptr5), which have diverged into two distinct groups based on sequence analysis and protein structure prediction. In vitro biochemical assays proved that Poptr1 and -5 are true DQD/SDHs, whereas Poptr2 and -3 instead have QDH activity with only residual DQD/SDH activity. Poplar DQD/SDHs have distinct expression profiles suggesting separate roles in protein and lignin biosynthesis. Also, the QDH genes are differentially expressed. In summary, quinate (secondary metabolism) and shikimate (primary metabolism) metabolic activities are encoded by distinct members of the same gene family, each having different physiological functions.  相似文献   

2.
The shikimate pathway synthesizes aromatic amino acids essential for protein biosynthesis. Shikimate dehydrogenase (SDH) is a central enzyme of this primary metabolic pathway, producing shikimate. The structurally similar quinate is a secondary metabolite synthesized by quinate dehydrogenase (QDH). SDH and QDH belong to the same gene family, which diverged into two phylogenetic clades after a defining gene duplication just prior to the angiosperm/gymnosperm split. Non‐seed plants that diverged before this duplication harbour only a single gene of this family. Extant representatives from the chlorophytes (Chlamydomonas reinhardtii), bryophytes (Physcomitrella patens) and lycophytes (Selaginella moellendorfii) encoded almost exclusively SDH activity in vitro. A reconstructed ancestral sequence representing the node just prior to the gene duplication also encoded SDH activity. Quinate dehydrogenase activity was gained only in seed plants following gene duplication. Quinate dehydrogenases of gymnosperms, represented here by Pinus taeda, may be reminiscent of an evolutionary intermediate since they encode equal SDH and QDH activities. The second copy in P. taeda maintained specificity for shikimate similar to the activity found in the angiosperm SDH sister clade. The codon for a tyrosine residue within the active site displayed a signature of positive selection at the node defining the QDH clade, where it changed to a glycine. Replacing the tyrosine with a glycine in a highly shikimate‐specific angiosperm SDH was sufficient to gain some QDH function. Thus, very few mutations were necessary to facilitate the evolution of QDH genes.  相似文献   

3.
Acinetobacter calcoaceticus LMD 79.41 produced significant amounts of pyrrolo-quinoline quinone (PQQ) in its culture medium when grown on quinic acid or shikimic acid. Studies with LMD 79.41 and PQQ--mutants of this strain demonstrated that this organism contains an NAD(P)-independent quinate dehydrogenase (QDH) (EC 1.1.99.-), catalyzing the first degradation step of these compounds, and that the enzyme contains PQQ as a cofactor, i.e. is a quinoprotein. Synthesis of QDH was induced by protocatechuate and the enzyme appeared to be particle-bound. Acinetobacter lwoffi RAG-1 produced a quinoprotein QDH apoenzyme since growth on quinic acid only occurred in the presence of PQQ. The results obtained with the PQQ--mutants of strain LMD 79.41 also provided some insight into the regulation of PQQ biosynthesis and assemblage of quinoprotein enzymes in the periplasmic space. Since two species of Pseudomonas also contained a quinoprotein QDH, it is assumed that bacterial NAD(P)-independent quinate dehydrogenase is a quinoprotein.Abbreviations DCPIP 2,6-dichlorophenolindophenol  相似文献   

4.
The quinate dehydrogenase (QDH) from Gluconobacter oxydans IFO3244 exhibits high affinity for quinate, suggesting its application in shikimate production. Nucleotide sequence analysis of the qdh gene revealed a full-length of 2475-bp encoding an 824-amino acid protein. The qdh gene has the unusual TTG translation initiation codon. Conserved regions and a signature sequence for the quinoprotein family were observed. Phylogenetic analysis demonstrated relatedness of QDH from G. oxydans to other quinate/shikimate dehydrogenases with the highest similarity (56%) with that of Acinetobacter calcoaceticus ADP1 and lower similarity (36%) with a membrane-bound glucose dehydrogenase of Escherichia coli. The function of the gene coding for QDH was confirmed by heterologous gene expression in pyrroloquinoline quinone-synthesizing Pseudomonas putida HK5.  相似文献   

