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 218 毫秒
1.
Galactomannans isolated from legume seed endosperms, including those of commercial interest, have been characterized by multidetection aqueous SEC. Galactomannans derived from seeds of the Faboideae subfamily had substantially higher Mw than those from Caesalpinioideae seeds (Mw,Fab = 2.4–3.1 × 106 g/mol, Mw,Caes. = 0.86–2.1 × 106 g/mol) and within the latter botanical subfamily, an apparent correlation between Mw and the degree of galactose substitution DG was found. The molar mass distributions were unimodal and differed primarily by a scale factor, with distributional widths narrower than a true Flory ‘most-probable distribution’; good fits to Schulz–Zimm model were obtained. Across subfamilies no differences were found in the exponents of [η]–M and RvM relationships (0.61 ± 0.02, 0.54 ± 0.01, respectively), the Flory chain stiffness ratio (C = 20 ± 1 (BSF analysis)), or the persistence length (Lp = 5.5 ± 0.2 nm) obtained from SEC fraction data. However, it was found that prefactors in the [η]–M and RvM relationships as well as the unperturbed parameter KΘ decrease in proportion to DG and therefore chain density. Generalized relationships incorporating galactose-dependent prefactors were therefore developed to model SEC fraction data of native galactomannans ([η]GM = (1800 ± 200) × Mo−1.61 × M0.61±0.02, Rv,GM = 0.63 ± 0.05 × Mo−0.54 × M0.54±0.01) as well as lower-M fractions obtained by ultrasonication ([η]GM = (730 ± 100) × Mo−1.71 × Mw0.71±0.02, Rv,GM = 0.49 ± 0.05 × Mo−0.57 × Mw0.57±0.01, M ≈ 1 × 105-native). As a consequence of this dependence and the observed patterns in molar mass variation, [η] varies within a narrow range for galactomannans as a whole despite substantial Mw differences.  相似文献   

2.
Xiaohua Wang  Lina Zhang   《Carbohydrate research》2009,344(16):2209-2216
Five fractions of lentinan, a β-(1→3)-d-glucan bearing β-(1→6)-d-glucopyranosyl branches, were treated with chlorosulfonic acid for 90 min at 60 °C in pyridine medium to synthesize water-soluble sulfated derivatives having the substitution degree of 1.44–1.76. The 13C NMR spectra of the sulfated β-glucans indicated that the C-6 position was preferentially substituted by the sulfate groups. The values of the weight-average molecular weight (Mw), radius of gyration (), and intrinsic viscosity ([η]) of the sulfated lentinan fractions were determined by size-exclusion chromatography with multi-angle laser light scattering (SEC–MALLS) and viscometry in 0.15 M aq NaCl at 25 °C, respectively. The dependence of [η] on Mw for the sulfated lentinan was found to be [η] = 8.93 × 10−3 (mL/g) in 0.15 M aq NaCl (for Mw ranging from 14.6 × 104 to 50.4 × 104). On the basis of the Yamakawa–Fujii–Yoshizaki (YFY) theory, the conformational parameters of the sulfated lentinan were calculated as 950 nm−1 for the molar mass per unit contour length (ML), 4.8 nm for the persistence length (q), and 13.9 for the characteristic ratio (C), indicating relatively extended single flexible chains in solution. The sulfated glucan fractions exhibited in vitro antiproliferative activities against sarcoma 180 (S-180) cells, and their inhibition ratios were lower than that of the triple-helix lentinan, but higher than that for the one with single random-coil lentinan chains.  相似文献   

