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.
2.
3.

Key message

This study identifies six UGT73Cs all able to glucosylate sapogenins at positions 3 and/or 28 which demonstrates that B. vulgaris has a much richer arsenal of UGTs involved in saponin biosynthesis than initially anticipated.

Abstract

The wild cruciferous plant Barbarea vulgaris is resistant to some insects due to accumulation of two monodesmosidic triterpenoid saponins, oleanolic acid 3-O-β-cellobioside and hederagenin 3-O-β-cellobioside. Insect resistance depends on the structure of the sapogenin aglycone and the glycosylation pattern. The B. vulgaris saponin profile is complex with at least 49 saponin-like metabolites, derived from eight sapogenins and including up to five monosaccharide units. Two B. vulgaris UDP-glycosyltransferases, UGT73C11 and UGT73C13, O-glucosylate sapogenins at positions 3 and 28, forming mainly 3-O-β-d-glucosides. The aim of this study was to identify UGTs responsible for the diverse saponin oligoglycoside moieties observed in B. vulgaris. Twenty UGT genes from the insect resistant genotype were selected and heterologously expressed in Nicotiana benthamiana and/or Escherichia coli. The extracts were screened for their ability to glycosylate sapogenins (oleanolic acid, hederagenin), the hormone 24-epibrassinolide and sapogenin monoglucosides (hederagenin and oleanolic acid 3-O-β-d-glucosides). Six UGTs from the UGT73C subfamily were able to glucosylate both sapogenins and both monoglucosides at positions 3 and/or 28. Some UGTs formed bisdesmosidic saponins efficiently. At least four UGT73C genes were localized in a tandem array with UGT73C11 and possibly UGT73C13. This organization most likely reflects duplication events followed by sub- and neofunctionalization. Indeed, signs of positive selection on several amino acid sites were identified and modelled to be localized on the UGT protein surface. This tandem array is proposed to initiate higher order bisdesmosidic glycosylation of B. vulgaris saponins, leading to the recently discovered saponin structural diversity, however, not directly to known cellobiosidic saponins.
  相似文献   

4.
Myosin is believed to act as the molecular motor for many actin-based motility processes in eukaryotes. It is becoming apparent that a single species may possess multiple myosin isoforms, and at least seven distinct classes of myosin have been identified from studies of animals, fungi, and protozoans. The complexity of the myosin heavy-chain gene family in higher plants was investigated by isolating and characterizing myosin genomic and cDNA clones from Arabidopsis thaliana. Six myosin-like genes were identified from three polymerase chain reaction (PCR) products (PCR1, PCR11, PCR43) and three cDNA clones (ATM2, MYA2, MYA3). Sequence comparisons of the deduced head domains suggest that these myosins are members of two major classes. Analysis of the overall structure of the ATM2 and MYA2 myosins shows that they are similar to the previously-identified ATM1 and MYA1 myosins, respectively. The MYA3 appears to possess a novel tail domain, with five IQ repeats, a six-member imperfect repeat, and a segment of unique sequence. Northern blot analyses indicate that some of the Arabidopsis myosin genes are preferentially expressed in different plant organs. Combined with previous studies, these results show that the Arabidopsis genome contains at least eight myosin-like genes representing two distinct classes.  相似文献   

