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.
Summary : Temperature-sensitive hydrogels undergo a volume phase transition (VPT) when heated above a critical temperature Tc. For the poly(N-isopropyl acrylamide) (PNIPA)-water system, Tc. = 34 °C. Below Tc the gels are transparent and highly swollen. On warming above Tc they promptly turn white and start to deswell. The rate of deswelling, however, can be orders of magnitude slower than that of swelling below Tc. The unstable intermediate structure above Tc, can retain the solvent and conserve the sample volume for may days, even with millimetre-sized samples. Light scattering observations of the internal structure of these gels above Tc are precluded by their strong turbidity. Small angle X-ray scattering measurements (SAXS), on the other hand, are less subject to multiple scattering as X-rays penetrate more easily into the bulk material. Conventional (incoherent) SAXS observations reveal intense scattering from smooth internal water-polymer interfaces with an estimated surface area of about 7 m2/g in the swollen gel. The dynamics in the off-equilibrium high temperature state, investigated by X-ray photon correlation spectroscopy (XPCS), displays a relaxation rate that is linearly proportional to the wavevector q, rather than to q2 as in diffusion processes. The physical origin of this relaxation is consistent with jamming, a phenomenon that is common in other disordered systems.  相似文献   

2.
The binary liquid mixture of triethylamine + water (TEA-W) has a lower consolute point at a critical composition of 32.27 mass % triethylamine. Starting at a temperature within the one-phase region, the electrical conductivity of a sample of this mixture with addition of (K+, Cl) ions was measured and found to be accurately described by the Vogel-Fulcher-Tammann (VFT) law. Before that, for the pure system, in a temperature range ΔT = Tc − T < 2 °C where Tc is the critical temperature, the electrical conductivity (σ) exhibits a monotonous deviation from the VFT behaviour. This anomaly is finite at Tc. The asymptotic behaviour of the electrical conductivity anomaly is described by a power law t1−α, where t is the reduced temperature |(TTc)/Tc| and α is the critical exponent of the specific heat anomaly at constant pressure. For the electrolyte mixtures, by combining the viscosity and the electrical conductivity data, the value of the computed Walden product has been determined and the salt dissociation degrees as well as the Debye screening length have been estimated.  相似文献   

3.
《Colloids and Surfaces》1988,29(1):71-87
The state of matter within a fluid layer near a critical point differs profoundly in many aspects from states further removed from the critical point. The interfacial tension tends to vanish, the interface thickens, and long-range concentration fluctuations exist. Because of these effects, critical phenomena have been investigated as possible sources of instability in thin films. Shear-field coalescence studies have been performed between phases of a simple ternary system containing no surfactant as a function of distance from a critical point. The coalescence efficiency was measured as a function of temperature through time dependent photomicrographic analysis of emulsion samples within a shear-field coalescence cell. The apparatus, procedure, and analysis are outlined. No evidence was found for the promotion of coalescence by critical phenomena for values of reduced temperature, Tr = (T − Tc)/Tc, down to 4×10−4.  相似文献   

4.
Intensity of light, I(q,t), scattered from homogeneous aqueous solutions, of nanoclay (Laponite) and protein (gelatin‐A), was studied to monitor the temporal and spatial evolution of the solution into a phase‐separated nanoclay–protein‐rich dense phase, when the sample temperature was quenched below spinodal temperature, Ts (=311 ± 3 K). The zeta potential data revealed that the dense phase comprised charge‐neutralized intermolecular complexes of nanoclay and protein chains of low surface charge. The early stage, t < 500 s, of phase separation could be described adequately through Cahn‐Hilliard theory of spinodal decomposition where the intensity grows exponentially, I(q, t) = I0 exp.(2R(q)t). The wave vector, q dependence of the growth parameter, R(q) exhibited a maxima independent of time. Corresponding correlation length, 1/qc = ξc was found to be ≈75 ± 5 nm independent of quench depth. In the intermediate regime, anomalous growth described by I(q, t) ~ tα with α = 0.1 ± 0.02 independent of q was observed. Rheological studies established that there was a propensity of network structures inside the dense phase. Isochronal temperature sweep studies of the dense phase determined the melting temperature, Tm = 312 ± 4 K, which was comparable with the spinodal temperature. The stress‐diffusion coupling prevailing in the dense phase when analyzed in the Doi‐Onuki model yielded a viscoelastic correlation length, ξv determined from low‐frequency storage modulus, G0kB T/ξ, which was ξv ≈ 35 ± 3 nm indicating 2ξv ≈ ξc. It is concluded that the early stage of phase separation in this system was sufficiently described by linear Cahn‐Hilliard theory, but the same was not true in the intermediate stage. © 2010 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 48: 555–565, 2010  相似文献   