3.
Soluble guanylate cyclase (sGC) has been purified from 100 L cell culture infected by baculovirus using the newer and highly effective titerless infected-cells preservation and scale-up (TIPS) method. Successive passage of the enzyme through DEAE, Ni2+-NTA, and POROS Q columns obtained approximately 100 mg of protein. The sGC obtained by this procedure was already about 90% pure and suitable for various studies which include high throughput screening (HTS) and hit follow-up. However, in order to obtain enzyme of greater homogeneity and purity for crystallographic and high precision spectroscopic and kinetic studies of sGC with select stimulators, the sGC solution after the POROS Q step was further purified by GTP-agarose affinity chromatography. This additional step led to the generation of 26 mg of enzyme that was about 99% pure. This highly pure and active enzyme exhibited a Mr = 144,933 by static light scattering supportive of a dimeric structure. It migrated as a two-band protein, each of equal intensity, on SDS–PAGE corresponding to the α (Mr 77,000) and β (Mr 70,000) sGC subunits. It showed an A430/A280 = 1.01, indicating one heme per heterodimer, and a maximum of the Soret band at 430 nm indicative of a penta-coordinated ferrous heme with a histidine as the axial ligand. The Soret band shifted to 398 nm in the presence of an NO donor as expected for the formation of a penta-coordinated nitrosyl-heme complex. Non-stimulated sGC had kcat/Km = 1.7 × 10−3 s−1 μM−1 that increased to 5.8 × 10−1 s−1 μM−1 upon stimulation with an NO donor which represents a 340-fold increase due to stimulation. The novel combination of using the TIPS method for co-expression of a heterodimeric heme-containing enzyme, along with the application of a reproducible ligand affinity purification method, has enabled us to obtain recombinant human sGC of both the quality and quantity needed to study structure–function relationships.  相似文献   

4.
The conformation and dilute solution properties of (2→1)-β-d-fructan in aqueous solution were studied by gel permeation chromatography, low-angle laser light-scattering photometry, viscometry, small-angle X-ray scattering and electron microscopy. Fractions covering a broad range of weight-average molecular weights (Mw) from 1.49 × 104 to 5.29 × 106 were obtained from a native sample by ultrasonic degradation and fractional precipitation. For Mw < 4 × 104, the intrinsic viscosity [η] varies with Mw0.71, indicating that the fructan chain behaves as a random coil expanded by an excluded-volume effect in this molecular weight region. For Mw > 105, [η] exhibits an unusually weak dependence on Mw and finally becomes almost independent of molecular weight. This behaviour is interpreted in terms of a globular conformation of the high-molecular-weight fructan molecules. Small-angle X-ray-scattering measurements and electron microscopic observations support this interpretation of the values of [η] observed.  相似文献   

5.
Dietary chitosans with different molecular weight Mw and the degree of deacetylation DDA (high molecular weight chitosan HCS with Mw 7.60 × 105 and DDA 85.5%, middle molecular weight chitosan MCS with Mw 3.27 × 104 and DDA 85.2%, chito-oligomer COS with Mw 0.99 × 103 and DDA 85.7% and water-soluble chitosan WSC with Mw 3.91 × 104 and DDA 52.6%) were used at the 1.05% level to feed mice for 90 days. Afterwards no pathological symptoms, clinical signs or deaths were observed. The body weight of mice in chitosan group and control group showed no significant difference. Although HCS, COS and WSC had no significant effect on the level of Fe, Zn and Cu in the tested mice’s liver, spleen, heart and kidney, MCS significantly increased the level of Fe, Zn and Cu in liver. Therefore dietary ingestion of chitosan did not depress the level of Fe, Zn and Cu in mice.  相似文献   

6.
In cultured bovine adrenal chromaffin cells expressing Nav1.7 sodium channel isoform, veratridine increased Ser473-phosphorylation of Akt and Ser9-phosphorylation of glycogen synthase kinase-3β by 217 and 195%, while decreasing Ser396-phosphorylation of tau by 36% in a concentration (EC50 = 2.1 μM)- and time (t1/2 = 2.7 min)-dependent manner. These effects of veratridine were abolished by tetrodotoxin or extracellular Ca2+ removal. Veratridine (10 μM for 5 min) increased translocation of Ca2+-dependent conventional protein kinase C-α from cytoplasm to membranes by 47%; it was abolished by tetrodotoxin, extracellular Ca2+ removal, or Gö6976 (an inhibitor of protein kinase C-α), and partially attenuated by LY294002 (an inhibitor of phosphatidylinositol 3-kinase). LY294002 (but not Gö6976) abrogated veratridine-induced Akt phosphorylation. In contrast, either LY294002 or Gö6976 alone attenuated veratridine-induced glycogen synthase kinase-3β phosphorylation by 65 or 42%; however, LY294002 plus Gö6976 completely blocked it. Veratridine (10 μM for 5 min)-induced decrease of tau phosphorylation was partially attenuated by LY294002 or Gö6976, but completely blocked by LY294002 plus Gö6976; okadaic acid or cyclosporin A (inhibitors of protein phosphatases 1, 2A, and 2B) failed to alter tau phosphorylation. These results suggest that Na+ influx via Nav1.7 sodium channel and the subsequent Ca2+ influx via voltage-dependent calcium channel activated (1) Ca2+/protein kinase C-α pathway, as well as (2) Ca2+/phosphatidylinositol 3-kinase/Akt and (3) Ca2+/phosphatidylinositol 3-kinase/protein kinase C-α pathways; these parallel pathways converged on inhibitory phosphorylation of glycogen synthase kinase-3β, decreasing tau phosphorylation.  相似文献   

