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 temperature dependence of the frequencies of121Sb nuclear quadrupole resonance in the spectra of M2Sb2SO4F6 (M = K, Rb, NH4) was studied in the temperature range from 77 to 340 K. A secondary phase transition was found above 320 K for (NH4)2SO4F6.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 1831–1833, October, 1994.This study was sponsored by the Russian Foundation for Basic Research (Grant No. 93-03-4394).  相似文献   

2.
The thermal decomposition of potassium bromate (KBrO3) has been studied as a function of particle size, in the range 53?C150???m, by isothermal thermogravimetry at different temperatures, viz. 668, 673, 678, and 683?K in static air atmosphere. The theoretical and experimental mass loss data are in good agreement for the thermal decomposition of all samples of KBrO3 at all temperatures studied. The isothermal decomposition of all samples of KBrO3 was subjected to both model fitting and model-free (isoconversional) kinetic methods of analysis. Isothermal model fitting analysis shows that the thermal decomposition kinetics of all the samples of KBrO3 studied can be best described by the contracting square equation. Contrary to the expected increase in rate followed by a decrease with decrease in particle size, KBrO3 shows a regular increase in rate with reduction in particle size, which, we suggest, is an impact of melting of this solid during decomposition.  相似文献   

3.
The persistence length lp of a polyelectrolyte chain can be represented as lp = lO + le where lO is the bare persistence and Ie is the electrostatic contribution coming from the effects of electrostatic chain self-interactions. Using a reparametrization-invariant path integral model of semiflexible polymers we find that Ie depends on the ionic strength I as IeI−1/2. This result accords with experimental observations and recent Monte Carlo simulations. Reparametrization-invariance is apparently an essential constrainet in selecting acceptable models of semiflexible polymers.  相似文献   

4.
A series of high quality 1-alkyl-3-methylimidazolium-based ionic liquids are synthesized and used for studying their surface tension. The capillary rise method is used for measuring the surface tension of I, Cl, PF6, and BF4 salts in the temperature range 298–393 K. The capillary apparatus is evacuated and sealed under vacuum. The experimental results show that surface tension of these compounds depend systematically on temperature.  相似文献   

5.
Aqueous polymerization of methyl methacrylate (MMA), initiated by the potassium bromate-thioglycollic acid (TGA) redox system, has been studied at 30 ± 0.2° C under positive pressure of nitrogen. The rate is given by K[MMA] [TGA] 0[KBrO3]x where × = 1 for lower KBrO3 concentrations and 0.5 for higher KBrO3 concentrations. The reaction has been studied over the 20–45°C range. The activation energy was found to be 65.72 kJ/mol (15.71 kcal/mol) in the investigated range of temperature. Inorganic electrolytes except MnSO4·4H2O and Na2C2O4 depress both the rate of polymerization and the maximum conversion. All the alcohols (viz., MeOH, EtOH, iso-PrOH, tert-BuOH) and acetone depress the rate of polymerization as well as the maximum conversion.  相似文献   

6.
Analysis of enthalpy of dilution data for the system NaCl/H2O at 298.15°K with correlating equations presently in widespread use shows that these do not satisfactorily fit the data at the lowest available concentrations. A new approach based on the extended Debye-Hückel theory is suggested. It is shown that a plot of the ratio of the apparent molar enthalpy of dilution to the change in ionic strength, L/I, versus (I i 1/2I f 1/2 )/I, should give a straight line at low enough concentration with a slope 2S H/3. The intercept is related to the Debye distance parameterA and the coefficient of the first virial correction term. The quantityS H is the limiting Debye slope as calculated from the properties of the pure solvent. These expectations are substantiated by the fits. Values of the parameters are compared with older estimates, and it is concluded that the choiceA=0 is reasonable.  相似文献   