5.
Seasonal changes and vertical distribution of fine (< 2 mm diameter) and coarse (2-10 mm diameter) root mass of Pinus kesiya and fine root and rhizome mass of herbaceous species, and root production were studied in the 6-, 15- and 23-year old Pinus kesiya forest stands at Shillong, in the Meghalaya state of north-east India. Maximum fine and coarse root mass of P. kesiya, and fine root and rhizome mass of the ground vegetation were recorded during the rainy season. The contribution of the tree fine roots in 0-10 cm soil layer declined from 51% in the 6-year old stand to about 33% in the older stands. The major proportion (63-88%) of herbaceous fine root and rhizome mass was concentrated in this soil layer in all the three stands. The majority (36-57%) of tree coarse roots were present in the 10-20 cm layer in all the stands. The biomass and necromass values in the case of fine roots were more or less equal in a given stand, but the coarse roots had 5 to 9 times more live than the dead mass. The proportion of herbaceous fine root mass to the total fine root mass declined from 54% in the 6-year old stand to 30-32% in the 15- and 23-year old stands. The mean total fine root mass (pine + herbaceous species) decreased from 417 g m–2 in the 6-year old stand to 302 in 15-year and 322 g m–2 in the 23-year old stand. Annual fine root production showed a marked decrease from 1055 g m–2 in the 6-year old stand to 743 g m–2 in the 23-year old stand, but coarse root production increased from 169 g m–2 in the 6-year to 466 g m–2  in the 23-year old stand; the total root production thus remained approximately constant.  相似文献   

6.
Quinate is produced and used by many plants in the biosynthesis of chlorogenic acids (CGAs). Chlorogenic acids are astringent and serve to deter herbivory. They also function as antifungal agents and have potent antioxidant properties. Quinate is produced at a branch point of shikimate biosynthesis by the enzyme quinate dehydrogenase (QDH). However, little information exists on the identity and biochemical properties of plant QDHs. In this study, we utilized structural and bioinformatics approaches to establish a QDH‐specific primary sequence motif. Using this motif, we identified QDHs from diverse plants and confirmed their activity by recombinant protein production and kinetic assays. Through a detailed phylogenetic analysis, we show that plant QDHs arose directly from bifunctional dehydroquinate dehydratase–shikimate dehydrogenases (DHQD‐SDHs) through different convergent evolutionary events, illustrated by our findings that eudicot and conifer QDHs arose early in vascular plant evolution whereas Brassicaceae QDHs emerged later. This process of recurrent evolution of QDH is further demonstrated by the fact that this family of proteins independently evolved NAD+ and NADP+ specificity in eudicots. The acquisition of QDH activity by these proteins was accompanied by the inactivation or functional evolution of the DHQD domain, as verified by enzyme activity assays and as reflected in the loss of key DHQD active site residues. The implications of QDH activity and evolution are discussed in terms of plant growth and development.  相似文献   

7.
An NAD-dependent D-2-hydroxyacid dehydrogenase (EC 1.1.1.) was isolated and characterized from the halophilic Archaeon Haloferax mediterranei. The enzyme is a dimer with a molecular mass of 101.4 ± 3.3 kDa. It is strictly NAD-dependent and exhibits its highest activity in 4 M NaCl. The enzyme is characterized by a broad substrate specificity 2-ketoisocaproate and 2-ketobutyrate being the substrates with the higher Vmax/Km. When pyruvate and 2-ketobutyrate were the substrates the optimal pH was acidic (pH 5) meanwhile for 2-ketoisocaproate maximum activity was achieved at basic pH between 7.5 and 8.5. The optimum temperature was 52 ºC and at 65 ºC there was a pronounced activity decrease. This new enzyme can be used for the production of D-2-hydroxycarboxylic acid.  相似文献   