7.
The molecular mass and size of five water-soluble polysaccharides isolated from Rhizoma Panacis Japonici (RPJ) were determined with laser light scattering (LLS), size-exclusion chromatography (SEC) combined with LLS (SEC–LLS), dynamic light scattering (DLS), as well as transmission electron microscope (TEM). Their weight-average molecular masses (Mw) were 3.5 × 104, 1.47 × 105, 1.24 × 106, 9.26 × 105 and 1.36 × 106, radii of gyration (<s2>z1/2) were 14.7, 31.7, 50.8, 41.8 and 40.4 nm, and hydrodynamic radii (Rh) were 19.9, 37.5, 66.2, 52.1, and 55.2 nm, respectively. The results showed that molecular masses and sizes of the polysaccharides were influenced by the pH and temperature of the extraction mediums. The conformation parameters were calculated from the above data according to polymer solution theory. The values of ρ (= <s2>z1/2/Rh) were from 0.7 to 0.8, exponents (ν) of <s2>z1/2 = k Mwν were from 0.31 to 0.43, and fractal dimension (df) were from 2.3 to 3.2, respectively. The results revealed that all of the polysaccharides existed as spheres in 0.15 M NaCl aqueous solution. TEM and atomic force microscope (AFM) further confirmed the spherical morphologies of these molecules. The spherical conformations of the polysaccharides were a result of their highly branched structures.  相似文献   

8.
A water-insoluble (1 → 3)-β-d-glucan (PCSG) isolated from the fresh sclerotium of Poria cocos was carboxymethylated to afford a water-soluble derivative coded as C-PCSG. The carboxymethylated (1 → 3)-β-d-glucan was fractionated to obtain eight fractions according to the nonsolvent addition method. The weight-average molecular mass (Mw), radius of gyration and intrinsic viscosity ([η]) of the fractions were determined by size-exclusion chromatography combined with laser light scattering (SEC-LLS) and viscometry in 0.2 M NaCl aqueous solution at 25 °C. The dependences of [η] and on Mw for C-PCSG were found to be , and (nm), respectively. Analysis of Mw and [η] in terms of the known theories for wormlike chain model yielded 633 nm−1 for molar mass per unit contour length (ML), 5.5 nm for persistence length (q), and 20.2 for characteristic ratio (C). These results indicated that C-PCSG exists as a relatively extended flexible chain in 0.2 M NaCl aqueous solution. Therefore, the introduction of the carboxymethyl groups into the β-glucan improved significantly the water solubility and enhanced the stiffness of the chains.  相似文献   

9.
High yield purification of soluble guanylate cyclase from bovine lung   总被引:1,自引:1,他引:0  
Soluble guanylate cyclase (sGC), the main target of nitric oxide (NO), is a cytosolic, heme-containing, heterodimeric enzyme that catalyzes the conversion of guanosine 5′-triphosphate (GTP) to 3,5′-cyclic guanosine monophosphate (cGMP) and pyrophosphate (PPi) in the presence of Mg2+. Cyclic GMP is then involved in transmitting the NO activating signals to a variety of downstream effectors such as cyclic-nucleotide-gated channels, protein kinases, and phosphodiesterases. In this work, sGC has been purified from bovine lung. The lungs were subjected to grinding and extraction with buffer at physiological pH followed by centrifugation. The resulting solution was subjected to successive column chromatography on DEAE- and Q-Sepharose, Ceramic Hydroxyapatite, Resource Q, and GTP–agarose. The purified enzyme migrated as a two-band protein on SDS–PAGE corresponding to sGC subunits α (Mr = 77,532) and β (Mr = 70,500) and had an A280nm/A430nm of 1 indicating one heme per heterodimer. The yield of enzyme was 8–10 mg from 4 to 5 kg bovine lungs. Vmax and Km of non-stimulated sGC were 22 nmol/mg/min and 180 μM, respectively. Upon stimulation with the NO donor 3-ethyl-3-(ethylaminoethyl)-1-hydroxy-2-oxo-1-triazene, the Vmax increased to 1330 nmol/mg/min while the Km dropped to 43 μM. The quality and quantity of enzyme make it suitable for studies to probe the structure and catalytic mechanism of this enzyme and for research related to drug discovery.  相似文献   

