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 156 毫秒
1.
On the basis of new spectroscopic evidence, structures are proposed for three amentoflavone glycosides and an apigenin di-C-glycoside previously isolated from Psilotum nudum. The major glycoside is identified as the 7,4′,4′“-tri-O-β-D-glucopyranoside, minor glycosides as the 4′,4′“-di-O-β-D-glucopyranoside and 7,4′“-di-O-β-D-glucopyranoside, and the apigenin di-C-glycoside as vicenin-2. The amentoflavone glucosides are all new natural products.  相似文献   

2.
The two purple-membrane glycolipids O-β-d-glucopyranosyl- and O-β-d-galactopyranosyl-(1→6)-O-α-d-mannopyranosyl-(1→2)-O-α-d-glucopyranosyl-(1→1)-2, 3-di-O-phytanyl-sn-glycerol were prepared by coupling O-(2,3,4-tri-O-acetyl-α-d-mannopyranosyl)-(1→2)-O-(3,4,6-tri-O-acetyl-α-d-glucopyranosyl)-(1→1)-2, 3-di-O-phytanyl-sn-glycerol (9) with 2,3,4,6-tetra-O-acetyl-α-d-glucopyranosyl bromide or 2,3,4,6-tetra-O-acetyl-α-d-mannopyranosyl bromide, respectively, followed by deacetylation. The glycolipid sulfate O-(β-d-glucopyranosyl 3-sulfate)-(1→6)-O-α-d-mannopyranosyl-(1→2)-O-α-d-glucopyranosyl-(1→1)-2,3-di-O-phytanyl-sn-glycerol was prepared by coupling of 9 with 2,4,6-tri-O-acetyl-3-O-trichloroethyloxycarbonyl-α-d-glucopyranosyl bromide in the presence of Hg(CN)2/HgBr2 followed by selective removal of the 3?-trichloroethyloxycarbonyl group, sulfation of HO-3?, and deacetylation. The suitably protected key-intermediate 9 could be prepared by two distinct approaches.  相似文献   

3.
We have investigated the phase characteristics of 1,2-bis(tricosa-10,12-diynoyl)-sn-glycero-3-phosphocholine (DC23PC), a phosphatidylcholine with diacetylenic groups in the acyl chains, and its saturated analog 1,2-ditricosanoyl-sn-glycero-3-phosphocholine (DTPC), using Fourier-transform infrared spectroscopy (FTIR). Previous studies on the phase behavior of DC23PC in H2O have shown that DC23PC exhibits: (1) formation of cylindrical structures (‘tubules’) by cooling fluid phase multilamellar vesicles (MLVs) through Tm (43° C), and 2) metastability of small unilamellar vesicles (SUVs) in the liquid-crystalline state some 40° C below Tm, with subsequent formation of a gel phase comprised of multilamellar sheets at 2° C. The sheets form tubules when heated and cooled through Tm. FTIR results presented here indicate that as metastable SUVs are cooled toward the transition to bilayer sheets, spectroscopic changes occur before the calorimetric transition as measured by a reduction in the CH2 symmetric stretch frequency and bandwidth. In spite of the vastly different morphologies, the sheet gel phase formed from SUVs is spectroscopically similar to the tubule gel phase. The C-H stretch region of DC23PC gel phase shows bands at 2937 and 2810 cm−1 not observed in the saturated analog of DC23PC, which may be related to perturbations in the acyl chains introduced by the diacetylenic moiety. The narrow CH2 scissoring mode at 1470 cm−1 and the prominent CH2 wagging progression indicate that DC23PC gel phase was highly ordered acyl chains with extended regions of all-trans methylene segments. In addition, the 13 cm−1 reduction in the C  O stretch frequency (1733–1720 cm−1) during the induction of DC23PC gel phase indicates that the interfacial region is dehydrated and rigid in the gel phase.  相似文献   

