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 78 毫秒
1.
A method for predicting an analytical equation of state for polymer mixtures and blends from surface tension and liquid state density at normal (ordinary) temperature (γn, ρn), as scaling constants, is presented. B2(T) follows a promising corresponding-states principle. Calculation of (T) and b(T), the two other temperature-dependent constants of the equation of state, are made possible by scaling. As a result, γn and ρn are sufficient for determination of thermophysical properties of polymer mixtures and blends.

We applied the procedure to predict liquid density of poly(ethylene glycol) (PEG-200) + 1-octanol solutions and poly(propylene glycol) (PPG) + poly(ethylene glycol) (PEG-200) blends at compressed state with temperature range from 298.15 to 338.15 K and pressures up to 40 MPa. In this work, the ISM EoS is extended to polymer mixtures and blends as well as pure case without proposing any mixing rule.  相似文献   


2.
Anion exchange membrane has been investigated in different electrolyte solutions by chronopotentiometry to explore the influence of co-ion and counterion of the exchange group of the membrane, on the transport phenomena. Chloride, nitrate, sulfate and acetate in sodium salts were used as counterions and sodium, potassium, calcium and ammonium in chloride salts were used as co-ions. The membrane showed a potential drop (E0) in all these electrolytes when a constant current was applied across it, which remained constant for a period less than τ, called the transition time and rose gradually to a maximum (Emax) value. The parameters such as τ, E0 and Emax and the potential jump (ΔE) and τ and the inflection zone (Δt) along the time axis have been measured and compared at an applied current density (I) of 10 mA cm−2 in 10 mM solutions. The values of τ1/2/zA[A0] or τ1/2/zC[C0], with or , E0 and ΔE with or (where rA and rC are the ionic radii of counter and co-ions, respectively) have been correlated. Permselectivity (P) and transference number of the membrane with respect to each one of the above electrolytes have been evaluated and discussed.  相似文献   

3.
The thermodynamic interactions in aqueous solutions of uncharged polymers were studied. Using a gel-deswelling method, the water activities (chemical potentials) in binary and ternary (two polymers in one solvent) solutions of methylcellulose (MC), polyvinyl alcohol (PVA) and polyvinyl pirrolidone (PVP), respectively were determined at various polymer volume fractions (1.0 × 10−2 < v2 < 1.0 × 10−1). On the theoretical basis of the Flory–Huggins approximations, the relevant solvent–segment (χ12 or χ13) and segment–segment pair interaction parameters (χ23) have been calculated.

The solvent activity curves (ln a1 versus polymer volume fraction) can be well described by a polynomial of third-degree in both the binary and the ternary solutions of the polymers. The solvent–segment interaction parameters exhibit a slight dependence on the polymer concentration. For each binary solution, the χ12v2 function can be fitted by a straight line wich has a small positive slope. In the mixtures of two polymers, the values of the segment–segment (χ23) interaction parameters were close to zero or sligthly negative (χ23 0 ± 0.03), indicating that under the studied conditions, the polymers in the ternary solutions are compatible.  相似文献   


4.
Gas electron diffraction is applied to determine the geometric parameters of the silacyclobutane molecule using a dynamic model where the ring puckering was treated as a large amplitude motion. The structural parameters and the parameters of the potential function were refined taking into account the relaxation of the molecular geometry estimated from ab initio calculations at the MP2/6-311+G(d, p) level of theory. The potential function has been described as V() = V0[(/e)2 − 1]2 with the following parameters V0 = 0.82 ± 0.60 kcal/mol and e = 33.5 ± 2.7°, where is a puckering angle of the ring.

The geometric parameters at the minimum V() (ra in Å, in degrees and uncertainties given as three times the standard deviations including a scale error) are: r(Si–Hax) = 1.467(96), r(Si–Heq) = 1.468(96), r(Si–C) = 1.885(2), r(C–C) = 1.571(3), r(C–H) = 1.100(3), CSiC = 77.2(9), HSiH = 108.3, SiCHeq = 123.5(16), SiCHax = 111.9(16), CC5Heq = 118.4(24), CC5Hax = 112.3(24), HC3H = 107.7, δ(HSiH) = 6.6, δ(HC3H) = 7.0, where the tilts δ, HSiH, and HC3H are estimated from ab initio constraints. The structural parameters are compared with those obtained for related compounds.  相似文献   


