首页 | 官方网站   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.
研究了聚碳酸酯(PC)和PC/ABS高分子材料的疲劳裂纹扩展规律,利用改进柔度法测量其裂纹扩展速率,采用扫描电子显微镜(SEM)观察其断口形貌,分析疲劳裂纹扩展机理.在较大裂纹扩展速率(10-6~10-3mm/cycle)范围内,PC/ABS的疲劳裂纹扩展速率可以用Paris公式da/dN=9·5587×10-5(ΔK)2·88381来描述.高分子材料PC的疲劳裂纹扩展速率约为高分子材料PC/ABS的3倍.高分子材料PC/ABS疲劳裂纹面上的特征以韧窝为主,较低裂纹扩展速率对应较小的韧窝,较高裂纹扩展速率对应较大的韧窝.高分子材料PC疲劳裂纹面有明显的不连续裂纹扩展带,其裂纹面相对较平.  相似文献   

2.
R. CurciF. Di Furia 《Tetrahedron》1972,28(14):3905-3913
The influence of OH concentration, in the solvent medium dioxane-water 40:60 at 25·0, on the oxidation rates of diphenylphosphine oxide (DPPO) by t-BuOOH, H2O2 and p-nitroperoxybenzoic acid, and of bis-p-tolylphosphine oxide by t-BuOOH has been investigated. For the oxidation of the phosphorus substrates by the hydroperoxides above, the rate law found (R = k2″[Ar2PHO] [ROO]) differs from the rate law confirmed to hold for the oxidation of DPPO by peroxyacids in alkaline media, i.e.: R = (k2″ + k3[OH−])[Ar2PHO] [RCO3]. This is interpreted on the basis of one common reaction mechanism involving formation of an intermediate of similar structure, wherein a change in the rate controlling step is likely to occur on passing from p-O2N·C6H4CO3 to t-BuOO. In the same solvent medium, pKa′ values for t-BuOOH and H2O2 have been estimated.  相似文献   

3.
The purpose of this work is to characterize the mechanical behavior of blends of polycarbonate (PC) and acrylonitrile-butadiene-styrene (ABS) during monotonic and cyclic loading. Compression experiments were performed using a SHIMADZU universal testing machine (10−4 to 10−2 s−1) and a split Hopkinson pressure bar (1600–5000 s−1), with, the test temperatures ranging from 293 to 353 K. The influence of the rate and temperature on the deformation of PC/ABS is discussed in detail. Based on the investigation of numerous constitutive models, a phenomenological model called DSGZ was chosen to describe the compression behavior of PC/ABS. This model could not accurately reproduce the deformation of polymers at high strain rates when utilizing the same material coefficients for the low and high strain–rate deformations. In addition, this model was unable to capture the deformation features during unloading and subsequent reloading when adopting the original stress–strain updating algorithm. Hence, some improvements to the model have been implemented to better predict the deformation. Finally, the model predictions are shown to be consistent with the experimental results.  相似文献   

4.
We report the synthesis of a novel bistriazene, 4,4′-bis(3-(4-phenylthiazol-2-yl)triazenyl)biphenyl (BPTTBP), and its highly sensitive color reaction with Hg2+. The new reagent was synthesized in good yield by coupling 2-amino-4-phenylthiazole with 4,4′-biphenyldiamine bisdiazonium salt. Using a blend of surfactants N-cetylpyridinium chloride (CPC) and polyethylene glycol n-octanoic phenyl ether (OP) as a micelle sensitizer, the red colored reagent assembles with Hg2+ in pH 9.8 borate buffer according to a 1:1 stoichiometry, forming a blue oligomeric/polymeric chelating complex with a high apparent stability constant (1.1 × 108 M−1). Whereas the maximum absorption of reagent occurs at 510 nm with an extinct coefficient of 1.35 × 104 M−1 cm−1, the complex absorbs at 611 nm, with an apparent extinct coefficient of 1.04 × 105 M−1 cm−1. Beer's law is obeyed in the range of 0-15 μg/25 mL Hg2+, and Sandell's sensitivity is 1.92 × 10−3 μg/cm2. In the presence of thiourea and Na4P2O7 as masking agents, the method was found free from interferences of foreign ions commonly occurring with mercury. The optimized protocol has been successfully applied to spectrophotometric determination of mercury in waste water samples. The features of the new reagent associated with its special structure were discussed, and an unprecedented “domino effect” was proposed to account for its unique chelating stoichiometry with Hg2+.  相似文献   

