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 46 毫秒
1.
Described here is a set of three-dimensional (3D) NMR experiments that rely on CACA-TOCSY magnetization transfer via the weak 3 \textJ\textCa\textCa ^{ 3} {\text{J}}_{{{\text{C}}\alpha {\text{C}}\alpha }} coupling. These pulse sequences, which resemble recently described 13C detected CACA-TOCSY (Takeuchi et al. 2010) experiments, are recorded in 1H2O, and use 1H excitation and detection. These experiments require alternate 13C-12C labeling together with perdeuteration, which allows utilizing the small 3 \textJ\textCa\textCa ^{ 3} {\text{J}}_{{{\text{C}}\alpha {\text{C}}\alpha }} scalar coupling that is otherwise masked by the stronger 1JCC couplings in uniformly 13C labeled samples. These new experiments provide a unique assignment ladder-mark that yields bidirectional supra-sequential information and can readily straddle proline residues. Unlike the conventional HNCA experiment, which contains only sequential information to the 1 3 \textCa ^{ 1 3} {\text{C}}^{\alpha } of the preceding residue, the 3D hnCA-TOCSY-caNH experiment can yield sequential correlations to alpha carbons in positions i1, i + 1 and i2. Furthermore, the 3D hNca-TOCSY-caNH and Hnca-TOCSY-caNH experiments, which share the same magnetization pathway but use a different chemical shift encoding, directly couple the 15N-1H spin pair of residue i to adjacent amide protons and nitrogens at positions i2, i1, i + 1 and i + 2, respectively. These new experimental features make protein backbone assignments more robust by reducing the degeneracy problem associated with the conventional 3D NMR experiments.  相似文献   

2.
Following Petoukhov and his collaborators, we use two length n zero-one sequences, α and β, to represent a length n genetic sequence ((a) || (b)){\alpha\choose\beta} so that the columns of ((a) || (b)){\alpha\choose\beta} have the following correspondence with the nucleotides: C ~ (0 || 0)C\sim{0\choose0} , U ~ (1 || 0)U\sim{1\choose0} , G ~ (1 || 1)G\sim{1\choose1} , A ~ (0 || 1)A\sim{0\choose1} . Using the Gray code ordering to arrange α and β, we build a 2 n ×2 n matrix C n including all the 4 n length n genetic sequences. Furthermore, we use the Hamming distance of α and β to construct a 2 n ×2 n matrix D n . We explore structures of these matrices, refine the results in earlier papers, and propose new directions for further research.  相似文献   

3.
The temporal profile of the phosphorescence of singlet oxygen endogenously photosensitized by photosystem II (PSII) reaction centre (RC) in an aqueous buffer has been recorded using laser excitation and a near infrared photomultiplier tube. A weak emission signal was discernible, and could be fitted to the functional form a[exp( - t/t2 ) - exp( - t/t1 )] a[\exp ( - t/\tau_{2} ) - \exp ( - t/\tau_{1} )] , with $ a > 0 $ a > 0 and $ \tau_{2} > \tau_{1} $ \tau_{2} > \tau_{1} . The value of t2 \tau_{2} decreased from 11.6 ± 0.5 μs under aerobic conditions to 4.1 ± 0.2 μs in oxygen-saturated samples, due to enhanced bimolecular quenching of the donor triplet by oxygen, whereas that of t1 \tau_{1} , identifiable with the lifetime of singlet oxygen, was close to 3 μs in both cases. Extrapolations based on the low amplitude of the emission signal of singlet oxygen formed by PSII RC in the aqueous buffer and the expected values of t1 \tau_{1} and t2 \tau_{2} in chloroplasts indicate that attempts to analyse the temporal profile of singlet oxygen in chloroplasts are unlikely to be rewarded with success without a significant advance in the sensitivity of the detection equipment.  相似文献   

4.
The difference equation f b :[0,1]–[0,1] defined by f b (x)=b x(1–x) is studied. In particular complete qualitative information is obtained for the parameter value b=3.83. For example the number of fixed points of (f b )i is given by
Ni = 1 + ( \frac1 + ?5 2 )i + ( \frac1 - ?5 2 )iN_i = 1 + \left( {\frac{{1 + \sqrt 5 }}{2}} \right)^i + \left( {\frac{{1 - \sqrt 5 }}{2}} \right)^i  相似文献   