10.
(1→3)-β-d-Glucan isolated from Poria cocos was phosphorylated to obtain a series of phosphorylated derivatives. Their structures, weight-average molecular weights (Mw), and chain conformation were studied by 13C NMR, 31P NMR, static laser light scattering and viscometry. The experimental results revealed that the phosphorylated glucan existed as relatively extended flexible chain in 0.15 M NaCl aqueous solution, and exhibited relatively strong inhibition against S-180 tumor cell in vitro and in vivo. In vivo, the fractions with relatively high molecular weight at low dosage exhibited stronger anti-tumor activities. The results revealed that the molecular weights and molecular conformation could influence the anti-tumor activities. The molecular weight ranging from 2.6 × 104 to 26.8 × 104 and the extended chain conformation were beneficial to enhance the anti-tumor activity, as a result of the increasing of the interaction between polysaccharide and immune system.  相似文献   

11.
Two kinds of water-insoluble (1 → 3)-α-d-glucan samples, ab-PCM3-I and ac-PCM3-I, isolated from different Poria cocos mycelia were sulfated, to produce two series of water-soluble derivatives ab-PCM3-I-S1–S5 and ac-PCM3-I-S1–S5, respectively. The derivatives having different weight-average molecular mass (Mw) were produced by changing reaction temperature and time as well as molar ratios between chlorosulfonic acid and number of hydroxyl groups in the glucan. The degrees of substitution (DS) of the sulfated derivatives were analyzed by elemental analysis (EA) to be 0.39–0.67 for ab-PCM3-I-S and 0.73–0.96 for ac-PCM3-I-S, respectively. The Mw and the intrinsic viscosity ([η]) of the samples ab-PCM3-I-S and the ac-PCM3-I-S were measured by size exclusion chromatography combined with laser light scattering (SEC–LLS) and viscometry in phosphate buffer solution (PBS) at 37 °C. The results indicated that their Mw ranged from 2.0 to 11.3 × 104 for the samples ab-PCM3-I-S, and 4.7 to 40.0 × 104 for the samples ac-PCM3-I-S. Moreover, the antitumor activities of the sulfated derivatives ab-PCM3-I-S and ac-PCM3-I-S against Sarcoma 180 tumor cell tested both in vitro and in vivo are significantly higher than those of the native α-d-glucans. Therefore, a moderate range of molecular mass from 2.0 × 104 to 40.0 × 104, relatively high chain stiffness and good water solubility of the sulfated derivatives are beneficial to the enhancement of their antitumor activities.  相似文献   

12.
Rapid method to determine the molecular weight of dextrins and dextrans   总被引:1,自引:0,他引:1  
A rapid method was developed to determine the molecular weight (Mn) of β-limit dextrin and dextrans (Leuconostoc mesenteroides) using a reducing power approach. The Mn of the β-limit dextrin was also estimated from high performance liquid chromatography (HPLC). Chromatograms were pre-calibrated with the dextrans. The three dextrins had a Mn of 2.09, 2.40 and 2.63 × 105 using the reducing method and 4.80, 5.90 and 2.80 × 105 by HPLC. The method could be employed to estimate Mn of dextrins where chromatographic systems were not available.  相似文献   

