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 15 毫秒
1.
We have performed 20?ns of fully atomistic molecular dynamics simulations of Hen Egg-White Lysozyme in 0, 10, 20, 30, and 100% by weight of glycerol in water to better understand the microscopic physics behind the bioprotection offered by glycerol to naturally occuring biological systems. The solvent exposure of protein surface residues changes when glycerol is introduced. The dynamic behavior of the protein, as quantified by the incoherent intermediate scattering function, shows a nonmonotonic dependence on glycerol content. The fluctuations of the protein residues with respect to each other were found to be similar in all water-containing solvents, but different from the pure glycerol case. The increase in the number of protein–glycerol hydrogen bonds in glycerol–water binary mixtures explains the slowing down of protein dynamics as the glycerol content increases. We also explored the dynamic behavior of the hydration layer. We show that the short length scale dynamics of this layer are insensitive to glycerol concentration. However, the long length scale behavior shows a significant dependence on glycerol content. We also provide insights into the behavior of bound and mobile water molecules.  相似文献   

2.
The radiowave dielectric dispersions of DNA in different water–organic co-solvent mixtures have been measured in the frequency range from 100 kHz to 100 MHz, where the polarization mechanism is generally attributed to the confinement of counterions within some specific lengths, either along tangential or perpendicular to the polyion chain. The dielectric dispersions have been analyzed on the basis of two partially different dielectric models, a continuum counterion fluctuation model proposed by Mandel and a discrete charged site model, proposed by Minakata. The influence of the quality of the solvent on the dielectric parameters has been investigated in water–methanol and water–glycerol mixtures at different composition, by varying the permittivity ?m and the viscosity η of the solvent phase. The analysis of the dielectric spectra in solvents where electrostatic and hydrodynamic interactions vary with the solvent composition suggests that both the two models are able, in principle, to account for the observed high-frequency dielectric behavior. However, while some certain assumptions are necessary about the polyion structure within the Mandel model, no structural prerequisite is needed within the Minakata model, where the polarization mechanism invoked considers a radial counterion exchange with the outer medium, which is largely independent of the local polyion conformation.  相似文献   

3.
Almond β-glucosidase was used to catalyze the transglycosylation of p-nitrophenyl-β-d-glucoside to alkyl glucosides, with hydrolysis to glucose as a side reaction. The conversions were carried out in alcohols with varying water contents below water saturation. Both the total reaction rate and the ratio between the transglycosylation and hydrolysis increased with increasing water activity, and at a fixed water activity in the different alcohols, rate and transglycosylation/hydrolysis ratio increased in the following order: 1-octanol<1-hexanol<1-butanol. Synthesis of alkyl glucosides by transglycosylation in monophasic alcohol media is thus most favorable for short chain alcohols, and should be carried out at high water content.  相似文献   

4.
1-Octyl -d-xylobioside and xyloside were prepared by direct transglycosylation reaction of xylan and 1-octanol using purified xylanase from Aureobasidium pullulans. 2-Ethylhexyl -d- xylobioside and xyloside were also prepared in the same way. The maximum yields of 2-ethylhexyl -d-xylobioside and 2-ethylhexyl -d-xyloside were 110 and 54 mg/g xylan, respectively. The proposed mechanism for production of octyl xylobioside and xyloside involved the reaction of xylan and octanol by xylanase to produce octyl xylobioside and xylotrioside, the latter of which was simultaneously hydrolyzed by xylanase into octyl xyloside and xylobiose.  相似文献   

5.
A rather widespread idea on the functional importance of sphingolipids in cell membranes refers to the occurrence of ordered domains enriched in sphingomyelin and ceramide that are largely assumed to exist irrespective of the type of N-acyl chain in the sphingolipid. Ceramides and sphingomyelins are the simplest kind of two-chained sphingolipids and show a variety of species, depending on the fatty acyl chain length, hydroxylation, and unsaturation. Abundant evidences have shown that variations of the N-acyl chain length in ceramides and sphingomyelins markedly affect their phase state, interfacial elasticity, surface topography, electrostatics, and miscibility, and that even the usually conceived “condensed” sphingolipids and many of their mixtures may exhibit liquid-like expanded states. Their lateral miscibility properties are subtlety regulated by those chemical differences. Even between ceramides with different acyl chain length, their partial miscibility is responsible for a rich two-dimensional structural variety that impacts on the membrane properties at the mesoscale level. In this review, we will discuss the miscibility properties of ceramide, sphingomyelin, and glycosphingolipids that differ in their N-acyl or oligosaccharide chains. This work is a second part that accompanies a previous overview of the properties of membranes formed by pure ceramides or sphingomyelins, which is also included in this Special Issue.  相似文献   

