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 glycolipid transfer protein (GLTP)-mediated movement of galactosylceramide from model membrane donor vesicles to acceptor vesicles is sensitive to the membrane environment surrounding the glycolipid. GLTP can catalyze the transfer of a fluorescently labeled GSL, anthrylvinyl-galactosylceramide (AV-GalCer), from vesicles composed of 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine and dipalmitoylphosphatidylcholine matrices, but not from vesicles prepared from N-palmitoylsphingomyelin, regardless of the cholesterol content of the vesicles. In this study, we have examined the structural features of sphingomyelin (SM) that are responsible for its inhibition of the rate of GLTP-catalyzed transfer of AV-GalCer. The rate of glycolipid transfer was enhanced when the N-palmitoyl chain of SM was replaced with an N-oleoyl chain. Analogs of N-palmitoyl-SM in which the 4,5-double bond of the long-chain base is reduced or the 3-hydroxy group is removed did not inhibit GLTP-catalyzed transfer of AV-GalCer. When the donor vesicles were prepared with phosphatidylcholines or ether-linked phosphatidylcholine analogs, the transfer rates of AV-GalCer increased with increasing degree of unsaturation. The rate of AV-GalCer transfer was strongly dependent on the unsaturation degree of the acyl and/or alkyl chains. For ester-linked PCs, the transfer rate increased in the order DPPC相似文献   

2.
The mammalian glycolipid transfer protein, GLTP, catalyzes the transfer in vitro of glycolipids between membranes. In this study we have examined on one hand the effect of the variations in the donor vesicle composition and on the other hand the effects of variations in the acceptor vesicle composition on the GLTP-catalyzed transfer kinetics of galactosylceramide between bilayer vesicles. For this purpose a resonance energy transfer assay was used, the energy donor being anthrylvinyl-galactosylceramide and the energy acceptor DiO-C16. First, we show that the transfer of anthrylvinyl-galactosylceramide from palmitoyl-oleoyl-phosphatidylcholine donor vesicles was faster than from dipalmitoyl-phosphatidylcholine vesicles, and that there is no transfer from palmitoyl-sphingomyelin vesicles regardless of the cholesterol amount. In this setup the acceptor vesicles were always 100% palmitoyl-oleoyl-phosphatidylcholine. We also showed that the transfer in general is faster from small highly curved vesicles compared to that from larger vesicles. Secondly, by varying the acceptor vesicle composition we showed that the transfer is faster to mixtures of sphingomyelin and cholesterol compared to mixtures of phosphatidylcholines and cholesterol. Based on these experiments we conclude that the GLTP mediated transfer of anthrylvinyl-galactosylceramide is sensitive to the matrix lipid composition and membrane bending. We postulate that a tightly packed membrane environment is most effective in preventing GLTP from accessing its substrates, and cholesterol is not required to protect the glycosphingolipid in the membrane from being transferred by GLTP. On the other hand GLTP can more easily transfer glycolipids to ‘lipid raft’ like membranes, suggesting that the protein could be involved in raft assembly.  相似文献   

3.
N-Acyl phosphatidylethanolamines are negatively charged phospholipids, which are naturally occurring albeit at low abundance. In this study, we have examined how the amide-linked acyl chain affected the membrane behavior of the N-acyl-1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylethanolamine (N-acyl-POPE) or N-acyl-dipalmitoyl-sn-glycero-3-phosphatidylethanolamine (N-acyl-DPPE), and how the molecules interacted with cholesterol. The gel → liquid crystalline transition temperature of sonicated N-acyl phosphatidylethanolamine vesicles in water correlated positively with the number of palmitic acyl chains in the molecules. Based on diphenylhexatriene steady state anisotropy measurements, the presence of 33 mol% cholesterol in the membranes removed the phase transition from N-oleoyl-POPE bilayers, but failed to completely remove it from N-palmitoyl-DPPE and N-palmitoyl-POPE bilayers, suggesting rather weak interaction of cholesterol with the N-saturated NAPEs. The rate of cholesterol desorption from mixed monolayers containing N-palmitoyl-DPPE and cholesterol (1:1 molar ratio) was much higher compared to cholesterol/DPPE binary monolayers, suggesting a weak cholesterol interaction with N-palmitoyl-DPPE also in monolayers. In bilayer membranes, both N-palmitoyl-POPE and N-palmitoyl-DPPE failed to form sterol-rich domains, and in fact appeared to displace sterol from sterol/N-palmitoyl-sphingomyelin domains. The present data provide new information about the effects of saturated NAPEs on the lateral distribution of cholesterol in NAPE-containing membranes. These findings may be of relevance to neural cells which accumulate NAPEs during stress and cell injury.  相似文献   

