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.
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   

2.
G Cohen  H Eisenberg 《Biopolymers》1966,4(4):429-440
Solutions of calf thymus NaDNA in 0.2M NaCl and CsDNA in 0.2M CsCl (fragmented by sonic irradiation to a molecular weight of about 5 × 105) were examined by electron microscopy, light scattering, and viscosity measurements. Electron microscopy showed that the molecular fragments are rodlike in shape and that the distribution of their lengths is sharp (Lw/Ln ~ 1.06). The weight-average number of nucleotides Zw per DNA molecule derived from light scattering was found to be in very good agreement with the value Lw derived from electron microscopy. The z-average length Lz derived from light scattering at 25°C. was found to be about 89% of the corresponding value derived by electron microscopy. Similarly, the value of Lw derived from intrinsic viscosity at 25°C. on the basis of a rodlike model was found to be about 87% of the corresponding value derived by electron microscopy. It is concluded that the molecules are slightly flexible and do not, assume their full contour length in solution because of the disorienting effect of Brownian motion. The intrinsie viscosity was found to increase linearly with decreasing temperature toward a limiting value corresponding to the fully stretched length of the macromolecules at 0°K.: a reasonable value for the modulus of elasticity could be calculated. It was also found that, no differences in conformation could be delected in t he two systems CsDNA-0.2M CsCl and NaDNA-0.2M NaCl.  相似文献   

3.
The polymerization of 1,2-anhydro-3,4,6-tri-O-benzyl-β-D -mannopyranose proceeds in the presence of Lewis acids, cationic coordination catalysts, and strong bases. Debenzylation of the products yields oligomeric saccharides or low polymers. Polymerization in toluene by means of potassium alkoxide complexed with crown ethers leads to essentially stereoregular (1 → 2)-α-D -mannopyranan. The original derivatives have been characterized by optical rotation, viscosity, molecular weight, gel permeation chromatography, and spectrometry. The free polysaccharides have been characterized by optical rotation, molecular weight, and 1H- and 13C-nmr spectrometry and compared to yeast mannan hydrolysate oligomers.  相似文献   

4.
A water soluble acidic heteropolysaccharide named WAF was isolated from Auricularia auricula‐judae by extracting with 0.9% NaCl solution. By using gas chromatography, gas chromatography‐mass spectrometry, and NMR, its chemical structure was determined to be composed of a backbone of α‐(1→3)‐linked D ‐mannopyranose residues with pendant side groups of β‐D ‐xylose, β‐D ‐glucose, or β‐D ‐glucuronic acid at position O6 or O2. Six fractions prepared from WAF with a weight‐average molecular mass (Mw) between 5.9 × 104 and 64.7 × 104 g/mol were characterized with laser light scattering and viscometry in 0.1M NaCl at 25°C. The dependence of intrinsic viscosity ([η]) and radius of gyration (Rg) on Mw for this polysaccharide were found to be [η] = 1.79 × 10?3Mw0.96 cm3 g?1 and Rg = 6.99 × 10?2 Mw0.54 nm. The molar mass per unit contour length (ML) and the persistence length (Lp) were estimated to be 1124 nm?1 and 11 nm, respectively. The WAF exhibited a semirigid character typical of linear polysaccharides. Molecular modeling was then used to predict the ordered and disordered states of WAF; the simulated ML and Lp were however much smaller than the experimental values. Taken altogether, the results suggested that WAF formed a duplex in solution. © 2010 Wiley Periodicals, Inc. Biopolymers 95: 217–227, 2011.  相似文献   

