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.
Physicochemical characterization of konjac glucomannan   总被引:1,自引:0,他引:1  
Four commercial konjac glucomannan (KGM) samples and a glucomannan derived from yeast were characterized by aqueous gel permeation chromatography coupled with multi angle laser light scattering (GPC-MALLS). Disaggregation of aqueous glucomannan solutions through controlled use of a microwave bomb facilitated reproducible molar mass distribution determination alleviating the need for derivatization of the polymer or the use of aggressive solvents. Further characterization was undertaken by use of capillary viscometry and photon correlation spectroscopy (PCS). The weight average molecular masses (M(w)) determined were in the region of 9.0 +/- 1.0 x 10(5) g mol(-1) for KGM samples and 1.3 +/- 0.4 x 10(5) g mol(-1) for the yeast glucomannan. The values determined for KGM in aqueous solution are in agreement with those reported for KGM in aqueous cadoxen. The degradation of samples observed upon autoclaving has been quantified by GPC-MALLS and intrinsic viscosity determination, allowing comparison with reported Mark-Houwink parameters. Shear flow experiments were undertaken for a range of KGM solutions of concentration 0.05 to 2.0% using a combination of controlled stress and controlled strain rheometers. The concentration dependence of the zero shear specific viscosity was determined by analysis of the data using the Ellis model. The dependence of the zero shear specific viscosity on the coil overlap parameter was defined and interpretation discussed in terms of the Martin and Tuinier equations.  相似文献   

2.
Hyaluronan (HA) has various biological functions that are strongly dependent on its chain length. In some cases, as in inflammation and angiogenesis, long and short chain-size HA effects are antagonistic. HA hydrolysis catalyzed by hyaluronidase (HAase) is believed to be involved in the control of the balance between longer and shorter HA chains. Our studies of native HA hydrolysis catalyzed by bovine testicular HAase have suggested that the kinetic parameters depend on the chain size. We thus used HA fragments with a molar mass ranging from 8x10(2) g mol(-1) to 2.5x10(5) g mol(-1) and native HA to study the influence of the chain length of HA on the kinetics of its HAase-catalyzed hydrolysis. The initial hydrolysis rate strongly varied with HA chain length. According to the Km and Vm/Km values, the ability of HA chains to form an efficient enzyme-substrate complex is maximum for HA molar masses ranging from 3x10(3) to 2x10(4) g mol(-1). Shorter HA chains seem to be too short to form a stable complex and longer HA chains encounter difficulties in forming a complex, probably because of steric hindrance. The hydrolysis Vm values strongly suggest that as the chain length decreases the HAase increasingly catalyses transglycosylation rather than hydrolysis. Finally, two HA chain populations, corresponding to HA chain molar masses lower and higher than approximately 2x10(4) g mol(-1), are identified and related to the bi-exponential character of the model we have previously proposed to fit the experimental points of the kinetic curves.  相似文献   

3.
Nine hyaluronan (HA) samples were fractionated by size-exclusion chromatography, and molar mass (M), radius of gyration (Rg), and intrinsic viscosity ([eta]) were measured in 0.15 M NaCl at 37 degrees C by on-line multiangle light scattering and viscometer detectors. Using such method, we investigated the Rg and [eta] molar mass dependence for HA over a very wide range of molar masses: M ranging from 4 x 10(4) to 5.5 x 10(6) g/mol. The Rg and the [eta] molar mass dependence found for HA showed a meaningful difference. The Rg = f(M) power law was substantially linear in the whole range of molar masses explored with a constant slope of 0.6. In contrast, the [eta] = f(M) power law (Mark-Houwink-Sakurada plot) showed a marked curve shape, and a linear regression over the whole range of molar masses does not make sense. Also the persistence length (stiffness) for HA was estimated. The persistence length derived by using both the Odijk's model (7.5 nm from Rg vs M data) and the Bohdanecky's plot (6.8 nm from [eta] vs M data) were quite similar. These persistence length values are congruent with a semistiff conformation of HA macromolecules.  相似文献   