8.
The shikimate dehydrogenase (SDH) family consists of enzymes with diverse roles in secondary metabolism. The two most widespread members of the family, AroE and YdiB, function in amino acid biosynthesis and quinate catabolism, respectively. Here, we have determined the crystal structure of an SDH homolog belonging to the RifI class, a group of enzymes with proposed roles in antibiotic biosynthesis. The structure of RifI2 from Pseudomonas putida exhibits a number of distinctive features, including a substantial C-terminal truncation and an atypical mode of oligomerization. The active site of the enzyme contains substrate- and cofactor-binding motifs that are significantly different from those of any previously characterized member of the SDH family. These features are reflected in the novel kinetic properties of the enzyme. RifI2 exhibits much lower activity using shikimate as a substrate than AroE, and a strong preference for NAD+ instead of NADP+ as a cofactor. Moreover, the enzyme has only trace activity using quinate, unlike YdiB. Cocrystallization of RifI2 with NAD+ provided the opportunity to determine the mode of cofactor selectivity employed by the enzyme. We complemented this analysis by probing the role of a strictly conserved residue in the cofactor-binding domain, Asn193, by site directed mutagenesis. This study presents the first crystal structure and formal kinetic characterization of a new NAD+-dependent member of the SDH family.  相似文献   

9.
Glucose-6-phosphate dehydrogenase (d-glucose-6-phosphate: NADP+ l-oxidoreductase EC 1.1.1.49) isolated from Paracoccus denitrificans grown on glucose/nitrate exhibits both NAD+-and NADP+-linked activities. Both activities have a pH optimum of pH 9.6 (Glycine/NaOH buffer) and neither demonstrates a Mg2+ requirement. Kinetics for both NAD(P)+ and glucose-6-phosphate were investigated. Phosphoenolpyruvate inhibits both activities in a competitive manner with respect to glucose-6-phosphate. ATP inhibits the NAD+-linked activity competitively with respect to glucose-6-phosphate but has no effect on the NADP+-linked activity. Neither of the two activities are inhibited by 100 M NADH but both are inhibited by NADPH. The NAD+-linked activity is far more sensitive to inhibition by NADPH than the NADP+-linked activity.  相似文献   

10.
A cardenolide-hydrolysing β-D-glucosidase was isolated from young leaves of Digitalis lanata. Since this enzyme differs from the cardenolide glucohydrolase (CGH) described and characterised previously, it was termed cardenolide glucohydrolase II (CGH II). CGH II was detected in various Digitalis tissue cultures as well as in young leaves of D. lanata. The latter source was used as the starting material for the isolation and purification of CGH II. The specific enzyme activity reached about 15 pkat·mg–1 protein in buffered leaf extracts. Optimal CGH II activity was seen at around pH 6.0 and 50 °C. CGH II was purified about 600-fold by anion exchange chromatography, size exclusion chromatography and hydroxyapatite chromatography. The apparent molecular mass of CGH II was 65 kDa as determined by SDS-PAGE. CGH II exhibited a high substrate specificity towards cardenolide disaccharides, especially to those with a 1-4-β-linked glucose-digitoxose moiety such as glucoevatromonoside. The Km- and Vmax-values for this particular substrate were calculated to be 101 μM and 19.8 nkat·mg–1 protein, respectively.  相似文献   