5.
Donald B. Siano 《Biopolymers》1978,17(12):2897-2908
Light-scattering studies on buffered aqueous solutions of the triple-stranded polyribonucleic acid poly(A)·2poly(U) were carried out at neutral pH and during titration. At pH 7.1 and 22°C, a sample of commercially available polymer in 0.005M phosphate buffer gave a Zimm plot which yielded values for the weight-average molecular weight, M w, of 874,000 ± 1800 g/mol, a root-mean-square radius, ρ of 930 ± 22 Å, and a second viral coefficient of 0.51 ± 0.05 × 10 ?3 cm3g?1 mol. The light-scattering data were also analyzed by serval linear and nonlinear least-squares programs which were devised to determine the model (e.g., rod, coil, or zigzag) which could best describe the shape of the molecule. It was found that a rodlike model, perhaps with a few bends, was in best overall agreement with the data. The assumption that the molecule is a thin rod leads to a value for the linear density of 206 g mol?1 Å?1 and a translation of 3.3 Å per residue. These values are also in close agreement with those expected for a triple-stranded, thin, base-stacked molecule. During titration from neutral pH with 0.1M HCl, the observed apparent molecular weight slowly increased until at about pH 3.5 a sudden, large increase (about 30-fold) occurred. The root-mean-square radius, on the other hand, after an initial small decrease (of about 25%), also exhibited a large increase (about 4-fold). Upon back titration with 0.1M NaOH, the molecular parameters did not retrace the original path, but instead exhibited hysteresis—the M w and ρ z are both larger on the basic branch than on the acid branch at a corresponding pH. A plot of long ρ z against log(M w) during the interval in which the high-moelcular-weight form was present (below pH 3.5 on the acid branch, and on the basic branch) gave a straight line with a slope of ?. This suggests that the aggregates were composed of some tens of rather open radom coils, presumably of poly(A)·poly(A), and that the hysteresis may be caused under conditions by the metastability of the entangled coils.  相似文献   

6.
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   

7.
Wang X  Zhang X  Xu X  Zhang L 《Biopolymers》2012,97(10):840-845
Lentinan (β‐(1→3)‐D ‐glucan) was found to be successfully fractionated by the mixture of dimethyl sulfoxide (DMSO) and lithium chloride (LiCl) as a solvent and acetone as a precipitant. Light scattering and viscosity measurements were made on solutions of fractionated samples in pure DMSO and 0.2M LiCl/DMSO in the range of the molecular weight Mw from 21.7 × 104 to 84.7 × 104. The values of Mw in both solvents were almost the same, but the remarkable difference between the values of intrinsic viscosity [η] demonstrated that the LiCl/DMSO solvent greatly enhances the stiffness of the lentinan backbone. The observed intrinsic viscosity [η] was analyzed by the Yoshizaki‐Nitta‐Yamakawa theory of a worm‐like chain, and the persistence length q and molecular weight per unit contour length ML were determined roughly as 6.0 nm and 890 g nm?1 in 0.2M LiCl/DMSO, and 5.1 nm and 890 g nm?1 in pure DMSO, respectively. This slightly larger persistent length in 0.2M LiCl/DMSO also confirmed the higher stiffness of lentinan enhanced by the LiCl/DMSO solvent. The enhancement of the chain stiffness was ascribed to the electrostatic repulsion because of the hydrogen bonding of the hydroxyl protons of lentinan with the chloride ion, which is in turn associated with the Li+(DMSO)n macrocation complex. © 2012 Wiley Periodicals, Inc. Biopolymers 97: 840–845, 2012.  相似文献   

8.
(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   

9.
Lina Zhang  Liqun Yang 《Biopolymers》1995,36(6):695-700
A Water-soluble glucan A was isolated from the fruit body of Auricularia auricula-judae. It is composed of a backbone chain of β-(1 → 3)-linked D -glucose residues, two out of three glucose residues being substituted at C-6 positions with a single glucose unit. The weight average molecular weight Mw, number average molecular weight Mn, and intrinsic viscosity [η] of the fractionated samples were studied at 25°C in water and in dimethylsulfoxide (DMSO). The Mark-Houwink equation was established as [η] = 6. 10 × 10?4 Mw1.14 for the glucan A having Mw ranging from 9 × 105 to 1.6 × 106 in water. The values of [η] in water are far higher than those in DMSO, but the values of Mn measured in water are the same as those in DMSO. Analysis of Mw and [η] in terms of the known theories for rods and wormlike chains yielded 1030 ± 100 nm?1, 90 ± 20 nm, 1.3 ± 0.3 nm, and 0.26 ± 0.03 nm for molar mass per unit contour length ML, persistence length q, diameter d, and contour length h per main-chain glucose residue, respectively. The present data suggest that glucan A dissolves in water as single-stranded helical chains and in DMSO as semiflexible chains. © 1995 John Wiley & Sons, Inc.  相似文献   