5.
6.
Glycyrrhizin, a sweet triterpenoid saponin found in the roots and stolons of Glycyrrhiza species (licorice), is an important active ingredient in traditional herbal medicine. We previously identified two cytochrome P450 monooxygenases, CYP88D6 and CYP72A154, that produce an aglycone of glycyrrhizin, glycyrrhetinic acid, in Glycyrrhiza uralensis. The sugar moiety of glycyrrhizin, which is composed of two glucuronic acids, makes it sweet and reduces its side‐effects. Here, we report that UDP‐glycosyltransferase (UGT) 73P12 catalyzes the second glucuronosylation as the final step of glycyrrhizin biosynthesis in Guralensis; the UGT73P12 produced glycyrrhizin by transferring a glucuronosyl moiety of UDP‐glucuronic acid to glycyrrhetinic acid 3‐O‐monoglucuronide. We also obtained a natural variant of UGT73P12 from a glycyrrhizin‐deficient (83‐555) strain of Guralensis. The natural variant showed loss of specificity for UDP‐glucuronic acid and resulted in the production of an alternative saponin, glucoglycyrrhizin. These results are consistent with the chemical phenotype of the 83‐555 strain, and suggest the contribution of UGT73P12 to glycyrrhizin biosynthesis in planta. Furthermore, we identified Arg32 as the essential residue of UGT73P12 that provides high specificity for UDP‐glucuronic acid. These results strongly suggest the existence of an electrostatic interaction between the positively charged Arg32 and the negatively charged carboxy group of UDP‐glucuronic acid. The functional arginine residue and resultant specificity for UDP‐glucuronic acid are unique to UGT73P12 in the UGT73P subfamily. Our findings demonstrate the functional specialization of UGT73P12 for glycyrrhizin biosynthesis during divergent evolution, and provide mechanistic insights into UDP‐sugar selectivity for the rational engineering of sweet triterpenoid saponins.  相似文献   

7.
From the roots of Gundelia tournefortii seven saponins have been isolated mainly by DCCC. The main saponins (A and B) were characterized, mainly by 13C and 1H NMR spectroscopy, as oleanolic acid 3-O-(2-[α-l-arabinopyranosyl(1 → 3) -β-d-gentiotriosyl(1 → 6) -β-d-glucopyranosyl]gb-d-xylopyranoside) (saponin A) and oleanolic acid 3-O-(2-[α-l-arabinopyranosyl] (1 → 3)-β-d-gentiobiosyl (1 → 6)-β-d-glucopyranosyl β-d-xylopyranoside) (saponin B). The other saponins are also derived from oleanolic acid and contain more sugar units. The saponin mixture and the saponins A and B possess strong molluscicidal activity against the schistosomiasis transmitting snail Biomphalaria glabrata.  相似文献   

8.
Betulin and oleanolic acids (pentacyclic triterpenoid secondary metabolites) have broad pharmacological activities and can be potentially used for the development of anti-cancer and anti-AIDS drugs. In this study, we detected the accumulation and the distribution characteristics of betulin and oleanolic acid in various organs of white birch at different ages. We also determined the expression of 4 OSC genes (LUS, β-AS, CAS1 and CAS2) involved in the triterpenoid synthesis pathways by real time RT-PCR. The result showed that the 1-year old birch can synthesize betulin and oleanolic acid. In addition, betulin and oleanolic acids were mainly distributed in the bark, while the content in the root skin and leaf was very low. The content of betulin and oleanolic acid in birch varied in different seasons. The content of betulin and oleanolic acid and their corresponding LUS and β-AS gene expression were very low in 1-year old birch. With increasing age of birch, betulin content was increased, while oleanolic acid was decreased. Similar changes were also observed for their corresponding synthesis genes LUS and β-AS. In the leaf of 1-year old plant, the highest expression of CAS1 and CAS2 occurred at end of September, while expression of LUS and the β-AS was low from June to October. In the stem skin,high expression of β-AS and the LUS genes occurred from the end of July to September. In the root, high expression of the β-AS gene was observed at the end of October. These results indicated that triterpenoid gene expression was similar to the triterpene accumulation. Expression of LUS gene and β-AS gene in birch with different ages were corresponding to the betulinic and oleanolic acid accumulation. Expression of CAS1 and CAS2 genes were elevated with increasing age of birch. This study provides molecular mechanisms of triterpenes synthesis in birch plants.  相似文献   

