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.
2.
Plants that contain high concentrations of the defense compounds of the phenylpropene class (eugenol, chavicol, and their derivatives) have been recognized since antiquity as important spices for human consumption (e.g. cloves) and have high economic value. Our understanding of the biosynthetic pathway that produces these compounds in the plant, however, has remained incomplete. Several lines of basil (Ocimum basilicum) produce volatile oils that contain essentially only one or two specific phenylpropene compounds. Like other members of the Lamiaceae, basil leaves possess on their surface two types of glandular trichomes, termed peltate and capitate glands. We demonstrate here that the volatile oil constituents eugenol and methylchavicol accumulate, respectively, in the peltate glands of basil lines SW (which produces essentially only eugenol) and EMX-1 (which produces essentially only methylchavicol). Assays for putative enzymes in the biosynthetic pathway leading to these phenylpropenes localized many of the corresponding enzyme activities almost exclusively to the peltate glands in leaves actively producing volatile oil. An analysis of an expressed sequence tag database from leaf peltate glands revealed that known genes for the phenylpropanoid pathway are expressed at very high levels in these structures, accounting for 13% of the total expressed sequence tags. An additional 14% of cDNAs encoded enzymes for the biosynthesis of S-adenosyl-methionine, an important substrate in the synthesis of many phenylpropenes. Thus, the peltate glands of basil appear to be highly specialized structures for the synthesis and storage of phenylpropenes, and serve as an excellent model system to study phenylpropene biosynthesis.  相似文献   

3.
ABSTRACT: BACKGROUND: Furanocoumarins are molecules with proven therapeutic properties and are produced in only a small number of medicinal plant species such as Ruta graveolens. In vivo, these molecules play a protective role against phytophageous insect attack. Furanocoumarins are members of the phenylpropanoids family, and their biosynthetic pathway is initiated from p-coumaroyl coA. The enzymes belonging to the CYP98A cytochrome P450 family have been widely described as being aromatic meta-hydroxylases of various substrates, such as p-coumaroyl ester derivatives, and are involved in the synthesis of coumarins such as scopoletin. In furanocoumarin-producing plants, these enzymes catalyze the step directly downstream of the junction with the furanocoumarin biosynthetic pathway and might indirectly impact their synthesis. RESULTS: In this work, we describe the cloning and functional characterization of the first CYP98A encoding gene isolated from R. graveolens. Using Nicotiana benthamiana as a heterologous expression system, we have demonstrated that this enzyme adds a 3-OH to p-coumaroyl ester derivatives but is more efficient to convert p-coumaroyl quinate into chlorogenic acid than to metabolize p-coumaroyl shikimate. Plants exposed to UV-B stress showed an enhanced expression level of the corresponding gene. The R. graveolens cyp98a22 open reading frame and the orthologous Arabidopsis thaliana cyp98a3 open reading frame were overexpressed in stable transgenic Ruta plants. Both plant series were analyzed for their production of scopoletin and furanocoumarin. A detailed analysis indicates that both genes enhance the production of furanocoumarins but that CYP98A22, unlike CYP98A3, doesn't affect the synthesis of scopoletin. CONCLUSIONS: The overexpression of CYP98A22 positively impacts the concentration of furanocoumarins in R. graveolens. This gene is therefore a valuable tool to engineer plants with improved therapeutical values that might also be more resistant to phytophageous insects.  相似文献   

4.
The 4- and 5-hydroxylations of phenolic compounds in plants are catalyzed by cytochrome P450 enzymes. The 3-hydroxylation step leading to the formation of caffeic acid from p-coumaric acid remained elusive, however, alternatively described as a phenol oxidase, a dioxygenase, or a P450 enzyme, with no decisive evidence for the involvement of any in the reaction in planta. In this study, we show that the gene encoding CYP98A3, which was the best possible P450 candidate for a 3-hydroxylase in the Arabidopsis genome, is highly expressed in inflorescence stems and wounded tissues. Recombinant CYP98A3 expressed in yeast did not metabolize free p-coumaric acid or its glucose or CoA esters, p-coumaraldehyde, or p-coumaryl alcohol, but very actively converted the 5-O-shikimate and 5-O-d-quinate esters of trans-p-coumaric acid into the corresponding caffeic acid conjugates. The shikimate ester was converted four times faster than the quinate derivative. Antibodies directed against recombinant CYP98A3 specifically revealed differentiating vascular tissues in stem and root. Taken together, these data show that CYP98A3 catalyzes the synthesis of chlorogenic acid and very likely also the 3-hydroxylation of lignin monomers. This hydroxylation occurs on depsides, the function of which was so far not understood, revealing an additional and unexpected level of networking in lignin biosynthesis.  相似文献   