10.
Six amylose‐2‐acetyl‐3,6‐bis(phenylcarbamate) (AAPC) samples ranging in weight‐average molar mass Mw from 1.8 × 104 g mol?1 to 1.1 × 106 g mol?1 have been prepared from enzymatically synthesized amylose samples. Static light scattering, small‐angle X‐ray scattering, sedimentation equilibrium, and viscosity measurements were made for the samples in 1,4‐dioxane (DIOX), 2‐ethoxyethanol (2EE), and 2‐butanone (MEK) all at 25°C to determine particle scattering functions, z‐average radii of gyration, intrinsic viscosities, as well as Mw. The data were analyzed in terms of the wormlike cylinder model mainly to yield the helix pitch per residue h and the Kuhn segment length λ?1, which corresponds to twice of the persistence length. The latter parameters (λ?1) in 2EE (11 nm) and MEK (12 nm) are quite smaller than those for amylose tris(phenylcarbamate) (ATPC) in the same solvent (16 nm in 2EE and 18 nm in MEK) whereas those for AAPC (21 nm) and ATPC (22 nm) in DIOX are essentially the same as each other. This indicates that the chain stiffness of AAPC is more strongly influenced by the solvents since the number of intramolecular H‐bonds of AAPC is more changeable than that for ATPC. © 2012 Wiley Periodicals, Inc. Biopolymers 97:1010–1017, 2012.  相似文献   

11.
A water‐soluble α‐(1→4)‐D ‐glucan heteropolysaccharide with 37% degree of branch extracted by base from Rhizoma Panacis Japonici, coded as RPS3, was fractionated into six fractions by the method of nonsolvent addition. Their weight‐average molecular mass (Mw), polydispersity index (Mw/Mn), and radius of gyration (〈s2z1/2) were determined with laser light scattering (LLS) and size exclusion chromatography combined with LLS. The structure of the fraction was determined by methylation analyses and 13C NMR. The dependences of intrinsic viscosity ([η]) and 〈s2z1/2 on Mw were established as [η] = 0.71 Mw0.27 ± 0.01 (cm3/g) and 〈s2z1/2 = 1.53 Mw0.27 ± 0.02 (nm) in the Mw range from 5.62 × 104 to 3.05 × 106 (g/mol) for RPS3 in 0.15M NaCl aqueous solution at 25°C. On the basis of the current theory of the polymer solution, the fractal dimension (df), unperturbed chain dimension (A), and characteristic ratio (C) were calculated to be 3.0, 1.48 Å, and 15.1, respectively. The results revealed that the RPS3 chains existed as spherical conformation in the aqueous solution. Transmission electron microscope further provided the evidence of the sphere shape of the RPS3 and its fractionated molecules in water. In vitro cytotoxicity assay indicated that the fractions could inhibit the tumor cells and showed no harm to normal cells at low dose. The bioactivity was relative with molecular mass of the samples. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 383–390, 2010. This article was originally published online as an acceptedpreprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office atbiopolymers@wiley.com  相似文献   

