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.
Conformation of the extracellular polysaccharide of Xanthomonas campestris.   总被引:2,自引:0,他引:2  
G Holzwarth 《Biochemistry》1976,15(19):4333-4339
The solution conformation of the extracellular polysaccharide of the bacterium Xanthomonas campestris is examined by optical rotation, viscometry, and potentiometric titration. Measurements of optical rotation vs. temperature for solutions of the polysaccharide at low ionic strength reveal a sharp transition to a denatured structure which is reversible if sufficient salt is present. The temperature Tm at the transition midpoint increases as log (Na+) or log (Ca2+). Viscosity-temperature profiles substantiate a structural change of the polysaccharide at Tm. The intrinsic viscosity of the native molecule at zero shear rate exceeds 5000 ml/g. This high figure is indicative of a stiff chain. The viscosity of the native molecule is relatively insensitive to salt, whereas the denatured molecule collapses if salt is present. Hydrogen-ion titration shows that the pKapp of the COO- groups of the polymer decreases from 3.2 in 0.01 M NaC1 to 2.6 in 0.2 M NaC1. All these data suggest that the native polysaccharide possesses ordered secondary structure stabilized by nonionic interactions outweighing the repulsion between adjacent COO- groups.  相似文献   

2.
The molar optical rotation at 220 nm and ellipticity values at 210 nm of both sodium hyaluronate and hyaluronic acid are greatly enhanced in comparison to the values for the monomeric units and oligosaccharides indicating a degree of preferred order. With increasing hydrogen ion concentration, there is no appreciable change in the 210 nm circular dichroic band, but the second circular dichroic band below pH 4 changes abruptly to the positive side and reaches a maximum value at pH 2·5. This positive circular dichroic band of hyaluronic acid is temperature and concentration dependent. The major change in sign and position of the second circular dichroic band of hyaluronic acid below pH 4 is attributed to the conformational change of a single polysaccharide chain or to a chain-chain interaction. The results indicate that increase in concentration or decrease in temperature and in the ionization of carboxyl group promotes the formation of ordered cross-link regions. The conformational changes found in solution have been interpreted as an order-disorder transition in the crosslink regions based on the interconversion of random coil and double helix.  相似文献   

3.
Mechanism and dynamics of conformational ordering in xanthan polysaccharide   总被引:4,自引:0,他引:4  
The thermally induced order-disorder transition of xanthan (extracellular bacterial polysaccharide from Xanthomonas campestris) has been investigated by optical rotation, differential scanning calorimetry, stopped-flow reaction kinetics and low-angle laser light scattering, and the results have been analysed in terms of Zimm -Bragg helix-coil transition theory. The reciprocal of the transition midpoint temperature (Tm) varies linearly with the logarithm of cation (K+) the salt dependence of Tm, is in agreement with Manning polyelectrolyte theory the ordered structure. The associated increase in cation binding, calculated from the salt dependence of tm, is in agreement with the Manning polyelectrolyte theory for one of the candidate structures from X-ray diffraction, a 5(1) single helix stabilized by packing of side-chains along the polymer backbone, but not for the alternative double-helix structure that has also been proposed. At each salt concentration, the two fundamental parameters of the Zimm -Bragg theory, s and sigma, were calculated. The equilibrium constant for growth of the ordered structure (s) is derived directly from calorimetric measurement of transition enthalpy (delta Hcal ), and sigma, which quantifies the relative instability of the helix nucleus, is derived from the ratio of delta Hcal to the apparent transition enthalpy (delta Happ ) obtained by van't Hoff analysis of the optical rotation data. The temperature course of conformational ordering calculated theoretically is in good quantitative agreement with experimental results from both optical rotation and scanning calorimetry. The calculated average length of stable, ordered chain-sequences increases with decreasing temperature, but equals or exceeds the total chain length from light scattering only at temperatures more than approximately equal to 70 K below Tm, suggesting that ordered and disordered regions may co-exist within the same xanthan molecule. Consistent with this interpretation, the observed rate of conformational ordering increases sharply under conditions where the starting solution for dynamic measurements is partially ordered, suggesting that ordered sequences within each chain may act as helix nuclei for adjacent disordered regions, so that helix growth, rather than the slower nucleation process, becomes rate limiting.  相似文献   