4.
Polyelectrolyte complexes of a synthetic polycation with either a genomic DNA or a synthetic poly(oxyethylene-block-sodium methacrylate), POE-b-PMANa, have been studied in aqueous solutions as a function of cation:anion ratio, the degree of polymerization of the polycation, the ionic strength, and temperature using dynamic light scattering and turbidity measurements. The polycation was a copolymer of methacryl oxyethyl trimethylammonium chloride and poly(oxyethylene) monomethyl ether monomethacrylate with 4-5 oxyethylene repeating units, PMOTAC-g-POE. The molar masses of the polycations in a homological series were 0.3, 0.9, and 2.1 x 10(6) g/ mol. The amount of comonomers with poly(oxyethylene) tails in the copolymers was 15 mol %. The molar mass of the POE-b-PMANa was 75000 g/mol and that of the POE-block was 5000 g/mol. The molar mass of the polycation was shown to have a dramatic effect on the stability and size of the complexes formed by either of the polyanions. An increase in the polycation molar mass shifts the cloud point toward the lower polycation content in the complexes, and a macro phase separation occurs in the solutions with the cation to anion molar ratios much below than 1:1. Increasing the ionic strength has a similar effect. Further addition of salt to turbid and phase-separated solutions results in dissociation of the complexes, and the polyions dissolve as individual macromolecules. The effect of POE on the stability of polyelectrolyte complexes is discussed as well.  相似文献   

5.
Characterization of poly-3-hydroxybutyric acid (PHB) and poly-3-hydroxybutyric-co-valeric acid (PHBV, 13% valerate) in chloroform was performed using size exclusion chromatography coupled to a multi-angle light scattering detector (SEC-MALS). Absolute molar mass averages, molar mass distribution, and the radius of gyration were determined. Three sample preparation methods were examined: dissolution in chloroform (1) at room temperature, (2) at 60 degrees C, and (3) after thermal pretreatment of samples (annealing at 180 degrees C with subsequent quenching in liquid nitrogen). Dissolution at 60 degrees C and dissolution of thermally pretreated samples gave molecularly dissolved PHB and PHBV. At 60 degrees C using acid free chloroform, there was no indication of degradation for up to 120 min dissolution time, whereas thermal degradation of polymers did take place during annealing at 180 degrees C. The degradation rate constants for number and weight average degree of polymerization at 180 degrees C were slightly higher for PHB (5.19 x 10(-5) min(-1), 4.95 x 10(-5) min(-1)) than for PHBV (4.99 x 10(-5) min(-1), 4.54 x 10(-5) min(-1)). The dependence of the radii of gyration on molar mass showed that both polymers form random coils in chloroform. The relationship between the absolute molar masses and relative SEC results was determined. DSC and NMR characterization also gave evidence of the progress of degradation.  相似文献   

6.
Bovine serum albumin (BSA)-dextran conjugates were prepared by using the Maillard reaction; depending on the ratio of dextran to BSA used, about 0.5-1 mol of dextran could be bound to 1 mol of native BSA. SDS-PAGE patterns revealed that BSA and dextran had been covalently bonded. Structural analyses by fluorescence spectroscopy and circular dichroism indicated that the BSA surface in each conjugate was covered with dextran without any great disruption of the native conformation. The conjugates could be grouped into two fractions on the basis of the weight-average molecular mass measured: the main fraction at 1.95-2.35 x 10(5) g/mol and a less-abundant fraction with aggregates greater than 1.50 x 10(6) g/mol. High-performance size-exclusion chromatography in conjunction with multi-angle laser light scattering detection revealed that the BSA-dextran conjugates prepared by using the Maillard reaction had various molar masses and radii.  相似文献   