13.
A biotinylated mannotriose (Man3-bio) was dispersively immobilized in the matrix of biotinylated lactose (Gal-Glc-bio) on a streptavidin-covered, 27-MHz quartz crystal microbalance (QCM), and binding kinetics of concanavalin A (Con A) to Man3-bio in the Gal-Glc-bio matrix could be obtained from frequency decreases (mass increases) of the QCM. Association constants (Ka) and binding and dissociation rate constants (kon and koff) could be determined separately as the 1:1 and 1:2 bindings of Con A to Man3-bio on the surface. When Man3-bio was immobilized with content of 1 to 5 mol% in the matrix, the 1:1 binding of Con A to Man3-bio was obtained as Ka = (4 ± 1) × 106 M−1, kon = (4 ± 1) × 104 M−1 s−1, and koff = (12 ± 2) × 10–3 s−1. On the contrary, when Man3-bio was immobilized with content of 20 to 100 mol% in the matrix, the 1:2 binding of Con A to Man3-bio was obtained as Ka = (14 ± 2) × 106 M−1, kon = (14 ± 2) × 104 M−1 s−1, and koff = (7 ± 2) × 10–3 s−1. Thus, Ka for the 1:2 binding was 10 times larger than that for the 1:1 binding, with a three times larger binding rate constant (kon) and a three times smaller dissociation rate constant (koff). This is the first example to obtain separate kinetic parameters for the 1:1 and 1:2 bindings of lectins to carbohydrates on the surface.  相似文献   

14.
The 31P nuclear magnetic resonance (NMR) characteristics, toxicity, and cellular penetration of five linear or cyclic α-aminophosphonate highly sensitive pH probes were investigated in Dictyostelium discoideum cells and isolated rat hearts and were compared with three phosphonic acid derivatives. The line width broadening at pH pKa, which was satisfactorily modelized for all compounds, was significantly limited in biological milieu for the new markers, affording a four- to sixfold better accuracy in pH determination. Cellular uptake or washout of nontoxic concentrations (<15 mM) of α-aminophosphonates occurred by rapid passive permeation, whereas standard probes required a much slower fluid-phase pinocytosis and transport processes that could ultimately lead to trapping. Using mild concentrations (<4 mM) three α-aminophosphonates having 6 < pKa < 7 allowed an easy and simultaneous 31P NMR determination of cytosolic, acidic, and extracellular compartments in anoxic–reoxygenated or starving D. discoideum.  相似文献   

15.
16.
We investigated the effects of reducing sarcoplasmic reticular (SR) Ca2+ stores using the Ca2+-ATPase inhibitor cyclopiazonic acid (CPA) in Langendorff-perfused mouse hearts exposed to different pro-arrhythmic agents all known to produce Ca2+-mediated arrhythmogenesis. CPA (100 and 150 nM) produced progressive (beginning over 1 min) and significant (P < 0.0001) reductions in peak amplitudes of Ca2+ transients evoked by regular stimulation in isolated Fluo-3 loaded myocytes from F/F0 = 3.2 ± 0.16 (n = 12 cells) to 1.62 ± 0.012 (n = 6 cells) and 1.53 ± 0.06 (n = 12 cells), respectively, consistent with previous reports describing reductions of store Ca2+ in other cell systems. The corresponding effects of CPA were then examined in intact hearts exposed to isoproterenol (100 nM), elevated extracellular [Ca2+] (5 mM) and caffeine (1 mM). All three agents produced ventricular tachycardia either when added alone or simultaneously with CPA during programmed electrical stimulation. However, arrhythmogenicity was not observed when such agents were added 10 min after introduction of CPA. CPA thus antagonized this Ca2+-mediated arrhythmogenesis but only under circumstances of SR Ca2+ depletion. These alterations in arrhythmogenic tendency took place despite an absence of alterations in electrogram and monophasic action potential characteristics. This was in sharp contrast to previous observations in murine, ΔKPQ-Scn5a (LQT3) and KCNE1−/− (LQT5), systems where re-entry has been implicated in arrhythmogenesis.  相似文献   

17.
In this study, we show that boronates, a class of synthetic organic compounds, react rapidly and stoichiometrically with peroxynitrite (ONOO) to form stable hydroxy derivatives as major products. Using a stopped-flow kinetic technique, we measured the second-order rate constants for the reaction with ONOO, hypochlorous acid (HOCl), and hydrogen peroxide (H2O2) and found that ONOO reacts with 4-acetylphenylboronic acid nearly a million times (k = 1.6 × 106 M− 1 s− 1) faster than does H2O2 (k = 2.2 M− 1 s− 1) and over 200 times faster than does HOCl (k = 6.2 × 103 M− 1 s− 1). Nitric oxide and superoxide together, but not alone, oxidized boronates to the same phenolic products. Similar reaction profiles were obtained with other boronates. Results from this study may be helpful in developing a novel class of fluorescent probes for the detection and imaging of ONOO formed in cellular and cell-free systems.  相似文献   