5.
An investigation of the influence of crystalline microstructure on fatigue crack propagation (FCP) in high-density polyethylene (HDPE) is reported. Various thermal histories were used to generate samples with the same crystallinity and supermolecular structure for three different molecular weight HDPEs. Estimation of tie chain densities were obtained from measurements of brittle fracture stress and predicted from the estimated chain dimensions of the polymers using the modified version of the approach originally taken by Huang and Brown. A significant decrease in FCP resistance and a clear transition to a more brittle fracture surface was observed with decreasing molecular weight. Detailed studies of damaged zones preceding the growing crack show a transition to a more highly branched crack structure for those samples associated with a higher FCP resistance. These results strongly suggest that the branched damaged zone structure improves the FCP resistance by enlarging and blunting the crack tip and, therefore, consuming more energy during the fatigue crack propagation. Additional efforts were made to prepare samples with the same crystallinity and tie chain density, but different supermolecular structure. However, in contrast to reports in the literature, no significant difference in FCP resistance was observed for specimens with different average spherulite sizes. This is probably because the propagating crack front is preceded by a significant zone of plastic deformation and is not expected to directly encounter the spherulites.  相似文献   

6.
The kinetics of the radical reactions of CH3 with HCl or DCl and CD3 with HCl or DCl have been investigated in a temperature controlled tubular reactor coupled to a photoionization mass spectrometer. The CH3 (or CD3) radical, R, was produced homogeneously in the reactor by a pulsed 193 nm exciplex laser photolysis of CH3COCH3 (or CD3COCD3). The decay of CH3/CD3 was monitored as a function of HCl/DCl concentration under pseudo-first-order conditions to determine the rate constants as a function of temperature, typically from 188 to 500 K. The rate constants of the CH3 and CD3 reactions with HCl had strong non-Arrhenius behavior at low temperatures. The rate constants were fitted to a modified Arrhenius expression k = QA exp (−Ea/RT) (error limits stated are 1σ + Students t values, units in cm3 molecule−1 s−1): k(CH3 + HCl) = [1.004 + 85.64 exp (−0.02438 × T/K)] × (3.3 ± 1.3) × 10−13 exp [−(4.8 ± 0.6) kJ mol−1/RT] and k(CD3 + HCl) = [1.002 + 73.31 exp (−0.02505 × T/K)] × (2.7 ± 1.2) × 10−13 exp [−(3.5 ± 0.5) kJ mol−1/RT]. The radical reactions with DCl were studied separately over a wide ranges of temperatures and in these temperature ranges the rate constants determined were fitted to a conventional Arrhenius expression k = A exp (−Ea/RT) (error limits stated are 1σ + Students t values, units in cm3 molecule−1 s−1): k(CH3 + DCl) = (2.4 ± 1.6) × 10−13 exp [−(7.8 ± 1.4) kJ mol−1/RT] and k(CD3 + DCl) = (1.2 ± 0.4) × 10−13 exp [−(5.2 ± 0.2) kJ mol−1/RT] cm3 molecule−1 s−1.  相似文献   

7.
Solvatochromic mixed-chelate copper(II) complexes, [Cu(Cl-acac)(diamine)]X (where Cl-acac = 3-chloroacetylacetonate ion, diamine = N,N′-dimethyl,N′-benzyl-1,2-diaminoethane and X = B(Ph)4, PF6, BF4 and ClO4), have been prepared. The complexes were characterized on the basis of elemental analysis, molar conductance, UV-Vis and IR spectroscopies. Single crystals of [Cu(Cl-acac)(diamine)(H2O)]PF6, complex 2, were also characterized by X-ray diffraction. The influence of the solvent polarity and counter ions on the νmax values of the d-d bands of the complexes have been investigated by means of visible spectroscopy. All the complexes demonstrated negative solvatochromism. A multi-parametric equation has been utilized to explain the solvent effect on the d-d transition of the complexes using SPSS/PC software. The stepwise multiple linear regression (SMLR) method demonstrated that the donor power of the solvent plays the most important role in the solvatochromism of the compounds. The relative donor power of the anions X was determined by visible spectra in the solvent dichloromethane.  相似文献   