4.
The conformational behaviour of the extracellular polysaccharides from Arthrobacter species of soil-borne bacteria has been investigated by nuclear magnetic resonance relaxation and optical rotation. Polysaccharides from A. stabilis, A. viscosus, and A. viscosus sp. n, in solution at room temperature, all show evidence of an ordered conformation which can be melted out on heating. The temperature course of this transition, however, shows considerable variation with bacterial species. Thus A. stabilis polysaccharide shows a very sharp conformational transition centred around 60°, whereas the transitions of the polysaccharides from both strains of A. viscosus occur over a much broader temperature-range. The transition for the polysaccharide of A. viscosus sp. n is again centred close to 60°, whereas, for A. viscosus, melting of the tertiary structure of the polysaccharide is incomplete at 100°. O-Deacetylation destroys the ordered conformation of both A. viscosus polysaccharides. The ordered structure of A. stabilis polysaccharide, by contrast, is stabilised by removal of acyl substituents (which here include succinic half-ester). Understanding of the conformational state of these materials affords considerable insight into their gelation behaviour and unusual solution rheology. The known solution interactions with certain plant polysaccharides suggest a possible biological role for Arthrobacter polysaccharides in relationships with components of plant root-systems.  相似文献   

5.
Arabinoxylan preparations from sugar cane show temperature-induced shifts of optical rotation in aqueous methyl sulphoxide solution, the sign and magnitude of which depend on the content of arabinofuranose side-groups. The shift has a sigmoidal form and shows a distinct hysteresis loop on heating and cooling. This evidence, together with the sign and magnitude, is interpreted in terms of a conformation change of the backbone of β-(1→4)-linked D-xylose residues, from the random coil to an ordered, ribbon-like conformation similar to that which is known to exist in the solid state. The driving force for this change is not intramolecular but derives from an aggregation that occurs simultaneously and can be detected by other methods. The arabinofuranose side-groups can be incorporated into the ordered assembly and contribute to the optical rotation shift, and they therefore have an unusual role compared with other polysaccharide side-chains. We conclude that the optical rotation shifts show a “melting” and re-formation of ordered associations which may imitate the natural biological cohesion between hemicellulose chains.  相似文献   

6.
The unusual increase in viscosity and pseudoplasticity often observed when salts are added to moderately concentrated aqueous solutions of xanthan gum is shown to arise from an increase in the extent of macromolecular association. The fractional change in viscosity on addition of KCl to salt-free 1% (w/v) solutions of purified polysaccharide in the K+ salt form is found to be positive only when the degree of pyruvate substitution (fraction of side chains which carry pyruvate ketal substituents) exceeds ≈0.31. Above this value, the fractional change in viscosity increases with further increase in the degree of pyruvate substitution. These differences cannot arise from different degrees of conformational ordering, since the magnitude of the thermally induced order disorder transition (monitored by optical rotation at low ionic strength) is independent of pyruvate content. The temperature at the transition midpoint, however, falls with increasing degree of pyruvate substitution. This is attributed to destabilization of the ordered structure by intramolecular electrostatic repulsion between pyruvate groups, and stabilization through apolar interactions of acetate methyl groups. Viscosity-concentration relationships show changes of slope which mark the onset of macromolecular association. Association commences at lower concentrations when ionic strength and degree of pyruvate substitution are high. It is suggested that once electrostatic repulsions have been diminished at high ionic strength, association is promoted by intermolecular apolar interactions of pyruvate methyl groups, which are suitably near the periphery of the helical conformation.  相似文献   

7.
The chemical structure and the rheological behavior of the Klebsiella polysaccharide ATCC 12657 was studied and compared with data described in the literature and obtained for similar polysaccharides. The acetylated polysaccharide presents in solution a normal viscoelastic behavior with no evidence of an ordered conformation whatever the experimental conditions are. The deacetylated form can induce the formation of physical gels, in the presence of salt excess or ethanol. Microcalorimetry, optical rotation, and rheology experiments demonstrate that a thermally reversible and highly cooperative conformational transition occurs at the same temperature than a sol-gel transition. The melting of the gel and the conformational transition temperatures are dependent on the nature of cations and ionic concentration, whereas the gel strength is only influenced by polymer concentration.  相似文献   