12.
Light scattering has been used to investigate the structure of human tracheobronchial mucin glycoproteins (HTBM) from the sputum of cystic fibrosis patients. The specimen was extracted using 6M guanidinium hydrochloride solution and fractionated by gel exclusion chromatography on Sephacryl S-1000. The fractionated HTBM was purified by density gradient ultracentrifugation. Purity of the resulting material was confirmed by SDS polyacrylamide gel electrophoresis and uv spectroscopy. Light scattering measurements on the fractionated mucins yield weight-average molecular weights Mw, and z-average radii of gyration Rg, z. The native cystic fibrosis HTBM consisted of a high molecular weight fraction with Mw = 9.3 × 106 daltons and a lower molecular weight fraction contanining partly degraded mucins. After reduction and carboxymethylation of the high molecular weight native fraction, the resulting material was separated into three pools with Mw values of 5.1 × 106, 1.6 × 106, and 400,000. The derived molecular weights for the protein cores Mp,w, and the experimental radii of gyration are found to be consistent with the Mp,wRg relation established previously for submaxillary, cervical, and gastric mucins. These results imply that HTBM has the same extended-coil conformation reported for other mucins and has a molecular structure consisting of subunits, linked into linear chains via covalent (disulfide) bonds.  相似文献   

13.
Thermal denaturation and circular dichroism (CD) properties of poly(L -lysine)–DNA complexes vary greatly when these complexes are prepared differently, that is, whether by NaCl-gradient dialysis starting from 2.0 M NaCl or by direct mixing at low salt. These differing properties were investigated in more detail by examining complexes, made by direct mixing in the presence of various concentrations of NaCl, both before and after the NaCl was dialyzed out of the complex solution. The precipitation curves of DNA due to polylysine binding indicate that such binding is noncooperative at zero salt; from 0.1 up to 1.0 M NaCl they exhibit varying degrees of cooperatively. Starting from zero salt, as the NaCl concentration used for complex formation is increased, both the CD and the melting properties of the complexes are shifted from those of directly mixed at zero salt to those of reconstitution: in the CD spectra there is a gradual shift from a B → C transition to a B → ψ transition; thermal denaturation results show a gradual increase in the melting temperatures of both free DNA (tm) and polylysine-bound DNA (tm). The progressive shift from B → C to B → ψ suggests a close relationship between these two transitions. Large aggregates of the complexes do not warrant the appearance of ψ-type CD spectra: ψ-spectra have been obtained in the supernatants of polylysine–DNA complexes made and measured at 1.0 M NaCl while slightly perturbed CD spectra in B → C transition have been observed in turbid solutions of fully covered complexes made at very low salt. If the complexes are made at intermediate salts and dialyzed to a very low salt, although up to 60% of the DNA is still bound by polylysine, the CD spectra of the complexes are shifted back to the B-type CD characteristic of pure DNA.  相似文献   

14.
Dynamic light scattering measurements were performed on dilute aqueous solutions of native ovalbumin (OA) and on those of linear OA aggregates induced by thermal denaturation at low ionic strength and neutral pH. The weight-average molecular weight Mw of four aggregates tested ranged from 1,700,000 to 5,500,000. The translational diffusion coefficient D0 of native OA at infinite dilution was estimated as 8.70 × 10 ?7 cm2/s, which gave 56.0 Å as the diameter of the rigid spherical particle. The intensity autocorrelation function of linear OA polymers was analyzed with the cumulant method to obtain the first cumulant Λe. The dependence of Λe on the scattering vector q at very low polymer concentration was found intermediate between those of a flexible chain and a rigid rod. The translational diffusion coefficient Dtr [≡ (Te/q2)q → 0] was in proportion to M, and the magnitude was in good agreement with a value calculated from the wormlike cylinder model with values of three parameters determined in an earlier study, ML = 1600 Å?1, d = 120 Å, and Q = 230 Å, where ML, d, and Q are the molecular weight per unit length, diameter, and persistence length, respectively. Based on these results, a new model, to be called as the dimer model, was proposed to interpret the formation mechanism of linear OA polymers induced by thermal denaturation. © 1993 John Wiley & Sons, Inc.  相似文献   