4.
The membrane-spanning segments of integral membrane proteins often are flanked by aromatic or charged amino acid residues, which may “anchor” the transmembrane orientation. Single spanning transmembrane peptides such as those of the WALP family, acetyl-GWW(LA)nLWWA-amide, furthermore adopt a moderate average tilt within lipid bilayer membranes. To understand the anchor residue dependence of the tilt, we introduce Leu-Ala “spacers” between paired anchors and in some cases replace the outer tryptophans. The resulting peptides, acetyl-GX2ALW(LA)6LWLAX22A-amide, have Trp, Lys, Arg, or Gly in the two X positions. The apparent average orientations of the core helical sequences were determined in oriented phosphatidylcholine bilayer membranes of varying thickness using solid-state 2H NMR spectroscopy. When X is Lys, Arg, or Gly, the direction of the tilt is essentially constant in different lipids and presumably is dictated by the tryptophans (Trp5 and Trp19) that flank the inner helical core. The Leu-Ala spacers are no longer helical. The magnitude of the apparent helix tilt furthermore scales nicely with the bilayer thickness except when X is Trp. When X is Trp, the direction of tilt is less well defined in each phosphatidylcholine bilayer and varies up to 70° among 1,2-dioleoyl-sn-glycero-3-phosphocholine, 1,2-dimyristoyl-sn-glycero-3-phosphocholine, and 1,2-dilauroyl-sn-glycero-3-phosphocholine bilayer membranes. Indeed, the X = Trp case parallels earlier observations in which WALP family peptides having multiple Trp anchors show little dependence of the apparent tilt magnitude on bilayer thickness. The results shed new light on the interactions of arginine, lysine, tryptophan, and even glycine at lipid bilayer membrane interfaces.  相似文献   

5.
Syntheses of 1,2-didodecanoyl-sn-glycero-3-phosphoryl-1′-(3′-O-L-lysyl)-sn-glycerol (IV) and 1,2-didodecanoyl-sn-glycero-3-phosphoryl-1′-(2′-O-L-lysyl)-sn-glycerol (VIII) as well as 1,2-didodecanoyl-sn-glycerol-3-phosphoryl-1′-sn-glycerol (XII) are described. 2′- and 3′-lysylphosphatidylglycerol are obtained as pure isomers and can be distinguished spectroscopically (infrared, 100 and 300 MHZ NMR). By these criteria a migration of the lysyl group from the 2′ to the 3′ position of the glycerol occurs in the presence of a strong acid catalyst such as HCl. On the other hand, a weak acid such as acetic acid appears ineffective in inducing lysyl migration, even at very high concentrations.Spectroscopic analysis furthermore demonstrated that lysylphosphatidylglycerol extracted from the Staphylococcus aureus membrane, is a 3′-isomer.  相似文献   

6.
Lipid composition of the membrane and rhodopsin packing density strongly modulate the early steps of the visual response of photoreceptor membranes. In this study, lipid-order and bovine rhodopsin function in proteoliposomes composed of the sn-1 chain perdeuterated lipids 14:0d27-14:1-PC, 16:0d31-16:1-PC, 18:0d35-18:1-PC, or 20:0d39-20:1-PC at rhodopsin/lipid molar ratios from 1:70 to 1:1000 (mol/mol) were investigated. Clear evidence for matching of hydrophobic regions on rhodopsin transmembrane helices and hydrophobic thickness of lipid bilayers was observed from 2H nuclear magnetic resonance order parameter measurements at low rhodopsin concentrations. Thin bilayers stretched to match the length of transmembrane helices observed as increase of sn-1 chain order, while thicker bilayers were compressed near the protein. A quantitative analysis of lipid-order parameter changes suggested that the protein adjusts its conformation to bilayer hydrophobic thickness as well, which confirmed our earlier circular-dichroism measurements. Changes in lipid order parameters upon rhodopsin incorporation vanished for bilayers with a hydrophobic thickness of 27 ± 1 Å, suggesting that this is the bilayer thickness at which rhodopsin packs in bilayers at the lowest membrane perturbation. The lipid-order parameter studies also indicated that a hydrophobic mismatch between rhodopsin and lipids triggers rhodopsin oligomerization with increasing rhodopsin concentrations. Both hydrophobic mismatch and rhodopsin oligomerization result in substantial shifts of the equilibrium between the photointermediates metarhodopsin I and metarhodopsin II; increasing bilayer thickness favors formation of metarhodopsin II while oligomerization favors metarhodopsin I. The results highlight the importance of hydrophobic matching for rhodopsin structure, oligomerization, and function.  相似文献   

