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 31 毫秒
1.
The solution properties of the B and Z forms of poly(dG-dC).poly(dG-dC) have been measured by static and dynamic laser light scattering. The radius of gyration, persistence length, translational and segmental diffusion coefficients, and the Rouse-Zimm parameters have been evaluated. The persistence length of the Z form determined at 3 M NaCl is about 200 nm compared to 84 and 61 nm respectively for the B forms of poly(dG-dC).poly(dG-dC), and calf thymus DNA, both determined at 0.1 M NaCl. The data on persistence length, diffusion coefficients and the Rouse-Zimm parameters indicate a large increase in the chain stiffness of Z DNA compared to the B form. These results are opposite to the ionic strength effects on random sequence native DNAs, for which the flexibility increases with ionic strength and levels off at about 1 M NaCl.  相似文献   

2.
Internal Brownian motions of clean ?29 and λ-DNAs have been studied using photon-correlation techniques at both visible (λ0 = 632.8 nm) and uv (λ0 = 363.8 nm) wavelengths. The present dynamic light scattering data, which extend to K2 = 19 × 1010 cm?2, can in every case be satisfactorily simulated by a Rouse-Zimm model polymer with an appropriate choice of the three model parameters. The effects of pH, salt concentration, single-strand breaks, and molecular weight on those model parameters have also been investigated. Intact clean DNAs exhibit surprisingly little variation with pH from 7.85 to 10.25, with salt concentration from 0.01 NaCl to 5.4M NH4Cl, or with molecular weight or GC content. The single-strand breaks have no effect at pH 9.46, but produce dramatic changes in the model parameters at pH 10.0 and 10.25, indicating the introduction of titratable joints at those pHs. The failure of either single-strand breaks or a large change in GC content to alter the model parameters in the neutral pH range is a strong indication that local denaturation is not required for those flexions and torsions that dominate the relaxation of fluctuations in the scattered light. The Langevin relaxation time for the slowest internal mode of a particular Rouse-Zimm model derived from the dynamic light scattering data is compared with pertinent literature data extrapolated to the same molecular weight. The present algorithm for determining model parameters from the light-scattering Dapp vs K2 curve actually yields a Langevin time in fairly good agreement with the literature value. For unknown reasons the light-scattering D0 values generally exceed those obtained from the molecular weight and sedimentation coefficient by about 20%.  相似文献   

3.
The influence of Escherichia coli single-strand binding (SSB) protein on the conformation and internal dynamics of pBR322 and pUC8 supercoiled DNAs has been investigated by using dynamic light scattering at 632.8 and 351.1 nm and time-resolved fluorescence polarization anisotropy of intercalated ethidium. SSB protein binds to both DNAs up to a stoichiometry that is sufficient to almost completely relax the superhelical turns. Upon saturation binding, the translational diffusion coefficients (D0) of both DNAs decrease by approximately 20%. Apparent diffusion coefficients (Dapp) obtained from dynamic light scattering display the well-known increase with K2 (K = scattering vector), leveling off toward a plateau value (Dplat) at high K2. For both DNAs, the difference Dplat - D0 increases upon relaxation of supercoils by SSB protein, which indicates a corresponding enhancement of the subunit mobilities in internal motions. Fluorescence polarization anisotropy measurements on free and complexed pBR322 DNA indicate a (predominantly) uniform torsional rigidity for the saturated DNA/SSB protein complex that is significantly reduced compared to the free DNA. These observations are all consistent with the notion that binding of SSB protein is accompanied by a gradual loss of supercoils and saturates when the superhelical twist is largely removed.  相似文献   

