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 46 毫秒
1.
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   

2.
Various flavonoid glycosides are found in nature, and their biological activities are as variable as their number. In some cases, the sugar moiety attached to the flavonoid modulates its biological activities. Flavonoid glycones are not easily synthesized chemically. Therefore, in this study, we attempted to synthesize quercetin 3-O-glucosyl (1→2) xyloside and quercetin 3-O-glucosyl (1→6) rhamnoside (also called rutin) using two uridine diphosphate-dependent glycosyltransferases (UGTs) in Escherichia coli. To synthesize quercetin 3-O-glucosyl (1→2) xyloside, sequential glycosylation was carried out by regulating the expression time of the two UGTs. AtUGT78D2 was subcloned into a vector controlled by a Tac promoter without a lacI operator, while AtUGT79B1 was subcloned into a vector controlled by a T7 promoter. UDP-xyloside was supplied by concomitantly expressing UDP-glucose dehydrogenase (ugd) and UDP-xyloside synthase (UXS) in the E. coli. Using these strategies, 65.0 mg/L of quercetin 3-O-glucosyl (1→2) xyloside was produced. For the synthesis of rutin, one UGT (BcGT1) was integrated into the E. coli chromosome and the other UGT (Fg2) was expressed in a plasmid along with RHM2 (rhamnose synthase gene 2). After optimization of the initial cell concentration and incubation temperature, 119.8 mg/L of rutin was produced. The strategies used in this study thus show promise for the synthesis of flavonoid diglucosides in E. coli.  相似文献   

3.
Molecular complexes of triterpene glycosides such as α-hederin (hederagenin 3-O-α-L-rhamnopyranosyl-(1 → 2)-O-α-L-arabinopyranoside) and hederasaponin C (hederagenin 3-O-α-L-rhamnopyranosyl-(1 → 2)-O-α-L-arabinopyranosyl-28-O-α-L-rhamnopyranosyl-(1 → 4)-O-β-D-glucopyranosyl-(1 → 6)-O-β-D-glucopyranoside) with β-cyclodextrin were synthesized. The complex formation was studied by FTIR spectroscopy. Toxic properties of the molecular complexes were examined.  相似文献   

4.
Dioscorea plants produce pharmaceutical diosgenin, which usually exists in plants in the form of saponins and has been a starting material for the production of steroids over seven decades. The first step of steroidal saponin biosynthesis from the corresponding aglycone is glycosylation by 3-O-sterol glycosyltransferase (S3GT), transferring the glycosyl from a sugar donor to the 3-OH position of the aglycone. In this study, a DzS3GT gene from Dioscorea zingiberensis was cloned and expressed in Escherichia coli, and the recombinant DzS3GT protein showed 3-O-sterol glycosyltransferase activity in vitro. Subcellular localization analysis revealed that the DzS3GT protein is located in the cytoplasm in rice protoplasts. The tissue profiles of DzS3GT differ from those reported SGT genes. DzS3GT is expressed strongly in leaves and very weakly in stems. The diosgenin 3-O-glucoside (trillin) content is much higher in the leaves than in other organs. The specificity of gene expression and saponins accumulation suggest that the biosynthesis of trillin may occur mainly in the leaves of D. zingiberensis. This is the first report of the cloning and biochemical characterization of a glycosyltransferase gene involved in the biosynthesis of diosgenin 3-O-glucoside in Dioscorea plants. In addition, the study provides a potential relevance to the biosynthesis and transport mechanism of steroidal saponins in Dioscorea plants.  相似文献   