8.
This paper concerns the study of the conformational transition of a new exopolysaccharide (YAS34) using experimental techniques such as optical rotation, conductimetric and microcalorimetric measurements as a function of temperature. The behaviors of this polysaccharide in the acid or sodium salt form are compared; a deacetylated sample is also prepared to demonstrate the role of substituents. For the native structure (never heated), a conformational transition is observed but the deacetylated polysaccharide exhibits no ordered conformation. Multidetection size exclusion chromatography (SEC) analyses and conductimetric experiments allowed to determine the nature of each conformation and the molecular dimensions. From these results, it is suggested that the native conformation is a double helix which by heating over T(m) (temperature corresponding to half conformational transition) dissociates into disordered single chains. In the acid and sodium salt forms, by cooling below T(m), an ordered conformation is restored. This conformation seems to be an intramolecular double helix 'hairpin-like turn' (called renatured conformation). Nevertheless an irreversible denaturation is obtained progressively in the sodium salt form when the time of heating over T(m) increases. The conformation of the deacetylated polysaccharide corresponds to that of a single flexible chain (disordered conformation). The conformational transition for the native conformation was studied also in relation to the polyelectrolytic character of the polysaccharide: stability as a function of salt nature and salt and polymer concentrations was investigated for the polymer initially in the sodium and acid forms.  相似文献   

9.
Deuterium oxide solutions of schizophyllan, a triple-helical polysaccharide, undergoing an order-disorder transition centered at 17 degrees C, were studied by optical rotation (OR) and heat capacity (C(p)) to elucidate the molecular mechanism of the transition and water structure in the solution and frozen states. The ordered structure at low temperature consisted of the side chains and water in the vicinity forming an ordered hydrogen-bonded network surrounding the helix core and was disordered at higher temperature. In the solution state appeared clearly defined transition curves in both the OR and C(p) data. The results for three samples of different molecular weights were analyzed theoretically, treating this transition as a typical linear cooperative transition from the ordered to disordered states and explained quantitatively if the molecular weight polydispersity of the sample was considered. The excess heat capacity C(EX)(p) defined as the C(p) minus the contributions from schizophyllan and D(2)O was estimated. In the frozen state it increased with raising temperature above 150 K until the mixture melted. This was compared with the dielectric increment observed in this temperature range and ascribed to unfreezable water. From the heat capacity and dielectric data, unfreezable water is mobile but more ordered than free water. In the solution state, the excess heat capacity originates from the interactions of D(2)O molecules as bound water and structured water, and so forth. Thus the schizophyllan triple helix molds water into various structures of differing orders in solution and in the solid state.  相似文献   

10.
Static and dynamic light scattering, viscosity, and optical rotation measurements have been made at eight different temperatures between 25 and 75 degrees C on two succinoglycan samples (sodium salt) with weight-average molecular weights M(w) of 7.14 x 10(5) and 3.54 x 10(5) (at 25 degrees C) in 0.01 M aqueous NaCl to investigate the thermally induced order-disorder conformation change of the polysaccharide. Additionally, viscometry and polarimetry have been performed for a sodium salt sample (M(w) = 4.55 x 10(5) at 25 degrees C) whose M(w), z-average radius of gyration (z)(1/2), and hydrodynamic radius R(H) in the aqueous salt had been determined previously. As the temperature increases, M(w), (z)(1/2), R(H), and the intrinsic viscosity for every sample sharply decrease around 55 degrees C where the specific rotation at 300 nm sigmoidally increases. In particular, M(w) at 25 degrees C (i.e., in the ordered helical state) is twice as large as that at 75 degrees C (i.e., in the disordered state). These findings substantiate that the ordered structure is composed of two chains and hence is a double helix. Data analysis shows that this helix at 25 degrees C is characterized by an unperturbed wormlike chain with a helix pitch of about 2 nm (per repeating unit) and a persistence length of about 50 nm and that upon heating, it dissociates directly (i.e., in all-or-none fashion) to disordered chains of a similar contour length but with a much smaller persistence length of about 10 nm. The temperature dependence of the light scattering second viral coefficient is discussed in relation to the association of disordered chains in the cooling process.  相似文献   