7.
The aqueous polymerization of methacrylamide (I) initiated by KBrO3–thioglycolic acid (TGA) has been studied at 30 ± 0.2°C in nitrogen. The rate is given by K[M]1.19 [thioglycolic acid]1 [KBrO3]0.53 for 10–15% conversion. Activation energy was found to be 53.96 kJ/mole (12.92 kcal/mole) in the investigated range of temperature 30–45°C. The role of addition of a series of aliphatic alcohols and some salts was also determined. The kinetics of polymerization was followed iodometrically.  相似文献   

8.
Thermal and dielectric properties of 2,2’-dihydroxybenzophenone were studied in relation with the potential progress of crystal nucleation and growth below the ordinary glass transition temperature, T. Differential scanning calorimetry was carried out in a range 100-350 K. The α glass transition was found to occur at T=239 K. Crystallization and fusion were observed to take place when the sample was cooled down to 103 K, but not observed when cooled to 203 K. Crystal nucleation was interpreted as having happened during annealing for a short time at 103 K which is much below the T. Heat capacities were measured in a range 7-350 K by an intermittent heating method with an adiabatic calorimeter. The temperature, enthalpy and entropy of fusion were determined to be 334.46 K, 20.07 kJ mol-1 and 60.01 J K-1mol-1, respectively. Crystal growth was found to proceed even at 220 K below the T, but no glass transition was detected below 220 K. Dielectric losses were measured in a temperature range of 100-250 K and a frequency range of 30Hz-10 kHz. β-Relaxation process was found dielectrically with the activation energy of 22.6 kJ mol-1, and the corresponding glass transition was expected to occur at 76.9 K. It is discussed, based on the “structurally ordered clusters aggregation” model for supercooled liquids and glasses, that the β process is potentially attributed to the crystal nucleation progressing at 103 K.  相似文献   

9.
Isothermal decomposition of KBrO3 has been studied as a function of concentration of the dopants, SO4 2- and Ba2+ by isothermal thermogravimetry in the temperature range 668 - 683 K. The rate law and the activation energies remained unaltered by doping. The results suggest a diffusion-controlled mechanism, the diffusing species being both K+ and BrO3 -.  相似文献   

10.
11.
An ionic liquid (IL) EPReO4 (N‐ethylpyridinium rheniumate) was prepared. The density and surface tension values of the IL were determined in the temperature range of 293.15–343.15 K. The ionic volume and surface entropy of the IL were estimated by extrapolation, respectively. In terms of Glasser's theory, the standard molar entropy and lattice energy of the IL were estimated, respectively. Using Kabo's and Rebelo's methods, the molar enthalpy values of vaporization of the IL, ΔglH0m (298 K), at 298 K and, ΔglH0m (Tb), at hypothetical normal boiling point were estimated, respectively. According to the interstice model, the thermal expansion coefficient of IL EPReO4 (α) was calculated and compared with experimental value, finding their magnitude order is in good agreement by 8.98%.  相似文献   

12.
Positron annihilation measurements as a function of temperature and time have been carried out on a poly(butadiene). The measurements were performed at several temperature points from 14 to 225 K. The measurement time was several hours to four days. The analysis of data shows the following features:
(i) the value of τ3 does not depend on the rate of cooling or time,
(ii) the value of I3 depends on the rate of cooling and the history of thermal treatment,
(iii) the dependence of I3 on time can be described by Debye function. But the rise in I3 is observed at very low temperatures,
(iv) the I3 decays to value of I3 observed during very slow cooling.