5.
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   

6.
A spirostane with an attached trisaccharide, (25R)-5α-spirostane-2α,3β,5α-triol 3-O-(O-α-l-rhamnopyranosyl-(1 → 2)-O-(β-d-galactopyranosyl-(1 → 3))-β-d-glucopyranoside), was isolated and identified from the aerial parts of Agapanthus africanus by activity-guided fractionation. Fungicidal properties of the crude extract, semi-purified fractions as well as the purified active saponin from A. africanus were screened in vitro against Fusarium oxysporum. At a concentration of 1 mg mL?1, the crude extract and semi-purified ethyl acetate and dichloromethane fractions showed significant antifungal activity. The purified saponin inhibited the in vitro mycelial growth of F. oxysporum completely (100 %) at a concentration of 125 µg mL?1. Furthermore, to verify previously observed induced resistance by crude extracts of A. africanus towards leaf rust, intercellular PR-protein activity was determined in wheat seedlings following foliar application of the purified saponin at 100 µg mL?1. In vitro peroxidase enzyme activity increased significantly (60 %) in wheat seedlings 48 h after treatment with the purified saponin, demonstrating its role as an elicitor to activate a defence reaction in wheat.  相似文献   

7.
Flavonoids are specialized plant secondary metabolites that are mainly present as glycoconjugates and function as attractants to pollinators and symbionts, UV protectants, allelochemicals, and have antimicrobial and antiherbivore activity for plant health. Because of the heterogeneity of UDP-glycosyltransferases (UGTs) for glycosylation in plants, their function in flavonoid glycosylation remains largely unknown in soybean and other legumes, particularly that of the UGT92 genes. Here, we identified 152 putative UGT92 genes across 48 plant species and elucidated their mode of duplication, expansion/deletion pattern, alignment, phylogenetic analysis, and genome-wide distribution. Two novel UGT-encoding genes Glyma14g04790 (UGT92G1) and Glyma15g03670 (UGT92G2) were isolated from soybean and their heterologous expression was optimized in Escherichia. coli. Both genes exhibited catalytic activity toward quercetin, kaempferol, and myricetin, with UDPglucose as the sugar donor and were characterized as flavanol-specific UGTs. High expression of both UGTs was observed in adaxial and abaxial parenchyma, suspensor cells, and adaxial and abaxial epidermis cells during seed development, suggesting that they are seed-specific flavanol glycosyltransferases in soybean. Co-expression analysis of UGT92 genes with their first and second neighborhood genes provided a basis for their network elucidation in soybean. We provide valuable information on the role of UGT92 in seed development via the glycosylation of multiple flavanols and the potential metabolic engineering of flavonoid compounds in both plants and E. coli.  相似文献   

8.
9.
Syringin, sinapyl alcohol 4-O-glucoside, is well known as a plant-derived bioactive monolignol glucoside. In Arabidopsis, recombinant chimeric protein UGT72E3/2 has been previously reported to lead to significantly higher syringin production than the parental enzymes UGT72E2 and UGT72E3. To enhance syringin content in Korean soybean (Glycine max L. ‘Kwangan’), we cloned the UGT72E3/2 gene under the control of the β-conglycinin or CaMV-35S promoter to generate β-UGT72E3/2 and 35S-UGT72E3/2 constructs, respectively, and then transformed them into soybean to obtain transgenic plants using the modified half-seed method. Real-time semi-quantitative PCR (RT-PCR) analysis showed that the UGT72E3/2 gene was expressed in the leaves of the β-UGT72E3/2 and 35S-UGT72E3/2 transgenic lines. HPLC analysis of the seeds and mature tissues of the T2 generation plants revealed that the β-UGT72E3/2 transgenic seeds accumulated 0.15 µmol/g DW of total syringin and 0.29 µmol/g DW of total coniferin, whereas coniferin and syringin were not detected in non-transgenic seeds. Moreover, coniferin and syringin also accumulated at high levels in non-seed tissues, particularly the leaves of β-UGT72E3/2 transgenic lines. In contrast, 35S-UGT72E3/2 lines showed no differences in the contents of coniferin and syringin between transgenic and non-transgenic soybean plants. Thus, the seed-specific β-conglycinin promoter might be an effective tool to apply to the nutritional enhancement of soybean crops through increased syringin production.  相似文献   