15.
Molecular mechanics and dynamics calculations were carried out on the disaccharides α-L-Rhap-(1 → 2)-α-L-Rhap-(1 → OMe) (1) and α-L-Rhap-(1 → 3)-α-L-Rhap-(1 OMe) (2), and the trisaccharide α-L-Rhap-(1 → 2)-α-L-Rhap-(1 → 3)-α-L-Rhap-(1 → OMe) (3). The semiflexible conformational behavior of these molecules was characterized by the occupation of a combination of different glycosidic linkage and side-chain conformational positions whose relative occupations were sensitive to dielectric screening. Molecular dynamics simulations of the trisaccharide 3 showed little difference between the linkage conformations in the trisaccharide and the component disaccharides 1 and 2. Experimental optical rotation data of 1 and 2 were obtained as a function of temperature in varying solvents. The molecular models were combined with the semiempirical theory of Stevens and Sathyanarayana to yield calculated optical rotations. Interpretation of the data of both 1 and 2 implied that a combination of conformations, both in glycosidic and side-chain positions, could explain the experimental data. Solvents effects were important in influencing the conformational mix and averaged optical rotation. Three-bond heteronuclear coupling constants 3JC, H were obtained for the glycosidic linkages of 1 and 2 in D2O and DMSO. Analysis of the coupling constants with a Karplus curve showed that small reductions in the glycosidic torsion angles of the conformations of the models used here of ca. 10°–15° in ϕ and 5°–10° in ψ were required to give better agreement with experiment; a combination of conformations for both 1 and 2 was consistent with the data. There was a negligible influence on the coupling constants of 1 on changing the solvent from D2O to DMSO. © 1997 John Wiley & Sons, Inc.  相似文献   

16.
The conformational properties of κ‐carrageenan in 0.2M LiI and ι‐carrageenan in 0.2M LiCl were analyzed by size exclusion chromatography combined with low‐angle laser light scattering. Fractionated samples with narrow molecular weight distributions (Mw/Mn ∼ 1.4) were used, and Mw in the disordered states were 35,000 (κ‐35) and 200,000 (κ‐200) for κ‐carrageenan and 65,000 (ι‐65) and 170,000 (ι‐170) for ι‐carrageenan, respectively. The analyses were performed across a temperature range where the conformational transitions occurred, and at extremely low concentrations (2–50 μg/mL) due to low amounts of samples injected and the subsequent dilution occurring during the separation. The results indicate that a twofold increase of the molecular weight (Mw) occurs for κ‐carrageenan upon inducing the ordered conformation. For ι‐carrageenan an additional increase in Mw may take place, which is attributed to the strong tendency for aggregation of ordered chains especially at high molecular weights. The results thus suggest that both κ‐ and ι‐carrageenan are double (or multiple) stranded in their ordered conformations, within the concentration range studied here. © 1999 John Wiley & Sons, Inc. Biopoly 49: 71–80, 1999  相似文献   

17.
Polyhydroxyalkanoates (PHAs) are naturally occurring biodegradable polymers with promising application in the formulation of plastic materials. PHAs are produced by numerous bacteria as energy/carbon storage materials from various substrates, including sugars and plant oils. Since these substrates compete as food sources, their use as raw material for industrial-scale production of PHA is limited. Therefore, efforts have been focused on seeking alternative sources for bacterial production of PHA. One substrate that seems to have great potential is the seed oil of Jatropha curcas plant. Among other favorable properties, J. curcas seed oil is non-edible, widely available, and can be cheaply produced. In this study, Pseudomonas oleovorans (ATCC 29347) was grown in a mineral salt medium supplemented with saponified J. curcas seed oil as the only carbon source under batch fermentation. Optimum PHA yield of 26.06% cell dry weight was achieved after 72 h. The PHA had a melting point (T m) between 150 and 160°C. Results of polymer analyses by gas chromatography/mass spectrometry (GC/MS) identified only the methyl 3-hydroxybutanoate monomeric unit. However, electrospray ionization–time of flight mass spectroscopy (ESI–TOF MS) confirmed that the PHA was a copolymer with the characteristic HB/HV peaks at m/z 1155.49 (HB) and 1,169, 1,184–1,194 (HV). The data were further supported by1H and 13C NMR analysis. Polymer analysis by gel permeation chromatography (GPC) indicated a peak molecular weight (MP) of 179,797, molecular weight (M W) of 166,838, weight number average mass (M n) of 131,847, and polydispersity (M w/M n) of 1.3. The data from this study indicate that J. curcas seed oil can be used as a substrate to produce the copolymer poly(3-hydroxybutyrate-co-3-hydroxyvalerate), poly(3HB-co-3HV).  相似文献   