9.
Cinnamic acid 4-hydroxylase (CA4H) is the second enzyme involved in phenylpropanoid biosynthesis and is a member of the cytochrome P-450 superfamily. Three CA4H homologous genes, cyp73a, cyp73b, and cyp73c, and a cDNA clone of cyp73a were isolated from a genomic library and a cDNA library of a hybrid aspen; Populus kitakamiensis, and were characterized. They might be interrupted by two introns each. cyp73a and cyp73b were very similar to each other not only in coding regions but also in non-coding regions. Southern blot analysis showed that four homologous genes for CA4H constructed a small gene family in the diploid genome of P. kitakamiensis. In the promoter regions, there were many common m-element-like sequences in phenylpropanoid biosynthesis genes.  相似文献   

10.
Streptococcus troglodytae TKU31 was isolated from the oral cavity of a chimpanzee (Pan troglodytes) and was found to be the most closely related species of the mutans group streptococci to Streptococcus mutans. The complete sequence of TKU31 genome consists of a single circular chromosome that is 2,097,874 base pairs long and has a G + C content of 37.18%. It possesses 2082 coding sequences (CDSs), 65 tRNAs and five rRNA operons (15 rRNAs). Two clustered regularly interspaced short palindromic repeats, six insertion sequences and two predicted prophage elements were identified. The genome of TKU31 harbors some putative virulence associated genes, including gtfB, gtfC and gtfD genes encoding glucosyltransferase and gbpA, gbpB, gbpC and gbpD genes encoding glucan‐binding cell wall‐anchored protein. The deduced amino acid identity of the rhamnose‐glucose polysaccharide F gene (rgpF), which is one of the serotype determinants, is 91% identical with that of S. mutans LJ23 (serotype k) strain. However, two other virulence‐associated genes cnm and cbm, which encode the collagen‐binding proteins, were not found in the TKU31 genome. The complete genome sequence of S. troglodytae TKU31 has been deposited at DDBJ/European Nucleotide Archive/GenBank under the accession no. AP014612.  相似文献   

11.
We have applied a two-gene system based on the sequences of nuclear genes encoding multi-domain plastid acetyl-CoA carboxylase (ACCase) and plastid 3-phosphoglycerate kinase (PGK) to study grass evolution. Our analysis revealed that these genes are single-copy in most of the grass species studied, allowing the establishment of orthologous relationships between them. These relationships are consistent with the known facts of their evolution: the eukaryotic origin of the plastid ACCase, created by duplication of a gene encoding the cytosolic multi-domain ACCase gene early in grass evolution, and the prokaryotic (endosymbiont) origin of the plastid PGK. The major phylogenetic relationships among grasses deduced from the nucleotide sequence comparisons of ACCase and PGK genes are consistent with each other and with the milestones of grass evolution revealed by other methods. Nucleotide substitution rates were calculated based on multiple pairwise sequence comparisons. On a relative basis, with the divergence of the Pooideae and Panicoideae subfamilies set at 60 million years ago (MYA), events leading to the Triticum/Aegilops complex occurred at the following intervals: divergence of Lolium (Lolium rigidum) at 35 MYA, divergence of Hordeum (Hordeum vulgare) at 11 MYA and divergence of Secale (Secale cereale) at 7 MYA. On the same scale, gene duplication leading to the multi-domain plastid ACCase in grasses occurred at 129 MYA, divergence of grass and dicot plastid PGK genes at 137 MYA, and divergence of grass and dicot cytosolic PGK genes at 155 MYA. The ACCase and PGK genes provide a well-understood two-locus system to study grass phylogeny, evolution and systematics.  相似文献   