10.
Chemical investigation of the freshwater microalga Chlorella sorokiniana led to the isolation of a monogalactosyldiacylglycerol (MGDG)-rich fraction possessing dose-dependent inhibitory activity against pancreatic lipase activity. The MGDG-rich fraction contains 12 MGDGs identified by LC/HRMS analysis. Among them, three MGDGs were new compounds, namely, (2S)-1-O-(7Z,10Z-hexadecadienoyl)-2-O-(7Z,10Z,13Z-hexadecatrienoyl)-3-O-β-D-galactopyranosylglycerol (1), (2S)-1-O-linoleoyl-2-O-(7Z,10Z-hexadecadienoyl)-3-O-β-D-galactopyranosylglycerol (6), and (2S)-1-O-oleoyl-2-O-(7Z,10Z-hexadecadienoyl)-3-O-β-D-galactopyranosylglycerol (8). The major galactolipids were isolated by semipreparative HPLC and tested for their effect toward pancreatic lipase inhibitory activity. All the tested MGDGs showed significant reduction of pancreatic lipase activity indicating possible beneficial use for management of lipase-related disorders such as obesity.  相似文献   

11.
12.
Epothilone A is a derivative of 16-membered polyketide natural product, which has comparable chemotherapeutic effect like taxol. Introduction of sialic acids to these chemotherapeutic agents could generate interesting therapeutic glycoconjugates with significant effects in clinical studies. Since, most of the organisms biosynthesize sialic acids in their cell surface, they are key mediators in cellular events (cell-cell recognition, cell-matrix interactions). Interaction between such therapeutic sugar parts and cellular polysaccharides could generate interesting result in drugs like epothilone A. Based on this hypothesis, epothilone A glucoside (epothilone A 6-O-β-D-glucoside) was further decorated by conjugating enzymatically galactose followed by sialic acids to generate epothilone A 7-O-β-D-glucopyranosyl, 4′-O-α-D-galactoside i.e., lactosyl epothilone A (lac epoA) and two sialosides of epothilone A namely epothilone A 7-O-β-D-glucopyranosyl, 4′-O-α-D-galactopyranosyl 3″-O-α-N-acetyl neuraminic acid and epothilone A 7-O-β-D-glucopyranosyl, 4′-O-α-D-galactopyranosyl 6″-O-α-N-acetylneuraminic acid i.e., 3′sialyllactosyl epothilone A: 3′SL-epoA, and 6′sialyllactosyl epothilone A: 6′SL-epoA, respectively. These synthesized analogs were spectroscopically analyzed and elucidated, and biologically validated using HUVEC and HCT116 cancer cell lines.  相似文献   

13.
Two new steroid glycosides were isolated from the Far East starfish Hippasteria kurilensis collected in the Sea of Okhotsk. They were characterized as (22E,24R)-3-O-(2-O-methyl-β-D-xylopyranosyl)-24-O-[2-O-methyl-β-D-xylopyranosyl-(1→5)-α-L-arabinofuranosyl]-5α-cholest-22-ene-3β,4β,6α,7α,8,15β,24-heptaol (kurilensoside I) and (24S)-3-O-(2-O-methyl-β-D-xylopyranosyl)-24-O-(α-L-arabinofuranosyl)-5α-cholestane-3β,4β,6β,15α,24-pentaol (kurilensoside J). In addition, the earlier known glycosides linkosides F and L1, leviusculoside G, forbeside L, desulfated echinasteroside, and granulatoside A were isolated and identified. The structures of the new compounds were established with the help of two-dimentional NMR spectroscopy and mass- spectrometry.  相似文献   

14.
Herbal medicinal products have proven to be safe, economical and effective alternatives to synthetic chemical pharmaceuticals. The green tea plant (Camellia sinensis) is of profound medicinal value due to the presence of potent bioactive constituents. The purpose of the present work is to investigate saponins from green tea seeds for potential use as anti-angiogenic, antimicrobial, and hemolytic agents. Green tea seed saponins were separated into six fractions by reverse phase HPLC. The presence of three aglycone chains in the saponins of each fraction was confirmed by acid hydrolysis. Anti-angiogenic activity was evaluated using saponin fractions at concentrations in the range of 2.5 ~ 25 μg/mL. Antimicrobial activity was evaluated by thin-layer chromatography (TLC) using E. coli; S. mutans, a zoonotic Salmonella species and the fungal strain, A. niger. Saponin fractions were more potent against E. coli (gram negative bacteria) than against S. mutans (gram positive bacteria) and strongly inhibited six strains of zoonotic Salmonella. Green tea saponins also showed potent anti-angiogenic effects. All saponin fractions exhibited hemolytic activity. Our results confirm that green tea saponins have antimicrobial, anti-angiogenic, and hemolytic activities; indicating their potential as natural pharmaceutical products.  相似文献   

