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.
The aim of this study was to present integrated mass spectrometric methods for the structural characterization and identification of flavonoid glycoconjugates. During the liquid chromatography/mass spectrometry analyses, TriVersa NanoMate chip‐based system with nanoelectrospray ionization and fraction collection was combined to a quadrupole time‐of‐flight mass spectrometer. In the extract samples prepared from green leaves of wheat plantlets, 41 flavonoid derivatives were recognized. Part of the target natural products had the full structure being characterized after the registration of mass spectra, where m/z values for protonated [M + H]+ and deprotonated molecules [M ? H]? were annotated. MS2 and pseudo‐MS3 experiments were performed for [M + H]+ or [M ? H]? and aglycone ions (Y0+/?‐type), respectively. It should be underlined that pseudo‐MS3 mass spectra were registered for aglycone product ions in the mass spectra of O‐glycosides present in the extract samples. In many cases, only tentative structural identification of aglycones was possible, mainly because of the presence of numerous C‐monoglycoside or C‐diglycoside in the samples. Acylation of the sugar moiety and/or methylation of the aglycone in the flavonoid glycosides under study was observed. The existence of isobaric and/or isomeric compounds was demonstrated in the extract studied. The collision‐induced dissociation mass spectra registered for C,O‐diglycosides and C,C‐diglycosides did not permit to draw complete structural conclusions about the compounds studied. For the investigated class of natural products, unambiguous classification of sugar moieties linked to the aglycones from the recorded mass spectra was not possible. Registration of the positive and negative ion mass spectra did not lead to the precise conclusion about the glycosylation position at C‐6 or C‐8, and O‐4′ or O‐7 atoms. It was possible, on the basis of the collected MS2 spectra, to differentiate between O‐glycosides and C‐glycosides present in the samples analyzed. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

2.
Kaempferol 3‐Oβ‐glucopyranoside, kaempferol 3‐Oβ‐galactopyranoside and higher glycosides of these two flavonoids with α‐rhamnose at C‐2 and/or C‐6 of the primary sugar were studied by negative ion electrospray ionisation and serial mass spectrometry in a three‐dimensional (3D) ion trap mass spectrometer. Kaempferol 3‐Oβ‐glucopyranoside and kaempferol 3‐Oα‐rhamnopyranosyl(1→6)‐β‐glucopyranoside could be distinguished from their respective galactose analogues by differences in the ratio of the radical aglycone ion [Y0 – H]?? to the rearrangement aglycone ion Y following MS/MS of the deprotonated molecules. Kaempferol 3‐O‐rhamnopyranosyl(1→2)‐β‐glucopyranoside and kaempferol 3‐Oα‐rhamnopyranosyl(1→2)[α‐rhamnopyranosyl(1→6)]‐β‐glucopyranoside could be distinguished from their respective galactose analogues by differences in the product ion spectra of the [(M – H) – rhamnose]? ion following serial mass spectrometry. In the triglycoside, it was deduced that this ion resulted from the loss of the rhamnose substituted at 2‐OH of the primary sugar by observing that MS/MS of deprotonated kaempferol 3‐Oβ‐glucopyranosyl(1→2)[α‐rhamnopyranosyl(1→6)]‐β‐glucopyranoside showed the loss of glucose and not rhamnose. Thus the class of sugar (hexose, deoxyhexose, pentose) at C‐2 and C‐6 of the primary sugar can be determined. These observations aid the assignment of kaempferol 3‐O‐glycosides, having glucose or galactose as the primary glycosidic sugar, in LC/MS analyses of plant extracts, and this can be done with reference to only a few standards. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

3.
Three new kaempferol glycosides, kaempferol 3‐Oβ‐D ‐glucopyranosyl‐(1→6)‐β‐D ‐galactopyranosyl‐7‐Oα‐L ‐rhamnopyranoside ( 1 ), kaempferol 3‐O‐β‐D ‐galactopyranosyl‐7‐Oβ‐D ‐glucopyranosyl‐(1→3)‐α‐L ‐rhamnopyranoside ( 2 ), and kaempferol 3‐Oβ‐D ‐glucopyranosyl‐(1→6)‐β‐D ‐galactopyranosyl‐7‐Oβ‐D ‐glucopyranosyl‐(1→3)‐α‐L ‐rhamnopyranoside ( 3 ), were isolated from the whole herbs of Cardamine leucantha, along with three known kaempferol glycosides, kaempferol 7‐Oα‐L ‐rhamnopyranoside, kaempferitrin, and kaempferol 3‐Oβ‐D ‐galactopyranosyl‐7‐Oα‐L ‐rhamnopyranoside. Their structures were elucidated on the basis of spectroscopic methods.  相似文献   

