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 484 毫秒
1.
New syntheses of C(2′)‐deuterated ribonucleosides have been accomplished starting either from 3,5‐di‐O‐benzyl‐1‐O‐methyl‐α,β‐D ‐ribofuranose ( 1b ) or 2,3‐O‐isopropylidene‐D ‐ribose ( 14 ), with >97 atom‐% D incorporation in both cases. The former is suited to the demands of multiple‐site deuteration or uniform 13C/multiple 2H double labeling of the ribofuranose moiety, whereas the latter is particularly appropriate for single‐site 2H labeling for mechanistic studies of enzyme reactions.  相似文献   

2.
Deuterium‐labeled sugars can be utilized as powerful tools for the architectural analyses of high‐sugar‐containing molecules represented by the nucleic acids and glycoproteins, and chiral building blocks for the syntheses of new drug candidates (heavy drugs) due to their potential characteristics, such as simplifying the 1H NMR spectra and the stability of C? D bonds compared with C? H bonds. We have established a direct and efficient synthetic method of deuterated sugars from non‐labeled sugars by using the heterogeneous Ru/C‐catalyzed H–D exchange reaction in D2O under a hydrogen atmosphere with perfect chemo‐ and stereoselectivities. The direct H–D exchange reaction can selectively proceed on carbons adjacent to the free hydroxyl groups, and the deuterium labeling of various pyranosides (such as glucose and disaccharides), as well as furanosides, represented by ribose and deoxyribose was realized. Furthermore, the desired number of deuterium atoms can be freely incorporated into selected positions by the site‐selective protection of the hydroxyl groups using acetal‐type protective groups because the deuterium exchange reaction never proceeds on positions adjacent to the protected hydroxyl groups.  相似文献   

3.
The conversion of 2‐phenoxyethanol to phenol and acetate by the anaerobic bacterium Acetobacterium sp. strain LuPhet1 proceeds through acetaldehyde with concomitant migration of a H‐atom from C(1) to C(2) of the glycolic moiety. Separate feeding experiments with (R)‐ and (S)‐2‐phenoxy(1‐2H)ethanol, prepared via chemoenzymatic syntheses, indicate that the H‐atom involved in the 1,2‐shift is the pro‐S one of the enantiotopic couple of the alcohol function.  相似文献   

4.
The condensation and polycondensations of terephthaldehyde ( 1 ) and methyl D ‐hexopyranosides (gluco‐, galacto‐ and mannopyranoside) are described. Methyl α‐D ‐glucopyranoside and methyl α‐D ‐galactopyranoside react with 1 to give mono‐ 5 a and 6 a and diacetals 5 b and 6 b . Their structures were confirmed by NMR and IR spectroscopy. The polycondensation of methyl α‐D ‐mannopyranoside ( 4 ) with 1 was studied in various solvents within the temperature range of 80–140°C. Regardless of the conversion or the initial comonomer feed ratios the composition of polycondensates depended on the reaction conditions leading to the formation of materials with diverse solubilities, molecular weights and optical properties. The regioselective polycondensation of 1 and 4 was examined by the 1H NMR spectroscopy of polymer 7 . In the case of five‐membered cyclic acetal units, mixtures of the endo‐H and exo‐H dioxolan‐2‐yl system diastereomers are formed. Experimental examples of functionalization via ester units in polymer molecules 8 are described and the efficiency of the reaction routes and procedures are evaluated. The molecular weight was estimated by size‐exclusion chromatography (SEC) measurements before and after the functionalization.  相似文献   

5.
Seventeen flavonoids, five of which are flavone C‐diosides, 1 – 5 , were isolated from the BuOH‐ and AcOEt‐soluble fractions of the leaf extract of Machilus konishii. Among 1 – 5 , apigenin 6‐Cβ‐D ‐xylopyranosyl‐2″‐Oβ‐D ‐glucopyranoside ( 2 ), apigenin 8‐Cα‐L ‐arabinopyranosyl‐2″‐Oβ‐D ‐glucopyranoside ( 4 ), and apigenin 8‐Cβ‐D ‐xylopyranosyl‐2″‐Oβ‐D ‐glucopyranoside ( 5 ) are new. Both 4 and 5 are present as rotamer pairs. The structures of the new compounds were elucidated on the basis of NMR‐spectroscopic analyses and MS data. In addition, the 1H‐ and 13C‐NMR data of apigenin 6‐Cα‐L ‐arabinopyranosyl‐2″‐Oβ‐D ‐glucopyranoside ( 3 ) were assigned for the first time. The isolated compounds were assayed against α‐glucosidase (type IV from Bacillus stearothermophilus). Kaempferol 3‐O‐(2‐β‐D ‐apiofuranosyl)‐α‐L ‐rhamnopyranoside ( 12 ) was found to possess the best inhibitory activity with an IC50 value of 29.3 μM .  相似文献   