15.
We present a computational study on two flavonols that were recently isolated from Loranthaceae family plant extracts: kaempferol 3-O-α-L-arabinofuranosyl-(1 → 3)-α-L-rhamnoside and quercetin 3-O-α-L-arabinofuranosyl-(1 → 3)-α-L-rhamnoside. Their structures and energetics have been investigated at the density functional level of theory, up to B3LYP/6-31+G(d,p), incorporating solvent effects with polarizable continuum models. In addition, their potential antioxidant activities were probed through the computation of the (i) bond dissociation enthalpies (BDEs), which are related to the hydrogen-atom transfer mechanism (HAT), and (ii) ionization potentials (IPs), which are related to the single-electron transfer mechanism (SET). The BDEs were determined in water to be 83.23 kcal/mol for kaempferol 3-O-α-L-arabinofuranosyl-(1 → 3)-α-L-rhamnoside and 77.49 kcal/mol for quercetin 3-O-α-L-arabinofuranosyl-(1 → 3)-α-L-rhamnoside. The corresponding IPs were obtained for both compounds as 133.38 and 130.99 kcal/mol, respectively. The BDEs and IPs are comparable to those probed for their parental molecules kaempferol and quercetin; this is in marked contrast to previous studies where glycosylation at the 3-position increases the corresponding BDEs, and, hence, decreases subsequent antioxidant activity. The BDEs and IPs obtained suggest both compounds are promising for antioxidant activity and thus further experimental tests are encouraged.  相似文献   

16.
3-Aminopropyl glycosides of α-D-mannopyranosyl-(1→2)-α-D-mannopyranosyl-(1→2)-α-D-mannopyranosyl-(1→2)-α-D-mannopyranose, α-D-mannopyranosyl-(1→3)-α-D-mannopyranosyl-(1→2)-α-D-mannopyranosyl-(1→2)-α-D-mannopyranose, and α-D-mannopyranosyl-(1→2)-[α-D-mannopyranosyl-(1→3)]-α-D-mannopyranosyl-(1→2)-α-D-mannopyranosyl-(1→2)-α-D-mannopyranose were efficiently synthesized starting from ethyl 2-O-acetyl(benzoyl)-3,4,6-tri-O-benzyl-1-thio-α-D-mannopyranoside, ethyl 4,6-di-O-benzyl-2-O-benzoyl-1-thio-α-D-mannopyranoside, ethyl 4,6-di-O-benzyl-2,3-di-O-benzoyl-1-thio-α-D-mannopyranoside, and 2,3,4,6-tetra-O-benzoyl-α-D-mannopyranosyl bromide. The oligosaccharide chains synthesized correspond to the three structural types of side chains of mannan from Candida albicans cell wall. A conjugate of the third pentasaccharide with bovine serum albumin was prepared using the squarate method.  相似文献   

17.
Six new natural compounds were isolated from two Far Eastern starfish species, Henricia aspera and H. tumida, collected in the Sea of Okhotsk. Two new glycosylated steroid polyols were obtained from H. aspera: asperoside A and asperoside B, which were shown to be (20R,24R, 25S)-3-O-(2,3-di-O-methyl-β -D-xylopyranosyl)-24-methyl-5α-cholest-4-ene-3β, 6β,8,15α,16β,26-hexaol and (20R, 24R,25S,22E)-3-O-(2,4-di-O-methyl-β-D-xylopyranosyl)-24-methyl-5α-cholest-22-ene-3β,4β,6β,8,15α,26-hexaol, respectively. Two other glycosylated polyols, tumidoside A, with the structure elucidated as (20R, 22E)-3-O-(2,4-di-O-methyl-β -D-xylopyranosyl)-26,27-dinor-24-methyl-5α-cholest-22-ene-3β,4β,6β,8,15α,25-hexaol, and tumidoside B, whose structure was elucidated as (20R,24S)-3-O-(2,3-di-O-methyl-β-D-xylopyranosyl)-5α-cholestan-3β,4β,6β,8,15α,24-hexaol, were isolated from the two starfish species. (20R, 24S)-5α-Cholestan-3β,6β,15α,24-tetraol and (20R, 24S)-5α-cholestan-3β,6β,8,15α,24-pentaol were identified only in H. tumida. The known monoglycosides henricioside H1 and laeviuscolosides H and G were also identified in both species.  相似文献   