4.
Negative electrospray ionization tandem quadrupole mass spectrometry was used to study the collision-induced dissociation (CID) of the O-glycosidic bond from different commercially available flavonoid glycosides. Depending on the structure, flavonoid glycosides can undergo both a collision-induced homolytic and heterolytic cleavage of the O-glycosidic bond producing deprotonated radical aglycone ((Y(0) - H)(-*)) and aglycone (Y(0) (-)) product ions. The relative abundance of the radical aglycone to the aglycone fragment from flavonol-3-O-glycosides increased with increasing number of hydroxyl substituents in the B ring and in the order kaempferol - 相似文献   

5.
From the whole plant of Morina nepalensis var. alba Hand.‐Mazz., two new acylated flavonoid glycosides ( 1 and 2 ), together with four known flavonoid glycosides ( 3–6 ), were isolated. Their structures were determined to be quercetin 3‐O‐[2″′‐O‐(E)‐caffeoyl]‐α‐L ‐arabinopyranosyl‐(1→6)‐β‐D ‐galactopyranoside (monepalin A, 1 ), quercetin 3‐O‐[2″′‐O‐(E)‐caffeoyl]‐α‐L ‐arabinopyranosyl‐(1→6)‐β‐D ‐glucopyranoside (monepalin B, 2 ), quercetin 3‐O‐α‐L ‐arabinopyranosyl‐(1→6)‐β‐D ‐galactopyranoside (rumarin, 3 ), quercetin 3‐O‐β‐D ‐galactopyranoside ( 4 ), quercetin 3‐O‐β‐D ‐glucopyranoside ( 5 ) and apigenin 4O‐β‐D ‐glucopyranoside ( 6 ). Their structures were determined on the basis of chemical and spectroscopic evidence. Complete assignments of the 1H and 13C NMR spectra of all compounds were achieved from the 2D NMR spectra, including H–H COSY, HMQC, HMBC and 2D HMQC‐TOCSY spectra. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

6.
The fragmentation behavior of hydroquinone glycosides involving one or two sugar groups from Fraxinus sieboldiana and their analogue arbutin was investigated systematically by electrospray ionization tandem mass spectrometry in negative ion mode. The characteristic fragmentation reaction of these compounds was through the homolytic and heterolytic cleavage of the O‐glycosidic bond to produce radical aglycone ion ([Y0 ? H]??) and aglycone ion (Y0?), respectively. Unambiguous differentiation between the mono‐O‐glycoside isomers which differ in glycosylation position was achieved by comparing the relative abundance of [Y0 ? H]?? and Y0? ions with the optimized collision energy. In the fragmentation of 1, 4‐di‐O‐glycosides, only the Y0? ion was produced when the first glucosyl residue was expelled. However, both the [Y0 ? H]?? and Y0? ions were present when the second glucosyl residue was eliminated. In addition, an interesting [Y0‐2H]? ion was present in the product ion spectra of hydroquinone glycosides with methoxy group(s) substituted at C‐3 or/and C‐5 positions of the benzene ring. The results of this study can facilitate the rapid determination of hydroquinone glycosides in crude plant extracts and also reveal that the systematic investigation and optimization of collision energy play an important role in the differentiation of isomers which have subtle differences in structures. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