12.
Genus Pinus is a widely dispersed genus of conifer plants in the Northern Hemisphere. However, the inadequate accessibility of genomic knowledge limits our understanding of molecular phylogeny and evolution of Pinus species. In this study, the evolutionary features of complete plastid genome and the phylogeny of the Pinus genus were studied. A total of thirteen divergent hotspot regions (trnk-UUU, matK, trnQ-UUG, atpF, atpH, rpoC1, rpoC2, rpoB, ycf2, ycf1, trnD-GUC, trnY-GUA, and trnH-GUG) were identified that would be utilized as possible genetic markers for determination of phylogeny and population genetics analysis of Pinus species. Furthermore, seven genes (petD, psaI, psaM, matK, rps18, ycf1, and ycf2) with positive selection site in Pinus species were identified. Based on the whole genome this phylogenetic study showed that twenty-four Pinus species form a significant genealogical clade. Divergence time showed that the Pinus species originated about 100 million years ago (MYA) (95% HPD, 101.76.35–109.79 MYA), in lateral stages of Cretaceous. Moreover, two of the subgenera are consequently originated in 85.05 MYA (95% HPD, 81.04–88.02 MYA). This study provides a phylogenetic relationship and a chronological framework for the future study of the molecular evolution of the Pinus species.  相似文献   

13.
We identified 411 processed sequences in the Arabidopsis thaliana genome based on the fact that they have lost their intron(s) and have a length that is at least 95% of the length of the gene that gave rise to them. These sequences were generated by 230 different genes and clearly originated from retrotranspositons events because most of them (91%) have a poly(A)-tail. They are composed of 376 sequences with frame shifts and/or premature stop codons (processed pseudogenes) and 35 sequences without disablements (processed genes). Eleven of these processed genes are likely functional retrotransposed genes because they have low Ka/Ks ratios and high Ks values, and their sequences match numerous Arabidopsis ESTs. Processed sequences are mostly randomly distributed in the Arabidopsis genome and their rate of accumulation has steadily been decreasing since it peaked some 50 MYA. In contrast with the situation observed in mammals, the processed sequences found in the Arabidopsis genome originate from genes with high copy numbers and not from highly expressed genes. The patterns of spontaneous mutations in Arabidopsis are slightly different than those of mammals but are similar to those observed in Drosophila. This suggests that methylated cytosine deamination is less frequent in Arabidopsis than in mammals. Electronic Supplementary Material Electronic Supplementary material is available for this article at and accessible for authorised users. [Reviewing Editor: Dr. Juergen Brosius]  相似文献   

14.
A Pseudomonas species [Pseudomonas sp. strain amino alkanoate catabolism (AAC)] was identified that has the capacity to use 12‐aminododecanoic acid, the constituent building block of homo‐nylon‐12, as a sole nitrogen source. Growth of Pseudomonas sp. strain AAC could also be supported using a range of additional ω‐amino alkanoates. This metabolic function was shown to be most probably dependent upon one or more transaminases (TAs). Fourteen genes encoding putative TAs were identified from the genome of Pseudomonas sp. AAC. Each of the 14 genes was cloned, 11 of which were successfully expressed in Escherichia coli and tested for activity against 12‐aminododecanoic acid. In addition, physiological functions were proposed for 9 of the 14 TAs. Of the 14 proteins, activity was demonstrated in 9, and of note, 3 TAs were shown to be able to catalyse the transfer of the ω‐amine from 12‐aminododecanoic acid to pyruvate. Based on this study, three enzymes have been identified that are promising biocatalysts for the production of nylon and related polymers.  相似文献   

15.

Background  

Myosins are molecular motors that carry cargo on actin filaments in eukaryotic cells. Seventeen myosin genes have been identified in the nuclear genome of Arabidopsis. The myosin genes can be divided into two plant-specific subfamilies, class VIII with four members and class XI with 13 members. Class XI myosins are related to animal and fungal myosin class V that are responsible for movement of particular vesicles and organelles. Organelle localization of only one of the 13 Arabidopsis myosin XI (myosin XI-6; At MYA2), which is found on peroxisomes, has so far been reported. Little information is available concerning the remaining 12 class XI myosins.  相似文献   