6.
Due to the possible role in human health, the number of analytical studies on lignans aimed at their quali‐ and quantitative analysis in plant extracts, biological fluids and foods is continuously increasing. However, helpful systematic mass spectrometric investigations on these compounds are few and rather limited to specific lignan sub‐classes. To increase the comprehension of the previously outlined picture of the gas‐phase properties of furofuran lignans, we extended the study to tetrahydrofuran lignans and here we reported the collision‐activated dissociation (CAD) fragmentation patterns of the alkali metal cation adducts, [M+Alk]+, and [M–H]? ions of three isomeric tetrahydrofuran lignans, (+)‐8′‐hydroxylariciresinol 4′‐O‐β‐D ‐glucopyranoside (1), (+)‐7′‐hydroxylariciresinol 7′‐O‐β‐D ‐glucopyranoside (2) and 4‐O‐β‐D ‐glucopyranosyloxy‐3,3′‐dimethoxy‐7,9′‐epoxylignan‐5′,8′,9‐triol (3) investigated by electrospray ionization triple quadrupole mass spectrometry (ESI‐TQMS). Hydrogen/deuterium (H/D) solution exchange experiments, allowing the selective H/D exchange of all the acidic hydrogen atoms, proved to be a very effective tool to obtain information on the nature of fragments generated during TQ/CAD processes. The [M+Na]+ CAD mass spectra of the three isomeric tetrahydrofurans revealed four different pathways involving the loss of the glucose moiety, which allowed the assignment of the glycosylation site. In the negative ion mode, the main fragmentation channel of the [M–H]? ions of O‐glucosylated lignans at the phenolic oxygen atoms is represented by the loss of 162 Da. When the sugar is bound to a benzylic OH group the loss of the sugar as a 180 Da unit occurs eventually following the loss of a water molecule involving both the C(9)H2OH chain and the sugar. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

7.
Two novel oligosaccharides, mono‐ and difructosyllactosucrose {[O‐β‐D ‐fructofuranosyl‐(2 → 1)]n‐β‐D ‐fructofuranosyl‐O‐[β‐D ‐galactopyranosyl‐(1 → 4)]‐α‐D ‐glucopyranoside, n = 1 and 2} were synthesized using 1F‐fructosyltransferase purified form roots of asparagus (Asparagus officinalis L.). Their 1H and 13C NMR spectra were assigned using several NMR techniques. The spectral analysis was started from two anomeric methines of aldose units, galactose and glucose, since they showed separate characteristic signals in their 1H and 13C NMR spectra. After assignments of all the 1H and 13C signals of two units of aldose, they were discriminated as galactose and glucose using proton–proton coupling constants. The HMBC spectrum revealed the galactose residue attached to C‐4 of glucose, fructose residue attached to the C‐1 of glucose, and further fructosyl fructose linkage extended from the glucosyl fructose residues. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

8.
Coenzyme B12 can assist radical enzymes that accomplish the vicinal interchange of a hydrogen atom with a functional group. It has been proposed that the Co? C bond homolysis of coenzyme B12 to cob(II)alamin and the 5′‐deoxyadenosyl radical is aided by hydrogen bonding of the corrin C19? H to the 3′‐O of the ribose moiety of the incipient 5′‐deoxyadenosyl radical, which is stabilized by 30 kJ mol?1 (B. Durbeej et al., Chem. Eur. J. 2009 , 15, 8578–8585). The diastereoisomers (R)‐ and (S)‐2,3‐dihydroxypropylcobalamin were used as models for coenzyme B12. A downfield shift of the NMR signal for the C19? H proton was observed for the (R)‐isomer (δ=4.45 versus 4.01 ppm for the (S)‐isomer) and can be ascribed to an intramolecular hydrogen bond between the C19? H and the oxygen of CHOH. Crystal structures of (R)‐ and (S)‐2,3‐dihydroxypropylcobalamin showed C19? H???O distances of 3.214(7) Å (R‐isomer) and 3.281(11) Å (S‐isomer), which suggest weak hydrogen‐bond interactions (?ΔG<6 kJ mol?1) between the CHOH of the dihydroxypropyl ligand and the C19? H. Exchange of the C19? H, which is dependent on the cobalt redox state, was investigated with cob(I)alamin, cob(II)alamin, and cob(III)alamin by using NMR spectroscopy to monitor the uptake of deuterium from deuterated water in the pH range 3–11. No exchange was found for any of the cobalt oxidation states. 3′,5′‐Dideoxyadenosylcobalamin, but not the 2′,5′‐isomer, was found to act as a coenzyme for glutamate mutase, with a 15‐fold lower kcat/KM than 5′‐deoxyadenosylcobalamin. This indicates that stabilization of the 5′‐deoxyadenosyl radical by a hydrogen bond that involves the C19? H and the 3′‐OH group of the cofactor is, at most, 7 kJ mol?1 (?ΔG). Examination of the crystal structure of glutamate mutase revealed additional stabilizing factors: hydrogen bonds between both the 2′‐OH and 3′‐OH groups and glutamate 330. The actual strength of a hydrogen bond between the C19? H and the 3′‐O of the ribose moiety of the 5′‐deoxyadenosyl group is concluded not to exceed 6 kJ mol?1 (?ΔG).  相似文献   