4.
1,2-Propanediol and 3-aryloxy/alkyloxy derivatives thereof are bulk commodities produced directly from glycerol. Glycosylation is a promising route for their functional diversification into useful fine chemicals. Regioselective glucosylation of the secondary hydroxyl in different 1,2-propanediols was achieved by a sucrose phosphorylase-catalyzed transfer reaction where sucrose is the substrate and 2-O-α-d-glucopyranosyl products are exclusively obtained. Systematic investigation for optimization of the biocatalytic synthesis included prevention of sucrose hydrolysis, which occurs in the process as a side reaction of the phosphorylase. In addition to ‘nonproductive’ depletion of donor substrate, the hydrolysis also resulted in formation of maltose and kojibiose (up to 45%) due to secondary enzymatic glucosylation of the glucose thus produced. Using 3-ethoxy-1,2-propanediol as the acceptor substrate (1.0 M), the desired transfer product was obtained in about 65% yield when employing a moderate (1.5-fold) excess of sucrose donor. Loss of the glucosyl substrate to ‘glucobiose’ by-products was minimal (<7.5%) under these conditions. The reactivity of other acceptors decreased in the order, 3-methoxy-1,2-propanediol > 1,2-propanediol > 3-allyloxy-1,2-propanediol > 3-(o-methoxyphenoxy)-1,2-propanediol > 3-tert-butoxy-1,2-propanediol. Glucosylated 1,2-propanediols were not detectably hydrolyzed by sucrose phosphorylase so that their synthesis by transglucosylation occurred simply under quasi-equilibrium reaction conditions.  相似文献   

5.
A new β-glucosidase gene (bglSp) was cloned from the ginsenoside converting Sphingomonas sp. strain 2F2 isolated from the ginseng cultivating filed. The bglSp consisted of 1344 bp (447 amino acid residues) with a predicted molecular mass of 49,399 Da. A BLAST search using the bglSp sequence revealed significant homology to that of glycoside hydrolase superfamily 1. This enzyme was overexpressed in Escherichia coli BL21 (DE3) using a pET21-MBP (TEV) vector system. Overexpressed recombinant enzymes which could convert the ginsenosides Rb1, Rb2, Rc and Rd to the more pharmacological active rare ginsenosides gypenoside XVII, ginsenoside C-O, ginsenoside C-Mc1 and ginsenoside F2, respectively, were purified by two steps with Amylose-affinity and DEAE-Cellulose chromatography and characterized. The kinetic parameters for β-glucosidase showed the apparent Km and Vmax values of 2.9 ± 0.3 mM and 515.4 ± 38.3 μmol min−1 mg of protein−1 against p-nitrophenyl-β-d-glucopyranoside. The enzyme could hydrolyze the outer C3 glucose moieties of ginsenosides Rb1, Rb2, Rc and Rd into the rare ginsenosides Gyp XVII, C-O, C-Mc1 and F2 quickly at optimal conditions of pH 5.0 and 37 °C. A little ginsenoside F2 production from ginsenosides Gyp XVII, C-O, and C-Mc1 was observed for the lengthy enzyme reaction caused by the side ability of the enzyme.  相似文献   

