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.
Summary The effect of water activity (aw) on the growth and end-product formation of Lactobacillus viridescens SMRICC 174, Lactobacillus SMRICC 173 (homofermentative) and Brochothrix thermosphacta ATCC 11509T was studied. All strains orginated from meat or meat products. The aw was adjusted in the range 0.94–0.99 with NaCl or glycerol. A greater reduction in growth rates was found for L. viridescens and B. thermosphacta when aw was regulated with NaCl rather than with glycerol, the opposite was true for Lactobacillus 173. L. viridescens grew at aw >-0.94. At 0.94 aw B. thermosphacta was totally inhibited when NaCl was the solute and Lactobacillus 173 when glycerol was the solute. Only minor variations in the end-product formation of the Lactobacillus spp. were found at different aw values. In aerobic culture B. thermosphacta produced less l-lactic acid and more acetic acid as the aw was decreased with NaCl, while the yields were unaffected when glycerol was used.  相似文献   

3.
由于荒漠生态系统植被覆盖度低、生产力低下,其在全球碳循环中的作用被长期忽视。为探讨荒漠生态系统碳收支各组分的变化规律,以腾格里荒漠红砂(Reaumuria soongorica Maxim.)-珍珠(Salsola passerina Beg.)群落为研究对象,采用静态箱式法研究了该群落的净生态系统CO2交换量(NEE)、生态系统呼吸、土壤呼吸的日变化规律,同时将该方法所获得的NEE结果与涡动相关法观测的结果进行了比较。结果表明:(1)红砂-珍珠群落NEE的日变化表现为,在6:00—9:00左右出现一个CO2吸收的高峰值,随后在12:00—15:00左右出现一个CO2释放高峰值。红砂种群、珍珠种群和整个群落NEE的平均值分别为0.018、0.020和0.028 mg CO2m-2s-1;(2)红砂种群、珍珠种群、土壤及整个群落生态系统呼吸速率的日变化规律一致,均表现为明显的单峰变化趋势,在12:00—15:00左右出现一个CO2释放的高峰值。红砂种群、珍珠种群、土壤和整个群落的生态系统呼吸的平均值分别为:0.121、0.062、0.029和0.040 mg CO2m-2s-1。以盖度为加权因子计算得到红砂种群、珍珠种群和土壤呼吸占生态系统呼吸的比例分别为:9%、21%和70%,由此可见,生态系统呼吸主要来源于土壤呼吸。(3)将箱式法和涡动相关法观测的NEE进行比较,结果表明两种方法观测的NEE变化规律基本一致,相关系数达到0.7。采用箱式法观测的NEE高于涡动相关法观测的结果,平均值分别0.028 mg CO2m-2s-1(箱式法)和0.015 mg CO2m-2s-1(涡动相关法),涡动相关法的观测结果与箱式法观测结果的比值为0.54。综上可得,荒漠生态系统土壤呼吸的变化速率决定了生态系统呼吸的变化规律,采用箱式法可能高估了荒漠生态系统CO2的释放量。  相似文献   

4.
Self-diffusion coefficients have been obtained over a range of concentrations for both polymer and solvent in amylopectin/dimethylsulfoxide and amylopectin/water systems. The measurements were made using pulsed-field gradient nuclear magnetic resonance (PFG-nmr). The implications of macromolecule polydispersity to the interpretion of the PFG-nmr experiment have been considered. Measurements of the self-diffusion coefficients of solute and solvent are shown to provide an effective probe of polymer size and shape. Such measurements demonstrate that in dimethylsulfoxide, wheat starch amylopectin molecules are highly planar with Mw of order 106. In contrast, amylopectin in water is an aggregate with a more spherical shape and has a volume some 400 times larger than the single molecule.  相似文献   