5.
The molecular structure of trichloronitromethane has been studied in the gas phase using electron diffraction data. The molecules are found to undergo low barrier rotation about the CN bond with a planar CNO2 moiety in agreement with HF/MP2/B3LYP/6-311G(d,p) calculations. The experimental data are consistent with a dynamic model using a potential function for the torsion of V = (V6/2)(1 − cos 6τ). The major geometrical parameters (rg and ) for the eclipsed form, obtained from least squares analysis of the data are as follows: r(NO3) = r(NO4) = 1.213(2) Å, r(CN) = 1.592(6) Å, r(CCl)av = 1.749(1) Å, Cl5CN/Cl6CN = 109. 6°/106.3°(2), O3NC/O4NC = 117. 6°/114.1°(4), τCl5C1N2O3 = 0.0°, and V6 = 0.20(25) kcal/mol.  相似文献   

6.
Since the discovery of superconductivity in Sr2CuO2F2+δ there has been an increased interest in ternary oxide-fluorides. Sr2CuO2F2+δ is prepared via low temperature (T = 220 °C) reaction routes. Low temperature fluorination induces an interesting structural rearrangement in the parent compound Sr2CuO3, which is a one-dimensional material containing linear chains of vertex sharing CuO4 squares along the crystallographic b axis. Upon fluorination, one oxide is substituted by two fluorides and Cu2+ becomes octahedrally coordinated by four oxides and two fluorides. The fluorinated compound Sr2CuO2F2+δ displays the T-type structure (La2CuO4). Insertion of excess fluorine, δ, also takes place and this fluorine occupies interstitial sites in the T structure. Although the starting material Ca2CuO3 is isostructural to Sr2CuO3, Ca2CuO2F2+δ displays the T′ (Nd2CuO4) structure due to the smaller radius of Ca2+ compared to that of Sr2+.

The alkaline-earth palladates with the general formula A2PdO3 (A = Ba, Sr) are isostructural with the A2CuO3(A = Ca, Sr) materials. We prepared the Ba2xSrxPdO3 (x = 0–2) series and performed low temperature fluorination, which led to the synthesis of the series Ba2xSrxPdO2F2+δ (0 ≤ x ≤ 1.5). All the compounds in the Ba2xSrxPdO2F2+δ series show T′ structure (Ca2CuO2F2+δ). Similarities and differences with Sr2CuO2F2+δ and Ca2CuO2F2+δ will be discussed.  相似文献   


7.
The radical copolymerization of (2,6-diphenyl) phenyl methacrylate (1) with methyl methacrylate in DMF with AIBN at 70°C has the reactivity ratios r1 = 0.071 and r2 = 1.42, from which Q1 = 1.45 and e1 = 1.20. The copolymers had Mns in the range of 10,000–40,000 and Tgs ranging from 406 to 480 K from which the hypothetical Tg for poly-1 was deduced as 500 K (227°C). Unlike 1, (2,6-diphenyl) phenyl acrylate could be polymerized to oligomers with Mn of the order of 2500.  相似文献   

8.
Fengling Liu 《大学化学》2020,35(9):168-172
A method for obtaining the delocalized π bonds πnm in a molecule has been discussed in this paper, and the delocalized π bonds πnm in linear, bent, planar trigonal, single cyclic conjugated, and polycyclic conjugated molecules have been studied. The reasons of π34 in NO2 and 2π32 in C3 molecules have been proposed. The delocalized π bonds 2π1818 in cyclo[18] carbon are analyzed.  相似文献   

9.
For a closed-shell MO configuration with 2n electrons which occupy n non-degenerate canonical MOs, it is deduced that the RHF energy, Σni=1[2H0nnj-1(2Jij-Kij)], may be expressed in Hückel-like form as 2Σni-1ε, −Σni-1[ji(λ+1)+1,(λ+2)] with λ=2(n-i). The li(λ+1) and Ii(λ+2) are the ionization potentials for the HOMO ψ, which arises after λ and λ+1 electrons have been successively removed from the initial configuration.  相似文献   

10.
The oxygen permeation properties of mixed-conducting ceramics SrFeCo0.5O3−δ (SFCO), Ba0.5Sr0.5Co0.8Fe0.2O3−δ (BSCFO), La0.2Sr0.8Co0.8Fe0.2O3−δ (LSCFO) and Ba0.95Ca0.05Co0.8Fe0.2O3−δ (BCCFO) were studied by thermogravimetric method in the temperature range 600–900 °C. The results show that the oxygen adsorption rate constants ka of all material are larger than oxygen desorption rate constants kd and both ka and kd are not strongly dependent on temperature in the studied temperature range. The oxygen vacancy contents δ(N2) and δ(O2) in nitrogen and oxygen and their difference Δδ = δ(N2) − δ(O2) play an important role in determining the temperature behavior of oxygen permeation flux JO2.  相似文献   