6.
The composition of the lipid bilayer is increasingly being recognised as important for the regulation of integral membrane protein folding and function, both in vivo and in vitro. The folding of only a few membrane proteins, however, has been characterised in different lipid environments. We have refolded the small multidrug transporter EmrE in vitro from a denatured state to a functional protein and monitored the influence of lipids on the folding process. EmrE is part of a multidrug resistance protein family that is highly conserved amongst bacteria and is responsible for bacterial resistance to toxic substances. We find that the secondary structure of EmrE is very stable and only small amounts are denatured even in the presence of unusually high denaturant concentrations involving a combination of 10 M urea and 5% SDS. Substrate binding by EmrE is recovered after refolding this denatured protein into dodecylmaltoside detergent micelles or into lipid vesicles. The yield of refolded EmrE decreases with lipid bilayer compositional changes that increase the lateral chain pressure within the bilayer, whilst conversely, the apparent rate of folding seems to increase. These results add further weight to the hypothesis that an increased lateral chain pressure hinders protein insertion across the bilayer. Once the protein is inserted, however, the greater pressure on the transmembrane helices accelerates correct packing and final folding. This work augments the relatively small number of biophysical folding studies in vitro on helical membrane proteins.  相似文献   

7.
In this study we have addressed the ability of the glycolipid transfer protein (GLTP) to transfer anthrylvinyl-galactosylceramide at different pH and sodium chloride concentrations, and the ability of three different mutants to transfer the fluorescently labeled galactosylceramide between donor and acceptor model membranes. We constructed single tryptophan mutants with site-directed mutagenesis where two of the three tryptophan (W) of wild-type human GLTP were substituted with phenylalanine (F) and named W85 GLTP (W96F and W142F), W96 GLTP (W85F and W142F) and W142 GLTP (W85F and W96F) accordingly. Wild-type GLTP and W96 GLTP were both able to transfer anthrylvinyl-galactosylceramide, but the two variants W85 GLTP and W142 GLTP did not show any glycolipid transfer activity, indicating that the tryptophan in position 96 is crucial for transfer activity. Tryptophan fluorescence emission showed a blue shift of the maximal emission wavelength upon interaction of glycolipid containing vesicle with wild-type GLTP and W96 GLTP, while no blue shift was recorded for the protein variants W85 GLTP and W142 GLTP. The quantum yield of tryptophan emission was highest for the W96 GLTP protein whereas W85 GLTP, W142 GLTP and wild-type GLTP showed a lower and almost similar quantum yield. The lifetime and anisotropy decay of the different tryptophan mutants also changed upon binding to vesicles containing galactosylceramide. Again wild-type GLTP and W96 GLTP showed similar behavior in the presence of vesicles containing glycolipids. Taken together, our data show that the W96 is involved not only in the activity of the protein but also in the interaction between the protein and glycolipid containing membranes.  相似文献   

8.
It was recently shown that the structure of the fluorophore attached to the acyl chain of phosphatidylcholine analogs determines their mechanism of transport across the plasma membrane of yeast cells (Elvington et al., J. Biol Chem. 280:40957, 2005). In order to gain further insight into the physical properties of these fluorescent phosphatidylcholine (PC) analogs, the rate and mechanism of their intervesicular transport was determined. The rate of spontaneous exchange was measured for PC analogs containing either NBD (7-nitrobenz-2-oxa-1,3-diazol-4-yl), Bodipy FL (4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene), Bodipy 530 (4,4-difluoro-5,7-diphenyl-4-bora-3a,4a-diaza-s-indacene), or Bodipy 581 (4,4-difluoro-5-(4-phenyl-1,3-butadienyl)-4-bora-3a,4a-diaza-s-indacene) attached to a five or six carbon acyl chain in the sn-2 position. The rate of transfer between phospholipid vesicles was measured by monitoring the increase in fluorescence as the analogs transferred from donor vesicles containing self-quenching concentrations to unlabeled acceptor vesicles. Kinetic analysis indicated that the transfer of each analog occurred by diffusion through the water phase as opposed to transfer during vesicle collisions. The vesicle-to-monomer dissociation rate constants differed by over four orders of magnitude: NBD-PC (kdis = 0.115 s− 1; t1/2 = 6.03 s); Bodipy FL-PC (kdis = 5.2 × 10− 4; t1/2 = 22.2 min); Bodipy 530-PC (kdis = 1.52 × 10− 5; t1/2 = 12.6 h); and Bodipy 581-PC (kdis = 5.9 × 10− 6; t1/2 = 32.6 h). The large differences in spontaneous rates of transfer through the water measured for these four fluorescent PC analogs reflect their hydrophobicity and may account for their recognition by different mechanisms of transport across the plasma membrane of yeast.  相似文献   