5.
It is shown that in the vicinity of the critical point of a binary polar mixture the electronic spectral linewidth increases when approaching the critical point. The linewidth obeys the law of Ao[(T-Tc)/Tc], where γ is the universal critical index of dispersion of concentration fluctuations.  相似文献   

6.
The electrical conductivity of highly concentrated binary ionic mixtures of ethylammonium nitrate in n-octanol at a critical salt mole fraction x = 0.766 and at an off-critical one x = 0.908 was measured over an extended temperature range above the critical consolute point. Far from the critical temperature T c, the conductivity is accurately described by the Vogel–Fulcher–Tammann (VFT) law. However, in a temperature range T = (TT c) 3 K, the conductivity exhibits a monotonous deviation from the VFT behavior. This anomaly is finite at T c and, for the critical mixture, its amplitude is 0.23% of (T c). The asymptotic behavior of the conductivity anomaly is described by a power law (1 – ), with = (TT c)/T c, the reduced temperature, and , the critical exponent of the specific heat anomaly at constant pressure. This critical anomaly is similar to the one observed in other highly concentrated critical electrolytes. The degree of dissociation of the salt for the critical mixture, diss 0.78 ± 0.04, is estimated from the value of the Walden product computed at T c, and accounts for the effective free ion concentration in the reduced critical coordinates of the system.  相似文献   

7.
The results of an experimental study of the kinetics of structural relaxation of amorphous poly(ethylene terephthalate) are reported. Samples were prepared by ultraquenching the melt on rotating stainless-steel discs. Two types of measurements by differential scanning calorimetry were made: (1) the dependence of the “fictive” (or “structural”) temperature Tf(q?) introduced by Tool, on the cooling rate q? and (2) the dependence of the glass transition temperature Tg on the heating rate q+. In this way the value x = 0.47 was obtained for the dimensionless parameter proposed by Narayanaswamy.  相似文献   

8.
For any particular fluid, the set of three critical constants (CC) – pressure Pc, temperature Tc and molar volume Vc – has a central importance in defining the physical behaviour of the fluid in the gas and liquid states. However, little attention seems to have been paid in the past to the relations between the CC of different substances. In the present paper, some simple and apparently novel relations have been found between the three CC for the set of four noble gases: Ne, Ar, Kr, Xe. Defining the critical quotient Qc ≡ RTc/Pc (where R is the Gas Constant) the correlations may be summarised by the dual equation: (Vc/cm3 mol−1) = 27 + 0.31 (Tc/K) = 3.3 + 0.280 (Qc/cm3 mol−1), which describes the CC data for the quartet Ne–Xe with an average uncertainty of 0.5%. Regarding the other two noble gases, the two isotopes of the lightest member, 3He and 4He, show the deviations from these relations that are expected from quantal effects and their low molar masses; while for the heaviest member, Rn, the correlations enable a value of 145(5) cm3 mol−1 to be estimated for Vc that is not otherwise well defined in the literature. By contrast, and contrary to the general assumption, the second lightest member, Ne, apparently does not show appreciable quantal effects in the area, so that Ne–Xe may be considered together as a group. These correlations are compared with the behaviour of a selection of polyatomic fluids; in these comparisons, the NG dual correlation equation provides a reference line defining the presumed simplest behaviour. This and related areas show a “Residual Volume Effect”, in that extrapolating the equivalent temperature and energy parameters to zero for the state of zero-mass point particles, referred to here as the hypothetical element zeronium (Ze), the system in each case still has a finite intercept; this intercept amounts to essentially 34% of the average volume for the present quartet Ne–Xe, rather than the zero volume expected for this condition.  相似文献   

9.
This paper presents measurements of Δ/E2 versus T for the critical solution nitroethane in cyclohexane. The Δ/E2 for the temperature near the critical temperature Tc has a positive value and satisfies the following dependence: Δ/E2 ∝ 1/(T ? Tc)λ where λ = 0.34.  相似文献   