5.
三氟氯氰菊酯对棉铃虫神经细胞钠及钙通道作用机理研究   总被引:13,自引:0,他引:13  
用膜片钳技术对比分析了棉铃虫三氟氯氰菊脂抗性品系(R)及其同源对照品系(S)幼虫了体培养中枢神经细胞Na^2 通道的门控特性及杀虫剂对R和S神经细胞Na^ 、Ca^ 通道门控过程的影响。结果表明,S神经细胞Na^ 通道电流(S-INa)在-50-40mV激活,-20mV左右达峰值,R神经细胞Na^2 通道电流(R-INa)在-40mV左右激活,-10-0mV达峰值,即R-INa激活电压与峰值电压均向正电位方向移动约10mV,提示二者Na^ 通道控特性不同,R神经细胞Na^ 通道功能发生了变异。三氟氯氰菊酯作用后,S-INgn R-ISs的I-V曲线均向负电位方向移动的10mV,S-INa在20min后基本消失,而R-INa被阻断需时约90min,延长近5倍,其幅值有减小再增大的现象。对Ca^2 通道分析表明,杀虫剂作用后,R及S神经细胞Ca^2 通道电流的I-V曲线均向负电位移动10-20mV,提示三氟氯氰菊酯对Ca^2 通道的门控过程也有影响。与R-INa幅值起伏变化相联系,可推知杀虫剂对神经细胞的毒性作用中,Na^2 、Ca^2 通道均受影响。  相似文献   

6.
Electric birefringence measurements of suspensions of T3 and T7 bacteriophages in 10?2 M phosphate buffer, pH 6.9, show that there is a difference in their rotational diffusion coefficient. The values corrected to 25°C and water viscosity are D25,w = 4630 ± 130 sec?1 and D25,w = 5290 ± 260 sec?1 for T3 and T7, respectively. The value obtained from shell model calculations (according to Filson and Bloomfield) is D25,w = 4500 ± 600 sec?1. The apparent permanent dipole moments are 4.5 × 10?26 C·m and 1.7 × 10?26 C·m for T3 and T7, respectively. For both phage particles the intrinsic optical anisotropy is +7.2 × 10?3. It is shown that this anisotropy is mainly due to the DNA molecule inside the head of the phage. Its positive value means that there exists an excess orientation of the DNA helix perpendicular to the symmetry axis of the particle. For T7 an unexpectedly large increase of Δns and Ksp occurs at a glycerol concentration of about 30% (v/v). This increase is interpreted as being caused by a change of the shape of the particle and/or a change in the secondary structure of the DNA inside the head of the bacteriophage.  相似文献   

7.
We have used two different approaches to determine hydrodynamic parameters for mucins secreted by guinea-pig tracheal epithelial cells in primary culture. Cells were cultured under conditions that promote mucous cell differentiation. Secreted mucins were isolated as the excluded fraction from a Sepharose CL-4B gel filtration column run under strongly dissociating conditions. Biochemical analysis confirmed the identity of the high molecular weight material as mucins. Analytical ultracentrifugation was used to study the physical properties of the purified mucins. The weight average molecular mass (M w ) for three different preparations ranged from 3.3×106 to 4.7×106 g/mol (corresponding to an average structure of 1 – 2 subunits), and the sedimentation coefficient from 25.5 to 35 S. Diffusion coefficients ranging from 4.5×10–8 to 6.4×10–8 cm2/s were calculated using the Svedberg equation. A polydispersity index (M z /M w ) of ∼1.4 was obtained. Diffusivity values were also determined by image analysis of mucin granule exocytosis captured by videomicroscopy. The time course of hydration and dissolution of mucin was measured and a relationship is presented which models both phases, each with first order kinetics, in terms of a maximum radius and rate constants for hydration and dissolution. A median diffusivity value of 8.05×10–8 cm2/s (inter-quartile range = 1.11×10–7 to 6.08×10–8 cm2/sec) was determined for the hydration phase. For the dissolution phase, a median diffusivity value of 6.98×10–9 cm2/s (inter-quartile range = 1.47×10–8 to 3.25×10–9 cm2/sec) was determined. These values were compared with the macromolecular diffusion coefficients (D 20,w ) obtained by analytical ultracentrifugation. When differences in temperature and viscosity were taken into account, the resulting D 37,g was within the range of diffusivity values for dissolution. Our findings show that the physicochemical properties of mucins secreted by cultured guinea-pig tracheal epithelial cells are similar to those of mucins of the single or double subunit type purified from respiratory mucus or sputum. These data also suggest that measurement of the diffusivity of dissolution may be a useful means to estimate the diffusion coefficient of mucins in mucus gel at the time of exocytosis from a secretory cell. Received: 10 March 1998 / Accepted: 27 March 1998  相似文献   