16.
Triterpenoid saponins are bioactive metabolites that have evolved recurrently in plants, presumably for defense. Their biosynthesis is poorly understood, as is the relationship between bioactivity and structure. Barbarea vulgaris is the only crucifer known to produce saponins. Hederagenin and oleanolic acid cellobioside make some B. vulgaris plants resistant to important insect pests, while other, susceptible plants produce different saponins. Resistance could be caused by glucosylation of the sapogenins. We identified four family 1 glycosyltransferases (UGTs) that catalyze 3-O-glucosylation of the sapogenins oleanolic acid and hederagenin. Among these, UGT73C10 and UGT73C11 show highest activity, substrate specificity and regiospecificity, and are under positive selection, while UGT73C12 and UGT73C13 show lower substrate specificity and regiospecificity and are under purifying selection. The expression of UGT73C10 and UGT73C11 in different B. vulgaris organs correlates with saponin abundance. Monoglucosylated hederagenin and oleanolic acid were produced in vitro and tested for effects on P. nemorum. 3-O-β-d-Glc hederagenin strongly deterred feeding, while 3-O-β-d-Glc oleanolic acid only had a minor effect, showing that hydroxylation of C23 is important for resistance to this herbivore. The closest homolog in Arabidopsis thaliana, UGT73C5, only showed weak activity toward sapogenins. This indicates that UGT73C10 and UGT73C11 have neofunctionalized to specifically glucosylate sapogenins at the C3 position and demonstrates that C3 monoglucosylation activates resistance. As the UGTs from both the resistant and susceptible types of B. vulgaris glucosylate sapogenins and are not located in the known quantitative trait loci for resistance, the difference between the susceptible and resistant plant types is determined at an earlier stage in saponin biosynthesis.Triterpenoid saponins are a heterogeneous group of bioactive metabolites found in many species of the plant kingdom. The general conception is that saponins are involved in plant defense against antagonists such as fungi (Papadopoulou et al., 1999), mollusks (Nihei et al., 2005), and insects (Dowd et al., 2011). Saponins consist of a triterpenoid aglycone (sapogenin) linked to usually one or more sugar moieties. This combination of a hydrophobic sapogenin and hydrophilic sugars makes saponins amphiphilic and enables them to integrate into biological membrane systems. There, they form complexes with membrane sterols and reorganize the lipid bilayer, which may result in membrane damage (Augustin et al., 2011).However, our knowledge of the biosynthesis of saponins, and the genes and enzymes involved, is limited. The current conception is that the precursor 2,3-oxidosqualene is cyclized to a limited number of core structures, which are subsequently decorated with functional groups, and finally activated by adding glycosyl groups (Augustin et al., 2011). These key steps are considered to be catalyzed by three multigene families: (1) oxidosqualene cyclases (OSCs) forming the core structures, (2) cytochromes P450 adding the majority of functional groups, and (3) family 1 glycosyltransferases (UGTs) adding sugars. This allows for a vast structural complexity, some of which probably evolved by sequential gene duplication followed by functional diversification (Osbourn, 2010). A major challenge is thus to understand the processes of saponin biosynthesis, which structural variants of saponins play a role in defense against biotic antagonists, and how saponin biosynthesis evolved in different plant taxa. This knowledge is also of interest for biotechnological production and the use of saponins as protection agents against agricultural pests as well as for pharmacological and industrial uses as bactericides (De Leo et al., 2006), anticancerogens (Musende et al., 2009), and adjuvants (Sun et al., 2009).Barbarea vulgaris (winter cress) is a wild crucifer from the Cardamineae tribe of the Brassicaceae family. It is the only species in this economically important family known to produce saponins. B. vulgaris has further diverged into two separate evolutionary lineages (types; Hauser et al., 2012; Toneatto et al., 2012) that produce different saponins, glucosinolates, and flavonoids (Agerbirk et al., 2003b; Dalby-Brown et al., 2011; Kuzina et al., 2011). Saponins of the one plant type make plants resistant to the yellow-striped flea beetle (Phyllotreta nemorum), diamondback moth (Plutella xylostella), and other important crucifer specialist herbivores (Renwick, 2002); therefore, it has been suggested to utilize such plants as a trap crop to diminish insect damage (Badenes-Perez et al., 2005). The other plant type is not resistant to these herbivores. B. vulgaris, therefore, is ideal as a model species to study saponin biosynthesis, insect resistance, and its evolution, as we can contrast genes, enzymes, and their products between closely related but divergent plant types.Insect resistance of the one plant type, called G because it has glabrous leaves, correlates with the content of especially hederagenin cellobioside, oleanolic acid cellobioside, 4-epi-hederagenin cellobioside, and gypsogenin cellobioside (Shinoda et al., 2002; Agerbirk et al., 2003a; Kuzina et al., 2009; Fig. 1). These saponins are absent in the susceptible plant type, called P because it has pubescent leaves, which contains saponins of unknown structures and function (Kuzina et al., 2011). The sapogenins (aglycones) of the resistance-causing saponins hederagenin and oleanolic acid cellobioside do not deter feeding by P. nemorum, which highlights the importance of glycosylation of saponins for resistance (Nielsen et al., 2010). Therefore, the presence or absence of sapogenin glycosyltransferases could be a determining factor for the difference in resistance between the insect resistant G-type and the susceptible P-type of B. vulgaris.Open in a separate windowFigure 1.Chemical structures of the four known G-type B. vulgaris saponins that correlate with resistance to P. nemorum and other herbivores. The cellobioside and sapogenin parts of the saponin are underlined, and relevant carbon positions are numbered.Some P. nemorum genotypes are resistant to the saponin defense of B. vulgaris (Nielsen, 1997b, 1999). Resistance is coded by dominant R genes (Nielsen et al., 2010; Nielsen 2012): larvae and adults of resistant genotypes (RR or Rr) are able to feed on G-type foliage and utilize B. vulgaris as host plant (de Jong et al., 2009), whereas larvae of the susceptible genotype (rr) die and adult beetles stop feeding on G-type foliage. Larvae and adults of all known P. nemorum genotypes can feed on P-type B. vulgaris (Fig. 2).Open in a separate windowFigure 2.Feeding behavior of adult P. nemorum that are either susceptible (ST) or resistant (AK) toward the saponin-based defense of G-type B. vulgaris; the P-type produces different saponins and is not resistant against P. nemorum. Potential feeding is shown by green arrows, and termination of feeding briefly after initiation is indicated by a red dashed arrow. Larvae of the ST line die if fed on G-type plants.In this study, we asked which enzymes are involved in glucosylation of sapogenins in B. vulgaris, whether saponins with a single C3 glucosyl group are biologically active, and whether the difference between the insect resistant and susceptible types of B. vulgaris is caused by different glucosyltransferases.We report the identification of two UDP-glycosyltransferases, UGT73C10 and UGT73C11, which have high catalytic activity and substrate specificity and regiospecificity for catalyzing 3-O-glucosylation of the sapogenins oleanolic acid and hederagenin. The products, 3-O-β-d-glucopyranosyl hederagenin and 3-O-β-d-glucopyranosyl oleanolic acid, are predicted precursors of hederagenin and oleanolic acid cellobioside, respectively. The expression patterns of UGT73C10 and UGT73C11 in different organs of B. vulgaris correlate with saponin abundance, and monoglucosylated sapogenins, especially 3-O-β-d-glucopyranosyl hederagenin, deter feeding by P. nemorum. Our results thus show that glucosylation with even a single glucosyl group activates the resistance function of these sapogenins. However, since the UGTs are present and active in both the insect-resistant and -susceptible types of B. vulgaris, we cannot explain the difference in resistance by different glucosylation abilities. Instead, the difference between the susceptible and resistant types must be determined at an earlier stage in saponin biosynthesis.  相似文献   