5.
Molecular dynamics simulations of the biphalin molecule, (Tyr-D-Ala-Gly-Phe-NH)2, and the active tetrapeptide hydrazide, Tyr-D-Ala-Gly-Phe-NH-NH2 were performed to investigate the cause of the increased μ and δ receptor binding affinities of the former over the latter. The simulation results demonstrate that the acylation of the two equal tetrapeptide fragments of biphalin produces the constrained hydrazide bridges C4a - C4¢- N9 - N10 {\hbox{C}}_4^{\alpha } - {{\hbox{C}}_4}\prime - {{\hbox{N}}_9} - {{\hbox{N}}_{{10}}} and N9 - N10 - C5¢- C5a {{\hbox{N}}_9} - {{\hbox{N}}_{{10}}} - {{\hbox{C}}_5}\prime - {\hbox{C}}_5^{\alpha } , which in turn increase the opportunity of conformations for binding to μ or δ receptors. Meanwhile, the connection of the two active tetrapeptide fragments of biphalin also results in the constrained side chain torsion angle χ2 at one of the two residues Phe. This constrained side chain torsion angle not only significantly increases the δ receptor binding affinity but also makes most of the δ receptor binding conformations of biphalin bind to the δ receptor through the fragment containing the mentioned residue Phe.  相似文献   

6.
Mammalian metallothioneins ( \textM7\textIIMTs {\text{M}}_7^{\text{IIMTs}} ) show a clustered arrangement of the metal ions and a nonregular protein structure. The solution structures of Cd3-thiolate cluster containing β-domain of mouse β-MT-1 and rat β-MT-2 show high structural similarities, but widely differing structure dynamics. Molecular dynamics simulations revealed a substantially increased number of \textNH - \textSg {\text{NH - }}{{\text{S}}^\gamma } hydrogen bonds in β-MT-2, features likely responsible for the increased stability of the Cd3-thiolate cluster and the enfolding protein domain. Alterations in the \textNH - \textSg {\text{NH - }}{{\text{S}}^\gamma } hydrogen-bonding network may provide a rationale for the differences in dynamic properties encountered in the β-domains of MT-1, -2, and -3 isoforms, believed to be essential for their different biological function.  相似文献   

7.
The rhizobacteria Azospirillum brasilense Sp245 produce immunochemically different lipopolysaccharides LPSI and LPSII, both containing identical pentasaccharides built from D-rhamnose residues as the repeating units of O-specific polysaccharides (OPS). In this study, we report the structure of the OPS from A. brasilense LPSILPSII mutant Sp245.5, which spontaneously lost the p85 and p120 plasmids upon the formation of a new 300-MDa megaplasmid after the long-term storage of the bacteria in a rich medium. The repeating unit of the OPS of A. brasilense Sp245.5 appeared to be a disaccharide consisting of residues of N-acetyl-D-galactosamine and N-acetyl-D-mannosaminuronic acid:
$ \to 6) - \alpha - D - GalpNAc - (1 \to 4) - \beta - D - ManpNAcA - (1 \to $ \to 6) - \alpha - D - GalpNAc - (1 \to 4) - \beta - D - ManpNAcA - (1 \to   相似文献   

8.
Two partially reconstructed karyotypes (RK1 and RK2) of Arabidopsis thaliana have been established from a transformant, in which four structurally changed chromosomes (α, β, γ, and δ) were involved. Both karyotypes are composed of 12 chromosomes, 2n = 1¢¢+ 3¢¢+ 4¢¢+ 5¢¢+ a¢¢+ g¢¢ = 12 {2}n = {1}\prime \prime + {3}\prime \prime + {4}\prime \prime + {5}\prime \prime + \alpha \prime \prime + \gamma \prime \prime = {12} for RK1 and 2n = 3¢¢+ 4¢¢+ 5¢¢+ a¢¢+ b¢¢+ g¢¢ = 12 {2}n = {3}\prime \prime + {4}\prime \prime + {5}\prime \prime + \alpha \prime \prime + \beta \prime \prime + \gamma \prime \prime = {12} for RK2, and these chromosome constitutions were relatively stable at least for three generations. Pairing at meiosis was limited to the homologues (1, 3, 4, 5, α, β, or γ), and no pairing occurred among non-homologous chromosomes in both karyotypes. For minichromosome α (mini α), precocious separation at metaphase I was frequently observed in RK2, as found for other minichromosomes, but was rare in RK1. This stable paring of mini α was possibly caused by duplication of the terminal tip of chromosome 1 that is characteristic of RK1.  相似文献   