10.
We measured the cloud-point curves of eight-arm star polystyrene (sPS) in methylcyclohexane (MCH) for polymer samples of three total molecular masses [weight-average molecular weight (Mw) × 10−3 = 77, 215, or 268]. We found a downward shift of 5–15 K in the critical temperature (Tc) of the star polymer solutions with respect to linear polystyrene (PS) solutions of the same Mw. The shift in Tc became smaller as Mw increased. The critical volume fraction for eight-arm sPS in MCH was equal within experimental uncertainty (10–40%) to that of linear PS in MCH. For sPS of Mw = 77,000 in MCH, we studied the mass density (ρ) as a function of temperature (T). As for linear polymers in solution, the difference in ρ between coexisting phases (Δρ) could be described over t = (TcT)/Tc for 1.1 × 10−4 < t < 4.7 × 10−3 with the Ising value of the exponent β in the expression Δρ = B tβ. Both ρ(T) above Tc and the average value of ρ below Tc were linear functions of temperature; no singular corrections were observed. The measurements of the shear viscosity (η) near Tc for sPS (Mw = 74,000) in MCH indicated a strong critical anomaly in η, but the data were not precise enough for a quantitative analysis. © 2003 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 129–145, 2004  相似文献   

11.
A method to define the Cubic Equation of State (CES) of a simple substance is presented in this work. CES is constructed with only three parameters of the fluid, namely, the critical compressibility ZcPcvc/RTc, the acentric factor ω ≡ − log  (P(sat)/Pc) − 1 (where P(sat) is the saturated vapor pressure), and the saturated vapor volume v(sat) at the temperature T(sat)/Tc = 0.7 (where Tc is the critical temperature, vc is the critical volume, and Pc is the critical pressure). The resulting CES is unique for each substance and, in general, it is different from other known CES in the literature.  相似文献   

12.
《Thermochimica Acta》1987,112(2):215-220
The Law-Lielmezs (L-L) modification of the Van der Waals equation of state: P = RT/(V-b)-a(T)/V2 where: a(T) = a(Tca(Tc·a(T1) and: a(T1) = 1 + pT1q has been extended to include unsaturated states in terms of a correcting function Cf(such that the α(T1) term becomes: a(T1) = 1 + pCfT1q The proposed extension has been compared with the results obtained by the use of the original Van der Waals equation of state.  相似文献   

13.
The solution of the exponential integral at linear heating for the general case that the activation energy linearly depends on temperature according toE(T)=E 0+RBT is
\fracAqò0T TB exp( - \fracE0 RT ) dT = \fracAq( \fracRTB + 2 E0 + (B + 2)RT ) exp( - \fracE0 RT ).\frac{A}{q}\int\limits_0^T {T^B \exp \left( { - \frac{{E_0 }}{{RT}}} \right) dT = \frac{A}{q}\left( {\frac{{RT^{B + 2} }}{{E_0 + (B + 2)RT}}} \right)} \exp \left( { - \frac{{E_0 }}{{RT}}} \right).  相似文献   

14.
The paper studies the steric eflect of terminal substituents on the temperature Tc of the phase transition nematic liquid crystal-isotropic liquid in mesogenic azomethins with one or two noninteracting or strongly interacting conformation degrees of freedom. The linear dependence Tc(Q) has been confirmed, where Q=(cos2 N) is the conformation paramefer of the molecular ensemble; N is the angle between the planes of the bridging group CH=N and the aniline ring in the benzylideneaniline fragment. Direct and indirect steric eflects of lateral substituents are shown to take place, which explain an anomalous variation of Tc in some compounds upon substitutions in them.L. V. Kirensky Institute of Physics, Siberian Branch, Russian Academy of Sciences. Translated fromZhurnal Strukturnoi Khimii, Vol. 34, No. 4, pp. 89–97, July–August, 1993.Translated by L. Smolina  相似文献   

15.
16.
Solid phases FexNi1-x(Htrz)3(NO3)2 · H2O (0.4 x 0.8$) and Ni(Htrz)3(NO3)2 · H2O were synthesized and studied. The phases were studied by means of magnetochemistry, powder Xray difraction analysis, and electronic and IR spectroscopy. The heterometallic phases are described by the stoichiographic method of differentiating dissolution (DD). The values of x were determined by two methods — atomic absorption and DD. Magnetochemical data showed that the solid phases exhibit a hightemperature 1A1 5T2 0.5 x 0.8 and disappears at x = 0.4. The spin transition is accompanied by thermochromism (color changed from pink to white at 0.6 x 1 and from pink to light lilac at x = 0.5). A decrease in x leads to a decrease in the temperature of the forward (under heating Tc ) and reverse (under cooling Tc ) transitions, a decrease in hysteresis value ( Tc), and a smearing of the spin transition.  相似文献   