17.
The lincomycin (LM)-production gene cluster of the overproducing strain Streptomyces iincolnensis 78-11 was cloned, analysed by hybridization, as well as by DNA sequencing, and compared with the respective genome segments of other lincomycin producers. The lmb/lmr gene cluster is composed of 27 open reading frames with putative biosynthetic or regulatory functions (lmb genes) and three resistance (lmr) genes, two of which, lmrA and lmrC, flank the cluster. A very similar overall organization of the lmb/lmr cluster seems to be conserved in four other LM producers, although the clusters are embedded in non-homologous genomic surroundings, in the wild-type strain (S. lincolnensis NRRL2936), the lmb/lmr-cluster apparently is present only in single copy. However, in the industrial strain S. lincolnensis 78-11 the non-adjacent gene clusters for the production of LM and melanin (melC) both are duplicated on a large (0.45-0.5 Mb) fragment, accompanied by deletion events. This indicates that enhanced gene dosage is one of the factors for the overproduction of LM and demonstrates that large-scale genome rearrangements can be a result of classical strain improvement by mutagenesis. Only a minority of the putative Lmb proteins belong to known protein families. These include members of the γ-glutamyl transferases (LmbA), amino acid acylases (LmbC), aromatic amino acid aminotransferases (LmbF), imidazoleglycerolphosphate dehydratases (LmbK), dTDP-glucose synthases (LmbO), dTDP-glucose 4,6-dehydratases (LmbM) and (NDP-) ketohexose (or ketocyclitol) aminotransferases (LmbS). In contrast to earlier proposals on the biosynthetic pathway of the C-8 sugar moiety (methylthiolincosaminide), this branch of the LM pathway actually seems to be based on nucleotide-activated sugars as precursors.  相似文献   