11.
Thermally reversible acid-induced gelation of low-methoxy pectin   总被引:1,自引:0,他引:1  
Gelation of low-methoxy pectin (DE 31.1) on cooling under acidic conditions in the absence of Ca2+ has been investigated by rheological measurements under low-amplitude oscillatory shear. The mechanical spectra obtained after 60 min at 5°C showed a progressive increase in solid-like response (increasing G′; decreasing tan δ; increasing frequency-dependence of η*) as the pH was reduced from 4.0 to 1.6, with formation of a critically crosslinked network at pH 3.0 (for a polymer concentration of 3.0 wt%). By extrapolation from X-ray fibre diffraction analysis of pectic acid, it is suggested that crosslinking occurs by association of three-fold helices. At pH values between 3.5 and 2.5 there is no detectable thermal hysteresis between the sol–gel transition on cooling and gel–sol transition on heating, and both are accompanied by a sigmoidal change in optical rotation (attributed to formation and melting of three-fold order). Substantial hysteresis is, however, observed at lower and higher pH, and is attributed to extensive aggregation as electrostatic repulsion is suppressed (below pH 2.5) and slow formation of intermolecular hydrogen bonds by protonated carboxyl groups (above pH 3.5), respectively. The transition enthalpy from DSC heating scans has a maximum value of ΔH≈11 J/g at pH 3.0, but decreases sharply at lower and higher pH, with accompanying loss of a detectable transition in optical rotation. It is suggested that the chain conformation in solution at low pH is predominantly three-fold with, therefore, little conformational change on adoption of the ordered, intermolecular structure, whereas at high pH the solution conformation is predominantly two-fold, with only limited conversion to the three-fold (acid) form on cooling.  相似文献   

12.
Measurements have been made of the shear modulus of calcium, potassium and sodium kappa carrageenate gels as a function of polysaccharide concentration and temperature. Under the experimental conditions used the efficiency of the cations in gelling the polysaccharide has been found to be Ca2+>K+>Na+. The relative gelling efficiencies of the cations is attributed to their extent of hydration which controls the solubility of the salt form of the polysaccharide. Gelation is attributed to ‘microcrystallite’ formation at localised sites on adjacent polysaccharide chains. The sharp decrease of the shear modulus on heating is attributed to localised melting of these ordered regions.  相似文献   

13.
For consideration of their conformations and interactions, carbohydrate chains can conveniently be divided into 3 classes on the basis of their covalent structure; namely periodic (a), interrupted periodic (b), and aperiodic (c) types. In aqueous solution carbohydrate chains often exist as highly disordered random coils. Under appropriate conditions, however, polysaccharides of types (a) and (b) can adopt a variety of ordered conformations. Physical methods, and in particular optical rotation, circular dichroism, and nuclear magnetic resonance, provide sensitive probes for the study of the mechanism and specificity of these disorder-order transitions in aqueous solution. Intermolecular interactions between such polysaccharide chains arise from co-operative associations of long structurally regular regions which adopt the ordered conformations. For acidic polysaccharides these cooperative associations may involve alignment of extended ribbons with cations sandwhiched between them. In other systems the interactions involve double belices which may then aggregate further, and geometric “matching” of different polysaccharide chains can also occur. These ordered, associated regions are generally terminated by deviations from structural regularity or by “kinks” which prevent complete aggregation of the molecules. The complex carbohydrate chains which occur at the periphery of animal cells have very different, aperiodic structures and although their conformations are as yet poorly understood, preliminary indications are considered.  相似文献   

14.
The conformational transitions of schizophyllan were studied in aqueous alkaline solutions by high-sensitivity differential scanning calorimetry (DSC) and optical rotation measurements. The temperature of half completion for reversible intramolecular conformational transition determined by DSC, centered at 7.4°C in water, increases to 37.2°C at 0.01M KOH with increasing alkaline concentration. The transition enthalpy per mole of the polysaccharide repeating unit is 2.62 ± 0.23 kJ mol−1 independent of the alkaline concentration. The cooperative unit size for the transition decreases with increasing alkaline concentration. Optical rotation was measured as a function of pH at 25 and 60°C. A sharp decrease in optical rotation was observed at pH = 13, which is ascribed to the triple helix-coil transition. From data obtained by DSC and optical rotation measurements, in combination with results reported previously, a phase diagram for the conformation of schizophyllan as a function of temperature and pH is proposed. The irreversibility of the triple helix to single coil transition, induced by strong alkali, was investigated as a function of polymer concentration by gel permeation chromatography and electron microscopy. The renatured samples at polymer concentrations < 1.0 mg/mL, which are prepared by dissolution in 0.25M KOH followed by neutralization with HCl, are observed as a mixture of globular, linear, and circular structures, and larger aggregates with less-defined morphology by electron microscopy. Higher concentrations lead to increased proportions of multichain clusters (aggregates). Subsequent annealing of the renatured samples at 115–120°C increases the proportion of circular species. The change in molecular weight distribution of samples that accompanies the renaturation and annealing mentioned above can be well interpreted in terms of the proportion of species having different morphology as observed by electron microscopy. © 1996 John Wiley & Sons, Inc.  相似文献   