7.
Treatment of 6,6′-di-O-trityl-trehalose (1) [2] with benzyl chloride in dioxane followed by acid hydrolysis and chromatography gave the chromatographically pure 2,3,4,2′,3′,4′-hexa-O-benzyl trehalose (2). Compound 2 was converted into the corresponding 6,6′-di-O-methane-sulphonyl derivative 3 in quantitative yield. Treatment of the latter compound with the potassium salts of 4-[p-(hexadecyloxy)-phenyl]butyric acid, corynomycolic acid and mycolic acid from Mycobacterium bovis afforded the corresponding benzylated-6,6′-di-O-acyl esters 4, 5 and 6 respectively. Catalytic hydrogenolysis of 4, 5, and 6 yielded 6,6′-di-O-4-[p-(hexadecyloxy)-phenyl] butyryl-trehalose 7; 6,6′-di-O-corynomycolyl-trehalose 8; and 6,6′-di-O-bovi-mycolyl-trehalose 9 respectively.  相似文献   

8.
Clostridium perfringens alpha-toxin degrades phosphatidylcholine (PC) in the bilayer of liposomes and destroys the membrane. The effect of the type and position of unsaturation in the fatty acyl chain of PC (18:0/18:1 PC) synthesized on the toxin-induced leakage of carboxyfluorescein (CF) from PC liposomes was examined. Differential scanning calorimetry showed that the phase transition temperature (Tm) was minimal when the triple bond was positioned at C (9) in the sn-2 acyl chain. The toxin-induced CF leakage decreased with the migration of the bond from C (9) to either end of the acyl chain in PC. The PC containing the cis-double bond had a similar Tm to that with the triple bond, but a lower value than the PC containing the trans-double bond. Furthermore, the toxin-induced leakage from liposomes composed of PC containing the cis-double bond resembled that with PC having the triple bond and was greater than that from liposomes with PC having the trans-double bond. The binding of a H148G mutant to PC liposomes showed a reciprocal relationship in terms of the Tm value of PC containing the triple bond. These results indicate that the toxin-induced membrane damage is closely related to membrane fluidity in liposomes.  相似文献   

9.
Modelin-5-CONH2, a synthetic antimicrobial peptide, was used to gain an insight into species-selective haemolytic activity. The peptide displayed limited haemolytic activity against sheep (12 %), human (2 %), and pig (2 %) erythrocytes. Our results show that Modelin-5-CONH2 had a disordered structure in the presence of vesicles formed from sheep, human, and pig erythrocyte lipid extract (<26 % helical) yet folded to form helices in the presence of a phosphatidylcholine (PC) membrane interface (e.g. >42 % in the presence of 1,2-dimyristoyl-sn-glycero-3-phosphocholine). Monolayer studies showed a strong correlation between anionic lipid content and monolayer insertion and lysis inducing surface pressure changes of 9.17 mN m?1 for 1,2-dimyristoyl-sn-glycero-3-phospho-l-serine compared with PC monolayers, which induced pressure changes of ca. 3 mN m?1. The presence of cholesterol in the membrane is shown to increase the packing density as the PC:sphingomyelin (SM) ratio increases so preventing the peptide from forming a stable association with the membrane. The data suggests that the key driver for membrane interaction for Modelin-5-CONH2 is the anionic lipid attraction. However, the key factors in the species-specific haemolysis level for this peptide are the differing packing densities which are influenced by the SM:PC:cholesterol ratio.  相似文献   