18.
3,28-Di-O-rhamnosylated oleanolic acid saponins, mimicking components of Chinese folk medicine Di Wu, have been designed and synthesized. One-pot glycosylation and ‘inverse procedure’ technologies have been applied thus significantly simplifying the preparation of desired saponins. The cytotoxic activity of compounds 3-O-[α-l-rhamnopyranosyl-(1→2)-β-d-xylopyranosyl]oleanolic acid 28-O-[α-l-rhamnopyranosyl-(1→4)-β-d-glucopyranosyl-(1→6)-β-d-glucopyranosyl] ester (3), 3-O-[α-l-rhamnopyranosyl]oleanolic acid 28-O-[α-l-rhamnopyranosyl- (1→4)-β-d-glucopyranosyl-(1→6)-β-d-glucopyranosyl] ester (4), 3-O-[α-l-rhamnopyranosyl]oleanolic acid 28-O-[α-l-rhamnopyranosyl] ester (5), and 3-O-[α-l-rhamnopyranosyl]oleanolic acid 28-O-[6-O-(α-l-rhamnopyranosyl)hexyl] ester (6) was preliminarily evaluated against HL-60 human promyelocytic leukemia cells. The natural saponin 3 and designed saponin 4 exhibited comparable moderate cytotoxic activity under our testing conditions.  相似文献   

19.
In this study, Nocardia iowensis was used to transform oleanolic acid (OA) into oleanane derivatives. The first derivative, which was found after 24 h of cultivation, was the known and already described OA methyl ester. After 1 week, two other derivatives (oleanonic acid methyl ester and an unknown metabolite) were identified as new products of a biotransformation by N. iowensis. These oleanane metabolites were characterized by HPLC, HPLC‐ESI‐MS, and HPLC‐1H NMR spectroscopy. The biotransformation was performed by suspended and immobilized cells (ICs) of N. iowensis. Cells immobilized in alginate beads were used in order to prepare a continuous process. The substrate uptake of free and ICs was similar, whereas the peak area of OA methyl ester of the ICs was only about 10% of the native cells. However, the final product (oleanonic acid methyl ester) concentrations were similar in both approaches, whereas the unknown metabolite 3 was only detected transiently in the medium of ICs. Based on these results, a new biosynthetic pathway for the biotechnological production of oleanonic acid methyl ester is proposed.  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