9.
A general synthesis of the four isomeric N7α‐D ‐, N7β‐D ‐, N9α‐D ‐, and N9β‐D ‐(purin‐2‐amine deoxynucleoside phosphoramidite) building blocks for DNA synthesis is described (Scheme). The syntheses start with methyl 3′,5′‐di‐O‐acetyl‐2′‐deoxy‐D ‐ribofuranoside ( 2 ) as the sugar component and the N2‐acetyl‐protected 6‐chloropurin‐2‐amine 1 as the base precursor. N7‐Selectivity was achieved by kinetic control, and N9‐selectivity by thermodynamic control of the nucleosidation reaction. The two N7‐(purin‐2‐amine deoxynucleosides) were introduced into the center of a decamer DNA duplex, and their pairing preferences were analyzed by UV‐melting curves. Both the N7α‐D ‐ and N7β‐D ‐(purin‐2‐amine nucleotide) units preferentially pair with a guanine base within the Watson‐Crick pairing regime, with ΔTms of −6.7 and −8.7 K, respectively, relative to a C⋅G base pair (Fig. 3 and Table 1). Molecular modeling suggests that, in the former base pair, the purinamine base is rotated into the syn‐arrangement and is able to form three H‐bonds with O(6), N(1), and NH2 of guanine, whereas in the latter base pair, both bases are in the anti‐arrangement with two H‐bonds between the N(3) and NH2 of guanine, and NH2 and N(1) of the purin‐2‐amine base (Fig. 4).  相似文献   

10.
The crystal structures of 1,2,3,4,6‐penta‐O‐acetyl‐α‐d ‐mannopyranose, C16H22O11, and 2,3,4,6‐tetra‐O‐acetyl‐α‐d ‐mannopyranosyl‐(1→2)‐3,4,6‐tri‐O‐acetyl‐α‐d ‐mannopyranosyl‐(1→3)‐1,2,4,6‐tetra‐O‐acetyl‐α‐d ‐mannopyranose, C40H54O27, were determined and compared to those of methyl 2,3,4,6‐tetra‐O‐acetyl‐α‐d ‐mannopyranoside, methyl α‐d ‐mannopyranoside and methyl α‐d ‐mannopyranosyl‐(1→2)‐α‐d ‐mannopyranoside to evaluate the effects of O‐acetylation on bond lengths, bond angles and torsion angles. In general, O‐acetylation exerts little effect on the exo‐ and endocyclic C—C and endocyclic C—O bond lengths, but the exocyclic C—O bonds involved in O‐acetylation are lengthened by ~0.02 Å. The conformation of the O‐acetyl side‐chains is highly conserved, with the carbonyl O atom either eclipsing the H atom attached to a 2°‐alcoholic C atom or bisecting the H—C—H bond angle of a 1°‐alcoholic C atom. Of the two C—O bonds that determine O‐acetyl side‐chain conformation, that involving the alcoholic C atom exhibits greater rotational variability than that involving the carbonyl C atom. These findings are in good agreement with recent solution NMR studies of O‐acetyl side‐chain conformations in saccharides. Experimental evidence was also obtained to confirm density functional theory (DFT) predictions of C—O and O—H bond‐length behavior in a C—O—H fragment involved in hydrogen bonding.  相似文献   