7.
A method coupling high‐performance liquid chromatography with hybrid ion trap time‐of‐flight mass spectrometry (TOFMS) using an electrospray ionization source was firstly used to characterize ten major pregnane glycosides including one novel compound auriculoside IV from the roots of Cynanchum auriculatum Royle ex Wight. In the MS/MS spectra, fragmentation reactions of the [M+Na]+ were recorded to provide abundant structural information on the aglycone and glycosyl moieties. Experiments using TOFMS allowed us to obtain precise elemental compositions of molecular ions and subsequent product ions with errors less than 6 ppm. The pregnane glycosides in C. auriculatum were classified into two major core groups: one is caudatin characterized by the neutral loss of one ikemamic acid molecule (128 Da) from the precursor ion, and the other is kidjoranin characterized by the neutral loss of cinnamic acid (148 Da) from the precursor ion. Meanwhile, a series of sugar‐chain fragment ions provided valuable information about the compositions of the sugar residues and the sequences of the sugar chain. Logical fragmentation pathways for pregnane glycosides have been proposed and are useful for the identification of these compounds in natural products especially when there are no reference compounds available. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
Flavonoid conjugates constitute several classes of plant phenolic secondary metabolites including many isomeric compounds differing in the hydroxylation pattern and substitution of their rings with different groups such as alkyls, acyls or sugars. These compounds occur in plant tissues mainly as glycosides and in many cases it is necessary to have reliable and detailed information concerning the structure of these natural products. Our results were obtained using leaf extracts of Arabidopsis thaliana and Lupinus angustifolius in which different glycosides of flavones, flavonols and isoflavones are present. Analysis of collision-induced dissociation (CID)/MS/MS spectra of protonated [M + H](+), sodiated [M + Na](+) or deprotonated [M - H](-) molecules recorded during HPLC runs may bring needed information in this respect. However, registration of mass spectra of [M + Na](+) ions with a good efficiency is possible only after post-column addition of a sodium acetate solution to the LC column eluate. The retention of sodium cation on the saccharidic parts of the molecule is observed after the CID fragmentation. In many cases, the location of this cation on the glycan attached to C-3 hydroxyl group of flavonol led to assignment of its structure. Additionally, the determination of the structure of the aglycone and of the sequence of the glycan part was made possible through the CID data obtained from the [M + H](+) and [M - H](-) ions. CID spectra show a different order of sugar elimination from hydroxyl groups at C-3 and C-7 in flavonol glycosides isolated from A. thaliana leaves and give sufficient information to discriminate flavonoid O-diglycosides from flavonoid di-O-glycosides.  相似文献   

9.
Four new flavonoid glycosides, 3‐O‐[α‐L ‐arabinopyranosyl‐(1→2)‐β‐D ‐galactopyranosyl]‐7‐Oβ‐D ‐glucopyranosylkaempferol ( 1 ), 3‐O‐(α‐L ‐arabinopyranosyl‐(1→2)‐{4‐O‐[(E)‐caffeoyl]‐β‐D ‐galactopyranosyl})‐7‐Oβ‐D ‐glucopyranosylquercetin ( 2 ), 3‐O‐{2‐O‐[(E)‐caffeoyl]‐α‐L ‐arabinopyranosyl‐(1→2)‐β‐D ‐galactopyranosyl}‐7‐Oβ‐D ‐glucopyranosylkaemperfol ( 3 ), and 3‐O‐{2‐O‐[(E)‐caffeoyl]‐α‐L ‐arabinopyranosyl‐(1→2)‐β‐D ‐galactopyranosyl}kaempferol ( 4 ), together with two known compounds were isolated from the aerial parts of Ranunculus chinensis Bge . The structures of the new glycosides were determined on the basis of spectroscopic analysis, including 1D‐ and 2D‐NMR, and ESI‐MS techniques, and chemical methods.  相似文献   

10.
Thirteen flavonoid glycosides ( 1‐7 , 11‐13 , 15 , 17 , and 18 ) were isolated from the EtOH extract of the leaves of Machilus philippinensis. Of these, kaempferol 3‐O‐(2‐O‐β‐D ‐apiofuranosyl)‐α‐L ‐rhamnopyranoside ( 1 ) and kaempferol 3‐O‐(2‐O‐β‐D ‐apiofuranosyl)‐a‐L ‐arabinofuranoside ( 2 ) are new natural products. By application of HPLC‐SPE‐NMR hyphenated technique, five additional flavonol glycosides were characterized ( 8‐10 , 14 , and 16 ). Their structures were elucidated based on spectroscopic analysis. Of these, quercetin 3‐O‐(6‐O‐α‐L ‐rhamnopyranosyl)‐β‐D ‐galactopyranoside ( 5 ) and kaempferol 3‐O‐α‐L ‐arabinopyranoside ( 15 ) showed moderate inhibitory activity against α‐glucosidase type IV from Bacillus stearothermophilus with the IC50 values of 19.5 and 19.0 μM, respectively.  相似文献   