11.
Plant homeobox genes play an important role in plant development, including embryogenesis. Recently, the function of a class I homeobox of knox 3 gene, HBK3, has been characterized in the conifer Picea abies (L.) Karst (Norway spruce) [8]. During somatic embryogenesis, expression of HBK3 is required for the proper differentiation of proembryogenic masses into somatic embryos. This transition, fundamental for the overall embryogenic process, is accelerated in sense lines over-expressing HBK3 (HBK3-S) but precluded in antisense lines (HBK3-AS) where the expression of this gene is experimentally reduced. Altered HBK3 expression resulted in major changes of ascorbate and glutathione metabolism. During the initial phases of embryogeny the level of reduced GSH was higher in the HBK3-S lines compared to their control counterpart. An opposite profile was observed for the HBK3-AS lines where the glutathione redox state, i.e. GSH/GSH + GSSG, switched towards its oxidized form, i.e. GSSG. Very similar metabolic fluctuations were also measured for ascorbate, especially during the transition of proembryogenic masses into somatic embryos (7 days into hormone-free medium). At this stage the level of reduced ascorbate (ASC) in the HBK3-AS lines was about 75% lower compare to the untransformed line causing a switch of the ascorbate redox state, i.e. ASC/ASC + DHA + AFR, towards its oxidized forms, i.e. DHA + AFR. Changes in activities of several ascorbate and glutathione redox enzymes, including dehydroascorbate reductase (EC 1.8.5.1), ascorbate free radical reductase (EC 1.6.5.4) and glutathione reductase (GR; EC 1.6.4.2) were responsible for these metabolic differences. Data presented here suggest that HBK3 expression might regulate somatic embryo yield through alterations in glutathione and ascorbate metabolism, which have been previously implicated in controlling embryo development and maturation both in vivo and in vitro.  相似文献   

12.
The effects of α-pinene, which is one of the major components of essential oils of several aromatic species, on energy metabolism of mitochondria isolated from maize (Zea mays L.) coleoptiles and primary roots were investigated. α-Pinene exerted similar effects on oxygen consumption irrespective of the source of mitochondria or of the substrate (L-malate, succinate or NADH). At concentrations lower than 250 μM, α-pinene stimulated respiration in state IV and inhibited respiration in state III. At higher concentrations the effect of α-pinene on state IV respiration was shifted toward inhibition. Complete suppression of respiratory control ratio was evident at α-pinene concentrations higher than 100 μM. When mitochondria were uncoupled with carbonyl cyanide 4-trifluoromethoxyphenyl-hydrazone (FCCP), α-pinene caused only inhibition of respiration. In the presence of α-pinene, the transmembrane potential was decreased as indicated by changes in the safranine binding by energized mitochondria. α-Pinene did not affect the activities of succinate dehydrogenase (EC 1.3.5.1) and L-malate dehydrogenase (L-malate:NAD+ oxidoreductase; EC 1.1.1.37). The results indicate that α-pinene acts by at least two mechanisms: uncoupling of oxidative phosphorylation and inhibition of electron transfer. Confirming the impairment of mitochondrial energy metabolism, α-pinene strongly inhibited mitochondrial ATP production. It is apparent that the actions of α-pinene on isolated mitochondria are consequences of unspecific disturbances in the inner mitochondrial membrane.  相似文献   

13.
5-Enolpyruvylshikimate 3-phosphate (EPSP) synthase (3-phosphoshikimate 1-carboxyvinyltransferase; EC 2.5.1.19), 3-dehydroquinate dehydratase (EC 4.2.1.10) and shikimate: NADP+ oxidoreductase (EC 1.1.1.25) were present in intact chloroplasts and root plastids isolated from pea seedling extracts by sucrose and modified-silica density gradient centrifugation. In young (approx. 10-d-old) seedling shoots the enzymes were predominantly chloroplastic; high-performance anion-exchange chromatography resolved minor isoenzymic activities not observed in density-gradientpurified chloroplasts. The initial enzyme of the shikimate pathway, 3-deoxy-d-arabino-heptulosonate 7-phosphate synthase (EC 4.1.2.15) was also associated with intact density-gradient-purified chloroplasts. 3-Dehydroquinate synthase (EC 4.6.1.3) and shikimate kinase (EC 2.7.1.71) were detected together with the other pathway enzymes in stromal preparations from washed chloroplasts. Plastidic EPSP synthase was inhibited by micromolar concentrations of the herbicide glyphosate.Abbreviations DAHP 3-deoxy-d-arabino-heptulosonate 7-phosphate - DEAE diethylaminoethyl - DHQase 3-dehydroquinate dehydratase - DTT dithiothreitol - EPSP 5-enolpyruvylshikimate 3-phosphate - SORase shikimate:NADP+ oxidoreductase  相似文献   