11.
A novel 30‐nortriterpenoid saponin, (3β)‐3‐hydroxy‐30‐noroleana‐12,20(29)‐dien‐28‐oic acid 3‐(β‐D ‐glucopyranosiduronic acid 6‐methyl ester) ( 1 ), and a known compound, (3β)‐oleanolic acid 3‐(β‐D ‐glucopyranosiduronic acid 6‐methyl ester) ( 2 ), were isolated from the aerial parts of Wedelia chinensis. The structures were established by their spectral data including 1H‐ and 13C‐NMR, 1H,1H‐COSY, HMBC, HSQC, NOESY, and HR‐FAB‐MS data.  相似文献   

12.
The syntheses, structures, and catalytic properties for lactones polymerization of eight novel yttrium complexes containing an amine‐bis(benzotriazole phenolate) ( C1NN BiBTP ) ligand are reported. A series of nitrophenolate (NP)‐type of ligands possessing R substituents with variable electronic properties (R = NO2, Cl, H, CH3) on ortho and/or para position attached to the phenolate rings have been selected and further reacted with C1NN BiBTP ‐H2 proligand and YCl3·6H2O. Two series of complexes, [Y( C1NN BiBTP )(TNP)(MeOH)2] ( 3 ), [Y( C1NN BiBTP )(2,4‐DNP)(MeOH)2] ( 4 ) and [Y( C1NN BiBTP )(2,5‐DNP)(MeOH)2] ( 5 ) with two MeOH molecules as initiators as well as [Y( C1NN BiBTP ‐H)(CNP)2] ( 6 ), [Y( C1NN BiBTP ‐H)(2‐NP)2] ( 7 ) and [Y( C1NN BiBTP ‐H)(MNP)2] ( 8 ) with two NP derivatives, were synthesized. Their ring‐opening polymerizations of L‐ lactide (L‐ LA) were investigated for all complexes in order to further understand the correlations between the inductive effect of substitutions and catalytic properties. Particularly, the activity and controllability of yttrium complexes 3 and 5 were improved dramatically comparing with the literature complex with the similar coordination environment, [Y( C1NN BiBTP )(NO3)(MeOH)2], which can be a successful example to enhance the catalytic properties by exchanging coordinate molecules. © 2019 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2019, 57, 2038–2047  相似文献   

13.
Methyl 2‐deoxy‐2‐[(1S)‐2,5‐dideoxy‐2,5‐imino‐L ‐ribitol‐1‐C‐yl)‐α‐D ‐glucopyranoside ((+)‐ 6 ) was obtained from the product of Nozaki‐Kishi coupling of 2,5‐{[(tert‐butoxy)carbonyl]imino}‐2,5‐dideoxy‐3,4‐O‐isopropylidene‐L ‐ribose ((−)‐ 9 ) and 4‐O‐benzyl‐6‐O‐[(benzyloxy)methyl]‐3‐deoxy‐2‐O‐[(trifluoromethyl)sulfonyl]‐α‐D ‐erythro‐hex‐2‐enopyranoside ((+)‐ 12 ). The alkenyl triflate (+)‐ 12 was derived from levoglucosenone ( 1 ).  相似文献   

14.
Efficient syntheses of 2‐C‐trifluoromethyl D‐ and L‐ribose and some O‐glycosides via trifluoromethylation of D‐ and L‐3,4‐O‐isopropylidene‐β‐erythro‐pentopyranosid‐2‐ulose with Ruppert's reagent are described.  相似文献   

15.
The total syntheses of tetrapeptides tubulysins D ( 1 b ), U ( 1 c ), and V ( 1 d ), which are potent tubulin polymerization inhibitors, are described. The synthesis of Tuv ( 2 ), an unusual amino acid constituent of tubulysins, includes an 1,3‐dipolar cycloaddition reaction of chiral nitrone D ‐ 6 derived from D ‐gulose with N‐acryloyl camphor sultam (?)‐ 9 employing the double asymmetric induction, whereas the synthesis of Tup ( 20 ), another unusual amino acid, involves a stereoselective Evans aldol reaction of (Z)‐boron enolate generated from (S)‐4‐isopropyl‐3‐propionyl‐2‐oxazolidinone with N‐protected phenylalaninal and a subsequent Barton deoxygenation protocol. We accomplished the total syntheses of tubulysins U ( 1 c ) and V ( 1 d ) by using these methodologies, in which the isoxazolidine ring was used as the effective protective group for γ‐amido alcohol functionality. Furthermore, to understand the structure‐activity relationship of tubulysins, we synthesized tubulysin D ( 1 b ) and cyclo‐tubulysin D ( 1 e ) from 2 ‐Me and 20 , and ent‐tubulysin D (ent‐ 1 d ) from ent‐ 2 ‐Me and ent‐ 20 , respectively. The preliminary results regarding their biological activities are also reported.  相似文献   