8.
Relaxation time measurements T1 and T2 of sodium in Halobacterium halobium pellets were carried out at two frequencies. From those measurements, combined with intensity measurements of the sodium in the system, estimation of the properties of the sodium ions in the system was carried out. It is suggested that three types of sodium ions are present in the bacterial pellet. (A) The extracellular sodium with properties of free solution and (B) sodium which is in the pericellular volume between the cell wall and the cell membrane. There is an exchange between type A and type B sodium. The type B sodium has (e2qQ)/h = 3.7 · 107 rad/s, τcB = 5.2 · 10−6 s and τB = 1 · 10−3 s. The sodium of type C is bound inside the cell and undetected. It's concentration inside the cell is assumed to be 1.9 M.  相似文献   

9.
Laser light-scattering has been used to investigate the size of native proteoglycan aggregates (PGA-aA1) from day-8 chick limb-bud chondrocyte cultures isolated under associative extraction and purification conditions in 0.4M guanidinium chloride (GdnHCl) solution. Dynamic light-scattering measurements yielded a hydrodynamic radius, Rs, of 244 ± 10 nm for PGA-aA1 in 0.4M GdnHCl, and a weight-average molecular weight (M w) of 150 ± 50 × 106 was obtained from a Zimm plot. Disaggregation in 4.0M GdnHCl aqueous solution yielded proteoglycan subunits (PGS) with Rs = 39 ± 2 nm, M w = 1.6 ± 0.3 × 106, which reassembled in 0.4M GdnHCl to form “reconstituted native” aggregates (PGA-raA1) with Rs = 121 ± 6 nm, M w = 17 ± 3 × 106. A second specimen of PGA-aA1 had Rs = 192 ± 10 nm, M w = 100 ± 10 × 106. The latter value was estimated from an empirical relationship between M w and Rs. After dissociation, this specimen reassembled to form PGA-raA1 with Rs = 85 ± 5 nm, M w = 12 ± 1 × 106. These data are compared with those for a specimen of reconstituted aggregate (PGA-A1) that had been extracted under dissociative conditions and then reaggregated by dialysis to 0.4M GdnHCl aqueous solution, for which Rs = 138 ± 9 nm, M w = 45 ± 8 × 106. From these values, we have calculated the weight-average number of subunits per aggregate Nw: 111 for PGA-aA1 and 12 for raA1 (70 and 7 for the second PGA-aA1 and PGA-raA1 specimen, respectively) as compared to 32 for PGA-A1. The numbers of subunits per aggregate were also determined from electron micrographs of spread specimens. The latter results show the same trends as those obtained by light scattering, but lead in each case to lower numbers of subunits per aggregate. These data demonstrate conclusively that PGA samples exhibit a higher degree of aggregation in solution than visualized in typical electron microscopy (EM) preparations, probably due to disaggregation during EM specimen preparation. Since Nw determined both by light scattering (LS) and by EM are larger for native versus reconstituted aggregate samples, our data point to a more compact aggregation of subunits along the hyaluronic acid (HA) chains in the former.  相似文献   

10.
The fractions obtained from the partially hydrolyzed branched Streptococcus salivarius levan were examined in solution. Sedimentation coefficients, S0, intrinsic viscosities, [η], weight-average molecular weights, M w, and radii of gyration were obtained from sedimentation velocity, viscosity, and light-scattering measurements. Double logarithmic plots of [η] vs M w and S0 vs M w each yielded two linear segments intersecting at M w ≈ 105. Hydrodynamic data suggest that fractions of M w > 105 behave as compact spheres, whereas for M w < 105, the particles are best characterized as linear random coils. Calculations based on theories of random coils and spheres support the above observations.  相似文献   

11.
The plant growth retardant paclobutrazol, (PP333) (2RS, 3RS)-1-(4-chlorophenyl)-4,4-dimethyl-2-(1,2,4-triazol-1-yl)pentan-3-ol, inhibits specifically the three steps in the oxidation of the gibberellin-precursorent-kaurene toent-kaurenoic acid in a cell-free system fromCucurbita maxima endosperm. The KI50 for this inhibition is 2×10–8 M. The KI50 values for the separated2S, 3S, and2R, 3R enantiomers of paclobutrazol in this system are 2×10–8 M and 7×10–7 M, respectively. A cell-free preparation from immatureMalus pumila embryos convertsent-kaurene to gibberellin A9, whereas no conversion occurs in a similar preparation fromMalus endosperm. The conversion ofent-kaurene by the embryo preparation is inhibited by paclobutrazol with KI50 values for the2S,3S and2R,3R enantiomers of 2×10–8 M and 6×10–8 M, respectively.  相似文献   