14.
In western Canada, oilseed rape (Brassica napus L. var. oleifera cv. Westar) is seeded during the early months of spring, when ambient temperatures are well below the optimum. This can result in poor seedling emergence. The objectives of the present study were to determine which developmental stages are sensitive to low temperature and whether the effects are thermal or developmental in nature. Seed was germinated at 22, 10 and 6 °C. Fresh weight changes and seedling growth were assessed on the basis of equal accumulated heat units, and the mobilization of storage reserves was assessed by employing antibodies against isocitrate lyase (ICL; EC 4.1.3.1), oleosin and cruciferin. Additionally, de novo protein synthesis was determined by quantifying the incorporation of methionine via in vivo labelling. Low temperature resulted in poor germination and early seedling growth with phase II of germination being most sensitive. At 10 °C, there was a temporal delay in germination that did not affect the overall success of germination. This was a thermal effect as seed at the lower temperatures required the equivalent of 16–24 degree days before germination occurred. Also, seedling growth at 10 °C was lower in comparison to seedlings grown at 22 °C. Seed at 6 °C displayed slow and incomplete germination and poor seedling growth as a result of both thermal and developmental effects.  相似文献   

15.
Natural rubber was identified for the first time in the latex of Ficus benghalensis, and the rubber biosynthetic activity in latex and rubber particles was investigated. 13C NMR analysis of samples prepared by successive extractions with acetone and benzene confirmed that the benzene-soluble residues were natural rubber, cis-1,4-polyisoprene. The rubber content in the latex of F. benghalensis was approximately 17 %. Gel permeation chromatography revealed that the molecular mass of the natural rubber from F. benghalensis was approximately 1 500 kDa. The high rubber content and large molecular size suggest that F. benghalensis is a good candidate for an alternative rubber source. Examination of latex serum from F. benghalensis by SDS-polyacrylamide gel electrophoresis revealed a small number of proteins with major proteins of 31 and 55 kDa in size. The 31-kDa protein was predominant in catalytically-active rubber particles. Determination of metal ion concentration in latex and a comparison of the effect of ethylenediamine-tetraacetic acid on in vitro rubber biosynthesis in F. benghalensis, F. carica and Hevea brasiliensis suggest that the divalent metal ion present in latex serum is an important physiological factor controlling the rubber biosynthetic activities in these plant species. Microscopic examination revealed that the rubber in F. benghalensis occurred in a series of laticifer cells located in concentric zones in the inner bark of stems and branches.  相似文献   

16.
The effect of light intensity and root nitrogen supply on the levels of leaf hypericins was examined for St. John’s wort (Hypericum perforatum L.) grown in a sand culture system with artificial lighting. Increasing the light intensity illuminating St. John’s wort plants from 106 to 402 μmol·m–2·s–1 resulted in a continuous increase in the level of leaf hypericins. Using a leaf dissection approach, the association of hypericins with the dark glands on the leaves was shown, and it was found that increasing light intensity resulted in a parallel increase in the number of dark glands. In this respect, a linear relationship was observed between leaf gland number and the level of leaf hypericins (R = 0.901). While a decrease in nitrogen supply to St. John’s wort plants also yielded an increase in the level of leaf hypericins, this response occurred in a discontinuous manner over the range of nitrogen levels tested and no significant effect upon the number of dark leaf glands was observed. Overall, these effects of increased light intensity and decreased nitrogen supply on leaf hypericins appear to be independent and additive, and may reflect differences in the sites and processes where these environmental parameters impact production of these phytochemicals.  相似文献   

17.
Clostridium difficile is the primary aetiological agent of antibiotic-associated diarrhoea. The faecal lactoferrin (FL) assay is a simple in vitro test which is highly sensitive to the presence of a marker of polymorphonuclear cells. We evaluated the use of the FL assay in conjunction with the C. difficile toxin assay in faecal samples obtained from 231 adult patients. The relationship between C. difficile toxin and FL in both negative and positive status was highly significant statistically (P < 0.001). Therefore, the FL assay performed simultaneously with the C. difficile toxin assay can help rule out asymptomatic carriage of C. difficile.  相似文献   