16.
Methanol extracts from Perezia hebeclada roots yielded the new 8‐β‐D ‐glucopyranosyloxy‐4‐methoxy‐5‐methylcoumarin ( 1 ) together with the known 4‐β‐D ‐glucopyranosyloxy‐5‐methylcoumarin ( 2 ). Their structures were determined and verified, respectively, by MS and NMR studies, including 1D and 2D experiments. Two 13C NMR signals of the sugar residue of 2 were reassigned. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

17.
D ‐Glyceraldehyde 3‐phosphate (=D ‐GAP; 2 ) was prepared by an improved chemical method (Scheme 2), and it was then employed to synthesize 1‐deoxy‐D ‐xylulose 5‐phosphate (=DXP; 3 ) which is enzymatically one of the key intermediates in the MEP ( 4 ) terpenoid biosynthetic pathway (Scheme 1). The recombinant DXP synthase of Rhodobacter capsulatus was used to catalyze the condensation of D ‐glyceraldehyde 3‐phosphate ( 2 ) and pyruvate (=2‐oxopropanoate; 1 ) to produce the sugar phosphate 3 (Scheme 2). The simple two‐step chemoenzymatic route described affords DXP ( 3 ) with more than 70% overall yield and higher than 95% purity. The procedure may also be used for the synthesis of isotope‐labeled DXP ( 3 ) by using isotope‐labeled pyruvate.  相似文献   

18.
A novel natural product indole, alkaloid, named rel‐pyricolluminol ( 1 ), was isolated from Aspidosperma pyricollum Müll .Arg . together with six known metabolites sitsirikine ( 2 ), aparicin ( 3 ), ulein ( 4 ), stemmadenine ( 5 ), lupeol ( 6 ), and (3β)‐sitoster‐3‐yl β‐D ‐glucopyranoside ( 7 ). These compounds were characterized on the basis of their spectral data, mainly 1D‐ (1H,13C‐DEPTQ) and 2D‐NMR (1H,1H‐COSY, NOESY, HSQC, and HMBC), and mass spectra (EI‐MS and HR‐ES‐MS), involving also comparison with data from the literature.  相似文献   

19.
In the context of Eschenmoser's work on pyranosyl‐RNA (‘p‐RNA’), we investigated the synthesis and base‐pairing properties of the 5‐methylisocytidine derivative. The previously determined clear‐cut restrictions of base‐pairing modes of p‐RNA had led to the expectation that a 5‐methylisocytosine β‐D ‐ribopyranosyl (= D ‐pr(MeisoC)) based (4′ → 2′)‐oligonucleotide would pair inter alia with D ‐pr(isoG) and L ‐pr(G) based oligonucleotides (D ‐pr and L ‐pr = pyranose form of D ‐ and L ‐ribose, resp.). Remarkably, we could not observe pairing with the D ‐pr(isoG) oligonucleotide but only with the L ‐pr(G) oligonucleotide. Our interpretation concludes that this – at first hand surprising – observation is caused by a change in the nucleosidic torsion angle specific for isoC.  相似文献   

20.
The syntheses and the fluorescence properties of 7H‐3,6‐dihydro‐1,2,3‐triazolo[4,5‐d]pyrimidin‐7‐one 2′‐deoxy‐β‐D ‐ribonucleosides (=2′‐deoxy‐8‐azainosine) 3 (N3), 15 (N2), and 16 (N1) as well as of 1,2,3‐benzotriazole 2′‐O‐methyl‐β‐ or ‐α‐D ‐ribofuranosides 6 (N1) and 24 (N1) are described. Also the fluorescence properties of 1,2,3‐benzotriazole 2′‐deoxy‐β‐D ‐ribofuranosides 4 (N1) and 5 (N2) are evaluated. From the nucleosides 3 – 6 , the phosphoramidites 19, 26a, 26b , and 28 are prepared and employed in solid‐phase oligonucleotide synthesis. In 12‐mer DNA duplexes, compound 3 shows similar ambiguous base‐pairing properties as 2′‐deoxyinosine ( 1 ), while the nucleosides 4 – 6 show strong pairing with each other and discriminate very little the four canonical DNA constituents.  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