12.
The effect of magnesium ions on the parameters of the DNA helix-coil transition has been studied for the concentration range 10?6–10?1M at the ionic strengths of 10?3M Na+. Special attention has been given to the region of low ion concentrations and to the effect of polyvalent metallic impurities present in DNA. It has been shown that binding with Mg++ increases the DNA stability, the effect being observed mainly in the concentration range 10?6–10?4M. At[Mg++]>10?2M the thermal stability of DNA starts to decrease. The melting range extends to concentrations ~10?5M and then decreases to 7–8°C at the ion content of 10?3M. Asymmetry of the melting curves is observed at low ionic strengths ([Na+] = 10?3M) and [Mg++] ? 10?5M. The results, analyzed in terms of the statistical thermodynamic theory of double-stranded homopolymers melting in the presence of ligands, suggest that the effects observed might be due to the ion redistribution from denatured to native DNA. An experimental DNA–Mg++ phase diagram has been obtained which is in good agreement with the theory. It has been shown that thermal denaturation of the system may be an efficient method for determining the ion-binding constants for both native and denatured DNA.  相似文献   

13.
We isolated eleven strains of the harmful algal bloom (HAB)-forming dinoflagellate Karlodinium veneficum during a bloom event in the NW Mediterranean coastal waters and we studied the inter-strain variability in several of their physiological and biochemical traits. These included autotrophic growth parameters, feeding capabilities (mixotrophy), lipid composition, and, in some cases, their responses to biotic and abiotic factors. The strains were found to differ in their growth rates (0.27–0.53 d−1) and in the maximum cell concentrations achieved during stationary phase (6.1 × 104–8.6 × 104 cells mL−1). Their ingestion performance, when offered Rhodomonas salina as prey, was also diverse (0.22–1.3 cells per K. veneficum per day; 8–52% of their daily ration). At least two strains survived for several months under strict heterotrophic conditions (no light, low inorganic nutrients availability, and R. salina as food source). These strains also showed very distinct fatty acid compositions, with very low contents of monounsaturated and polyunsaturated fatty acids. According to a Bray Curtis similarity analysis, three or four strain groups able to perform different roles in bloom development were identified. We further analyzed one strain from each of the two most distinct groups with respect to prey concentration, light intensity, nutrient availability, and we determined the functional responses (growth and feeding rates) to food concentration. Taken together, the results served to highlight the role of mixotrophy and clone variability in the formation of HABs.  相似文献   

14.
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   

15.
该研究采用红外气体分析法(IRGA)于2013年3–12月原位测定了北京市东升八家郊野公园中2个主要阔叶树种(槐(Sophora japonica)、旱柳(Salix matsudana))3个高度上的枝干呼吸(Rw)日进程,旨在量化Rw的种间差异,探索种内Rw及其温度敏感系数(Q10)的时间动态和垂直分布格局。研究结果显示:(1)Rw在不同树种之间差异明显,相同月份(4月份除外)槐Rw是旱柳的1.12(7月)–1.79倍(5月)。两树种枝干表面CO2通量速率均表现出明显的单峰型季节变化,峰值分别出现在7月((5.13±0.24)μmol·m–2·s–1)和8月((3.85±0.17)μmol·m–2·s–1)。同一树种在生长月份内的平均呼吸水平显著高于非生长季,但其Q10值季节变化趋势与之相反。(2)RW随测量高度的增加而升高,并在3个高度上表现出不同的日变化规律:其中,树干基部及胸高位置为单峰格局,而一级分枝处的呼吸速率在一天内存在两个峰值,中间出现短暂的"午休"现象。温度是造成一天内呼吸变化的主要原因。此外,顶部Rw及其对温度的敏感程度明显高于基部。温度本身和Q10值差异可在一定程度上解释RW的垂直梯度变化。(3)在生长月份,单位体积木质组织的日累积呼吸速率(mmol·m–3·d–1)与受测部位直径倒数(D–1)呈极显著正相关关系。单位面积(μmol·m–2·s–1)可准确表达两树种在生长期间的RW水平,能合理有效地比较不同个体的呼吸差异及同一个体的时空变异。这些结果表明,采用局部通量法上推至树木整体呼吸时,应全面考虑Rw的时、空变异规律,并选择恰当的表达单位,以减小估测误差。  相似文献   