11.
Only the dried stigma of the saffron, a flower deemed as the most valuable spice globally, is utilized for industrial production. Hence, there exists a growing interest in utilizing saffron floral bio-residues. The anti-hyperuricemic activity of a flavonoid extract from saffron floral bio-residues was assessed in potassium oxonate-induced hyperuricemia mice. In addition, an ultra-high performance liquid chromatography-triple quadrupole mass spectrometry method was established and validated to determine the pharmacokinetics of five main flavonoids and three phase-II metabolites in rat plasma after oral administration of the flavonoid extract for the first time. Compared with pharmacokinetic parameters of kaempferol-3-O-sophoroside, the most abundant flavonoid in the extract, and its aglycone kaempferol, we observed that coexisting compounds significantly reduced the absorption, accelerated the excretion of kaempferol-3-O-sophoroside, while significantly increasing the absorption and prolonging the residence time of kaempferol in the flavonoid extract. These results suggest the promising potential of the flavonoid extract from saffron floral bio-residues as an anti-hyperuricemic agent. Kaempferol was absorbed in plasma at high concentrations owing to the biotransformation of kaempferol glycosides in vivo.  相似文献   

12.
Kale is a member of the Brassicaceae family and has a complex profile of flavonoid glycosides. Therefore, kale is a suitable matrix to discuss in a comprehensive study the different fragmentation patterns of flavonoid glycosides. The wide variety of glycosylation and acylation patterns determines the health‐promoting effects of these glycosides. The aim of this study is to investigate the naturally occurring flavonoids in kale. A total of 71 flavonoid glycosides of quercetin, kaempferol and isorhamnetin were identified using a high‐performance liquid chromatography diode‐array detection/electrospray ionization multi‐stage mass spectrometry (HPLC‐DAD/ESI‐MSn) method. Of these 71 flavonol glycosides, 27 were non‐acylated, 30 were monoacylated and 14 were diacylated. Non‐acylated flavonol glycosides were present as mono‐, di‐, tri‐ and tetraglycosides. This is the first time that the occurrence of four different fragmentation patterns of non‐acylated flavonol triglycosides has been reported in one matrix simultaneously. In addition, 44 flavonol glycosides were acylated with p‐coumaric, caffeic, ferulic, hydroxyferulic or sinapic acid. While monoacylated glycosides existed as di‐, tri‐ and tetraglycosides, diacylated glycosides occurred as tetra‐ and pentaglycosides. To the best of our knowledge, 28 compounds in kale are reported here for the first time. These include three acylated isorhamnetin glycosides (isorhamnetin‐3‐O‐sinapoyl‐sophoroside‐7‐O‐D‐glucoside, isorhamnetin‐3‐O‐feruloyl‐sophoroside‐7‐O‐diglucoside and isorhamnetin‐3‐O‐disinapoyl‐triglucoside‐7‐O‐diglucoside) and seven non‐acylated isorhamnetin glycosides. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

13.
A comprehensive characterization of C‐glycosyl flavones in wheat germ has been conducted using multi‐stage high resolution mass spectrometry (HRMSn) in combination with a mass defect filtering (MDF) technique. MDF performed the initial search of raw data with defined C‐glycosyl flavone mass windows and mass defect windows to generate the noise‐reduced data focusing on targeted flavonoids. The high specificity of the exact mass measurement permits the unambiguous discrimination of acyl groups (nominal masses of 146, 162 and 176.) from sugar moieties (rhamnose, glucose or galactose and glucuronic acid). A total of 72 flavone C‐glycosyl derivatives, including 2 mono‐C‐glycosides, 34 di‐C‐glycosides, 15 tri‐glycosides, 14 acyl di‐C‐glycosides and 7 acyl tri‐glycosides, were characterized in wheat germ, some of which were considered to be important marker compounds for differentiation of whole grain and refined wheat products. The 7 acylated mono‐O‐glycosyl‐di‐C‐glycosyl flavones and some acylated di‐C‐glycosyl flavones are reported in wheat for the first time. The frequent occurrence of numerous isomers is a remarkable feature of wheat germ flavones. Both UV and mass spectra are needed to maximize the structure information obtained for data interpretation. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