9.
Glycosphingolipids (GSLs) are important constituents of lipid rafts and caveolae, are essential for the normal development of cells, and are adhesion sites for various infectious agents. One strategy for modulating GSL composition in lipid rafts is to selectively transfer GSL to or from these putative membrane microdomains. Glycolipid transfer protein (GLTP) catalyzes selective intermembrane transfer of GSLs. To enable effective use of GLTP as a tool to modify the glycolipid content of membranes, it is imperative to understand how the membrane regulates GLTP action. In this study, GLTP partitioning to membranes was analyzed by monitoring the fluorescence resonance energy transfer from tryptophans and tyrosines of GLTP to N-(5-dimethyl-aminonaphthalene-1-sulfonyl)-1,2-dihexadecanoyl-sn-glycero-3-phospho-ethanolamine present in bilayer vesicles. GLTP partitioned to POPC vesicles even when no GSL was present. GLTP interaction with model membranes was nonpenetrating, as assessed by protein-induced changes in lipid monolayer surface pressure, and nonperturbing in that neither membrane fluidity nor order were affected, as monitored by anisotropy of 1,6-diphenyl-1,3,5-hexatriene and 6-dodecanoyl-N,N-dimethyl-2-naphthylamine, even though the tryptophan anisotropy of GLTP increased in the presence of vesicles. Ionic strength, vesicle packing, and vesicle lipid composition affected GLTP partitioning to the membrane and led to the following conclusion: Conditions that increase the ratio of bound/unbound GLTP do not guarantee increased transfer activity, but conditions that decrease the ratio of bound/unbound GLTP always diminish transfer. A model of GLTP interaction with the membrane, based on the partitioning equilibrium data and consistent with the kinetics of GSL transfer, is presented and solved mathematically.  相似文献   

10.
To investigate the effect of lipid structure upon the membrane topography of hydrophobic helices, the behavior of hydrophobic peptides was studied in model membrane vesicles. To define topography, fluorescence and fluorescence quenching methods were used to determine the location of a Trp at the center of the hydrophobic sequence. For peptides with cationic residues flanking the hydrophobic sequence, the stability of the transmembrane (TM) configuration (relative to a membrane-bound non-TM state) increased as a function of lipid composition on the order: 1:1 (mol:mol) 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC):1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine ∼ 6:4 POPC:cholesterol < POPC ∼ dioleoylphosphatidylcholine (DOPC) < 1,2-dioleoyl-sn-glycero-3-[phospho-rac-(1-glycerol)] sodium salt (DOPG) ≤ 1,2-dioleoyl-sn-glycero-3-[phospho-l-serine] sodium salt (DOPS), indicating that the anionic lipids DOPG and DOPS most strongly stabilized the TM configuration. TM stabilization was near maximal at 20-30 mol% anionic lipid, which are physiologically relevant values. TM stabilization by anionic lipid was observed for hydrophobic sequences with a diverse set of sequences (including polyAla), diverse lengths (from 12 to 22 residues), and various cationic flanking residues (H, R, or K), but not when the flanking residues were uncharged. TM stabilization by anionic lipid was also dependent on the number of cationic residues flanking the hydrophobic sequence, but was still significant with only one cationic residue flanking each end of the peptide. These observations are consistent with TM-stabilizing effects being electrostatic in origin. However, Trp located more deeply in DOPS vesicles relative to DOPG vesicles, and peptides in DOPS vesicles showed increased helix formation relative to DOPG and all other lipid compositions. These observations fit a model in which DOPS anchors flanking residues near the membrane surface more strongly than does DOPG and/or increases the stability of the TM state to a greater degree than DOPG. We conclude that anionic lipids can have significant and headgroup structure-specific effects upon membrane protein topography.  相似文献   