5.
An enzyme preparation from wheat seedlings containing p-coumaroyl:CoA ligase activity was used to synthesize caffeoyl-CoA, p-coumaroyl-CoA, and feruloyl-CoA. The same enzyme preparation also contains caffeic acid-3-O-methyl transferase and caffeoyl-CoA-3-O-methyl transferase activities. The maximum activity was found in enzyme preparation from 2-day-old seedlings, where 15-20% of the hydroxy cinnamic acid could be converted into the corresponding thioester. This yield is a result of an equilibrium between the ligase and a thioesterase also present in the crude enzyme preparation. The activity of caffeic acid 3-O-methyl transferase and caffeoyl-CoA 3-O-methyl transferase enables the production of (14)C-labeled feruloyl-CoA when using S-adenosyl-l-[methyl-(14)C]-methionine as methyl donor. The produced thioesters can be purified by reverse phase HPLC using a phosphoric acid-acetonitrile gradient.  相似文献   

6.
To elucidate the genetic and biochemical regulation of elicitor-induced p-coumaraldehyde accumulation in plants, we undertook a multifaceted approach to characterize the metabolic flux through the phenylpropanoid pathway via the characterization and chemical analysis of the metabolites in the p-coumaryl, coniferyl, and sinapyl alcohol branches of this pathway. Here, we report the identification and characterization of four cinnamyl alcohol dehydrogenases (CADs) from cucumber (Cucumis sativus) with low activity toward p-coumaraldehyde yet exhibiting significant activity toward other phenylpropanoid hydroxycinnamaldehydes. As part of this analysis, we identified and characterized the activity of a hydroxycinnamoyl-coenzyme A:shikimate hydroxycinnamoyl transferase (HCT) capable of utilizing shikimate and p-coumaroyl-coenzyme A to generate p-coumaroyl shikimate. Following pectinase treatment of cucumber, we observed the rapid accumulation of p-coumaraldehyde, likely the result of low aldehyde reductase activity (i.e. alcohol dehydrogenase in the reverse reaction) of CsCAD enzymes on p-coumaraldehyde. In parallel, we noted a concomitant reduction in the activity of CsHCT. Taken together, our findings support the hypothesis that the up-regulation of the phenylpropanoid pathway upon abiotic stress greatly enhances the overall p-coumaryl alcohol branch of the pathway. The data presented here point to a role for CsHCT (as well as, presumably, p-coumarate 3-hydroxylase) as a control point in the regulation of the coniferyl and sinapyl alcohol branches of this pathway. This mechanism represents a potentially evolutionarily conserved process to efficiently and quickly respond to biotic and abiotic stresses in cucurbit plants, resulting in the rapid lignification of affected tissues.  相似文献   

7.
8.
Rosmarinic acid and chlorogenic acid are caffeic acid esters widely found in the plant kingdom and presumably accumulated as defense compounds. In a survey, more than 240 plant species have been screened for the presence of rosmarinic and chlorogenic acids. Several rosmarinic acid-containing species have been detected. The rosmarinic acid accumulation in species of the Marantaceae has not been known before. Rosmarinic acid is found in hornworts, in the fern family Blechnaceae and in species of several orders of mono- and dicotyledonous angiosperms. The biosyntheses of caffeoylshikimate, chlorogenic acid and rosmarinic acid use 4-coumaroyl-CoA from the general phenylpropanoid pathway as hydroxycinnamoyl donor. The hydroxycinnamoyl acceptor substrate comes from the shikimate pathway: shikimic acid, quinic acid and hydroxyphenyllactic acid derived from l-tyrosine. Similar steps are involved in the biosyntheses of rosmarinic, chlorogenic and caffeoylshikimic acids: the transfer of the 4-coumaroyl moiety to an acceptor molecule by a hydroxycinnamoyltransferase from the BAHD acyltransferase family and the meta-hydroxylation of the 4-coumaroyl moiety in the ester by a cytochrome P450 monooxygenase from the CYP98A family. The hydroxycinnamoyltransferases as well as the meta-hydroxylases show high sequence similarities and thus seem to be closely related. The hydroxycinnamoyltransferase and CYP98A14 from Coleus blumei (Lamiaceae) are nevertheless specific for substrates involved in RA biosynthesis showing an evolutionary diversification in phenolic ester metabolism. Our current view is that only a few enzymes had to be “invented” for rosmarinic acid biosynthesis probably on the basis of genes needed for the formation of chlorogenic and caffeoylshikimic acid while further biosynthetic steps might have been recruited from phenylpropanoid metabolism, tocopherol/plastoquinone biosynthesis and photorespiration.  相似文献   