18.
19.
3-Aminopropyl glycoside of 3,2′-di-O-α-L-fucosyl-N-acetyllactosamine (Ley tetrasaccharide) was synthesized. The glycosyl donor, 2-O-acetyl-2,4,6-tri-O-benzoyl-α-D-galactopyranosyl bromide, was coupled with glycosyl acceptor, 1,6-anhydro-2-acetamido-2-deoxy-β-D-glucopyranose or its 3-O-acetyl derivative, to give the corresponding N-acetyllactosamine derivatives in 20 and 71% yields, respectively. The glycosyl donor was synthesized from 1,2-di-O-acetyl-3,4,6-triO-benzoyl-D-galactopyranose, which was obtained by the treatment of benzobromogalactose with sodium borohydride to yield 1,2-O-benzylidene derivative and subsequent removal of benzylidene group and acetylation. Acidic methanolysis of the disaccharide derivatives resulted in the selective removal of one or both acetyl groups to give the disaccharide acceptor bearing hydroxy groups at C3 of the glucosamine residue and C2 of the galactose residue. The introduction of fucose residues in these positions by the treatment with tetrabenzylfucopyranosyl bromide resulted in a tetrasaccharide derivative, which was converted into 3,2′-di-O-α-L-fucopuranosyl-1,6-anhydro-N-acetyllactosamine peracetate after substitution of acetyl groups for benzoyl and benzyl groups. Opening of the anhydro ring by acetolysis resulted in peracetate, which was then converted into the corresponding oxazoline derivative by two steps. Glycosydation of the oxazoline derivative with 3-trifluoroacetamidopropan-1-ol and removal of O-acetyl and N-trifluoroacetyl protective groups resulted in a free spacered Ley tetrasaccharide.  相似文献   

20.
Group A saponins are thought to be the cause of bitter and astringent tastes in processed foods of soybean (Glycine max), and the elimination of group A saponins is an important breeding objective. The group A saponins include two main Aa and Ab types, controlled by codominant alleles at the Sg-1 locus that is one of several key loci responsible for saponin biosynthesis in the subgenus Glycine soja. However, A0 mutant lacking group A saponin is a useful gene resource for soybean quality breeding. Here, eight Chinese wild soybean A0 accessions were sequenced to reveal the mutational mechanisms, and the results showed that these mutants were caused by at least three kinds of mechanisms involving four allelic variants (sg-10-b2, sg-10-b3, Sg-1b-0, and Sg-1b-01). The sg-10-b2 had two nucleotide deletions at positions +?72 and +?73 involving in the 24th and 25th amino acids. The sg-10-b3 contained a stop codon (TGA) at the 254th residue. The Sg-1b-0 and Sg-1b-01 were two novel A0-type mutants, which likely carried normal structural alleles, and nevertheless did not encode group A saponin due to unknown mutations beyond the normal coding regions. In addition, to reveal the structural features, allelic polymorphism, and mechanisms of the abiogenetic absence of group A (i.e., A0 phenotype), nucleotide sequence analysis was performed for the Sg-1 locus in wild soybean (Glycine soja). The results showed that Sg-1 alleles had a lower conservatism in the coding region; as high as 18 sequences were found in Chinese wild soybeans in addition to the Sg-1a (Aa) and Sg-1b (Ab) alleles. Sg-1a and Sg-1b alleles were characterized by eight synonymous codons and nine amino acid substitutions. Two evolutionarily transitional allelic sequences (Sg-1a7 and Sg-1b2) from Sg-1a toward Sg-1b were detected.  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