4.
Amiloride, a sodium channel blocker, is known to suppress NaCl responses of the chorda tympani (CT) nerve in various mammalian species. In mice, the NaCl suppressing effect of amiloride is reported to differ among strains. In C57BL mice, amiloride inhibits NaCl responses to about 50% of control, whereas no such clear suppression was evident in prior studies with 129 mice. However, evidence from behavioral studies is not entirely consistent with this. Recently, it has been found that genetic backgrounds of 129 mice differ within substrains. 129X1/SvJ (formerly 129/SvJ) mice differ from the 129P3/J (formerly 129/J) strain by 25% of sequence length polymorphisms. Therefore, we examined possible substrain difference between 129P3/J and 129X1/SvJ mice in the amiloride sensitivity of electrophysiologically recorded NaCl responses. Amiloride significantly suppressed CT responses to NaCl without affecting responses to KCl both in 129P3/J and 129X1/SvJ mice. However, the magnitude of the amiloride inhibition was significantly larger (approximately 50% of control in response to 0.01-1.0 M NaCl by 100 microM amiloride) in 129X1/SvJ than in 129P3/J mice (approximately 20% of control in response to 0.03-0.3 M NaCl by 100 microM amiloride). Threshold amiloride concentration for suppression of responses to 0.3 M NaCl was 30 microM in 129P3/J mice, which was higher than that in 129X1/SvJ mice (10 microM). In 129X1/SvJ mice, the threshold amiloride concentration eliciting inhibition of NaCl responses and the magnitude of the inhibition were comparable with those in C57BL/6 mice. These results suggest that amiloride sensitivity of NaCl responses differs even among the 129 substrains, 129P3/J and 129 X1/SvJ, and the substrain difference of 129 mice in amiloride sensitivity is as large as that between two inbred strains (129P3/J and C57BL/6).  相似文献   

5.
Langowski J  Hammermann M  Klenin K  May R  Tóth K 《Genetica》1999,106(1-2):49-55
We present here recent results on the structure of superhelical DNA and its changes with salt concentration between 0.01 and 1.5 M NaCl. Scattering curves of two different superhelical DNAs were determined by static light scattering. The measured radii of gyration do not change significantly with salt concentration. Small-angle neutron scattering, together with calculations from a Monte Carlo model, allows to determine the superhelix diameter. Measured and simulated scattering curves agreed almost quantitatively. Experimentally we find that the diameter decreases from 16.0±0.9 nm at 10 mM to 9.0±0.7 nm at 100 mM NaCl. The superhelix diameter from the simulated conformations decreased from 18.0±1.5 nm at 10 mM to 9.4±1.5 nm at 100 mM NaCl. At higher salt concentrations up to 1.5 M NaCl, the diameter stays constant at 9 nm.This revised version was published online in October 2005 with corrections to the Cover Date.  相似文献   

6.
T Piekenbrock  E Sackmann 《Biopolymers》1992,32(11):1471-1489
In the first part of this work we report quasielastic light scattering (QELS) studies of the internal dynamics of transient actin networks over a time range of 10(-6)-10(-2) s, scattering angles between zeta = 20 degrees and 150 degrees, and a concentration range of 0.015 (0.3) to 0.7 mg/mL (15 microM). We confirm our previous result that (1) the dynamic structure factor g(q,t) is determined by the thermally excited undulations of the actin filaments and (2) that the initial decay of g(q, t) scales as g(q, t) varies; is directly proportional to exp(-q alpha t) while the long time decay scales as g(q, t) varies; is directly proportional to exp [-(Aq alpha t) 2/3] with alpha = 2.75. The deviation of alpha from the theoretical value of alpha = 3 predicted for Rouse-Zimm chains is similar to that found for high molecular weight macromolecular solutions by QELS. A refined analysis of the dynamic structure factor showed that it can be interpreted in terms of three relaxation processes (besides the contribution of the residual monomer diffusion): (1) the dominant Rouse-Zimm dynamics, which comprises between 65 (at high concentrations) and 85% of the signal; (2) a fast relaxation process with a decay constant of gamma = 9 x 10(3) s-1, which contributes at all concentrations with the same amplitude; and (3) a nonexponential ultraslow contribution of the form g(us) varies; is directly proportional to exp [(-gamma ust)]1/4. The third contribution appears only at high concentrations and increases strongly with decreasing scattering angles. It is thus attributed to fluctuations of the mesh size of the transient actin network. In the second part we show that high sensitivity QELS may be applied to follow the actin polymerization process at low temperatures (10 degrees C). The apparent diffusion coefficient and the static scattering intensity of the actin filaments were determined as functions of polymerization time tpol. We show that the process consists of the rapid growth of a few filaments that become very long (approximately 10 microns; even at actin concentrations of 0.04 micrograms/mL) near the critical growth concentration of 0.012 micrograms/mL, as is expected for a growth process determined by nucleation. Finally, we studied actin networks polymerized in the presence of complexes of gelsolin with actin. By application of the CONTIN program we could determine the length distribution of the filaments.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