6.
7.
8.
Diverse variations in membrane properties are observed in binary phosphatidylcholine/cholesterol mixtures. These mixtures are nonideal, displaying single or phase coexistence, depending on chemical composition and other thermodynamic parameters. When compared with pure phospholipid bilayers, there are changes in water permeability, bilayer thickness and thermomechanical properties, molecular packing and conformational freedom of phospholipid acyl chains, in internal dipolar potential and in lipid lateral diffusion. Based on the phase diagrams for DMPC/cholesterol and DPPC/cholesterol, we compare the equivalent polarity of pure bilayers with specific compositions of these mixtures, by using the Py empirical scale of polarity. Besides the contrast between pure and mixed lipid bilayers, we find that liquid-ordered (?o) and liquid-disordered (?d) phases display significantly different polarities. Moreover, in the ?o phase, the polarities of bilayers and their thermal dependences vary with the chemical composition, showing noteworthy differences for cholesterol proportions at 35, 40, and 45 mol%. At 20 °C, for DMPC/cholesterol at 35 and 45 mol%, the equivalent dielectric constants are 21.8 and 23.8, respectively. Additionally, we illustrate potential implications of polarity in various membrane-based processes and reactions, proposing that for cholesterol containing bilayers, it may also go along with the occurrence of lateral heterogeneity in biological membranes.  相似文献   

9.
Signaling through the Wnt/β-catenin pathway is a crucial determinant of hepatic zonal gene expression, liver development, regeneration, and tumorigenesis. Transgenic mice with hepatocyte-specific knockout of Ctnnb1 (encoding β-catenin) have proven their usefulness in elucidating these processes. We now found that a small number of hepatocytes escape the Cre-mediated gene knockout in that mouse model. The remaining β-catenin-positive hepatocytes showed approximately 25% higher cell volumes compared to the β-catenin-negative cells and exhibited a marker protein expression profile similar to that of normal perivenous hepatocytes or hepatoma cells with mutationally activated β-catenin. Surprisingly, the expression pattern was observed independent of the cell's position within the liver lobule, suggesting a malfunction of physiological periportal repression of perivenously expressed genes in β-catenin-deficient liver. Clusters of β-catenin-expressing hepatocytes lacked expression of the gap junction proteins Connexin 26 and 32. Nonetheless, β-catenin-positive hepatocytes had no striking proliferative advantage, but started to grow out on treatment with phenobarbital, a tumor-promoting agent known to facilitate the formation of mouse liver adenoma with activating mutations of Ctnnb1. Progressive re-population of Ctnnb1 knockout livers with wild-type hepatocytes was seen in aged mice with a pre-cirrhotic phenotype. In these large clusters of β-catenin-expressing hepatocytes, perivenous-specific gene expression was re-established. In summary, our data demonstrate that the zone-specificity of a hepatocyte's gene expression profile is dependent on the presence of β-catenin, and that β-catenin provides a proliferative advantage to hepatocytes when promoted with phenobarbital, or in a pre-cirrhotic environment.  相似文献   

10.
Guaiacol (2-methoxyphenol) and curcumin [1E,6E-1,7-di(4-hydroxy-3-methoxy-phenyl)-1,6-heptadiene-3,5-dione] were converted into their corresponding glucosides using glucose and an amyloglucosidase from Rhizopus. Guaiacol-α-D-glucoside yields ranged from 3 to 52% with the highest at pH 7.0. Curcumin-bis-α-D-glucoside yields ranged from 3 to 48% with the highest at pH 4.0 with 50% (w/w D-glucose) of enzyme. The phenolic hydroxyl group of guaiacol and both phenolic hydroxyl groups of curcumin were glucosylated at the C1 carbon of α-D-glucose indicating that the enzymatic reaction is stereospecific. Both guaiacol-α-D-glucoside and curcumin-bis-α-D-glucosides had antioxidant activities.  相似文献   