Article Outline

1. Introduction
2. Experiments
3. Results
4. Discussion
5. Conclusions
6. Uncited Reference
Acknowledgements
References

1. Introduction

If a glass is formed by rapid cooling of a super-cooled liquid to a temperature below the glass–liquid transition temperature, Tg, its properties will not be static, but will relax toward values characteristic of the corresponding “equilibrium” supercooled liquid as extrapolated from above to below Tg. This process named as structural relaxation or “physical aging” is of great practical importance because of its relevance to the designing and engineering of amorphous materials with desired properties. The relaxation property and transport phenomena of disordered polymers can be explained within the free-volume concept (Ferry, 1980). However, an unsettled problem is a way of quantifying the free-volume properties, such as the free-volume fraction, the average and the distribution of the free-volume size. In the last decade, the positron annihilation lifetime spectroscopy (PALS) technique has been recognised as a useful method to detect atomic scale free-volume holes of polymers ( Schrader and Jean, 1988). This technique involves using a positron source, mostly 22Na, to emit positrons into the sample. But these positrons and the accompanying gamma–quanta have sufficient energy (average positron energy 200 keV, gamma 890 keV) to induce radiation effects, and the positron probe can thus affect the sample being investigated during PALS experiments.The basic assumption of positron annihilation lifetime spectroscopy (PALS) data interpretation in terms of the free-volume concept is the proportionality of the intensity of long-lived ortho-positronium (o-Ps) component, I3, to the concentration of free-volume holes (Kobayashi et al., 1989). However, there are different findings regarding the influence of external factors on the “true” intrinsic value of I3. Its variation with the measurement time is regarded as a manifestation of the relaxation of free-volume fraction. On the other hand, the decrease in I3 with PALS measurement time is related to the activity of the positron source and the chemical processes in the positron spur, e.g., formation of free radicals. There are PALS measurements on semi-crystalline samples (Suzuki et al., 1996), observing the I3 increase with elapsed time when the temperature of the sample is below Tg.All these reports indicate that the o-Ps formation in polymers is more complicated and the basic assumption of PALS interpretation may be questionable.In this work, PALS results will be presented on the amorphous cistrans-1,4-poly(butadiene), cistrans-1,4-PBD, in a wide temperature range from 14 to 350 K. The aim of this paper is the study of the influence of temperature, time and sample history on the intensity I3, life time of o-Ps, τ3, as well as the S-parameter from Doppler broadening measurements.

2. Experiments

The PALS experiments were conducted using a conventional fast–fast coincidence system having a time resolution of ca. 320 ps (FWHM). Cistrans-1,4-PBD has a molecular weight of Mw = 2 × 104, the glass transition temperature Tg = 178 K (Zorn et al., 1995). The isomer composition was 41% cis, 52% trans and 7% vinyl form. This isomer composition was chosen to avoid a crystallisation process on the PBD sample (Zorn et al., 1995).The positron source, consisting of 2 MBq 22N a sealed between two 3.5 μm Ni foils, was sandwiched between polymer discs, each of about 3 mm thick and with a diameter of 10 mm. At a chosen temperature, each spectrum was accumulated for 1 h, resulting in a total number of counts of about 1.14 mil. At least, two such spectra were recorded at each temperature point.The 22Na source–sample assembly was mounted on a closed cycle helium gas refrigerator. The assembly was kept in a rotary pump vacuum of about 4 Pa. Automatic temperature regulation was used during all the measurements and the temperature was controlled within ±1 K. Several different temperature scans on the specimens were performed. The first sequence (heating) was the following: I3, τ3 were first evaluated at room temperature of 300 K immediately after the source installation. Then, fast cooling to the temperature of 40 K at a rate 4 K/min was performed and the temperature increased in steps of 10 K. The second sequence (cooling) started at 300 K, then the temperature decreased to 14 K in steps of 10 K.For the PALS measurement as a function of time, the PBD was annealed in the chamber at 300 K for several hours, then cooled to the measurement temperature and the measurement began immediately.The positron life-time spectra were measured as a function of the elapsed time at 14 different temperature points below and above Tg.The PALS data were also accumulated during heating of the samples to 300 K and cooling of PBD to chosen temperature below 300 K. The total irradiation time of 1080 h was divided between PALS and calibration (Bi) measurements. To clearly describe the thermal history of the experiment, the time dependence of I3 and τ3 is shown in Fig. 1 and Fig. 2, respectively. The values of I3 and τ3 at room temperature were the same despite the long irradiation time and complicated thermal history. This indicates that a possible irradiation damage does not influence the annihilation observables.  相似文献   