7.
The following physical parameters of the legumin from Vicia faba were determined by means of small-angle X-ray scattering, quasi-elastic light scattering and circular dichroism: molar mass, M = 3.5 X 10(5) g/mol; radius of gyration, Rg = 4.45 nm; maximum dimension, L = 13 nm; translational diffusion coefficient, D0(20),w = 3.38 X 10(-7) cm2 X s-1; alpha-helix content about 15%; content of beta-sheets 10%; dihedral point group symmetry of the molecule 32.  相似文献   

8.
The molecular weights, subunit dissociation, and conformation in solution of the hemocyanins of three species of octopi were investigated by light-scattering, ultracentrifugation, absorbance, and circular dichroism methods. The molecular weights of the hemocyanins of Octopus bimaculoides, Octopus bimaculatus, and Octopus rubescens obtained by light scattering were 3.3 X 10(6), 3.4 X 10(6), and 3.5 (+/- 0.3) X 10(6), respectively. The average molecular weights of the fully dissociated hemocyanins of the same octopi, investigated at alkaline pH and in the presence of 8 M urea and 6 M guanidinium chloride (GdmCl), were found to be close to one-tenth of those of the parent proteins, with average molecular masses of 3.4 X 10(5), 3.3 X 10(5), and 3.3 (+/- 0.3) X 10(5). These findings confirm the earlier observations of van Holde and co-workers with other cephalopod hemocyanins that the basic cylindrical assembly of molluscan hemocyanins consists of 10 subunits. Circular dichroism and absorbance measurements suggest that the dissociated subunits at alkaline pH and in concentrated urea solutions retain their native, multidomain folding. Fairly concentrated GdmCl above 3-4 M is necessary to unfold fully the dissociated hemocyanin chains. Molecular weight measurements studied as a function of reagent concentration with the urea and Hofmeister salt series as dissociating agents show that the ureas are very effective dissociating agents, while the salts are ineffective to moderately effective reagents for octopus hemocyanin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Using laser light scattering, we have measured the static and dynamic structure factor of two different superhelical DNAs, p1868 (1868 bp) and simian virus 40 (SV40) (5243 bp), in dilute aqueous solution at salt concentrations between 1 mM and 3 M NaCl. For both DNA molecules, Brownian dynamics (BD) simulations were also performed, using a previously described model. A Fourier mode decomposition procedure was used to compute theoretical light scattering autocorrelation functions (ACFs) from the BD trajectories. Both measured and computed autocorrelation functions were then subjected to the same multiexponential decomposition procedure. Simulated and measured relaxation times as a function of scattering angle were in very good agreement. Similarly, computed and measured static structure factors and radii of gyration agreed within experimental error. One main result of this study is that the amplitudes of the fast-relaxing component in the ACF show a peak at 1 M salt concentration. This nonmonotonic behavior might be caused by an initial increase in the amplitudes of internal motions due to diminishing long-range electrostatic repulsions, followed by a decrease at higher salt concentration due to a compaction of the structure.  相似文献   

10.
Dynamic light scattering (DLS) measurements were performed on self-assembled solutions of RecA as a function of assembly time under strand exchange ionic strength conditions (10 mM MgCl2, 65 mM NaCl, 10 mM Tris-HCl, pH = 7.5, 1 mM DTT, 3-4 microM RecA) in the absence of ATP. These measurements yield distributions of the translational diffusion coefficients of the changing populations of assembling protein species. Interpretations of results of DLS measurements are made in terms of model hydrodynamic calculations that indicate, under the solution conditions employed, the smallest fundamental quaternary subunit of RecA is a hexamer in a toroidal or lock-washer configuration. Interactions of M13mp19 circular single strand DNA (ssDNA) with RecA assembled to different stages were also investigated. Additions of ssDNA to self-assembled solutions of RecA acts to dissociate the associated structures into hexamer subunits. However, the effect of ssDNA on assembled RecA is highly dependent on the RecA self-assembly state. The longer the assembly time, the less reversible the self-assembled structures of RecA become. Binding isotherms of titrated mixtures of ssDNA with RecA self-assembled to various stages were also determined. Evaluated dissociation constants of RecA/ssDNA complexes were found to increase with increases of the associated state of RecA. These results strongly suggest that, under the solvent conditions employed, the active ssDNA binding form of RecA is a hexamer.  相似文献   