11.
The processes of dehydration and rehydration of β-cyclodextrin were studied by analysis of the (1)H NMR (nuclear magnetic resonance) line shape. Dehydration was carried in an open ampoule as a function of temperature and above 400 K total dehydration of β-cyclodextrin was observed. This result was confirmed by the thermogravimetry (TG) measurements. Rehydration was studied as a function of time at room temperature. After 40 days, β-cyclodextrin was found to absorb eight water molecules. The analysis of temperature changes in the shape of the (1)H NMR line of β-cyclodextrin kept in a closed ampoule and its dielectric measurements provided information on the mobility of water molecules. The water molecules were found to perform complex molecular motions, that is, reorientational jumps below 200K and additionally, translational motion (diffusion) above 200K.  相似文献   

12.
Novel biocomposites were fabricated by impregnating β-tricalcium phosphate (β-TCP)/zirconia particles into the polymers matrix. The composite materials were characterized using thermo-gravimetric analysis (TGA), X-ray diffraction (XRD), Fourier Transform Infrared (FT-IR) analyzes and Scanning Electron Microscopy (SEM). The results confirmed the conversion of hydroxyapatite (HA) to β-TCP at a sintering temperature of 1150 °C with or without zirconia powder. The in vitro behavior was assessed via measurement of calcium and phosphorus ions in SBF (simulated body fluid). FT-IR and SEM of the composites were performed pre and post immersion in SBF. The results prove that the bone like apatite layer formation was enhanced on the β-TCP-Z20/polymeric composite surface more than that on the β-TCP-Z10/polymeric composite. Therefore, the data confirmed that zirconia plays an important role in the enhancement of the apatite formation. The conclusions proved that the β-TCP-Z20/polymeric biocomposites, containing 20% of zirconia, are promising for bone remodeling applications.  相似文献   

13.
A mutant strain of Escherichia coli in which β-glucoside transport is resistant to catabolite inhibition by methyl α-glucoside was characterized. The mutation was probably within the gene, bglC, coding for the β-glucoside enzyme II. The mutant organism is shown to transport the β-glucoside substrate, salicin, in preference to methyl α-glucoside or fructose. Salicin also caused inducer exclusion of lactose in the mutant strain.  相似文献   

14.
Accumulating evidence suggests that Aβ1–42–membrane interactions may play an important role in the pathogenesis of Alzheimer’s disease. However, the mechanism of this structural transition remains unknown. In this work, we have shown that submicellar concentrations of sodium dodecyl sulfate (SDS) can provide a minimal platform for Aβ1–42 self-assembly. To further investigate the relation between Aβ1–42 structure and function, we analyzed peptide conformation and aggregation at various SDS concentrations using circular dichroism (CD), Fourier transform infrared spectroscopy, and gel electrophoresis. These aggregates, as observed via atomic force microscopy, appeared as globular particles in submicellar SDS with diameters of 35–60 nm. Upon sonication, these particles increased in disc diameter to 100 nm. Pyrene I 3/I 1 ratios and 1-anilinonaphthalene-8-sulfonic acid binding studies indicated that the peptide interior is more hydrophobic than the SDS micelle interior. We have also used Forster resonance energy transfer between N-terminal labeled pyrene and tyrosine (10) of Aβ1–42 in various SDS concentrations for conformational analysis. The results demonstrate that SDS at submicellar concentrations accelerates the formation of spherical aggregates, which act as niduses to form large spherical aggregates upon sonication. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

15.
Abstract

Speciation of binary complexes of Co(II), Ni(II) and Cu(II) with L-aspartic acid in (0-60% v/v) propylene glycol-water mixtures was studied pH metrically at 303.0±0.1 K and at an ionic strength of 0.16 mol L-1. The binary species refined were ML, ML2, ML2H2, ML2H3 and ML2H4. The stabilities of the complexes followed the Irving-Williams order i.e.Co(II) <Ni(II) < Cu(II). The linear variation of stability constants as a function of dielectric constant of the medium indicated the dominance of electrostatic forces over non-electrostatic forces. Some species were stabilised due to electrostatic interactions and some were destabilised due to the decreased dielectric constant. The order of ingredients influencing the magnitudes of stability constants due to incorporation of errors in their concentrations was alkali > acid > ligand > metal. Equilibria for the formation of binary complexes were proposed based on the forms of the ligand and their existence at different pH values.  相似文献   