15.
Breads with a specific volume up to, and beyond, that of conventional wheat bread can be made with rice flour (which does not contain gluten) by incorporation of hydroxypropylmethylcellulose (HPMC) and ispaghula husk (isabgol) from Plantago ovata Forsk. Isabgol disperses to a fibrillar ‘weak gel’ network stable over the temperature range of proving (fermentation with yeast) and initial heating in the baking oven. Weakening of the isabgol network at higher temperature is accompanied by thermal gelation of HPMC, so that the overall network strength (storage modulus, G′) remains virtually constant, and the gascell structure developed during proving is retained. Gelation of HPMC (and of methylcellulose) involves two distinct processes which we suggest correspond to partial dissociation and subsequent aggregation of cellulosic ‘bundles’ present in the solution state at low temperature, a mechanism directly analogous to the thermal gelation of globular proteins.  相似文献   

16.
A new extracellular microbial polysaccharide, XM-6, has been isolated from cultures of an Enterobacter species and shows unusual gelation properties of potential technological significance. The polysaccharide contains d-glucose, l-fucose and d-glucuronate in the approximate molar ratio 3:1:1. No significant amounts of acetate or pyruvate were detected. d-Glucuronate and some d-glucose are destroyed on periodate oxidation, but l-fucose and some d-glucose may be recovered intact, indicating the presence of some 1,3 linkages in the primary structure. The major oligosaccharide isolated from autohydrolysates was an aldobiuronic acid containing equal amounts of d-glucose and l-fucose.Thermally-reversible gels are formed on addition of salt to solutions of the polysaccharide. A preliminary investigation of the mechanism of gelation by optical rotation, circular dichroism, high resolution n.m.r. and mechanical spectroscopy suggests interchain association through conformationally ordered ‘junction zones’, with specific incorporation of site-bound cations within the ordered structures. In the sol state the polysaccharide shows the shear-rate and temperature dependence of viscosity typical of a disordered (‘random coil’) polymer solution. Divalent cations are, in general, more effective than monovalent cations in promoting gelation of XM-6, while trivalent cations normally cause precipitation. Within Groups I and II, optimum gelation is achieved with Na+ and Ca2+ (ionic radius ? 0·1 nm), with larger and smaller ions becoming progressively less effective. Both gel strength and melting temperature increase with increasing salt concentration.XM-6 forms gels of reasonable strength at unusually low concentrations of the polysaccharide. For example, gels comparable to those required for normal industrial or food applications may be obtained using 0·3% w/v XM-6 and 1% w/v NaCl. Gel strength increases with increasing polymer concentration but there is no systematic variation in melting point. The sol-gel transition of XM-6 is unusually sharp and, by suitable adjustment of salt concentration, can be made to occur just below body temperature (e.g. 30–35°C), with obvious implications for biomedical or food applications.  相似文献   

17.
T Hirao  T Sato  A Teramoto  T Matsuo  H Suga 《Biopolymers》1990,29(14):1867-1876
A triple helical polysaccharide schizophyllan in aqueous solution exhibited a highly cooperative transition between ordered and disordered states associated with the conformation of its side chains and nearby water molecules. The transition was followed by optical rotation and calorimetry using water containing additives such as NaOH and DMSO as solvents. The ordered state was stabilized or destabilized depending on the kind and amount of the additive employed; in particular, the addition of DMSO had a remarkable stabilizing effect. This effect was analyzed by means of a statistical mechanical theory of linear cooperative transitions, where DMSO was assumed to interact favorably with the ordered side chains. A small amount of NaOH in a solvent mixture stabilized the ordered state and made the transition curve very gradual. No molecular mechanism was elucidated to account for the role of NaOH.  相似文献   