18.
In order to investigate the polymorphism of α-globin chain of hemoglobin amongst caprines, the linked Iα and IIα globin genes of Barbary sheep (Ammotragus lervia), goat (Capra hircus), European mouflon (Ovis aries musimon), and Cyprus mouflon (Ovis aries ophion) were completely sequenced, including the 5′ and 3′ untranslated regions. European and Cyprus mouflons, which do not show polymorphic α globin chains, had almost identical α globin genes, whereas Barbary sheep exhibit two different chains encoded by two nonallelic genes. Four different α genes were observed and sequenced in goat, validating previous observations of the existence of allelic and nonallelic polymorphism. As in other vertebrates, interchromosomal gene conversion appears to be responsible for such polymorphism. Evaluation of nucleotide sequences at the level of molecular evolution of the Iα-globin gene family in the caprine taxa suggests a closer relationship between the genus Ammotragus and Capra. Molecular clock estimates suggest sheep-mouflon, goat-aoudad, and ancestor-caprine divergences of 2.8, 5.7, and 7.1 MYBP, respectively.  相似文献   

19.
The immunologic cross-reactivity of the α and α+ forms of the large subunit and the β subunit of the (Na+ + K+)-ATPase from brain and kidney preparations was examined using rabbit antiserum prepared against the purified holo lamb kidney enzyme. As previously reported by Sweadner ((1979) J. Biol. Chem. 254, 6060–6067) phosphorylation of the large subunit of the (Na+ + K+)-ATPase in the presence of Na+, Mg2+, and [γ-32P]ATP revealed that dog and, very likely, rat brain contain two forms of the large subunit (designated α and α+) while dog, rat, and lamb kidney contain only one form (α). The cross-reactivity of the α and α+ forms in these preparations was investigated by resolving the subunits by SDS-polyacrylamide gel electrophoresis. The separated polypeptides were transferred to unmodified nitrocellulose paper, and reacted with rabbit anti-lamb kidney serum, followed by detection of the antigen-antibody complex with 125I-labeled protein A and autoradiography. By this method, the α and α+ forms of rat and dog brain, as well as the α form found in kidney, were shown to cross-react. In addition, membranes from human cerebral cortex were shown to contain two immunoreactive bands corresponding to the α and α+ forms of dog brain. In contrast, the brain of the insect Manduca sexta contains only one immunoreactive polypeptide with a molecular weight intermediate to the α and α+ forms of dog brain. The β subunit from lamb, dog and rat kidney and from dog and rat brain cross-reacts with anti-lamb kidney (Na+ + K+)-ATPase serum. The mobility of the β subunit from dog and rat brain on SDS-polyacrylamide electrophoresis gels is greater than the mobility of the β subunit from lamb, rat or dog kidney.  相似文献   

20.
Summary The requirement for Na+ and Cl in the bathing media to obtain a maximal HCO 3 secretory flux ( ) across isolated short-circuitedAmphiuma duodenum was investigated using titration techniques and ion substitution. Upon substitution of media Na+ with choline, HCO 3 secretion was markedly reduced. Replacement of media Cl produced a smaller reduction of . The presence of Cl enhanced HCO 3 secretion only if Na+ was also in the media. Elevation of media Na+ or Cl in the presence of the other ion produced a saturable increase of . In the presence of Na+, Cl stimulated when added to the mucosal but not the serosal medium. In the presence of Cl, Na+ elevated when added to the serosal but not the mucosal medium. The ability of mucosal Cl to stimulate was not apparently dependent on mucosal Na+. Simultaneous addition of 10mm Cl to the Na+-free mucosal medium and 10mm Na+ to the Cl-free serosal medium stimulated above levels produced by serosal Na+ alone. In conclusion, intestinal HCO 3 secretion required mucosal Cl and serosal Na+ and did not involve mucosal NaCl cotransport. The results are consistent with a mucosal Cl absorptive mechanism in series with parallel basolateral Na+–H+ and Cl–HCO 3 exchange mechanisms.  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