11.
We have monitored the composition of supported phospholipid bilayers during phospholipase A2 hydrolysis using specular neutron reflection and ellipsometry. Porcine pancreatic PLA2 shows a long lag phase of several hours during which the enzyme binds to the bilayer surface, but only 5 ± 3% of the lipids react before the onset of rapid hydrolysis. The amount of PLA2, which resides in a 21 ± 1 Å thick layer at the water-bilayer interface, as well as its depth of penetration into the membrane, increase during the lag phase, the length of which is also proportional to the enzyme concentration. Hydrolysis of a single-chain deuterium labelled d31-POPC reveals for the first time that there is a significant asymmetry in the distribution of the reaction products between the membrane and the aqueous environment. The lyso-lipid leaves the membrane while the number of PLA2 molecules bound to the interface increases with increasing fatty acid content. These results constitute the first direct measurement of the membrane structure and composition, including the location and amount of the enzyme during hydrolysis. These are discussed in terms of a model of fatty-acid mediated activation of PLA2.  相似文献   

12.
A novel cytochrome ba complex was isolated from aerobically grown cells of the thermoacidophilic archaeon Acidianus ambivalens. The complex was purified with two subunits, which are encoded by the cbsA and soxN genes. These genes are part of the pentacistronic cbsAB-soxLN-odsN locus. The spectroscopic characterization revealed the presence of three low-spin hemes, two of the b and one of the as-type with reduction potentials of + 200, + 400 and + 160 mV, respectively. The SoxN protein is proposed to harbor the heme b of lower reduction potential and the heme as, and CbsA the other heme b. The soxL gene encodes a Rieske protein, which was expressed in E. coli; its reduction potential was determined to be + 320 mV. Topology predictions showed that SoxN, CbsB and CbsA should contain 12, 9 and one transmembrane α-helices, respectively, with SoxN having a predicted fold very similar to those of the cytochromes b in bc1 complexes. The presence of two quinol binding motifs was also predicted in SoxN. Based on these findings, we propose that the A. ambivalens cytochrome ba complex is analogous to the bc1 complexes of bacteria and mitochondria, however with distinct subunits and heme types.  相似文献   

13.
The influence of a mammalian sterol cholesterol and a plant sterol β-sitosterol on the structural parameters and hydration of bilayers in unilamellar vesicles made of monounsaturated diacylphosphatidylcholines (diCn:1PC, n = 14-22 is the even number of acyl chain carbons) was studied at 30 °C using small-angle neutron scattering (SANS). Recently published advanced model of lipid bilayer as a three-strip structure was used with a triangular shape of polar head group probability distribution (Ku?erka et al., Models to analyze small-angle neutron scattering from unilamellar lipid vesicles, Physical Review E 69 (2004) Art. No. 051903). It was found that 33 mol% of both sterols increased the thickness of diCn:1PC bilayers with n = 18-22 similarly. β-sitosterol increased the thickness of diC14:1PC and diC16:1PC bilayers a little more than cholesterol. Both sterols increased the surface area per unit cell by cca 12 Å2 and the number of water molecules located in the head group region by cca 4 molecules, irrespective to the acyl chain length of diCn:1PC. The structural difference in the side chain between cholesterol and β-sitosterol plays a negligible role in influencing the structural parameters of bilayers studied.  相似文献   