8.
Na2[(VIVO)2(ttha)]·8 H2O (ttha = triethylenetetraamine–N,N,N′,N″,N′″,N′″–hexaacetate ion), prepared by treating [VO(H2O)5][(VO)2(ttha)]·4 H2O with Na6(ttha), has been characterized by single crystal X-ray diffraction, infrared spectroscopy, UV–Vis absorption spectroscopy, electron spin resonance spectroscopy, and modeled by density functional theory (DFT). The X-ray structure revealed a distorted octahedral geometry around each vanadium center. The electronic absorption spectrum of [(VO)2(ttha)]2− (aq) features absorptions at ca. 200 nm (ε > 13900 L mol−1 cm−1), 255 nm (ε = 3480 L mol−1 cm−1), 586 nm (ε = 33 L mol−1 cm−1), and 770 nm (ε = 38 L mol−1 cm−1). The time-dependent density functional theory (TDDFT) calculated electronic absorption spectrum was remarkably similar to the actual spectrum, and TDDFT predicts absorption peaks at 297, 330, 458, 656, and 798 nm. TDDFT assigned the peak at 798 nm to be the α spin HOMO → LUMO transition. Hence, the peak at 770 nm in the actual spectrum is most likely the α spin HOMO → LUMO transition. Moreover, the TDDFT calculations revealed that the α spin HOMO and LUMO are partly comprised of d orbitals on both vanadium centers, and the first derivative electron spin resonance spectrum also suggests that the two unpaired electrons in [(VO)2(ttha)]2− are localized near the vanadium centers.  相似文献   

9.
The standard molar heat capacity C°p,m of adenine(cr) has been measured using adiabatic calorimetry over the range 6<(T/K)<310 and the results used to derive thermodynamic functions for adenine(cr) at smoothed temperatures. At T=298.15 K, C°p,m=(142.67±0.29) J · K−1 · mol−1 and the third law entropy S°m=(145.62±0.29) J · K−1 · mol−1. The standard molar Gibbs free energy of formation ΔfG°m at T=298.15 K for crystalline adenine was calculated, using the standard molar enthalpy of formation for the compound and entropies of the elements from the literature, and found to be ΔfG°m=(301.4±1.0) kJ · mol−1. The results were combined with solution calorimetry and solubility measurements from the literature to yield revised values for the standard molar thermodynamic properties of aqueous adenine at T=298.15 K: ΔfG°m=(313.4±1.0) kJ · mol−1, ΔfH°m=(129.5±1.4) kJ · mol−1, and Sm°=(217.68±0.44) J · K−1 · mol−1.  相似文献   

10.
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   

11.
A hyphenated ion-pair (tetrabutylammonium chloride—TBACl) reversed phase (C18) HPLC-ICP-MS method (High Performance Liquid Chromatography Inductively Coupled Plasma Mass Spectroscopy) for anionic Rh(III) aqua chlorido-complexes present in an HCl matrix has been developed. Under optimum chromatographic conditions it was possible to separate and quantify cationic Rh(III) complexes (eluted as a single band), [RhCl3(H2O)3], cis-[RhCl4(H2O)2], trans-[RhCl4(H2O)2] and [RhCln(H2O)6−n]3−n (n = 5, 6) species. The [RhCln(H2O)6−n]3−n (n = 5, 6) complex anions eluted as a single band due to the relatively fast aquation of [RhCl6]3− in a 0.1 mol L−1 TBACl ionic strength mobile phase matrix. Moreover, the calculated t1/2 of 1.3 min for [RhCl6]3− aquation at 0.1 mol kg−1 HCl ionic strength is significantly lower than the reported t1/2 of 6.3 min at 4.0 mol kg−1 HClO4 ionic strength. Ionic strength or the activity of water in this context is a key parameter that determines whether [RhCln(H2O)6−n]3−n (n = 5, 6) species can be chromatographically separated. In addition, aquation/anation rate constants were determined for [RhCln(H2O)6−n]3−n (n = 3-6) complexes at low ionic strength (0.1 mol kg−1 HCl) by means of spectrophotometry and independently with the developed ion-pair HPLC-ICP-MS technique for species assignment validation. The Rh(III) samples that was equilibrated in differing HCl concentrations for 2.8 years at 298 K was analyzed with the ion-pair HPLC method. This analysis yielded a partial Rh(III) aqua chlorido-complex species distribution diagram as a function of HCl concentration. For the first time the distribution of the cis- and trans-[RhCl4(H2O)2] stereoisomers have been obtained. Furthermore, it was found that relatively large amounts of ‘highly’ aquated [RhCln(H2O)6−n]3−n (n = 0-4) species persist in up to 2.8 mol L−1 HCl and in 1.0 mol L−1 HCl the abundance of the [RhCl5(H2O)]2− species is only 8-10% of the total, far from the 70-80% as previously proposed. A 95% abundance of the [RhCl6]3− complex anion occurs only when the HCl concentration is above 6 mol L−1. The detection limit for a Rh(III) species eluted from the column is below 0.147 mg L−1.  相似文献   