9.
10.
Members of the Zingiberaceae such as turmeric (Curcuma longa L.) and ginger (Zingiber officinale Rosc.) accumulate at high levels in their rhizomes important pharmacologically active metabolites that appear to be derived from the phenylpropanoid pathway. In ginger, these compounds are the gingerols; in turmeric these are the curcuminoids. Despite their importance, little is known about the biosynthesis of these compounds. This investigation describes the identification of enzymes in the biosynthetic pathway leading to the production of these bioactive natural products. Assays for enzymes in the phenylpropanoid pathway identified the corresponding enzyme activities in protein crude extracts from leaf, shoot and rhizome tissues from ginger and turmeric. These enzymes included phenylalanine ammonia lyase, polyketide synthases, p-coumaroyl shikimate transferase, p-coumaroyl quinate transferase, caffeic acid O-methyltransferase, and caffeoyl-CoA O-methyltransferase, which were evaluated because of their potential roles in controlling production of certain classes of gingerols and curcuminoids. All crude extracts possessed activity for all of these enzymes, with the exception of polyketide synthases. The results of polyketide synthase assays showed detectable curcuminoid synthase activity in the extracts from turmeric with the highest activity found in extracts from leaves. However, no gingerol synthase activity could be identified. This result was explained by the identification of thioesterase activities that cleaved phenylpropanoid pathway CoA esters, and which were found to be present at high levels in all tissues, especially in ginger tissues. These activities may shunt phenylpropanoid pathway intermediates away from the production of curcuminoids and gingerols, thereby potentially playing a regulatory role in the biosynthesis of these compounds.  相似文献   

11.
Auxotrophic mutants of Escherichia coli W or K12 blocked before shikimic acid in the aromatic biosynthetic pathway grew poorly on shikimic acid as sole aromatic supplement. This poor growth response was correlated with a relatively poor ability to transport shikimic acid. If citrate was present in the growth medium (as it is in some commonly used basal media) the growth of some of the E. coli K12 mutants on shikimate was further reduced. Mutants were derived from pre-shikimate auxotrophs which grew rapidly on media containing shikimic acid. These derivatives all had an increased ability to transport shikimic acid. Thus, it is proposed that the growth on shikimate observed in the parent cells is restricted by their relatively poor uptake of shikimate from the medium and that this restriction may be removed by a mutation which enhances shikimate transport. Transduction analysis of the mutations which enhanced utilization and transport of shikimic acid by E. coli K12 strains indicated at least two classes. Class 1 was about 20% cotransduced with the histidine region of the E. coli K12 chromosome and appeared to be coincident with a known shikimate transport locus, shiA. Class 2 was not cotransduced with his. The locus (or loci) of this class is unknown. Kinetic measurements suggested that both classes had shikimate uptake systems derived from the wild-type system. Two class 1 mutants had increased levels of otherwise unaltered wild-type transport while one class 2 mutant had an altered Michaelis constant (Km) for shikimate transport.  相似文献   

12.
Multiple adaptations were necessary when plants conquered the land. Among them were soluble phenylpropanoids related to plant protection and lignin necessary for upright growth and long‐distance water transport. Cytochrome P450 monooxygenase 98 (CYP98) catalyzes a rate‐limiting step in phenylpropanoid biosynthesis. Phylogenetic reconstructions suggest that a single copy of CYP98 founded each major land plant lineage (bryophytes, lycophytes, monilophytes, gymnosperms and angiosperms), and was maintained as a single copy in all lineages but the angiosperms. In angiosperms, a series of independent gene duplications and losses occurred. Biochemical assays in four angiosperm species tested showed that 4‐coumaroyl‐shikimate, a known intermediate in lignin biosynthesis, was the preferred substrate of one member in each species, while independent duplicates in Populus trichocarpa and Amborella trichopoda each showed broad substrate ranges, accepting numerous 4‐coumaroyl‐esters and ‐amines, and were thus capable of producing a wide range of hydroxycinnamoyl conjugates. The gymnosperm CYP98 from Pinus taeda showed a broad substrate range, but preferred 4‐coumaroyl‐shikimate as its best substrate. In contrast, CYP98s from the lycophyte Selaginella moellendorffii and the fern Pteris vittata converted 4‐coumaroyl‐shikimate poorly in vitro, but were able to use alternative substrates, in particular 4‐coumaroyl‐anthranilate. Thus, caffeoyl‐shikimate appears unlikely to be an intermediate in monolignol biosynthesis in non‐seed vascular plants, including ferns. The best substrate for CYP98A34 from the moss Physcomitrella patens was also 4‐coumaroyl‐anthranilate, while 4‐coumaroyl‐shikimate was converted to lower extents. Despite having in vitro activity with 4‐coumaroyl‐shikimate, CYP98A34 was unable to complement the Arabidopsis thaliana cyp98a3 loss‐of‐function phenotype, suggesting distinct properties also in vivo.  相似文献   