14.
Apicultural products have been widely used in diet complements as well as in phytotherapy. Bee pollen from Echium plantagineum was analysed by high‐performance liquid chromatography/photodiode‐array detection coupled to ion trap mass spectrometry (HPLC‐PAD‐MSn) with an electrospray ionisation interface. The structures have been determined by the study of the ion mass fragmentation, which characterises the interglycosidic linkage in glycosylated flavonoids and differentiates positional isomers. Twelve non‐coloured flavonoids were characterised, being kaempferol‐3‐O‐neohesperidoside the major compound, besides others in trace amounts. These include quercetin, kaempferol and isorhamnetin glycosides, with several of them being isomers. Acetylated derivatives are also described. This is the first time that non‐coloured flavonoids are reported from this pollen, with MS fragmentation proving to be most useful in the elucidation of isomeric structures. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

15.
Sceptridium ternatum is a medicinal herb with multiple health benefits. However, its antioxidant activity and active components have not been clarified. In this study, the antioxidant capacity of S. ternatum was comprehensively investigated using multiple colorimetric methods and 1,1‐diphenyl‐2‐picrylhydrazyl–high‐performance liquid chromatography analysis. First, the phenolic content, flavonoid content, and radical scavenging ability of S. ternatum were parallelly determined using colorimetric methods performed in 96‐well microplates. The flavonoid content, rather than the phenolic content, was highly correlated with its antioxidant activity. Sceptridium ternatum was shown to be a rich source of flavonoids, with a highest flavonoid yield of 3.44 ± 0.11 mg/g. Subsequently, 1,1‐diphenyl‐2‐picrylhydrazyl–high‐performance liquid chromatography experiment and quadrupole time‐of‐flight mass spectrometry analyses were carried out for rapid screening of the individual antioxidants. A total of 14 O‐glycosyl flavonoids with quercetin or kaempferol aglycone have been characterized. Particularly, quercetin 3‐O‐rhamnoside‐7‐O‐glucoside exhibited the most potent antioxidant ability. Its half‐maximal effective concentrations for scavenging 1,1‐diphenyl‐2‐picrylhydrazyl and 2,2?‐azino‐bis (3‐ethylbenzthiazoline‐6‐sulfonic acid) radicals were 70.55 ± 2.69 and 106.90 ± 1.76 µg/mL, respectively, which were comparable with those of l ‐ascorbic acid. Our results indicated that the combined colorimetric and chromatographic methods provided a practical strategy for the discovery of bioactive compounds from natural products.  相似文献   

16.
The influence of the glycosylation site on the fragmentation behavior of 18 flavonoid glycoside standards was studied using positive and negative electrospray ionization mass spectrometry in combination with collision-induced dissociation and tandem mass spectrometry. The glycosylation position is shown to affect the relative abundance of the radical aglycone ions that can be observed in the [M-H]- collision-induced dissociation spectra. In particular, the radical aglycone ions are very abundant for deprotonated flavonol 3-O-glycosides. Collisional activation of the radical aglycone ions produced from positional isomers revealed minor differences: m,nB0- product ions are pronounced for 7-O-glycosides, whereas m,nA0- product ions are relatively more abundant for 4'-O-glycosides. In addition, the ratio between the radical aglycone and the regular aglycone ions in the [M+Na]+ high-energy collision-induced dissociation spectra gives an indication about the glycosylation site. This ion ratio allows the differentiation between flavonoid 3-O- and 7-O-glycosides or can be useful in the comparison of unknown compounds with standards. Unambiguous differentiation between O-glycosylation at the common positions of flavonoid O-glycosides, i.e. the 3-, 4'- and 7-positions, is achieved by collisional activation of sodiated molecules at high collision energy. The presence of a B-ring product ion containing the sugar residue indicates 4'-O-glycosylation, whereas the loss of the B-ring part from the aglycone product ion is characteristic of 3-O-glycosylation and the loss of the B-ring part from both the [M+Na]+ precursor ion and the aglycone product ion points to 7-O-glycosylation.  相似文献   