7.
Two different ethyl(hydroxyethyl) cellulose (EHEC) samples were characterized by size-exclusion chromatography (SEC) with multiangle light scattering (MALS) detection and high-performance anion-exchange chromatography (HPAEC) with pulsed amperometric detection (PAD). The aim of the study was to investigate the molar mass distribution and the heterogeneity of the substituent distribution, factors that are thought to affect the functional properties of EHEC. The presence of blocks of unsubstituted glucose units was studied by enzymic degradation of EHEC by two different endoglucanases from Trichoderma reesei. The SEC-MALS analysis of the hydrolysis products showed that both enzymes were strongly inhibited by the large number of substituents along the cellulose chain. However, as the weight-average molar mass was reduced from approximately 360,000 to 80,000 g/mol in one of the polymers and from 770,000 to 60,000 g/mol in the other polymer, it was suggested that both samples were composed of some unsubstituted regions where the enzymes got access to the glucosidic bonds. The amount of glucose released upon endoglucanase hydrolysis was determined by HPAEC-PAD, which gave information on the homogeneity of the substituent distribution. The production of unsubstituted glucose units indicated that one of the polymers had a more uneven distribution compared with the other. It was demonstrated that chemical characterization of EHEC is a complex task, which requires an analytical approach involving numerous different methods and techniques.  相似文献   

8.
Copaifera langsdorfii (Desf.) Kuntze (copaiba) seeds are abundantly produced and have not yet been characterized. The seed oil presents a characteristic odor of coumarin (250.1+/-6.57 mg/g determined through LC). The fatty acid composition of the oil was determined through CG/FID, being 45.3% linoleic acid, 32.3% monounsaturated, and 22.4% saturated fat. For the lipid-free seeds, the total carbohydrate, protein and moisture were 75.4%, 6.8% and 14.8%, respectively. The C. langsdorfii xyloglucan had an intrinsic viscosity of 804 mL/g, and the average molar mass (Mw) was 7.82 x 10(5)g/mol and Rg of 65 nm. The degree of polydispersion was 1.7, indicating the polydisperse family of polysaccharides. Its homogeneity, low degree of polymer contaminants and high intrinsic viscosity and molecular mass, represent good potential as a thickening agent. The presence of coumarin and xyloglucan as major components of C. langsdorfii seeds denotes its potential for use in the cosmetic or pharmaceutical industries.  相似文献   

9.
Hardy LW  Kirsch JF 《Biochemistry》1984,23(6):1275-1282
The Bacillus cereus beta-lactamase I catalyzes the hydrolysis of a wide variety of penicillins and cephalosporins with values of k(cat)/K(m) varying over several orders of magnitude. The values of this parameter for the most reactive of these compounds, benzylpenicillin, I, and furylacryloyl-penicillin, II (k(cat)/K(m) = 2.43 x 10(7) M(-1) s(-1) and 2.35 x 10(7) M(-1) s(-1), respectively, at pH 7.0 in potassium phosphate buffer containing 0.17 M KCl, I(c) = 0.63, 25 degrees C) are decreased markedly by increasing viscosity in sucrose- or glycerol-containing buffers. The relative sensitivities to viscosity of k(cat)/K(m) values for I and for cephaloridine, III, were found to be virtually unchanged at pH 3.8 from those observed at pH 7.0. The differential effects of viscosity on the reactive vs. the sluggish [e.g., cephalothin (IV), k(cat)/K(m) = 1 x 10(4) M(-1) s(-1)] substrates support the contention that the rates of reaction of the former with the enzyme are in part diffusion controlled. Quantitative analysis gives values for the association rate constants, k(1), of 7.6 x 10(7) M(-1) s(-1), 4 x 10(7) M(-1) s(-1), and 1.1 x 10(7) M(-1) s(-1) for I, II, and III, respectively. As both reactive and sluggish substrates associate with the active site of the enzyme with relatively similar rate constants, the variation in k(cat)/K(m) values is primarily due to the variation in the partition ratios k(-1)/k(2), for the ES complex, which are 2.3, 0.77, and 30 for I, II, and III, respectively. The preceding analysis is based on direct application of the Stokes-Einstein diffusion law to enzyme kinetics. The range of applicability of this law to the diffusion of substrate size molecules and the mechanics of diffusion of ionic species through viscous solutions of sucrose vs. polymers are explored.  相似文献   