13.
Auxotrophic mutants of Escherichia coli W or K12 blocked before shikimic acid in the aromatic biosynthetic pathway grew poorly on shikimic acid as sole aromatic supplement. This poort growth response was correlated with a relatively poor ability to transport shikimic acid. If citrate was present in the growth medium (as it is in some commonly used basal media) the growth of some of the E. coli K12 mutants on shikimate was further reduced.Mutants were derived from pre-shikimate auxotrophs which grew rapidly on media containing shikimic acid. These derivatives all had an increased ability to transport shikimic acid. Thus, it is proposed that the growth on shikimate observed in the parent cells is restricted by their relatively poor uptake of shikimate from the medium and that this restriction may be removed by a mutation which enhances shikimate transport.Transduction analysis of the mutations which enhanced utilization and transport of shikimic acid by E. coli K12 strains indicated at least two classes. Class 1 was about 20% contransduced with the histidine region of the E. coli K12 chromosome and appeared to be coincident with a known shikimate transport locus, shiA. Class 2 was not contransduced with his. The locus (or loci) of this class is unknown. Kinetic measurements suggested that bot classes had shikimate uptake systems derived from the wild-type system. Two class 1 mutants had increased levels of otherwise unaltered wild-type transport while one class 2 mutant had an altered Michaelis constant (Km) for shikimate transport.  相似文献   

14.
15.
Plants of Ocimum basilicum L. grown under glass were exposed to short treatments with supplementary UV-B. The effect of UV-B on volatile essential oil content was analysed and compared with morphological effects on the peltate and capitate glandular trichomes. In the absence of UV-B, both peltate and capitate glands were incompletely developed in both mature and developing leaves, the oil sacs being wrinkled and only partially filled. UV-B was found to have two main effects on the glandular trichomes. During the first 4 d of treatment, both peltate and capitate glands filled and their morphology reflected their 'normal' mature development as reported in the literature. During the following days there was a large increase in the number of broken oil sacs among the peltate glands as the mature glands broke open, releasing volatiles. Neither the number of glands nor the qualitative or quantitative composition of the volatiles was affected by UV-B. There seems to be a requirement for UV-B for the filling of the glandular trichomes of basil.  相似文献   

16.
Among the large number of plant O-methyltransferases that are involved in secondary metabolism, only a few have been enzymatically characterized, and little information is available on the structure of their substrate binding site and the mechanism which determines their substrate specificity and methylation regiospecificity. We have previously reported the isolation of two O-methyltransferases, S-adenosyl-l-methionine:(iso)eugenol O-methyltransferase (IEMT) and S-adenosyl-l-methionine:caffeic acid O-methyltransferase (COMT) from Clarkia breweri, an annual plant from California. While IEMT and COMT (which methylate eugenol/isoeugenol and caffeic acid/5-hydroxyferulic acid, respectively) share 83% identity at the amino acid level, they have distinct substrate specificity and methylation regiospecificity. We report here that seven amino acids play a critical role in discriminating between eugenol/isoeugenol and caffeic acid/5-hydroxyferulic acid. When these amino acids in IEMT were replaced by the corresponding residues of COMT, the hybrid protein showed activity only with caffeic acid/5-hydroxyferulic acid. Conversely, when these amino acids in COMT were replaced by corresponding IEMT residues, the hybrid protein had activity only with eugenol/isoeugenol. These results provide strong evidence that O-methyltransferase substrate preference could be determined by a few amino acid residues and that new OMTs with different substrate specificity could begin to evolve from an existing OMT by mutation of a few amino acids. Phylogenetic analysis confirms that C. breweri IEMT evolved recently from COMT.  相似文献   

17.