11.
A series of CexPr1−xO2−δ mixed oxides were synthesized by a sol–gel method and characterized by Raman, XRD and TPR techniques. The oxidation activity for CO, CH3OH and CH4 on these mixed oxides was investigated. When the value x was changed from 1.0 to 0.8, only a cubic phase CeO2 was observed. The samples were greatly crystallized in the range of the value x from 0.99 to 0.80, which is due to the formation of solid solutions caused by the complete insertion of Pr into the CeO2 crystal lattices. Raman bands at 465 and 1150 cm−1 in CexPr1−xO2−δ samples are attributed to the Raman active F2g mode of CeO2. The broad band at around 570 cm−1 in the region of 0.3 ≤ x ≤ 0.99 can be linked to oxygen vacancies. The new band at 195 cm−1 may be ascribed to the asymmetric vibration caused by the formation of oxygen vacancies. The TPR profile of Pr6O11 shows two reduction peaks and the reduction process is followed: . The reduction temperature of CexPr1−xO2−δ mixed oxides is lower than those of Pr6O11 or CeO2. TPR results indicate that CexPr1−xO2−δ mixed oxides have higher redox properties because of the formation of CexPr1−xO2−δ solid solutions. The presence of the oxygen vacancies favors CO and CH3OH oxidation, while the activity of CH4 oxidation is mostly related to reduction temperatures and redox properties.  相似文献   

12.
With the aim of understanding the influence of donor solvents on the reactivity of the amine complexes [RuCl2(PPh3)2(piperidine)] (1) and [RuCl2(PPh3)2(imidazole)2] (2) in the presence of ethyldiazoacetate, and on the properties of the resulting polymer, a ring opening metathesis polymerization of norbornene was carried out in the presence of small amounts of common solvents such as additives (isopropanol, THF, N,N-dimethylformamide, 2,6-lutidine, isopropanethiol, acetonitrile, dimethyl sulfoxide, NEt3, NH2Me and pyridine). From observations, typical coordinating solvents like DMSO, NEt3, NH2Me and pyridine, hardly affected the yields when either complex was employed. With other additives, the major advantage was the decrease in the polydispersity indices. On using complex 1 with 2,6-lutidine, observed values of Mw/Mn were as low as 1.3, while the yield decreased from 99% to about 20–30% at RT for 1 min in pure solution. In the case of complex 2, which is almost inactive to ROMP (19% at 50 °C for 5 min with Mw/Mn = 6.30), the yield was three-fold (60% at 50 °C for 5 min with Mw/Mn = 1.95) compared to that of without THF. Further, the Mw/Mn was observed to decrease to 1.34 with 200 eq. of THF.  相似文献   

13.
This study examines a linear variation of the specific heat CP with the frequency shifts 1/ν(∂ν/∂T) for the Brillouin frequencies of the L-mode [010], [001] and [100] in the ferroelectric phase of NaNO2 according to our spectroscopically modified Pippard relation. We obtain this linear relationship for those modes studied and calculate dTC/dP in the ferroelectric phase of NaNO2. Our calculated values of dTC/dP for the [001] and [100] modes are in good agreement with the values given in the literature.  相似文献   

14.
A quantum mechanical and molecular mechanical (QM + MM) direct dynamics classical trajectory simulation is used to study energy transfer and fragmentation in the surface-induced dissociation (SID) of N-protonated diglycine, (gly)2H+. The peptide ion collides with the hydrogenated diamond {111} surface. The Austin Model 1 (AM1) semiempirical electronic structure theory is used for the (gly)2H+ intramolecular potential and molecular mechanical functions are used for the diamond surface potential and peptide/surface intermolecular potential. The simulations are performed at collision energies Ei of 30, 50, 70, and 100 eV and collision angle of 0° (perpendicular to the surface). The percent energy transfer to the peptide ion is nearly independent of Ei, while energy transfer to the surface increases with increase in Ei. A smaller percent of the energy remains in peptide translation as Ei is increased. These trends in energy transfer are consistent with previous trajectory simulations of SID. At each Ei the most likely initial pathway leading to fragmentation is rupture of the +H3NCH2---CONHCH2COOH bond. Fragmentation occurs by two general mechanisms. One is the traditional Rice-Ramsperger-Kassel-Marcus (RRKM) model in which the peptide ion is activated by its collision with the surface, “bounces off”, and then dissociates after undergoing intramolecular vibrational energy redistribution (IVR). The other mechanism is shattering in which the ion fragments as it collides with the surface. Shattering is the origin of the large increase in number of product channels with increase in Ei, i.e., 6 at 30 eV, but 59 at 100 eV. Shattering becomes the dominant dissociation mechanism at high Ei.  相似文献   