10.
Losartan (DuP 753) and PD123177 are nonpeptide angiotensin (ANG) receptor ligands for subtypes of ANG II receptors ANG II-1 and ANG II-2, respectively. We examined the effects of losartan and PD123177 on dose - mean arterial pressure (MAP) response curves for ANG II and ANG III in eight groups (n = 6 each) of conscious rats. Saline (0.9% NaCl), losartan (1 x 10(-6) and 9 x 10(-6) mol/kg), and PD123177 (2 x 10(-5) mol/kg) were i.v. bolus injected 15 min before the construction of ANG II dose - response curves in groups I, II, III, and IV, respectively. Groups V-VIII were treated similarly to I-IV except that ANG III was given in place of ANG II. Losartan dose dependently shifted the dose-response curves of ANG II and ANG III to the right with similar dissociation constants (-log KI of 6.6 +/- 0.7 and 6.6 +/- 0.1 mol/kg, respectively) and no change in the maxima. PD123177 affected neither maximum MAP nor ED50 values for ANG II or ANG III. Our results show that losartan but not PD123177 is a competitive antagonist of the MAP effects of ANG II and ANG III.  相似文献   

11.
The exopolysaccharide (EPS) "viilian" was isolated from a large-batch fermentation of Lactococcus lactis subsp. cremoris SBT 0495. After applying a newly developed purification procedure, pure viilian with a weight-averaged molar mass of 2.64 x 10(3) kg/mol was obtained in a yield of 0.6 g/L culture broth. The native EPS, as well as lower molar mass fractions obtained by sonication of the native polymer, were studied by capillary viscometry and size-exclusion chromatography (SEC) coupled to multiangle laser light scattering detection (MALLS). From the viscosity data at various ionic strengths, we extracted a Mark-Houwink-Kuhn-Sakurada exponent a = 0.79, and a Smidsrod B value of 0.03. By application of the Hearst, Bohdanecky, and Odijk models for stiff polymer coils, in connection to the experimental viscosity data, we established the characteristic ratio to be C(infinity) = 44 and the intrinsic persistence length q(0) = 11.5 nm. The rms radii of gyration predicted from each of the models were in good agreement with the experimental radii (e.g., (1/2)(w) = 162 nm for native viilian in 0.2M NaNO(3)), as determined by SEC-MALLS. In addition, the Odijk model predicts correct ionic strength-linear charge density dependence of the rms radius of gyration. From the combined viscosity and SEC-MALLS experiments we concluded that, in dilute aqueous solutions, viilian behaves as an intermediately stiff, random coil polyelectrolyte system.Copyright 2000 John Wiley & Sons, Inc.  相似文献   

12.
A water soluble galactomannan isolated from Leucaena leucocephala seeds gave an intrinsic viscosity of 3.5dl/g and viscosity average molecular mass, M(v), of 6.98×10(5)g/mol. This was in reasonably good agreement with the value of the weight average molecular mass, M(w), of 5.44±0.20×10(5)g/mol determined by GPC-MALLS coupled to RI. The onset of polymer coil overlap occurred at c*[η] of 2.1, with slope of 3.0 above and 1.3 below the point of polymer coil overlap. The shear viscosity of the polysaccharide was temperature dependent and decreased with increasing temperature. The activation energy for viscous flow of 3.0% polysaccharide concentration obtained by Arrhenius plot of zero shear viscosity as a function of temperature was 26.4kJ/mol. Both the storage modulus (G') and loss modulus (G″) showed strong dependence on frequency indicating the presence of entangled coils. The Cox-Merz plot gave close superimposition of the complex and shear viscosities.  相似文献   

13.
Goh KK  Pinder DN  Hall CE  Hemar Y 《Biomacromolecules》2006,7(11):3098-3103
Polysaccharides isolated from flaxseed meals using ethanol consisted of a soluble ( approximately 7.5% w/w) and an insoluble fraction (2% w/w). The soluble fraction was dialyzed in various salt concentrations and characterized using viscometry and light scattering techniques. Observations using a size-exclusion column coupled to a multiangle laser light scattering (SEC-MALLS) revealed three molecular weight fractions consisting of a small amount ( approximately 17%) of large molecular weight species (1.0 x 10(6)) and a large amount ( approximately 69%) of small molecular weight species (3.1 x 10(5) Da). Dynamic light scattering measurements indicated the presence of very small molecules (hydrodynamic radius approximately 10 nm) and a very large molecular species (hydrodynamic radius in excess of 100 nm); the latter were probably aggregates. The intrinsic viscosity, [eta], of the polysaccharide in Milli-Q water was 1030 +/- 20 mL/g. The viscosity was due largely to the large molecular weight species since viscosity is influenced by the hydrodynamic volume of molecules in solution. The Smidsrod parameter B obtained was approximately 0.018, indicating that the molecules adopted a semi-flexible conformation. This was also indicated by the slope ( approximately 0.56) from the plot of root-mean-square (RMS) radius versus molar mass (M(w)).  相似文献   