In light of climate change and risks of food insecurity, it is becoming increasingly important to preserve plant germplasm in genebanks. Storage of seeds, particularly via cryopreservation, is one of the most proficient methods for ex situ plant germplasm conservation. Whilst seed cryo-banking can have little, to no, or even beneficial effects on subsequent seedling vigor in some species, it can lead to a number of plant abnormalities (morphological and physiological). This study investigated the effects of maize seed cryopreservation on seedling growth (until 14 d) and levels of selected amino acids produced in the shikimate pathway, a major link between primary and secondary metabolism. Seed cryopreservation reduced FW in recovered seedlings, reduced caffeic acid (2.5-fold decrease), and increased levels of all other shikimate pathway–related compounds assessed: phenylalanine (2.9-fold increase), tyrosine (2.6-fold increase), and shikimic (2.1-fold increase) and protocathecuic (3.1-fold increase) acids in cotyledons. Our results suggest that maize seed cryopreservation results in seedlings that exhibit signs of an ‘overly’ efficient and caffeic acid–deficient shikimate pathway, possibly related to their reduced growth during a highly vulnerable growth stage. However, these metabolic abnormalities manifested most severely in the maternal (cotyledonary), as opposed to vegetative (roots, stems, and leaves), tissues and hence are likely to disappear when the seedlings shed the cotyledons and become completely autotrophic.

  相似文献   

18.
Red clover (Trifolium pratense) leaves accumulate several μmol of phaselic acid [2-O-caffeoyl-l-malate] per gram fresh weight. Post-harvest oxidation of such o-diphenols to o-quinones by endogenous polyphenol oxidases (PPO) prevents breakdown of forage protein during storage. Forages like alfalfa (Medicago sativa) lack both foliar PPO activity and o-diphenols. Consequently, breakdown of their protein upon harvest and storage results in economic losses and release of excess nitrogen into the environment. Understanding how red clover synthesizes o-diphenols such as phaselic acid will help in the development of forages utilizing this natural system of protein protection. We have proposed biosynthetic pathways in red clover for phaselic acid that involve a specific hydroxycinnamoyl-CoA:malate hydroxycinnamoyl transferase. It is unclear whether the transfer reaction to malate to form phaselic acid involves caffeic acid or p-coumaric acid and subsequent hydroxylation of the resulting p-coumaroyl-malate. The latter would require a coumarate 3′-hydroxylase (C3′H) capable of hydroxylating p-coumaroyl-malate, an activity not previously described. Here, a cytochrome P450 C3′H (CYP98A44) was identified and its gene cloned from red clover. CYP98A44 shares 96 and 79% amino acid identity with Medicago truncatula and Arabidopsis thaliana C3′H proteins that are capable of hydroxylating p-coumaroyl-shikimate and have been implicated in monolignol biosynthesis. CYP98A44 mRNA is expressed in stems and flowers and to a lesser extent in leaves. Immune serum raised against CYP98A44 recognizes a membrane-associated protein in red clover stems and leaves and cross-reacts with C3′H proteins from other species. CYP98A44 expressed in Saccharomyces cerevisiae is capable of hydroxylating p-coumaroyl-shikimate, but not p-coumaroyl-malate. This finding indicates that in red clover, phaselic acid is likely formed by transfer of a caffeoyl moiety to malic acid, although the existence of a second C3′H capable of hydroxylating p-coumaroyl-malate cannot be definitively ruled out.  相似文献   

19.
Two enzymes thought to be involved in the biosynthesis of chlorogenic acid have been separated and purified by ion exchange chromatography and their properties studied. These two enzymes, p-coumarate CoA ligase and hydroxycinnamyl CoA: quinate hydroxycinnamyl transferase, acting together catalyse the conversion of p-coumaric acid to 5′-p-coumarylquinic acid and of caffeic acid to chlorogenic acid. The ligase has a higher affinity for p-coumaric than for caffeic acid and will in addition activate a number of other cinnamic acids such as ferulic, isoferulic and m-coumaric acids but not cinnamic acid. The transferase shows higher activity and affinity with p-coumaryl CoA than caffeyl CoA. It also acts with ferulyl CoA but only very slowly. The enzyme shows high specificity for quinic acid; shikimic acid is esterified at only 2% of the rate with quinic acid and glucose is not a substrate. The transferase activity is reversible and both chlorogenic acid and 5′-p-coumarylquinic acids are cleaved in the presence of CoA to form quinic acid and the corresponding hydroxycinnamyl CoA thioester.  相似文献   

20.
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