9.
Using primary hepatocytes in culture, various 2-acetamido-2-deoxy-D-glucose (GlcNAc) analogs were examined for their effects on the incorporation of D-[3H]glucosamine, [35S]sulfate, and L-[14C]leucine into cellular glycoconjugates. A series of acetylated GlcNAc analogs, namely methyl 2-acetamido-3,4,6-tri-O-acetyl-2-deoxy-α-(3) and β-D-glucopyranoside (4) and 2-acetamido-1,3,4,6-tetra-O-acetyl-2-deoxy-D-glucopyranose (5), exhibited a concentration-dependent reduction of D-[3H]glucosamine, but not of [35S]sulfate incorporation into isolated glycosaminoglycans (GAGs), without affecting L-[14C]leucine incorporation into total protein synthesis. These results suggest that analogs 3–5 exhibit an inhibitory effect on D-[3H]glucosamine incorporation into isolated GAGs by diluting the specific activity of cellular D-[3H]glucosamine and by competing for the same metabolic pathways. In the case of the corresponding series of 4-deoxy-GlcNAc analogs, namely methyl 2-acetamido-3,6-di-O-acetyl-2,4-dideoxy-α-(6) and β-D-xylo-hexopyranoside (7) and 2-acetamido-1,3,6-tri-O-acetyl-2,4-dideoxy-D-xylo-hexopyranose (8), compound 8 at 1.0 mM exhibited the greatest reduction of D-[3H]glucosamine and [35S]sulfate incorporation into isolated GAGs, namely to ∼7% of controls, and a moderate inhibition of total protein synthesis, namely to 60% of controls. Exogenous uridine was able to restore the inhibition of total protein synthesis by compound 8 at 1.0 mM. Isolated GAGs from cultures treated with compound 8 were shown to be smaller in size (∼40 kDa) than for control cultures (∼77 kDa). These results suggest that the inhibitory effects of compound 8 on cellular GAG synthesis may be mediated by the incorporation of a 4-deoxy moiety into GAGs resulting in premature chain termination and/or by its serving as an enzymatic inhibitor of the normal sugar metabolites. The inhibition of total protein synthesis from cultures treated with compound 8 suggests a uridine trapping mechanism which would result in the depletion of UTP pools and cause the inhibition of total protein synthesis. A 1-deoxy-GlcNAc analog, namely 2-acetamido-3,4,6-tri-O-acetyl-1,5-anhydro-2-deoxy-D-glucitol (9), also exhibited a reduction in both D -[3H]glucosamine and [35S]sulfate incorporation into isolated GAGs by 19 and 57%, of the control cells, respectively, at 1.0 mM without affecting total protein synthesis. The inability of compound 9 to form a UDP-sugar and, hence, be incorporated into GAGs presents another metabolic route for the inhibition of cellular GAG synthesis. Potential metabolic routes for each analog's effects are presented.  相似文献   

10.
A simple method of the synthesis of P 1-(11-phenoxyundecyl)-P 2-(2-acetamido-2-deoxy-α-D-galactopyranosyl) diphosphate, which is a synthetic lipid acceptor for glycosyl transferases participating in the biosynthesis of O-antigenic polysaccharides of Gram-negative bacteria, is suggested.  相似文献   

11.
The chlorophyll a-specific absorption coefficient ( a\textph* ( l) a_{\text{ph}}^{*} \left( \lambda \right) ) in a highly eutrophic lake can show characteristics distinct from that in the ocean due to the differences in the structure and composition of phytoplankton. In this study, investigated the variation of a\textph* ( l) a_{\text{ph}}^{*} \left( \lambda \right) in Lake Kasumigaura, a highly eutrophic lake in Japan, in association with the package effect and the effect of accessory pigments, and carried out the parameterization of a\textph* ( l) a_{\text{ph}}^{*} \left( \lambda \right) . Although a\textph* ( l) a_{\text{ph}}^{*} \left( \lambda \right) did not vary spatially, it did show significant temporal variation, with a particularly high value after spring-bloom. This high a\textph* ( l) a_{\text{ph}}^{*} \left( \lambda \right) in spring was attributed to a lower package effect and a higher proportion of carotenoid than the other samples. Although the value of a\textph* ( l) a_{\text{ph}}^{*} \left( \lambda \right) was correlated with the concentration of chlorophyll-a (Chl-a), the correlation coefficient was lower than those reported in the ocean. Some lake-water samples showed variations of the package effect and the effect of accessory pigments that were independent of the concentration of Chl-a, and these independent variations resulted in the weak correlation between a\textph* ( l) a_{\text{ph}}^{*} \left( \lambda \right) and the concentration of Chl-a. Together, these results suggest that the factors controlling a\textph* ( l) a_{\text{ph}}^{*} \left( \lambda \right) in highly eutrophic lakes are distinct from that in ocean samples.  相似文献   