18.
A water-insoluble polysaccharide (PCS3-II) extracted from sclerotium of Poria cocos was identified as a linear (1  3)-β-d-glucan by 13C NMR and gas chromatography. Aqueous 0.5 M NaOH/0.2 M urea was a good solvent for PCS3-II and the dependence of intrinsic viscosity ([η]) on weight-average molecular weight (Mw) was established in the Mw range from 7.68 × 104 to 5.14 × 105 to be [η] = 3.39 × 10?2 MW0.62cm3g-1 at 25 °C by using laser light scattering and viscometry. The chain conformation parameters of PCS3-II in the 0.5 M NaOH/0.2 M urea solution was 2.3 (± 0.3) nm for persistence length (q), 580 g mol?1 nm?1 for molar mass per unit contour length (ML), 0.8 (± 0.2) nm for the diameter of the chain (d) and 3.63 for limited characteristic ratio (C). The results revealed, for the first time, that PCS3-II existed as a flexible chain in 0.5 M NaOH/0.2 M urea aqueous solution.  相似文献   

19.
Native calf thymus DNA was sheared by sonication in a viscous solvent to the molecular-weight range from 3 × 104 to 3 × 105 daltons, and fractionated by gel chromatography. Number and weight average molecular weights (M?n and M?w) were determined for individual fractions by electron microscopy; the ratio M?w/M?n for the peak fraction is approximately 1.1. Sedimentation coefficients (s020,w) of these fractionated samples show an approximately linear dependence on the logarithm of the molecular weight M?w. This behavior is that expected for rodlike molecules, and is in quantitative agreement with the theory of Yamakawa and Fujii [(1973) Macromolecules 6 , 407–415] for the sedimentation coefficient of a wormlike chain with a persistence length of 625 Å, a diameter of 25 Å, and a mass per unit length of 195 daltons/Å. It appears that the wormlike coil model, without excluded volume, can represent the sedimentation behavior of DNA over the entire conformational range from rigid rod to flexible coil, using the above parameters. Equilibrium melting curves were determined for various fractions in aqueous 2.4 M tetraethylammonium bromide. A substantial broadening of the transition and decrease of the melting temperature were observed with decreasing molecular weight. Empirical expressions have been obtained relating both the transition temperature and breadth in this solvent to molecular weight.  相似文献   

20.
The state of aggregation of recombinant desulfatohirudin (r-HV1) in solution under physiological conditions (pH 7.5, 0.15N NaCl) was investigated by sedimentation equilibrium. The weight-average molecular weight ¯M w determined by sedimentation equilibrium was found to be 6914±76 Da compared to 6964 Da expected from the amino acid sequence. The ¯M z /¯M w ratio was found to be 1.03, which demonstrates that under the conditions studied hirudin exists in solution as a monomer. This result is in agreement with the relative molecular weight (M r ) of recombinant hirudin variant 3 reported by Otto and Seckler [(1991),Eur. J. Biochem. 202, 67–73], who also used equilibrium ultracentrifugation, but not with the molecular weight estimated from gel permeation chromatography of natural hirudin (51,300 Da) [Konnoet al. (1988),Arch. Biochem. Biophys. 267, 158–166]. Knowledge of the state of aggregation is essential for understanding the mechanism of interaction of thrombin and hirudin under physiological conditions.Abbreviations ¯M w weight-average molecular weight - ¯M z Z-average molecular weight - M r relative molecular weight - NTSB 2-nitro-5-thiosulfobenzoic acid - Tris Tris(hydroxymethyl)aminomethane - r-HV1 recombinant desulfatohirudin - M molar extinction coefficient  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