10.
The 6-O-mesyl, 6-O-tosyl, 6-bromo-6-deoxy, and 6-deoxy-6-iodo derivatives of 1,4-anhydro-DL-allitol were obtained by treatment of the corresponding 1,6-di-substituted derivatives (2, 3, 6, 4) of 2,3,4,5-tetra-O-acetylallitol with hot, methanolic hydrogen chloride. Compounds 2 and 3 were prepared by the acetolysis of the 1,6-di-O-mesyl and 1,6-di-O-tosyl derivatives (8 and 11) of di-O-benzylideneallitol. Iodide displacement on 2 gave 4, and detritylation-bromination of 2,3,4,5-tetra-O-acetyl-1,6-di-O-tritylallitol (5) gave 6. The acetal residues of di-O-benzylideneallitol have been shown to span the secondary carbon atoms.  相似文献   

11.
To prepare a new antiviral duplex drug linking Zidovudine (AZT) and Foscarnet (PFA) via a lipophilic octadecylglycerol residue we condensed 1-O-4-monomethoxytrityl-3-O-octadecyl-sn-glycerol-2-hydrogenphosphonate obtained from 3-O-octadecyl-sn-glycerol with AZT by the phosphonate method. The purified condensation product was de-tritylated resulting in 3′-azido-3′-deoxythymidylyl-(5′  2-O)-3-O-octadecyl-sn-glycerol, followed by treatment with (ethoxycarbonyl)phosphoric dichloride. The resulting 3′-azido-3′-deoxy-thymidylyl-(5′  2)-3-O-octadecyl-sn-glycerol-1-O-(ethoxycarbonyl)phosphonate was purified by preparative RP-18 column chromatography. The antiviral duplex drug 3′-azido-3′-deoxythymidylyl-(5′  2-O)-3-O-octadecyl-sn-glycerol-1-O-phosphonoformate trisodium salt (AZT–lipid–PFA) was obtained after alkaline cleavage of the phosphonoformate ethylester residue. The overall yield of the five step synthesis performed at gram scale was about 30%. According to a supposed pathway AZT–lipid–PFA could be cleaved to yield a mixture of different antiviral compounds such as AZT, AZT-5′-monophosphate, octadecylglycerol–AZT, PFA and octadecylglycerol–PFA, possibly producing additive and/or synergistic antiviral effects. In vitro studies showed that the duplex drug exhibits antiviral activities against HIV and especially against drug-resistant strains and clinical isolates of HSV and HCMV. The E50 values of AZT–lipid–PFA against HIV ranged between 170 and 200 nM. The half-maximal inhibitory doses (IC50) against highly acyclovir (ACV)-resistant HSV isolates determined by a plaque reduction assay ranged between 1.87 and 4.59 μM. Using ganciclovir (GCV)-sensitive, GCV resistant and drug cross-resistant HCMV strains the IC50-values of AZT–lipid–PFA were between 2.78 and 1.18 μM. With regard to PFA, the IC50-value of AZT–lipid–PFA determined on a multi-drug-resistant HCMV strain was about 90-fold lower than that of PFA, demonstrating the superior antiviral effect of the duplex-drug.  相似文献   

12.
Stereo- and regio-selective synthesis of 3-O-(2-acetamido-2-deoxy-3-O-β-d- galactopyranosyl-β-d-galactopyranosyl)-1,2-di-O-tetradecyl-sn-glycerol by use of 1,2-di-O-tetradecyl-3-O-(3,4,6-tri-O-acetyl-2-deoxy-2-phthalimido-β-d-galactopyranosyl)-sn-glycerol as a key intermediate is described.  相似文献   

13.
Stable nanosized bilayer disks were prepared from either 1,2-distearoyl-sn-glycero-3-phosphocholine (DSPC) and cholesterol, or lipid mixtures with a composition reflecting that of the porcine brush border membrane. Two different polyethylene glycol (PEG)-grafted lipids, the negatively charged 1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy (polyethylene glycol)-5000] (DSPE-PEG5000) and the neutral N-palmitoyl-sphingosine-1-[succinyl (methoxy (polyethylene glycol) 5000] (Ceramide-PEG5000), were used to stabilize the disks. The disks were employed as model membranes in drug partition studies based on a fast chromatography method. Results show that the lipid composition, as well as the choice of PEG-lipid, have an important influence on the partition behavior of charged drugs. Comparative studies using multilamellar liposomes indicate that bilayer disks have the potential to generate more accurate partition data than do liposomes. Further, initial investigations using bacteriorhodopsin suggest that membrane proteins can be reconstituted into the bilayer disks. This fact further strengthens the potential of the bilayer disk as an attractive model membrane.  相似文献   