14.
Inositol phosphorylceramides (IPCs) are a class of anionic sphingolipids with a single inositol-phosphate head group coupled to ceramide. IPCs and more complex glycosylated IPCs have been identified in fungi, plants and protozoa but not in mammals. IPCs have also been identified in detergent resistant membranes in several organisms. Here we report on the membrane properties of the saturated N-palmitoyl-IPC (P-IPC) in one component bilayers as well as in complex bilayers together with 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine and cholesterol. The membrane properties of P-IPC were shown to be affected by calcium. According to anisotropy changes reported by DPH, the gel-to-liquid transition temperature (Tm) of P-IPC was 48 °C. Addition of 5 mM CaCl2 during vesicle preparation markedly increased the Tm (65 °C). According to fluorescence quenching experiments in complex lipid mixtures, P-IPC formed sterol containing domains in an otherwise fluid environment. The P-IPC containing domains melted at a lower temperature and appeared to contain less sterol as compared to domains containing N-palmitoyl-sphingomyelin. Calcium further reduced the sterol content of the ordered domains and also increased the thermal stability of the domains. Calcium also induced vesicle aggregation of unilamellar vesicles containing P-IPC, as was observed by 4D confocal microscopy and dynamic light scattering. We believe that IPCs and the calcium induced effects could be important in numerous membrane associated cellular processes such as membrane fusion and in membrane raft linked processes.  相似文献   

15.
Guanylate cyclase-activating protein-2 (GCAP-2) is a retinal Ca2+ sensor protein. It is responsible for the regulation of both isoforms of the transmembrane photoreceptor guanylate cyclase, a key enzyme of vertebrate phototransduction. GCAP-2 is N-terminally myristoylated and full activation of its target proteins requires the presence of this lipid modification. The structural role of the myristoyl moiety in the interaction of GCAP-2 with the guanylate cyclases and the lipid membrane is currently not well understood. In the present work, we studied the binding of Ca2+-free myristoylated and non-myristoylated GCAP-2 to phospholipid vesicles consisting of dimyristoylphosphatidylcholine or of a lipid mixture resembling the physiological membrane composition by a biochemical binding assay and 2H solid-state NMR. The NMR results clearly demonstrate the full-length insertion of the aliphatic chain of the myristoyl group into the membrane. Very similar geometrical parameters were determined from the 2H NMR spectra of the myristoyl group of GCAP-2 and the acyl chains of the host membranes, respectively. The myristoyl chain shows a moderate mobility within the lipid environment, comparable to the acyl chains of the host membrane lipids. This is in marked contrast to the behavior of other lipid-modified model proteins. Strikingly, the contribution of the myristoyl group to the free energy of membrane binding of GCAP-2 is only on the order of − 0.5 kJ/mol, and the electrostatic contribution is slightly unfavorable, which implies that the main driving forces for membrane localization arises through other, mainly hydrophobic, protein side chain-lipid interactions. These results suggest a role of the myristoyl group in the direct interaction of GCAP-2 with its target proteins, the retinal guanylate cyclases.  相似文献   

16.
Lim YR  Yeom SJ  Kim YS  Oh DK 《Bioresource technology》2011,102(5):4277-4280
The optimum conditions for the production of l-arabinose from debranched arabinan were determined to be pH 6.5, 75 °C, 20 g l−1 debranched arabinan, 42 U ml−1 endo-1,5-α-l-arabinanase, and 14 U ml−1 α-l-arabinofuranosidase from Caldicellulosiruptor saccharolyticus and the conditions for sugar beet arabinan were pH 6.0, 75 °C, 20 g l−1 sugar beet arabinan, 3 U ml−1 endo-1,5-α-l-arabinanase, and 24 U ml−1 α-l-arabinofuranosidase. Under the optimum conditions, 16 g l−1l-arabinose was obtained from 20 g l−1 debranched arabinan or sugar beet arabinan after 120 min, with a hydrolysis yield of 80% and a productivity of 8 g l−1 h−1. This is the first reported trial for the production of l-arabinose from the hemicellulose arabinan by the combined use of endo- and exo-arabinanases.  相似文献   