18.
To decrease the cost of in vitro conservation of banana cv. Karpura Chakkarakeli (AAB; Mysore subgroup) without any adverse effects on cultures, expensive components of medium such as sucrose and gelling agents, i.e. phytagel or agar (90% of the total cost of the medium), were replaced with inexpensive alternates such as market sugar and isabgol, respectively (Experiment 1). In general, no significant effects of isabgol and market sugar were observed on shoot (1.0–1.3 shoots/shoot explant) and root (1.5–2.0 roots/shoot explant) regeneration. Up to 12 months, 100% of cultures survived on isabgol-media, which was significantly higher than that on agar-media (79–83%) and on phytagel-media (51–57%). Isabgol-media with or without other constituents of medium were also tested for survival of banana cultures (Experiment 2); significant differences were observed for survival of cultures (20–100%). Slow growth of the cultures on isabgol-media was attributed to low availability of free water and consequently slower rate of transport of nutrients from isabgol matrix to the plantlets than that of other media tested, as evidenced by significantly lower relative matric potentials (0.801 and 0.804) of isabgol-media. In vitro conservation-derived plants grown in the field exhibited no significant morphological variations. The total cost of medium used for in vitro conservation of banana was decreased by 59% by using isabgol as an alternate gelling agent to agar and phytagel.  相似文献   

19.
Nongelling solutions of structurally regular chain segments of agarose sulphate show disorder–order and order–disorder transitions (as monitored by the temperature dependence of optical rotation) that are closely similar to the conformational changes that accompany the sol–gel and gel–sol transitions of the unsegmented polymer. The transition midpoint temperature (Tm) for formation of the ordered structure on cooling is ~25 K lower than Tm for melting. Salt-induced conformational ordering, monitored by polarimetric stopped-flow, occurs on a millisecond time scale, and follows the dynamics expected for the process 2 coil ? helix. The equilibrium constant for helix growth (s) was calculated as a function of temperature from the calorimetric enthalpy change for helix formation (ΔHcal = ?3.0 ± 0.3 kJ per mole of disaccharide pairs in the ordered state), measured by differential scanning calorimetry. The temperature dependence of the nucleation rate constant (knuc), calculated from the observed second-order rate constant (kobs) by the relationship kobs = knuc(1 ? 1/s) gave the following activation parameters for nucleation of the ordered structure of agarose sulphate (1 mg mL?1; 0.5M Me4NCl or KCl): ΔH* = 112 ± 5 kJ mol?1; ΔS* = 262 ± 20 J mol?1 K?1; ΔG*298 = 34 ± 6 kJ mol?1; (knuc)298 = (7.5 ± 0.5) × 106 dm3 mol?1 s?1. The endpoint of the fast relaxation process corresponds to the metastable optical rotation values observed on cooling from the fully disordered form. Subsequent slow relaxation to the true equilibrium values (i.e., coincident with those observed on heating from the fully ordered state) was monitored by conventional optical rotation measurements over several weeks and follows second-order kinetics, with rate constants of (2.25 ± 0.07) × 10?4 and (3.10 ± 0.10) × 10?4 dm3 mol?1 s?1 at 293.7 and 296.2 K, respectively. This relaxation is attributed to the sequential aggregation processes helix + helix → dimer, helix + dimer → trimer, etc., with depletion of isolated helix driving the much faster coil–helix equilibrium to completion. Light-scattering measurements above and below the temperature range of the conformational transitions indicate an average aggregate size of 2–3 helices.  相似文献   

20.
The effect of a number of tetramethylammonium salts on the equilibria and kinetics of the disorder to order transition in the polysaccharide κ-carrageenan have been investigated. Data from the temperature dependence of optical rotation show that anion stabilization of the ordered form follows the lyotropic series I? > Br? > NO > Cl? > F?. Stopped-flow polarimetry was used to study the kinetics of conformational ordering following a rapid increase in salt concentration. The transition to the new equilibrium position was shown to be biphasic for all of the tetramethylammonium salts studied. The rate equation for the fast phase and the temperature dependence of the observed forward rate constant accord with a cooperative dimerization process. Activation parameters for helix nucleation, ΔH* and ΔS*, vary with both salt concentration and (at constant ionic strength) the anion type, increasing through the lyotropic series from I? to F?. The slow phase shows second-order kinetics, and is interpreted as further stabilization of the ordered form either through limited aggregation or annealing. The rate constant for the slow phase also follows the lyotropic series. Thus we have shown that both the growth and nucleation processes are anion dependent.  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