12.
A facile synthesis method is described for transforming the reducing-end residue of chitooligosaccharides (DP 2–4) into lactone. The desired 4-O-β-N-acetylchitooligosyl lactones (GNnL) were conveniently prepared from chitooligosaccharides by consecutive dehydration and oxidation reactions to afford 4-O-β-tri-N-acetylchitotriosyl 2-acetamido-2,3-dideoxydidehydro-gluconolactone (GN3L), 4-O-β-di-N-acetylchitobiosyl 2-acetamido-2,3-dideoxydidehydro-gluconolactone (GN2L), and 4-O-β-2-acetamido-2-deoxy-D-glucopyranosyl 2-acetamido-2,3-dideoxydidehydro-gluconolactone (GNL). The resulting lactone derivatives exhibited considerable suppression (42.6–54.3% at a concentration of 400 µM) in umu gene expression of the SOS response in Salmonella typhimurium TA1535/pSK1002 against the mutagen, 2-(2-furyl)-3-(5-nitro-2-furyl)acrylamido (AF-2). Lactonization of the chitooligosaccharides was found to be essential for their suppression of the SOS-inducing activity.  相似文献   

13.
14.
Existing correlations of Power law consistency index with Penicillium chrysogenum biomass concentration and morphology were revised using a microscope magnification of 50 times to characterize the latter, rather than the 80 times used previously. This allowed tests of the correlations on broths of Aspergillus oryzae and Aspergillus niger, which have such large mycelial sizes that a lower magnification is required for accurate morphological analysis. The new correlations were successful at predicting the rheology of A. oryzae broths but not A. niger broths, which may be because of a change in the exponent on the biomass concentration in the correlations for the latter. Because the mean maximum dimension of clumps is magnification independent, the preferred correlation was:
K = C\textm2 ×[4×10-5 D-9 ×10-4] K = C_{\text{m}}^{2} \times \left[{4\times 10^{-5}\,D-9 \times 10^{-4}}\right]  相似文献   

15.
Host-guest principles are put into action advantageously by cyclodextrins that give remarkable sensor responses to halogenated hydrocarbons on mass-sensitive devices such as QMB (quartz micro balance) and SAW (surface acoustic wave) resonators. Modifications of the structure can tune the compound to different analytes and molecular modeling allows us to understand and predict specific host-guest interactions. The FT-IR analysis of CDCl3, incorporated into a partially methylated cyclodextrine, yields an astonishingly strong band shift of about 50 cm-1 to lower wave numbers, in contrast to the permethylated product that shows a displaced band of 10 cm-1 for the C-D stretching vibration. The explanation can be given by semiempirical methods and force field calculations. Two CDCl3 binding sites are revealed, one being a multicentered inclusion ([(n)\tilde] ~ 2195  cm-1{\tilde\nu}\sim 2195\;cm^{-1} of CDCl3 at the upper rim of the #-cyclodextrine cone, whereas the other is incorporated within the cavity ([(n)\tilde] ~ 2240  cm-1{\tilde\nu}\sim 2240\;cm^{-1}.  相似文献   

16.
Abstract Three kinds of trisaccharides were prepared by digesting fucoidan from the brown alga Kjellmaniella crassifolia, with the extracellular enzymes of the marine bacterium Fucobacter marina. Their structures were determined as Δ4,5GlcpUA1-2(L-Fucp(3-O-sulfate)α1-3)D-Manp, Δ4,5GlcpUA1-2(L-Fucp(3-O-sulfate)α1-3)D-Manp(6-O-sulfate), and Δ4,5GlcpUA1-2(L-Fucp(2,4-O-disulfate)α1-3)D-Manp(6-O-sulfate), which indicated the existence of a novel polysaccharide in the fucoidan and a novel glycosidase in the extracellular enzymes. In order to determine the complete structure of the polysaccharide and the reaction mechanism of the glycosidase, the fucoidan was partially hydrolyzed to obtain glucuronomannan, which is the putative backbone of the polysaccharide, and its sugar sequence was determined as (-4-D-GlcpUAβ1-2D-Manpα1-)n, which disclosed that the main structure of the polysaccharide is (-4-D-GlcpUAβ1-2(L-Fucp(3-O-sulfate)α1-3)D-Manpα1-)n. Consequently, the glycosidase was deduced to be an endo-α-D-mannosidase that eliminatively cleaves the α-D-mannosyl linkage between D-Manp and D-GlcpUA residues in the polysaccharide and produces the above trisaccharides. The novel polysaccharide and glycosidase were tentatively named as sulfated fucoglucuronomannan (SFGM) and SFGM lyase, respectively.  相似文献   