14.
Guar gum was cross-linked with glutaraldehyde and characterized by GPC, rheology, WADX, SEM and TGA. This guar gum is a galactomannan polysaccharide, that contains small amount of arabinose, glucose and uronic acid, besides galactose and mannose. The polymer has high molar mass, with Mw, Mn and Mv values of 2.0x10(6), 1.2x10(6) and 1.9x10(6)g/mol, respectively. The reticulation follows a slow process and lead to a viscosity increase of 40 times compared with the original gum solution. The final viscosity was similar to that of Hylan G-F 20, a hyaluronate derivative, commercially used in viscosupplementation treatment. The gel contains 95.6% of water and the amount of residual glutaraldehyde is much lower than the LD-50. Porous structure was detected by SEM and thermal stability was improved by the cross-linking. The low viscosity, the small amount of remained glutaraldehyde, and the thermal stability indicates that the guar hydrogel has potential to be applied as biomaterial with specific rheological requirements.  相似文献   

15.
A soluble methane monooxygenase (sMMO: EC 1.14.13.25) was purified from a type II obligate methanotroph, Methylocystis sp. M. Ion exchange chromatography elution separated the sMMO into three components, I, II, and III. Components II and III were purified to homogeneity and were essential for the sMMO activity. Components II and III had molecular masses of approximately 233,000 and 39,000, respectively. Component II consisted of three subunits with molecular masses of 55,000, 44,000, and 21,000, which appeared to be present in stoichiometric amounts, suggesting a (αβγ)2 configuration in the native protein. Component II contained 1–4 mol of iron and was considered to be a hydroxylase. Component III was a flavoprotein, which contained 1 mol of FAD as well as 1–2mol of iron. It catalyzed the reduction of K3Fe(CN)6 and 2,6-dichloroindophenol by NADH. Component I, which was partially purified and not essential for sMMO activity, stimulated the activity by about 11-fold. Its stimulation could be replaced by addition of Fe2+. The molecular mass of the partially purified component I was estimated to be from 35,000 to 40,000 based on gel filtration, which suggested the presence of a new type of regulatory protein of sMMO.  相似文献   

16.
Firefly luciferase (EC 1.13.12.5) (FL) is the key enzyme in the firefly bioluminescence method (FB), which is widely used to determine the viability of living cells. The FB method can also be applied to monitoring the influence of different pollutants, such as pesticides. Firefly luciferase is a hydrophobic enzyme and its activity depends on the type of solvent, pH and substances present in the reaction mixture. The influence of three aromatic pesticides, including fenoxaprop-p-ethyl (I), diclofop-methyl (II) and metsulfuron methyl (III), on the enzyme activity was indirectly evaluated through the measurement of emitted light in the bioluminescence reaction, expressed in relative luminescence units (RLU). The reaction mixture used in the bioluminescence measurements consisted of: Tris buffer (pH 7.75), adenosine triphosphate (ATP) and ATP monitoring reagent, where FL is present. Ethanol-water solutions of each pesticide were then added at concentrations of 2.4 x 10(-4)-2.4 x 10(-8) mol/L. The FL activity inhibition factors (FL In%) were determined. The FL activity was maximally inhibited in the presence of all pesticides under study at a concentration of 2.4 x 10(-4) mol/L and was lowered by about 15-26% for pesticide I at concentrations of 2.4 x 10(-5)-2.4 x 10(-8) mol/L, whereas pesticides II and III, applied in the same concentration range, showed smaller FL inhibition values (5.3-20%). The pesticide degradation products (obtained after a 1 month period), measured in the same experimental conditions, in most cases exhibited a much less inhibitory effect on the enzyme activity than the corresponding initial pesticide.  相似文献   