11.
We have analyzed the equilibrium and nonequilibrium properties of the complex of the single stranded DNA binding protein of Escherichia coli (EcoSSB) and circular single stranded DNA of filamentous phages M13mp8 and F1 using static and dynamic light scattering, analytical ultracentrifugation and electron microscopy. Upon binding to the single stranded DNA the EcoSSB tetramer replaces an equivalent volume of water trapped within the coiled single stranded DNA and hinders the folding of the single stranded DNA into secondary structures at all salt concentrations. The salt dependent compaction of the stoichiometric complex can be described assuming a flexible polyelectrolyte chain. The solution structure of the macromolecular complex is a random coil and in the electron microscope a beaded flexible structure of the complex with a bead diameter of 6 nm appears at all salt concentrations used. The internal motions of the stoichiometric complex can be described by the Rouse-Zimm model of polymer dynamics. The segmental mobility of the complex can be correlated with changes in the binding site size of the EcoSSB tetramer; it indicates the presence of interactions between EcoSSB tetramers bound to single stranded DNA.  相似文献   

12.
Three commercial chitosans with a degree of acetylation of 25–30% were studied by light scattering (static and dynamic), analytical ultracentrifugation (sedimentation velocity and sedimentation equilibrium), and capillary viscometry in 0.02 M acetate buffer/0.1 M NaCl, pH 4.5. The molecular masses obtained by sedimentation equilibrium measurements or sedimentation and diffusion coefficients according to the Svedberg equation agreed well or fairly well with those from static light scattering whereas the molecular masses calculated via the Scheraga–Mandelkern equation were found too low by almost 50%. The various Mark–Houwink type relationships suggested a nearly free-draining flexible worm-like chain. A prolate ellipsoid of revolution with an axial ratio a/b25 was shown to be a hydrodynamically equivalent body of the flexible worm-like chain that had been derived from static light scattering. The findings illustrate the fact that a hydrodynamically strongly asymmetric shape need not mean a strongly elongated shape of the molecules in reality.  相似文献   

13.
R P Hertzberg  P B Dervan 《Biochemistry》1984,23(17):3934-3945
The synthesis of methidiumpropyl-EDTA (MPE) is described. The binding affinities of MPE, MPE.Ni(II), and MPE.Mg(II) to calf thymus DNA are 2.4 X 10(4) M-1, 1.5 X 10(5) M-1, and 1.2 X 10(5) M-1, respectively, in 50 mM NaCl, pH 7.4. The binding site size is two base pairs. MPE.Mg(II) unwinds PM2 DNA 11 +/- 3 degrees per bound molecule. MPE.Fe(II) in the presence of O2 efficiently cleaves DNA and with low sequence specificity. Reducing agents significantly enhance the efficiency of the cleavage reaction in the order sodium ascorbate greater than dithiothreitol greater than NADPH. At concentrations of 0.1-0.01 microM in MPE.Fe(II) and 10 microM in DNA base pairs, optimum ascorbate and dithiothreitol concentrations for DNA cleavage are 1-5 mM. Efficient cleavage of DNA (10 microM in base pairs) with MPE.Fe(II) (0.1-0.01 microM) occurs over a pH range of 7-10 with the optimum at 7.4 (Tris-HCl buffer). The optimum cleavage time is 3.5 h (22 degrees C). DNA cleavage is efficient in a Na+ ion concentration range of 5 mM to 1 M, with the optimum at 5 mM NaCl. The number of single-strand scissions on supercoiled DNA per MPE.Fe(II) under optimum conditions is 1.4. Metals such as Co(II), Mg(II), Ni(II), and Zn(II) inhibit strand scission by MPE. The released products from DNA cleavage by MPE.Fe(II) are the four nucleotide bases. The DNA termini at the cleavage site are 5'-phosphate and roughly equal proportions of 3'-phosphate and 3'-(phosphoglycolic acid). The products are consistent with the oxidative degradation of the deoxyribose ring of the DNA backbone, most likely by hydroxy radical.  相似文献   