17.
An extracellular polysaccharide elaborated by a new species of Beijerinckia indica, named TX-1, was composed of D-glucose, L-fucose, D-glycero-D-manno-heptose, and D-glucuronic acid in a molar ratio of 5.0:1.0:2.0:0.9, in addition to 16.2% of the acetyl group. Among the polysaccharides of the Beijerinckia species, the present polysaccharide might be the first acidic type having an L-fucose residue. A methylation analysis, Smith degradation study and fragmentation analysis show that this polysaccharide consisted of non-reducing terminal D-glucose, O-4 substituted D-glucose, O-2 substituted D-glycero-D-manno-heptose, O-4 substituted D-glucuronic acid, O-3 and O-4 substituted D-glucose, and O-3 substituted L-fucose residues. A D-glucuronic acid residue was linked to the O-3 position of the L-fucose residue by an α-glycosidic linkage. Most of the D-glucose residues in the backbone chain were substituted at the O-3 position, with the side chain having non-reducing terminal D-glucose residues. It is suggested by the reaction with Con A that the anomeric configuration of the terminal D-glucose residues was β.  相似文献   

18.
Delipidated cell walls from Aureobasidium pullulans were fractionated systematically.

The cell surface heteropolysaccharide contains D-mannose, D-galactose, D-glucose, and D-glucuronic acid (ratio, 8.5:3.9:1.0:1.0). It consists of a backbone of (1→6)-α-linked D-mannose residues, some of which are substituted at O-3 with single or β-(1→6)-linked D-galactofuranosyl side chains, some terminated with a D-glucuronic acid residue, and also with single residues of D-glucopyranose, D-galactopyranose, and D-mannopyranose.

This glucurono-gluco-galactomannan interacted with antiserum against Elsinoe leucospila, which also reacted with its galactomannan, indicating that both polysaccharides contain a common epitope, i.e., at least terminal β-galactofuranosyl groups and also possibly internal β-(1→6)-linked galactofuranose residues.

It was further separated by DEAE-Sephacel column chromatography to gluco-galactomannan and glucurono-gluco-galactomannan.

The alkali-extracted β-D-glucan was purified by DEAE-cellulose chromatography to afford two antitumor-active (1→3)-β-D-glucans. One of the glucans (Mr, 1–2 × 105) was a O-6-branched (1→3)-β-D-glucan with a single β-D-glucosyl residue, d.b., 1/7, and the other (Mr, 3.5–4.5 × 105) had similar branched structure, but having d.b., 1/5. Side chains of both glucans contain small proportions of β-(1→6)-and β-(1→4)-D-glucosidic linkages.  相似文献   

19.
An acidic polysaccharide, termed gordonan, was isolated from the culture medium of Gordonia sp. as an inducer of cell aggregation in an insect cell line, BM-N4. Gordonan had an average molecular weight of 5×106 and its structure was identified as →3)-4-O-(1-carboxyethyl)-β-D-Manp-(1→4)-β-D-GlcAp-(1→4)-β-D-Glcp-(1→ mainly by acid hydrolysis experiments and NMR analysis. It induces cell aggregation at the concentration of 4 μg/ml. A partially hydrolyzed polysaccharide derived from gordonan with a molecular weight of 5×105 showed weak activity, while any fragment molecules with lower molecular weights prepared from gordonan showed no activity.  相似文献   

20.
The structure of the capsular polysaccharide of Type XIX Streptococcus pneumoniae (S-XIX) has been elucidated by 1H- and 13C-n.m.r. spectroscopy. Mild hydrolysis of S-XIX with acid yielded a major oligosaccharide, the repeating unit of S-XIX, which was shown to be O-2-acetamido-2-deoxy-β-d-mannopyranosyl-(1→4)-O-α-d-glucopyranosyl-(1→2)-l-rhamnose 4′′-phosphate. Phosphoric acid forms a diester linkage in the S-XIX molecule, which explains the instability of S-XIX towards acid or alkali. The phosphodiester linkages in S-XIX join HO-1 of α-l-rhamnose and HO-4 of the 2-acetamido-2-deoxy-d-mannopyranosyl residue in the next repeating-unit. Treatment of S-XIX with alkali or alkaline-NaBH4 produced the repeating units in a lower yield. The proposed structure of S-XIX is
  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