15.
Møller-Plesset MP2/6-31G method was used to examine the gas-phase elimination of 2-substituted alkyl ethyl N,N-dimethylcarbamates. The results of these calculations support a concerted non-synchronous six-membered cyclic transition state mechanism for carbamates containing a Cβ–H bond at the alkyl side of the ester. These substrates produce the N,N-dimethylcarbamic acid and the corresponding olefin. The unstable intermediate, N,N-dimethylcarbamic acid, rapidly decomposes through a four-membered cyclic transition state to dimethylamine and CO2 gas. Correlation of the logarithm of theoretical rate coefficients against original Taft's σ* values gave an approximate straight line (ρ*=−1.39, r=0.9558 at 360 °C). In addition to this fact, when log krel is plotted against the theoretical log krel for 2-substituted ethyl N,N-dimethylcarbamates a reasonable straight line (r=0.9919 at 360 °C) is obtained, suggesting similar mechanism.  相似文献   

16.
The interactions between some acrylic and sulphonic polyanions and some protonated amines (diamines NH2-(CH2)x-NH2, x=2,…,10; linear tri-, tetra-, penta- and hexa-amines) were studied potentiometrically in aqueous solution, at 25°C. For both types of polyanions AL2Hi (L, monomer of polyanion, A, amine) species are formed, with i=1,…,n (n=number of amino groups in the amine). The stability of these species is strictly dependent on the polyammonium cation charge, and fairly independent of the type of amine (in diamine species maximum stability is observed for x=4, 5). Acrylic and sulphonic polyanion complexes are considerably stronger than analogous species formed by low molecular weight anions. Mean stability can be expressed as log K=2.87ζ2/3, for polyacrylic anions and log K=2.42ζ2/3 for polysulphonic anions (ζ=absolute value for charge product of reactants).  相似文献   

17.
The X-ray structure of tetraethylammonium hydrogenselenate, [N(C2H5)4]HSeO4, was determined at room temperature. The crystal belongs to the P space group of triclinic system, Z=2, a=8.290(2), b=9.073(2), c=9.517(2) Å, =76.75(3), β=74.31(3) and γ=63.92(3)°. The hydrogenselenate anions are joined into cyclic dimers by two identical (equivalent by Ci) strong hydrogen bonds O(2)–H(02)O(1a); the O(2)O(1a) distance equals 2.611(5) Å. Powder IR and Raman spectra are discussed with respect to the crystal structure. The DSC reveals two phase transitions at 328 and 358 K.  相似文献   

18.
A hydrothermal reaction of 1,3-dicyanobenzene and Ca(OH)2 yielded a triple helical calcium-based coordination polymer of the formula, C20H25Ca2.50O18.50 (1). The 1,3-benzenecarboxylate anion, found in the final product was generated in situ during the synthesis by the hydrolysis of 1,3-dicyanobenzene. X-ray diffraction study shows that the complex 1 crystallizes in the monoclinic system, C2/c space group, a=15.5701(5), b=21.4445(7), c=17.1601(6) Å, β=111.7400(7)°, V=5322.1(3) Å3, Z=8, Dc=1.651 Mg/m3. The calcium atoms show differences in the coordination environments. Complex 1 emits strong blue fluorescent light (λem(max)=419 nm) when it is excited by UV light (λex(max)=316 nm) in the solid state at room temperature.  相似文献   

19.
Twelve series of linear oligomers of four different degrees of polymerization (xn = 8.77−41.55), having a common perfluorinated random copolymeric chain as molecular body and two equal foreign end units of one of the types listed in Table 1, have been synthesized by derivatization of base samples of one of them having a diolic---CH2OH functionality. The glass transition temperature Tg of all the series was measured and thus examined as a function of xn. A clear end unit effect is observed, dominantly determined in every series by chemical nature and structure of the end units, quantitatively expressed at any xn by different positive or negative Tg deviations from the common asymptotic Tg value. The results are also discussed in terms of copolymer end effect and of relation between Tg and end copolymeric composition.  相似文献   

20.
The vibrational excitation of a Morse oscillator in a collinear collision is determined analytically using a new dynamical algebra. The dependence of V → T processes on the initial vibrational state is examined. At finite coupling strengths the decline of Pnn−1/n with n is faster than that derived for weak coupling.  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