16.
17.
Ten C-glycosyl β2- and β/β2-peptides with three to eight amino acid residues have been prepared. Solution and solid-phase peptide syntheses were employed to assemble β2-amino acids in which C-glycosylic substituents are attached to the C-2 position of β-amino acids. Conformational analysis of the C-glycosyl β2-peptides using NMR and CD spectra indicates that the tripeptide can form a helical secondary structure. Besides, helix directions of the C-glycosyl β/β2-peptides are governed by the configuration at the α-carbon of the peptide backbone that originates from the stereocenter of the C-glycosyl β2-amino acids.  相似文献   

18.
Upregulated gene 11 (URG11), recently identified as a new HBx-upregulated gene that may activate β-catenin and Wnt signaling, was found to be upregulated in a human tubule cell line under low oxygen. Here, we investigated the potential role of URG11 in hypoxia-induced renal tubular epithelial-to-mesenchymal (EMT). Overexpression of URG11 in a human proximal tubule cell line (HK2) promoted a mesenchymal phenotype accompanied by reduced expression of the epithelial marker E-cadherin and increased expression of the mesenchymal markers vimentin and α-SMA, while URG11 knockdown by siRNA effectively reversed hypoxia-induced EMT. URG11 promoted the expression of β-catenin and increased its nuclear accumulation under normoxic conditions through transactivation of the β-catenin promoter. This in turn upregulated β-catenin/T-cell factor (TCF) and its downstream effector genes, vimentin, and α-SMA. In vivo, strong expression of URG11 was observed in the tubular epithelia of 5/6-nephrectomized rats, and a Western blot analysis demonstrated a close correlation between HIF-1α and URG11 protein levels. Altogether, our results indicate that URG11 mediates hypoxia-induced EMT through the suppression of E-cadherin and the activation of the β-catenin/TCF pathway.  相似文献   

19.
Flexible and fully disordered protein regions that fold upon binding mediate numerous protein-protein interactions. However, little is known about their mechanism of interaction. One such coupled folding and binding occurs when a flexible region of neuronal nitric oxide synthase adopts a β-finger structure upon binding to its protein ligand, a PDZ [PSD-95 (postsynaptic density protein-95)/Discs large/ZO-1] domain from PSD-95. We have analyzed this binding reaction by protein engineering combined with kinetic experiments. Mutational destabilization of the β-finger changed mainly the dissociation rate constant of the proteins and, to a lesser extent, the association rate constant. Thus, mutation affected late events in the coupled folding and binding reaction. Our results therefore suggest that the native binding interactions of the β-finger are not present in the rate-limiting transition state for binding but form on the downhill side in a cooperative manner. However, by mutation, we could destabilize the β-finger further and change the rate-limiting step such that an initial conformational change becomes rate limiting. This switch in rate-limiting step shows that multistep binding mechanisms are likely to be found among flexible and intrinsically disordered regions of proteins.  相似文献   

20.
Despite the important functions of protein transmembrane domains, their structure and dynamics are often scarcely known. The SNARE proteins VAMP/synaptobrevin and syntaxin 1 are implicated in membrane fusion. Using different spectroscopic approaches we observed a marked sensitivity of their transmembrane domain structure in regard to the lipid/peptide ratio. In the dilute condition, peptides corresponding to the complete transmembrane domain fold into an α-helix inserted at ∼ 35° to the normal of the membranes, an observation in line with molecular simulations. Upon an increase in the peptide/lipid ratio, the peptides readily exhibited transition to β-sheet structure. Moreover, the insertion angle of these β-sheets increased to 54° and was accompanied by a derangement of lipid acyl chains. For both proteins the transition from α-helix to β-sheet was reversible under certain conditions by increasing the peptide/lipid ratio. This phenomenon was observed in different model systems including multibilayers and small unilamellar vesicles. In addition, differences in peptide structure and transitions were observed when using distinct lipids (DMPC, DPPC or DOPC) thus indicating parameters influencing transmembrane domain structure and conversion from helices to sheets. The putative functional consequences of this unprecedented dynamic behavior of a transmembrane domain are discussed.  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