17.
As an aid for structure elucidation of new steviol glycosides, reversed‐phase C18 high‐performance liquid chromatography method was developed with several previously characterized diterpene glycosides, to identify known and detect novel aglycone‐C13 oligosaccharide moieties and indirectly identify C‐19 interlinkages. Elution order of several diterpene glycosides and their aglycone‐C13 oligosaccharide substituted with different sugar arrangements were also summarized. Comparison of the retention time of a product obtained after alkaline hydrolysis with the aglycone‐C‐13 portions of known compounds reported herein allowed us to deduce the exact positions of the sugars in the C‐13 oligosaccharide portion. The elution position of several steviol glycosides with an ent‐kaurene skeleton was helpful to describe an identification key. Two previously uncharacterized diterpene glycosides together with two known compounds were isolated from a commercial Stevia rebaudiana leaf extract. One was found to be 13‐[(2‐O‐β‐d ‐xylopyranosyl3‐O‐β‐d ‐glucopyranosyl‐β‐d ‐glucopyranosyl)oxy]ent‐kaur‐16‐en‐19‐oic acid‐(2‐O‐β‐d ‐glucopyranosyl‐β‐d ‐glucopyranosyl) ester (rebaudioside V), whereas the other was determined to be 13‐[(2‐O‐β‐d ‐xylopyranosyl3‐O‐β‐d ‐glucopyranosyl‐β‐d‐ glucopyranosyl)oxy]ent‐kaur‐16‐en‐19‐oic acid‐(2‐O‐α‐l ‐rhamnopyranosyl‐3‐O‐β‐d ‐glucopyranosyl‐β‐d ‐glucopyranosyl) ester (rebaudioside W). Previously reported compounds were isolated in gram quantities and identified as rebaudioside J and rebaudioside H. In addition, a C‐19 sugar‐free derivative was also prepared from rebaudioside H to afford rebaudioside H1. Chemical structures were partially determined by the high‐performance liquid chromatography method and unambiguously characterized by using one‐dimensional and two‐dimensional nuclear magnetic resonance experiments.  相似文献   

18.
Five new flavonol glycosides characterized as syringetin 3‐O‐α‐rhamnoside‐7‐O‐β‐glucoside, syringetin 3‐O‐α‐rhamnoside‐7,4′‐di‐O‐β‐glucoside, quercetin‐7‐O‐β‐galactosyl (1→3)‐β‐galactoside, myricetin 3‐O‐α‐rhamnosyl (1→4)‐β‐galactoside and myricetin 3‐O‐β‐glucosyl (1→2)‐β‐glucoside‐7‐O‐β‐glucosyl‐(1→4)‐α‐rhamnoside have been isolated from a methanolic extract of Embelia keniensis leaves. Known flavonols isolated from the same extract included myricetin, quercetin, kaempferol, myricetin 3‐O‐α‐rhamnoside, myricetin 3‐O‐β‐glucoside, quercetin 3‐O‐α‐rhamnoside, quercetin 3‐O‐β‐glucoside, quercetin 3‐O‐β‐xyloside, isorhamnetin 3‐O‐α‐rhamnoside and myricetin 3‐O‐rutinoside. Their structures were established from extensive spectroscopic and chemical studies and by comparison with authentic samples.  相似文献   

19.
Positive-ion fast atom bombardment (FAB) mass spectra of flavonol and steroid glycosides with sodium chloride added showed well known characteristic features; of the appearance of [M + Na]+ peaks, disappearance of [M + H]+ peaks and a significant decrease in the peak heights of fragment ions. Compared with the features in the FAB mass spectra of crown ethers with addition of salt, and above features suggest a complexation between Na+ and the glycosides in matrix solution. The B/E-constant linked scanning technique was used to obtain structural information of the [M + Na]+ ion of the glycosides. The B/E spectra gave the daughter-ion peaks, suggesting that coordination of Na+ with the biosides and triosides occurs at the sugar moiety, whereas the coordination with the monoglycosides occurs at the aglycone moiety, except for monoglycosides in which the aglycone moiety does not contain significant oxygen functional groups such as OH and CO.  相似文献   

20.

Lanthanum trifluoromethanesulfonate‐catalyzed solvent‐free per‐O‐acetylation with stoichiometric acetic anhydride proceeds in high yield (95%–99%) to afford exclusively pyranose products as anomeric mixtures. Subsequent anomeric substitution employing borontrifluoride etherate and thiols or alcohols furnished the corresponding 1,2‐trans‐linked thioglycosides and O‐glycosides, respectively, in good to excellent overall yield (75%–85%). Alternatively, reaction of free sugars in neat alcohol employing the same catalyst at elevated temperature gives the corresponding 1,2‐cis‐linked O‐glycosides (along with 1,2‐trans‐linked glycosides as minor product) in good yield (73%–80%). Anomeric mixtures of compounds thus produced were characterized as their per‐O‐acetylated derivatives.   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