14.
Synthetic routes for the following mannooligosylglycerolipids of biological interest were developed by using regioselectively protected monosaccharide synthons and l,2-di-O-alkyl-sn-glycerol; 3-O-(2-O-α-D-mannopyranosyl-α-D-mannopyranosyl)-l,2-di-O-tetradecyl-sn-glycerol; 3-O-[2-O-(2-O-α-D-mannopyranosyl-α-D-mannopyranosyl)-α-D-mannopyranosyl]-l,2-di-O-tetradecyl-sn-glycerol; 3-O-(6-O-α-D-mannopyranosyl-α-D-mannopyranosyl)-l,2-di-O-tetradecyl-sn-glycerol; and 3-O-(3,6-di-O-α-D-mannopyranosyl-α-D-mannopyranosyl)-1,2-di-α-tetradecyl-sn-glycerol.  相似文献   

15.
The bilayer phase transitions of palmitoylstearoyl-phosphatidylcholine (PSPC), diheptadecanoyl-PC (C17PC) and stearoylpalmitoyl-PC (SPPC) which have the same total carbon numbers in the two acyl chains were observed by differential scanning calorimetry and high-pressure optical method. As the temperature increased, these bilayers exhibited four phases of the subgel (Lc), lamellar gel (Lβ′), ripple gel (Pβ′) and liquid crystal (Lα), in turn. The Lc phase was observed only in the first heating scan after cold storage. The temperatures of the phase transitions were almost linearly elevated by applying pressure. The temperature-pressure phase diagrams and the thermodynamic quantities associated with the phase transitions were compared among the lipid bilayers. For all the bilayers studied, the pressure-induced interdigitated gel (LβI) phase appeared above the critical interdigitation pressure (CIP) between the Lβ′ and Pβ′ phases. The CIPs for the PSPC, C17PC and SPPC bilayers were found to be 50.6, 79.1 and 93.0 MPa, respectively. Contribution of two acyl chains to thermodynamic properties for the phase transitions of asymmetric PSPC and SPPC bilayers was not even. The sn-2 acyl chain lengths of asymmetric PCs governed primarily the bilayer properties. The fluorescence spectra of Prodan in lipid bilayers showed the emission maxima characteristic of bilayer phases, which were dependent on the location of Prodan in the bilayers. Second derivative of fluorescent spectrum exhibited the original emission spectrum of Prodan to be composed of the distribution of Prodan into multiple locations in the lipid bilayer. The F497/F430 value, a ratio of second derivative of fluorescence intensity at 497 nm to that at 430 nm, is decisive evidence whether bilayer interdigitation will occur. With respect to the Lβ′/LβI phase transition in the SPPC bilayer, the emission maximum of Prodan exhibited the narrow-range red-shift from 441 to 449 nm, indicating that the LβI phase in the SPPC bilayer has a less polar “pocket” formed by a space between uneven terminal methyl ends of the sn-1 and sn-2 chains, in which the Prodan molecule remains stably.  相似文献   

16.
Chlorhexidine (CHX) is an effective anti-bacterial agent whose mode of action is thought to be the disruption of the cell membrane. It is known to partition into phospholipid bilayers of aqueous model-membrane preparations. Neutron diffraction data taken at 36 °C on the location of CHX in phosphatidylcholine (PC) bilayers is presented. The center of mass of the deuterated hydrocarbon chain of CHX is found to reside 16 Å from the center of the bilayer in 1,2-dimyristoyl-sn-glycero-3-phosphatidylcholine (14:0–14:0 PC). This places the drug near the glycerol backbone of the lipid, and suggests a mode of action whereby the molecule is bent in half and inserts wedge-like into the lipid matrix. This mechanism is distinct from detergent-like mechanisms of membrane disruption and more similar to some anti-microbial peptide action, where peptides insert obliquely into the bilayer headgroup region to disrupt its structure.  相似文献   