14.
Using a combination of spectroscopic techniques, quasi-elastic laser light scattering (QLS), and electron microscopy (EM), we have been able to show that the B to Z transition of poly(dG-m5dC) X poly(dG-m5dC) is accompanied by extensive condensation of the DNA in both low and high ionic strength buffers. At low concentrations of NaCl (2 mM Na+), an intermediate rodlike form, which exhibits a circular dichroism (CD) spectrum characteristic of an equimolar mixture of B and Z forms, is observed. This is produced by the orderly self-association of about four molecules of the polymer after prolonged incubation of a concentrated solution at 4 degrees C. On addition of 5 microM Co(NH3)63+, the CD spectrum of the intermediate changes to that of the Z form, which is visualized as a dense population of discrete toroids on an EM grid stained with uranyl acetate. On the other hand, addition of NaCl to a solution of poly(dG-m5dC) X poly(dG-m5dC) in the absence of any multivalent ion condenses the polymer to toroidal structures at the midpoint (0.75 M NaCl) of the B to Z transition. Further addition of NaCl unfolds these toroids to rodlike structures, which show characteristic Z-form CD spectra. These results show that Z DNA can take up a variety of tertiary structural forms and indicate that its inverted CD spectrum is due to its left-handed helical sense rather than to differential scattering artifacts.  相似文献   

15.
The thermal depolymerization of porcine submaxillary mucin   总被引:2,自引:0,他引:2  
The time dependence of the molecular weight, radius of gyration, and hydrodynamic size distribution for porcine submaxillary mucin (PSM) in solution have been studied using static and dynamic light scattering. The weight average molecular weight (Mw) of PSM in 6 M guanidine HCl, pH 7, is initially 3 X 10(6) and decreases with time in three phases: rapidly from 3-2 X 10(6), less rapidly from 2-0.9 X 10(6), and slowly below 0.9 X 10(6). The rates of decrease are much greater at pH 2. The energy of activation associated with each phase is 20 kcal/mol, which is similar to that reported for peptide bond cleavage at an aspartic acid residue. Addition of mercaptoethanol to PSM in 6 M guanidine HCl leads to a rapid decrease in Mw to 0.9 X 10(6), followed by a very slow further decrease. These results suggest that native PSM consists of subunits (Mw = 0.9 X 10(6] that are linked by disulfide bonds to form dimers (Mw = 2 X 10(6] and then higher aggregates. This cross-linking appears to occur at unglycosylated regions of the protein core, which are believed to be richer in aspartic acid than the rest of the molecule.  相似文献   

16.
The reflection coefficient (sigma) and permeability (P) of urea and ethylene glycol were determined by fitting the equations of Kedem and Katchalsky (1958) to the change in light scattering produced by adding a permeable solute to a red cell suspension. The measurements incorporated three important modifications: (a) the injection artifact was eliminated by using echinocyte cells; (b) the use of an additional adjustable parameter (Km), the effective dissociation constant at the inner side of the membrane; (c) the light scattering is not directly proportional to cell volume (as is usually assumed) because refractive index and scattering properties of the cell depend on the intracellular permeable solute concentration. This necessitates calibrating for known changes in refractive index (by the addition of dextran) and cell volume (by varying the NaCl concentration). The best fit was for sigma = 0.95, Po = 8.3 X 10(-4) cm/s, and Km = 100 mM for urea and sigma = 1.0, Po = 3.9 X 10(-4) cm/s, and Km = 30 mM for ethylene glycol. The effects of the inhibitors copper, phloretin, p- chloromercuriphenylsulfonate, and 5,5'-dithiobis (2-nitro) benzoic acid on the urea, ethylene glycol, and water permeability were determined. The results suggest that there are three separate, independent transport systems: one for water, one for urea and related compounds, and one for ethylene glycol and glycerol.  相似文献   

17.
Abstract

A high molecular weight ‘cryogel’ was obtained as insoluble complexes by cold incubation at near-freezing temperatures from heparinized plasma of patients with rheumatoid arthritis. After the cryogel was solubilized at 37°C, 1:1 complex of fibrinogen and fibronectin was purified at room temperature by affinity chromatography on a gelatin-Sepharose 4B. Hydrodynamic properties of the complex were investigated as a function of temperature and NaCl concentration using a dynamic light scattering. The diffusion coefficients of the complex at 20°C decreased with increasing of NaCl concentration as free fibronectin. The complex appears to be a more compact form at low ionic concentration, which is associated with conformational changes of fibronectin. The diffusion coefficient of the complex at 20°C in 0.05 M Tris- HCl(pH7.4) containing 0.5 M NaCl was estimated as 8.5× 10?8 cm2s?1. The complex did not dissociate over the temperature range from 20 to 37°C. The diffusion coefficients of the complex decreased significantly at 12°C and 40°C. The thermal denaturation of fibrinogen molecule in the complex was observed at 40°C. The CONTIN analysis of the light scattering data showed that the complex associated to form higher aggregates at 15°C, but not at near- freezing temperature. The equilibrium between the complex and higher aggregates appeared reversible.  相似文献   