17.
In this paper, the lytic activity of two variants of Serratia marcescens against promastigotes of Leishmania braziliensis was studied. In vitro assays showed that S. marcescens variant SM365 lyses L. braziliensis promastigotes, while the variant DB11 did not. Scanning electron microscopy (SEM) revealed that S. marcescens SM365 adheres to all cellular body and flagellum of the parasite. Several filamentous structures were formed and identified as biofilms. After 120 min incubation, they connect the protozoan to the developing bacterial clusters. SEM also demonstrated that bacteria, adhered onto L. braziliensis promastigote surface, formed small filamentous structures which apparently penetrates into the parasite membrane. d-mannose protects L. braziliensis against the S. marcescens SM365 lytic effect in a dose dependent manner. SM365 variant pre cultivated at 37 °C did not synthesize prodigiosin although the adherence and lysis of L. braziliensis were similar to the effect observed with bacteria cultivated at 28 °C, which produce high concentrations of prodigiosin. Thus, we suggest that prodigiosin is not involved in the lysis of promastigotes and that adherence promoted by bacterial mannose-sensitive (MS) fimbriae is a determinant factor in the lysis of L. braziliensis by S. marcescens SM365.  相似文献   

18.
The iridoid glycosides, genipin 1-O-β-d-isomaltoside (1) and genipin 1,10-di-O-β-d-glucopyranoside (2), together with six known iridoid glycosides, genipin 1-O-β-d-gentiobioside (3), geniposide (4), scandoside methyl ester (5), deacetylasperulosidic acid methyl ester (6), 6-O-methyldeacetylasperulosidic acid methyl ester (7), and gardenoside (8) were isolated from an EtOH extract of Gardeniae Fructus. The structures and relative stereochemistries of the metabolites were elucidated on the basis of 1D- and 2D-NMR spectroscopic techniques, high-resolution mass spectrometry, and chemical evidence. Geniposide (4), one of the main compounds of Gardeniae Fructus, was tested for treatment of ankle sprain using an ankle sprain model in rats. From the second to fifth day, the geniposide (4) (100 mg/ml) treated group exhibited significant differences (p < 0.01) with ∼21-34% reduction in swelling ratio compared with those of the vehicle treated control group. This indicated the potential effect of geniposide (4) for the treatment of disorders such as ankle sprain.  相似文献   

19.
β-d-Galactopyranosyl-(1→3)-2-acetamido-2-deoxy-d-galactose (galacto-N-biose, GNB) is an important core structure in functional sugar chains such as T-antigen disaccharide and the core 1 sugar chain in mucin glycoproteins. We successfully developed a one-pot enzymatic production of GNB from sucrose and GalNAc by the concomitant action of four enzymes: sucrose phosphorylase, UDP-glucose-hexose 1-phosphate uridylyltransferase, UDP-glucose 4-epimerase, and galacto-N-biose/lacto-N-biose I phosphorylase in the presence of UDP-glucose and phosphate, by modifying the method of lacto-N-biose I production [Nishimoto, M.; Kitaoka, M., Biosci. Biotechnol. Biochem., 2007, 71, 2101-2104]. The reaction yield of GNB was 88% from GalNAc. GNB was isolated from the reaction mixture by crystallization after yeast treatment to obtain approximately 45 g of GNB in 95% purity from a 280-mL reaction mixture.  相似文献   

20.
Trichoderma asperellum produces two extracellular 1,3-β-d-glucanase upon induction with cell walls from Rhizoctonia solani. A minor 1,3-β-d-glucanase was purified to homogeneity by ion exchange chromatography on Q-Sepharose and gel filtration on Sephacryl S-100. A typical procedure provided 13.8-fold purification with 70% yield. SDS-PAGE of the purified enzyme showed a single protein band of molecular weight 27 kDa. The enzyme exhibited optimum catalytic activity at pH 3.6 and 45 °C. It was thermostable at 40 °C, and retained 75% activity after 60 min at 45 °C. The Km and Vmax values for 1,3-β-d-glucanase, using laminarin as substrate, were 0.323 mg ml−1 and 0.315 U min−1, respectively. The enzyme was strongly inhibited by Hg2+ and SDS. The enzyme was only active toward glucans containing β-1,3-linkages. Peptide sequences showed similarity with two endo-1,3(4)-β-d-glucanases from Aspergillus fumigatus Af293when compared against GenBank non-redundant database.  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