17.
The major flavonoid of Marchantia berteroana is hypolaetin 8-O-β-d-glucuronide. This is accompanied by apigenin and luteolin, isoscutellarein (8-hydroxyapigenin) 8-O-β-d-glucuronide, the 7-O-β-d-glucuronide and -galacturonide of apigenin and luteolin, luteolin 3′-O-β-d-glucuronide and -galacturonide, luteolin 7,3′-di-O-β-d-glucuronide and -galacturonide, luteolin 3′,4′-di-O-β-d-glucuronide and -galacturonide, luteolin 7,4′-di-O-β-d-glucuronide, and hypolaetin 8,4′-di-O-β-d-glucuronide. The isoscutellarein and hypolaetin glucuronides, and the galacturonide flavones are all new natural products.  相似文献   

18.
2-Methyl-(2-acetamido-3,4,6-tri-O-benzyl-1,2-dideoxy-α-D-glucopyrano)-[2,1-d]-2-oxazoline,2-methyl-(2-acetamido-6-O-acetyl-3,4-di-O-benzyl-1,2-dideoxy-α-D-glucopyrano)-[2,1-d]-2-oxazoline,and 2-methyl-(2-acetamido-4-O-acetyl-3,6-di-O-benzyl-1,2-dideoxy-α-D-glucopyrano)-[2,1-d]-2-oxazoline were synthesized from the allyl 2-acetamido-3,4,6-tri-O-benzyl-2-deoxy-D-glucopyranosides, and from the 3,4-di-O-benzyl or 3,6-di-O-benzyl analogs, respectively, both the α and β anomer being used in each case. The preparation of allyl 2-acetamido-3,4,6-tri-O-benzyl- and 3,6-di-O-benzyl-2-deoxy-β-D-glucopyranoside is also described. Treatment of the tri-O-benzyl oxazoline with dibenzyl phosphate gave a pentabenzylglycosyl phosphate, from which all the benzyl groups were removed by catalytic hydrogenation, giving 2-acetamido-2-deoxy-α-D-glucopyranosyl phosphate. The corresponding β anomer was not detectable. Treatment of the 3,4-, or 3,6-, di-O-benzyl oxazoline with allyl 2-acetamido-3,4-di-O-benzyl-α-D-glucopyranoside readily gave disaccharide products from which the protecting groups were removed, to give the (1→6)-linked isomer of di-N-acetylchitobiose. Under both acidic and basic conditions, this isomer was less stable than the (1→4)-linked compound.Attempts to employ the 3,6-di-O-benzyl oxazoline for the formation of (1→4)-linked disaccharides, by treatment with either anomer of allyl 2-acetamido-3,6-di-O-benzyl-2-deoxy-D-glucopyranoside, were not very successful, presumably owing to hindrance by the bulky benzyl groups.  相似文献   

19.
Synthetic 2′-hydroxy-3,4′,6′-trimethoxy-4-benzyloxychalcone (I) affords (±)-7,3′-di-O-methyleriodictyol (II) and 7,3′-di-O-methylluteolin (or velutin, VII) identical with natural samples. Similarly synthetic 2′-hydroxy-4,4′,6′-trimethoxy-3-benzyloxychalcone (X) gives natural (±)-7,4′-di-O-methyleriodictyol (XI) and 7,4′-di-O-methylluteolin (or pilloin, IX). However, attempts to partially etherify II with one mole of prenyl bromide to obtain the natural prenyl ether failed; only the corresponding diprenyloxychalcone (IV) was obtained.  相似文献   

20.
The major flavonoids of Marchantia polymorpha var. polymorpha and aquatica are the 7-O-β-d-glucuronides of apigenin and luteolin, luteolin 3′-O-β-d-glucuronide, luteolin 7,3′-di-O-β-d-glucuronide, and the 7,4′-di-O-β-d-glucuronides of apigenin and luteolin. These are accompanied by minor amounts of apigenin, luteolin, luteolin 3′,4′-di-O-β-d-glucuronide and luteolin 7,3′,4′-tri-O-β-d-glucuronide. All the luteolin di- and triglucuronides except the 3′,4′-di- substituted compound are new natural products.  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