17.
Length scale hierarchy in gelatin sol, gel, and coacervate (induced by ethanol) phases, having same concentration of gelatin in aqueous medium (13% w/v), has been investigated through small angle neutron scattering and rheology measurements. The static structure factor profile, I(q) versus wave vector q, was found to be remarkably similar for all these samples. This data could be split into three distinct q‐regimes: the low‐q regime, Iex(q) = Iex(0)/(1+q2ζ2)2 valid for q < 3Rg?1; the intermediate q‐regime, I(q) = I(0)/(1+q2ξ2) for 3Rg?1 < q < ξ?1; and the asymptotic regime, I(q) = (c/q) exp(?Rc2q2/2) for q > ξ?1. Consequently, three distinct length scales could be deduced from structure factor data: (a) inhomogeneity of size, ζ = 20 ± 1 nm for all the three phases; (b) average mesh size, ξ0 = 2.6 ± 0.2 nm for sol and gel, and smaller mesh size, ξos = 1.2 ± 0.2 nm for coacervate; and (c) cross section of gelatin chains, Rc = 0.35 ± 0.04 nm. In addition, the structure factor data obtained from coacervating solution analyzed in the Guinier region, I(q) = exp(?q2Rg2/3), yielded value of typical radius of gyration of clusters, Rg ≈ 69 nm that indicated existence of triple‐helices of length, L ≈ 239 nm; (d) Frequency and temperature sweep measurements conducted on coacervate samples revealed two other length scales: (e) viscoelastic length, ξve = 14 ± 2 nm and (f) correlation length at melting, ξT = 500 ± 70 nm. Thus, existence of six distinct length scales, (a–f), ranging from 1.2 to 500 nm has been established in the coacervate phase of gelatin–ethanol–water system. Results are discussed within the framework of Landau‐Ginzburg treatment of dynamically asymmetric systems (Prog Theor Phys 1977, 57, 826; Phys Rev A 1991, 44, R817; J Phys II (France) 1992, 2, 1631). © 2006 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 44: 1653–1667, 2006  相似文献   

18.
The vapour pressure and orthobaric molar volumes of tetramethylsilane have been messured from 373 K to the critical temperature, and the critical temperature (IPTS-68: To = 448.64 K), critical pressure (pc = 2821 kPa), and critical molar volume (Vmc = 361 cm3 mol?1) have been determined.  相似文献   

19.
The discovery of superconductivity in H3S at 203 K marked an advance towards room-temperature superconductivity and demonstrated the potential of H-dominated compounds to possess a high critical temperature (Tc). There have been numerous reports of the H-S system over the last five years, but important questions remain unanswered. It is crucial to verify whether the Tc was determined correctly for samples prepared from compressed H2S, since they are inevitably contaminated with H-depleted byproducts. Here, we prepare stoichiometric H3S by direct in situ synthesis from elemental S and excess H2. The Im m phase of D3S samples exhibits a Tc significantly higher than previously reported values (ca. 150 K), reaching a maximum Tc of 166 K at 157 GPa. Furthermore, we confirm that the sharp decrease in Tc below 150 GPa is accompanied by continuous rhombohedral structural distortions and demonstrate that the Cccm phase is non-metallic, with molecular H2 units in the crystal structure.  相似文献   

20.
The liquid–liquid phase‐separation (LLPS) behavior of poly(n‐methyl methacrylimide)/poly(vinylidene fluoride) (PMMI/PVDF) blend was studied by using small‐angle laser light scattering (SALLS) and phase contrast microscopy (PCM). The cloud point (Tc) of PMMI/PVDF blend was obtained using SALLS at the heating rate of 1 °C min?1 and it was found that PMMI/PVDF exhibited a low critical solution temperature (LCST) behavior similar to that of PMMA/PVDF. Moreover, Tc of PMMI/PVDF is higher than its melting temperature (Tm) and a large temperature gap between Tc and Tm exists. At the early phase‐separation stage, the apparent diffusion coefficient (Dapp) and the product (2Mk) of the molecules mobility coefficient (M) and the energy gradient coefficient (k) arising from contributions of composition gradient to the energy for PMMI/PVDF (50/50 wt) blend were calculated on the basis of linearized Cahn‐Hilliard‐Cook theory. The kinetic results showed that LLPS of PMMI/PVDF blends followed the spinodal decomposition (SD) mechanism. © 2008 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 46: 1923–1931, 2008  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