12.
A systematic thermodynamic and kinetic study of the entire SFxCl (x = 0-5) series has been carried out. High-level quantum chemical composite methods have been employed to derive enthalpy of formation values from calculated atomization and isodesmic energies. The resulting values for the SCl, SFCl, SF2Cl(C1), SF3Cl(Cs), SF4Cl(Cs) and SF5Cl molecules are 28.0, −36.0, −64.2, −134.3, −158.2 and −237.1 kcal mol−1. A comparison with previous experimental and theoretical values is presented. Statistical adiabatic channel model/classical trajectory, SACM/CT, calculations of selected complex-forming and recombination reactions of F and Cl atoms with radicals of the series have been performed between 200 and 500 K. The reported rate coefficients span over the normal range of about 6 × 10−12 and 5 × 10−11 cm3 molecule−1 s−1 expected for this type of barrierless reactions.  相似文献   

13.
Absolute rate constants for the reaction of t-butylperoxy radicals with hexaphenylditin and hexaphenyldilead have been found to fit the equations log (kp/M−1·s−1) = (5.5±0.15) − (3800±150)/Θ and log (kp/M−1·s−1) (9.5±0.3) − (6000±250)/Θ, where Θ = 2.303 RT cal mol−1, respectively.  相似文献   

14.
This paper describes a highly sensitive, selective catalytic-kinetic-spectrophotometric method for the determination of copper(II) concentration as low as 6 ng ml−1. The method is based on the catalytic effect of copper(II) on the oxidation of citric acid by alkaline hexacyanoferrate(III). The reaction was followed by measuring the decrease in absorbance of hexacyanoferrate(III) at 420 nm (λmax of [Fe(CN)6]3−,  = 1020 dm3 mol−1 cm−1). The dependence of rate of the indicator reaction on the reaction variables has been studied and discussed to propose a suitable mechanism to get a relation between the reaction rate and [Cu2+]. A fixed time procedure has been used to obtain a linear calibration curve between the initial rate and lower [Cu2+] or log[Cu2+] in the range 1 × 10−7 to 4 × 10−4 mol l−1 (6.35-25,400 ng ml−1). The detection limit has been calculated to be 4 ng ml−1. The maximum average error is 3.5%. The effect of the presence of various cations, commonly associated with copper(II) and some anions has also been investigated and discussed. The proposed method is sensitive, accurate, rapid and inexpensive compared to other techniques available for determination of copper(II) in such a large range of concentration. The new method has been successfully applied for the determination of copper(II) in various samples.  相似文献   

15.
Hoshino M  Kamino S  Mitani S  Asano M  Yamaguchi T  Fujita Y 《Talanta》2011,85(5):2339-2343
Spectrophotometric determination of titanium(IV) was accomplished with o-carboxyphenylfluorone (OCPF) in the presence of hexadecyltrimethyl ammonium chloride (HTAC) under strongly acidic media. In the determination of titanium(IV), Beer's law was obeyed in the range of 24-340 ng mL−1 with an effective molar absorption coefficient (at 530 nm) and relative standard deviation of 2.24 × 105 dm3 mol−1 cm−1 and 0.64% (n = 8), respectively. The severe interference of iron ions was easily eliminated by the addition of ethylenediaminetetraacetic acid (EDTA); the effects of other foreign substances were low. Equilibrium and kinetic studies under analytical conditions were investigated to quantitatively evaluate the reaction mechanism. The obtained orange complex is considered to be Ti(OCPF)4. Its stability log Kf and rate constant Kobs are 16.88 and 1.65 × 10−2 s−1, respectively. It is suggested that the color of the complex is related to the species of OCPF in the solution.  相似文献   