17.
Cultured chicken embryo fibroblasts synthesize two distinct molecular size classes of hyaluronic acid. The high molecular weight material (form I, 2.98 x 10(6) is the predominant species synthesized by transformed cells, whereas form II (1.42 x 10(5)) is the major product of non-transformed cells. A shift to synthesis of predominantly form I hyaluronic acid is an early transformation event in cells infected with LA24 Rous sarcoma virus and maintained at the permissive temperature for transformation (35 degrees C). Form I hyaluronic acid exhibits greater binding to preparations of cellular fibronectin and to both normal and transformed cells than does form II. Both forms bind more to transformed cells than to normal, uninfected cells. Hyaluronic acid (predominantly form I) isolated from transforming cells stimulates proliferation in growth-retarded, non-transformed cells.  相似文献   

18.
Blood viscosity and cardiac output in acute experimental anemia.   总被引:3,自引:0,他引:3  
The significance of blood viscosity alterations during anemia was evaluated in dogs under morphine-chloralose anesthesia. In group I, anemia (mean hematocrit 18.1 +/- 1.3 vol %) was produced by exchange transfusion with clinical dextran (avg mol wt 70,000). In group II, anemia was produced (mean hematocrit 19.9 +/- 0.88 vol %) with 500,000 molecular weight dextran, thus preventing the decrease in blood viscosity in group I. The cardiac output increase in group I (93.4%) with low-viscosity anemia was significantly greater than in group II (43.3%) with unchanged blood viscosity. Group III animals were transfused with a clinical dextran-red cell mixture, and group IV animals received a 500,000 mol wt dextran-red cell mixture. In group III, blood viscosity and cardiac output did not change. In group IV, blood viscosity rose and cardiac output fell significantly. The results suggest that a change in blood viscosity exerts a significant effect upon cardiac output, especially during acute dextran-exchange anemia.  相似文献   

19.
A new Naringenin Schiff-base ligand (H3L) and its complex, [La(H2L)2(NO3).3H2O], have been synthesized and characterized on the basis of elemental analyses, molar conductivities, mass spectra, 1H NMR, thermogravimetry/differential thermal analysis (TG-DTA), UV spectra, and IR spectra. Spectrometric titrations, ethidium bromide displacement experiments, and viscosity measurements indicate that the two compounds, especially the La(III) complex, strongly bind with calf-thymus DNA, presumably via an intercalation mechanism. The intrinsic binding constants of the La(III) complex and ligand with DNA were 1.83 x 10(7) and 9.46 x 10(5) M(-1), respectively. Comparative cytotoxic activities of the La(III) complex and ligand were also determined by MTT [3-(4,5-dimethyl-2-thiazoyl)-2,5-diphenyl-2H-tetrazolium bromide] and SRB (sulforhodamine B) methods. The results showed that the La(III) complex had significant cytotoxic activity against the tested cells.  相似文献   

20.
When submerged cultured Pseudomonas fluorescens NCIMB 11761 was fed-batch supplemented with alpha-pinene oxide, a rapid formation of 2,6-dimethyl-5-methylene-hept-(2Z)-enal (I) (isonovalal) was observed. Biotransformation and isomerisation of (I) to the (2E)-isomer (II) (novalal) were enhanced by Lewatit OC 1064, a macroporous polystyrene adsorbent. Accelerated isomerisation in the presence of an amino donor (glycine) at pH 7.3 pointed to a merely chemical mechanism. A maximum yield of 48 g of aldehydesl(-1) was achieved, but quantitative analysis of the volatile fraction showed that the molar conversion of the pinene oxide substrate reached no more than 67%. To fill this gap of the mass balance, the acidic fraction was isolated. It contained several compounds which suggested a beta-oxidation-like catabolism starting from 2,6-dimethyl-5-methylene-hept-(2E)-enoic acid (III) (novalic acid). Using [2H7]-2,5,6-dimethyl-hept-(2E)-enoic acid as a conversion substrate and gas chromatography coupled to atomic emission detection and mass spectrometry a degradation pathway via labelled 3,4-dimethylpentenoic and methylpropanoic acids was evidenced. This pathway may play a predominant role in isoprenoid degradation by soil bacteria.  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