13.
Kinetics and mechanism of the Os(VIII) catalysed oxidation of crotonic acid (CA) by KBrO3 in alkaline medium have been investigated. Zero order dependence in [KBrO3] was observed, while first order with respect to CA in its lower concentration range tends to zero order at its higher concentration range. The order in [Os(VIII)] was found to be unity and a positive effect of [OH] was observed. Variation of the ionic strength (μ) and dielectric constant of the medium and addition of Hg(OAc)2 (used as Br scavenger) had an insignificant effect on the rate of reaction. Thermodynamic parameters have also been calculated and reported. A suitable mechanism consistent with the observed kinetic results has been suggested and the related rate law deduced.  相似文献   

14.
The superionic properties of the compounds RbAg4I5, KAg4I5 and KCu4I5 have been investigated by powder neutron diffraction and complex impedance spectroscopy. RbAg4I5 and KAg4I5 have room-temperature ionic conductivities of σ=0.21(6) and 0.08(5) Ω−1 cm−1, respectively, which increase gradually on increasing temperature. KCu4I5 is only stable in the temperature range between 515(5) K and its melting point of 605 K, and its ionic conductivity is σ=0.61(8) Ω−1 cm−1, at T=540 K. At lower temperatures, KCu4I5 disproportionates into KI+4CuI and the ionic conductivity falls by over three orders of magnitude. Least-squares refinements of the powder neutron diffraction data for RbAg4I5 at ambient temperature confirm the reported structure (space group P4132, Z=4, a=11.23934(3) Å), though with some differences in the preferred locations of the mobile Ag+. KAg4I5 and KCu4I5 are found to adopt the same basic structure as RbAg4I5, with the I− forming a β-Mn-type sublattice, with the K+ located in a distorted octahedral environment and the Ag+(Cu+) predominantly distributed over two sites which are tetrahedrally co-ordinated to I. The implications for the conduction mechanism within these compounds are discussed, using a novel maximum entropy difference Fourier technique to map the distribution of the Ag+(Cu+) within the unit cell.  相似文献   

15.
The rate constant for the reaction I(2P1/2) + CH3I → I2 + CH3 has been reevaluated taking into account both collisional deactivation of excited iodine atoms and loss of I2 by I2 + CH3 → I + CH3I. The reevaluation is based upon data obtained (R. T. Meyer), J. Chem. Phys., 46 , 4146 (1967) from the flash photolysis of CH3I using time-resolved mass spectrometry to measure the rate of I2 formation. Computer simulations of the complete kinetic system and a closed-form solution of a simplified set of the differential equations yielded a value of 6(± 4) × 106 1./mole-sec for the excited iodine atom reaction in the temperature region of 316 to 447 K. A slight temperature dependence was observed, but an activation energy could not be evaluated quantitatively due to the small temperature range studied. An upper limit for the collisional deactivation of I(2P1/2) with CH3I was also determined (2.4 × 107 1./mole-sec).  相似文献   

16.
Relative enthalpies for low-and high-temperature modifications of Na3FeF6 and for the Na3FeF6 melt have been measured by drop calorimetry in the temperature range 723–1318 K. Enthalpy of modification transition at 920 K, δtrans H(Na3FeF6, 920 K) = (19 ± 3) kJ mol−1 and enthalpy of fusion at the temperature of fusion 1255 K, δfusH(Na3FeF6, 1255 K) = (89 ± 3) kJ mol−1 have been determined from the experimental data. Following heat capacities were obtained for the crystalline phases and for the melt, respectively: C p(Na3FeF6, cr, α) = (294 ± 14) J (mol K)−1, for 723 = T/K ≤ 920, C p(Na3FeF6, cr, β) = (300 ± 11) J (mol K)−1 for 920 ≤ T/K = 1233 and C p(Na3FeF6, melt) = (275 ± 22) J (mol K)−1 for 1258 ≤ T/K ≤ 1318. The obtained enthalpies indicate that melting of Na3FeF6 proceeds through a continuous series of temperature dependent equilibrium states, likely associated with the production of a solid solution.   相似文献   