18.
Decomposition rates of Phragmites australis, Carex riparia, Nuphar luteum and Salvinia natans and benthic processes were measured from December 2003 to December 2004 in a shallow wetland (Paludi di Ostiglia, Northern Italy) by means of litter bags and intact cores incubations. Decay rate was highest for N. luteum (k = 0.0152 d−1), intermediate for S. natans (k = 0.0041 d−1) and similar for P. australis (k = 0.0027 d−1) and C. riparia (k = 0.0028 d−1).Benthic metabolism followed a seasonal pattern with summer peaks of O2 demand and TCO2, CH4 and NH4+ efflux whilst soluble reactive phosphorus (SRP) fluxes were negligible also under hypoxic conditions, indicating that P was mainly retained by sediment. The initial C:P ratio was similar in N. luteum and S. natans (170) and significantly lower than that of P. australis and C. riparia (360). During the detritus decay P was progressively lost by N. luteum and S. natans tissues, whereas, after an initial leaching, it was probably re-used during the microbial decomposition of the more refractory P. australis and C. riparia detritus. Nuphar luteum, P. australis and S. natans had comparable initial C:N mass ratio (15), significantly lower than that of C. riparia (26). The C:N ratio was rather constant for N. luteum (12.9 ± 1.5) and S. natans (14.6 ± 0.9), decreased slightly to below 20 for C. riparia and increased up to 30 for P. australis. Overall, differences among species were likely due to the recalcitrance of decomposing detritus, whilst process rates were controlled by limitation of microbial processes by nutrients and electron acceptor availability.  相似文献   

19.
A novel polysaccharide designated EPS-1A with an average molecular weight around 40 kDa was fractionated and purified by anion-exchange and gel-filtration chromatography from the crude exopolysaccharide (EPS) isolated from fermentation broth of Cs-HK1, a Tolypocladium sp. fungus isolated from wild Cordyceps sinensis. The structural characteristics of EPS-1A were determined with various methods (e.g. GC, GC–MS, FT-IR, 1H NMR and 13C NMR) and through acid hydrolysis, methylation, periodate-oxidation and Smith degradation. The results suggested that EPS-1A was composed of glucose, mannose and galactose at 15.2:3.6:1.0 M ratio. EPS-1A was a slightly branched polysaccharide and its backbone was composed of (1 → 6)-α-d-glucose residues (77%) and (1 → 6)-α-d-mannose residues (23%). Branching occurred at O-3 position of (1 → 6)-α-d-mannose residues of the backbone with (1 → 6)-α-d-mannose residues and (1 → 6)-α-d-glucose residues, and terminated with β-d-galactose residues.  相似文献   

20.
A complex trisaccharide β-d-GalpNAcA-(1 → 4)-β-d-GlcpNAc-(1 → 4)-d-ManpNAc (3) was prepared in a good yield (35%) in a transglycosylation reaction catalyzed by β-N-acetylhexosaminidase from Talaromyces flavus using p-nitrophenyl 2-acetamido-2-deoxy-β-d-galacto-hexodialdo-1,5-pyranoside (1) as a donor followed by the in situ oxidation of the aldehyde functionality by NaClO2. The disaccharide β-d-GlcpNAc-(1 → 4)-d-ManpNAc (2) was used as galactosyl acceptor. A disaccharide β-d-GalpNAcA-(1 → 4)-d-GlcpNAc (4; 39%) originated as a by-product in the reaction. Oligosaccharides comprising a carboxy moiety at C-6 are shown to be very efficient ligands to natural killer cell activation receptors, particularly to human receptor CD69. Thus, oxidized trisaccharide 3 is the best-known oligosaccharidic ligand to this receptor, with IC50 = 2.5 × 10−9 M. The presented method of introducing a β-d-GalpNAcA moiety into carbohydrate structures is versatile and can be applied in the synthesis of other complex oligosaccharides.  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