18.
The hydrodynamic properties of large homodisperse single stranded DNAs complexed with the helix destabilizing protein of phage T4, the product of gene 32 (GP32), have been measured. The results suggest a size of the binding site between 8 and 10 nucleotides/GP32 molecule, in reasonable agreement with earlier work on a complex between GP32 and single stranded 145 base DNA. From static light scattering experiments it is concluded that the persistence length of these complexes is about 30 nm, distinctly smaller than the generally accepted value for double stranded DNA. The quasi-elastic light scattering properties of the DNA-GP32 complexes were determined. The variation of the apparent translation diffusion coefficient Dapp with the scattering vector q was analyzed using the discrete ISMF and Rouse-Zimm models [S.C. Lin et al., Biopolymers 17 (1978) 425]. The model parameters that followed from the fit of Dapp versus q2 and from an extensive global analysis of the actually measured autocorrelation functions agreed with the notion that these DNA-protein complexes are indeed rather flexible. The continuous Soda model [K. Soda, Macromolecules 17 (1984) 2365] could successfully explain the variation of Dapp versus q2, assuming a persistence length of 30 nm and a base-base distance in the complex of 0.44 nm.  相似文献   

19.
A myosin phosphatase has been purified to homogeneity from bovine aortic smooth muscle. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of the enzyme eluted from nondenaturing gels revealed two subunits (Mr = 67,000 and 38,000). Densitometric scans of the subunits indicated a molar ratio of 1:1. Several phosphoproteins were substrates for the phosphatase including histone II-A, isolated 20,000-dalton smooth muscle myosin light chains, phosphorylase a, and smooth muscle myosin. In the presence of 0.25 M NaCl and a substrate concentration of 2 microM, myosin was preferentially dephosphorylated. The specific activity of the phosphatase for myosin at a concentration of 10 microM was found to be 5 mumol/mg/min. The phosphatase required Mn2+ or Co2+ ions for activity. Mg2+, Ca2+, or Mg-ATP would not substitute for Mn2+ or Co2+ at equimolar concentrations. This phosphatase may play an important role in regulating actin-myosin interaction in smooth muscle by serving to dephosphorylate myosin.  相似文献   

20.
Complexation of human serum albumin (HSA) with poly(N-isopropylacrylamide) (PNIPA) ranging in molecular weight (M(PNIPA)) from 2.1 x 10(4) to 1.72 x 10(6) was studied in an aqueous system (pH 3) containing NaCl as a supporting salt. Dynamic light scattering, static light scattering, electrophoretic light scattering, and dialyzing techniques were used as the experimental tool in a suitable combination. The measurements were performed mainly at 25 degrees C and at 0.01 M NaCl as a function of mixing ratio (r(m), molar ratio of PNIPA to HSA). The results of DLS and ELS evidently demonstrated the formation of a water-soluble complex through mixing of HSA and PNIPA. A detailed analysis of SLS data with the aid of dialysis data revealed that the resulting complex is an "intramolecular" complex consisting of a PNIPA chain with several of bound HSA molecules. Both hydrodynamic radius (R(h)) and radius gyration (R(g)) of intramolecular complexes decreased as r(m) was increased. This result correlated well to the fact that the number (n) of bound proteins per polymer decreases with increasing r(m). The size and the molar mass of the complex became large depending on M(PNIPA), but the increase of M(PNIPA) led to a decrease in n at r(m) < 1. The increase in NaCl concentration from 0.01 to 0.3 M brought about the increase in the size and the molar mass of an intramolecular HSA-PNIPA complex prepared at r(m) = 1.1. This was found to be due to an increase of n. A similar trend was observed when temperature rose from 25 to 32 degrees C (close to lower critical solution temperature of PNIPA). However, the effect of temperature on the increase of was strong in comparison with that of ionic strength. On the basis of these results obtained, the complexation mechanism was discussed in detail.  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