17.
Pyrolysis of cyclohexane was conducted with a plug flow tube reactor in the temperature range of 873-973 K. Based on the experimental data, the mechanism and kinetic model of cyclohexane pyrolysis reaction were proposed. The kinetic analysis shows that overall conversion of cyclohexane is a first order reaction, of which the rate constant increased from 0.0086 to 0.0225 to 0.0623 s-1 with the increase of temperature from 873 to 923 to 973 K, and the apparent activation energy was determined to be 155.0±1.0 kJ mol-1. The mechanism suggests that the cyclohexane is consumed by four processes:the homolysis of C-C bond (Path I), the homolysis of C-H bond (Path II) in reaction chain initia- tion, the H-abstraction of various radicals from the feed molecules in reaction chain propagation (Path III), and the process associated with coke formation (Path IV). The reaction path probability (RPP) ratio of XPath I:XPath II : XPath III : XPath IV was 0.5420:0.0045:0.3897:0.0638 at 873 K, and 0.4336 : 0.0061 : 0.4885 : 0.0718 at 973 K, respectively.  相似文献   

18.
In order to study the atomic jump motions in the high-temperature solid phase of LiBH4, we have measured the 1H and 11B nuclear magnetic resonance (NMR) spectra and the 1H, 7Li and 11B spin–lattice relaxation rates in this compound over the resonance frequency range of 14–34.4 MHz. In the temperature range 384–500 K, all the spin–lattice relaxation data are satisfactorily described in terms of a thermally activated jump motion of Li ions with the pre-exponential factor τ0=1.1×10−15 s and the activation energy Ea=0.56 eV. The observed frequency dependences of the spin–lattice relaxation rates in this temperature range exclude a presence of any distributions of the Li jump rate or any other jump processes on the frequency scale of 107–1010 s−1. The strong narrowing of the 1H and 11B NMR lines above 440 K is consistent with the onset of diffusive motion of the BH4 tetrahedra.  相似文献   

19.
In pH 3.8 acetic acid‐sodium acetate (HAC‐NaAC) buffer solution, laccase exhibited a strong catalytic effect on the H2O2 oxidation of I ? to form I2, and I2 combined with excess I ? to form I3? that reacted with cationic surfactants of tetradecyl dimethylbenzyl ammonium chloride (TDMAC) to produce the (TDMAC‐I3)n association complex particles, which exhibited a strong resonance scattering (RS) peak at 468 nm. Under the chosen conditions, as the concentration of laccase activity increased, the RS intensity at 468 nm (I468 nm) increased linearly. The increased RS intensity ΔI468 nm was linear to laccase activity in the range of 0.08–0.96 U/mL, with a regression equation of ΔI468 nm?88.8U?1.9, and a detection limit of 0.02 U/mL laccase. This proposed method was applied to detect laccase activity in waste water, with satisfactory results.  相似文献   

20.
Heat capacities in the liquid phase C l of methylbenzeneamines and heat capacities in the solid phase C s of benzenediols and of 4-methylbenzeneamine were measured by commercial Setaram heat conduction and power compensated calorimeters. Results obtained cover the following temperature range (depending on the compound and state of aggregation): 2-methylbenzeneamine 313 to 371 K, 3-methylbenzeneamine 263 to 371 K, 4-methylbenzeneamine 133 to 353 K, 1,2-benzenediol 153 to 353 K, 1,3-benzenediol 173 to 353 K, 1,4-benzenediol 133 to 403 K. The heat capacity data obtained in this work were merged with experimental data from literature, critically assessed and sets of recommended data were developed by correlating selected data as a function of temperature. This revised version was published online in August 2006 with corrections to the Cover Date.  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