16.
Ruhela R  Sharma JN  B S Tomar  Hubli RC  Suri AK 《Talanta》2011,85(2):1217-1220
A precise, sensitive and selective method for the spectrophotometric determination of palladium (II) using N,N,N′,N′-tetra(2-ethylhexyl) thiodiglycolamide T(2EH)TDGA as an extractant is described. Palladium (II) forms yellow colored complex with T(2EH)TDGA which exhibits an absorption maximum at ∼300 nm. The colored complex obeys Beer's law in the concentration range 1.0-15.0 μg ml−1 of palladium with a molar absorptivity of 1.29 × 105 M−1 cm−1. The effects of various experimental parameters have been studied to establish the optimum conditions for the extraction and determination of palladium. The precision of the method has been evaluated and the relative standard deviation has been found to be less than 0.5%. The method has been successfully applied to the determination of palladium in simulated high level liquid waste (SHLW) solution.  相似文献   

17.
Using ozonolysis of the acid-catalyzed cyclized products of (−)-nidorellol and air-autoxidation as the key steps, (+)-ambrox was obtained in 53% overall yield. In the course of our synthesis, we discovered that (−)-nidorellol provided (+)-ambrox instead of the expected product, (−)-ambrox. Thus the absolute configuration of (−)-nidorellol was proved to be trans-(5R,7R,8R,9S,10R)-labda-12,14-diene-7α,8β-diol, which is opposite to that illustrated in a previous report.  相似文献   

18.
The vibration-vibration energy transfer of HBr gas initially excited to the first vibrational state has been observed. Collisional pumping to the V = 2 level is measured by monitoring the fluorescence of the 2−1 transition. The rate constant for the process: HBr(V=0) + HBr(V=2) → 2HBr(V=1) is found to be 1.4×105 sec−1 torr−1.  相似文献   

19.
The collisional broadening and shift rate coefficients of the “forbidden“ 6p2 3P0 → 6p2 3P1 transition in lead were determined by diode laser absorption measurements performed simultaneously in two resistively heated hot-pipes. One hot-pipe contained Pb vapor and noble gas (Ar or He) at low pressure, while the other was filled with Pb and noble gas at variable pressure. The measurements were performed at temperatures of 1220 K and 1290 K, i.e., lead number densities of 4.8 × 1015 cm− 3 and 1.2 × 1016 cm− 3. The broadening rates were obtained by fitting the experimental collisionally broadened absorption line shapes to theoretical Voigt profiles. The shift rates were determined by measuring the difference between the peak absorption positions in the spectra measured simultaneously in the heat pipe filled with noble gas at reference pressure and the one with noble gas at variable pressure. The following data for the broadening and shift rate coefficients due to collisions with Ar and He were obtained: γBAr = (3.4 ± 0.1) × 10− 10 cm3 s− 1, γBHe = (3.8 ± 0.1) × 10− 10 cm3 s− 1, γSAr = (− 7.3 ± 0.8) × 10− 11 cm3 s− 1, γSHe = (− 6.5 ± 0.7) × 10− 11 cm3 s− 1.  相似文献   

20.
The acid-catalysed hydrolysis of sulphilimines of XC6H4(Me)SNTs and MePhSNSO2C6H4Y type has been studied by a kinetic method in moderately concentrated (1–6 M) aqueous H2SO4 and HClO4 solutions. The rate law: rate = kψ[sulphilimine] is valid for hydrolysis leading to sulphoxides and sulphonamides. The dependence of kψ on acidity, temperature and substituents X and Y has been measured and interpreted, ?X, ?Y and ΔS data (+ 1·19, + 1·00 and −18·7- - 22·6 e.u., resp) show that the nucleophilic attack of water on the positively polarized S(IV) atom of protonated sulphilimines can be regarded as the rate-determining step of the hydrolysis. From φ parameters (0·94−1·5) calculated for the hydrolysis of MePhSNTs it follows that water participates in the reaction as a nucleophile and proton-transfer agent.  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-23

京公网安备 11010802026262号