16.
The translational diffusion coefficient D 20,w 0 , of monomeric human immunoglobulin G (IgG) has been studied by photon-correlation spectroscopy as a function of pH and protein concentration. At pH 7.6, we find D 20,w 0 =3.89×10–7±0.02 cm2/sec, in good agreement with the value determined by classic mehods. This value corresponds to an effective hydrodynamic radius R, of 55.1±0.3 Å. As pH is increased to 8.9; with the same ionic strength, the molecule appears to expand slightly (3.5% increase in hydrodynamic radius). The concentration dependence of the IgG diffusion constant is interpreted in terms of solution electrostatic effects and shows that long-range repulsive interactions are negligible in the buffer used. The diffusion coefficient for dimeric IgG has also been determined to be D20,w=2.81×10–7±0.04 cm2/sec at 1.6 mg/ml, which corresponds to a hydrodynamic radius of 75 Å. For light-scattering studies of protein molecules in the dimension range of 5–10 nm (Mr=105–107) we find monomeric horse spleen ferritin well suited as a reference standard. Ferritin is a spherical molecule with a hydrodynamic radius R of 6.9±0.1 nm and is stable for years in our standard Tris-HCl-NaCl buffer even at room temperature.  相似文献   

17.
Iwao Satake  Jen Tsi Yang 《Biopolymers》1975,14(9):1841-1846
The conformational phase diagram of poly(L -lysine) (4.6 × 10?4 M, residue) in sodium dodecyl sulfate (1.6 × 10?2 M) solution was constructed from circular dichroism results at various temperatures and pH's. Poly(L -lysine)–sodium dodecyl sulfate complexes undergo a β–helix transition upon raising the pH of the solution. The transition pH tends to shift downward at elevated temperatures. No helix–β transition can be detected for poly(L -lysine) in sodium dodecyl sulfate solution (pH > 11) even after 1-hr heating at 70°C. This is in marked contrast with uncharged poly(L -lysine) solution without sodium dodecyl sulfate, which is converted into the β-form upon mild heating of the solution above 50°C.  相似文献   

18.
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   

19.
Polymeric excipients are often the least well-characterized components of pharmaceutical formulations. The aim of this study was to facilitate the QbD approach to pharmaceutical manufacturing by evaluating the inter-grade and inter-batch variability of pharmaceutical-grade polymeric excipients. Sodium alginate, a widely used polymeric excipient, was selected for evaluation using appropriate rheological methods and test conditions. The materials used were six different grades of sodium alginate and an additional ten batches of one of the grades. To compare the six grades, steady shear measurements were conducted on solutions at 1%, 2%, and 3% w/w, consistent with their use as thickening agents. Small-amplitude oscillation (SAO) measurements were conducted on sodium alginate solutions at higher concentrations (4–12% w/w) corresponding to their use in controlled-release matrices. In order to compare the ten batches of one grade, steady shear and SAO measurements were performed on their solutions at 2% w/w and 8% w/w, respectively. Results show that the potential interchangeability of these different grades used as thickening agents could be established by comparing the apparent viscosities of their solutions as a function of both alginate concentration and shear conditions. For sodium alginate used in controlled-release formulations, both steady shear behavior of solutions at low concentrations and viscoelastic properties at higher concentrations should be considered. Furthermore, among batches of the same grade, significant differences in rheological properties were observed, especially at higher solution concentrations. In conclusion, inter-grade and inter-batch variability of sodium alginate can be determined using steady shear and small-amplitude oscillation methods.  相似文献   

20.
The DNA helix–coil transition has been studied in the presence of high concentrations of manganese ions (about 10?3M), which corresponds to the conditions close to equal stability of the A+T and G+C pairs, at the ionic strengths of 10?1, 10?2, and 1.6 × 10?3M Na+. With the Mn2+ ion effect, the transition range is significantly reduced to not more than 0.2°C at 1.2 × 10?3M Mn2+ and 1.6 × 10?3M Na+. The melting curves display a sharp kink at the end of the helix–coil transition, which is interpreted as an indication of the second-order phase transition. It is shown that the melting curves obtained can be approximated by a simple analytical expression 1 – θ = exp[–a(tc - t)], where θ is the DNA helix fraction, tc is the phase transition temperature, and a is an empirical parameter characterizing the breadth of the melting range and responsible for the magnitude of a jump of the helicity derivative with respect to the temperature at the phase transition point.  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