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.
The radical polymerizations of 2-, 3-, and 4-(trimethylsilylethynyl)styrenes (1 a – c) and copolymerizations of 1 a – c (M1) with styrene (M2) have been studied. Copolymerization parameters were determined as r1 = 1.22 and r2 = 0.54 for 1 a, 1 = 1.10 and r2 = 0.90 for 1 b, and r1 = 1.42 and r2 = 0.38 for 1 c. The deprotection of the trimethylsilyl groups in poly[(trimethylsilylethynyl)styrene] (2 a – c) and poly[(trimethylsilylethynyl)styrene-co-styrene] (4 a – c) using (C4H9)4NF smoothly proceeded to yield poly(ethynylstyrene) (3 a – c) and poly(ethynylstyrene-co-styrene) (5 a – c), respectively, which underwent curing reactions at elevated temperature to form crosslinking polystyrenes. Received: 31 March 1997/Revised: 2 June 1997/Accepted: 3 June 1997  相似文献   

2.
Summary The radical polymerizations of 2-, 3-, and 4-(phenylethynyl)styrenes (1a–c) and the copolymerizations of 1a–c (M1) with styrene (M2) were carried out using AIBN as the initiator in toluene at 60°C. The number-average molecular weights (M ns) were extremely low for poly(2-phenylethynylstyrene) (2a) and poly[(phenylethynyl)styrene-co-styrene] (3a), and increased in the order of 2a, 3a << 2b, 3b < 2c, 3c. Monomer reactivity ratios were determined as r 1= 1.80 and r 2= 0.51 for 1a, r 1= 1.72 and r 2= 0.53 for 1b, and r 1= 3.17 and r 2= 0.24 for 1c. Polymers 2a–c and 3a–c underwent an exothermic reaction at elevated temperature to form organic solvent-insoluble polymers. Although the decomposition of 2a was observed from 200°C, 2b and 2c exhibited a high heat-resistance property in both nitrogen and air atmospheres, in particular, 2b showed no significant weight loss below 450°C. Received: 28 January 1998/Accepted: 5 March 1998  相似文献   

3.
Summary The dynamic mechanical behaviour of fully cured off-stoichiometric epoxy systems prepared from poly(oxypropylene)diamine (Jeff D-400) or -triamine (Jeff T-403) and diglycidyl ether of Bisphenol A (DGEBA) with initial ratios of reactive amine (NH2) and epoxide (E) groups, r H= (2[NH2])/[E] ranging from 2.4 to 4.3 was investigated in the gelation threshold region. The evolution of the dynamic behaviour of two stoichiometric samples with r H= 1 with reaction time, t r , was also studied. The critical ratios for gelation, r c H, of fully cured samples were determined from extraction experiments (r c H= 2.54 for Jeff D-400/DGEBA and r c H= 4.15 for Jeff T-403/DGEBA systems, respectively). For both stoichiometric critical gel (CG) structures obtained by changing the curing time t r , a power-law rheological behaviour (G′∼ G″∼ωn, G′ and G″ are the storage and loss moduli, respectively, ω is angular frequency and n is a critical exponent) with the loss tangent, tan δ= G″/G′ independent of frequency, was found. On the other hand, both CG off-stoichiometric systems with r c H ratios show a small dependence of tan δ on ω, so that the critical power-law behaviour is not exactly obeyed. More complex CG structure in these samples, formed due to differences in the structure growth process, was suggested to account for violation of the power-law behaviour. Received: 4 December 1998/Accepted: 10 March 1999  相似文献   

4.
Dipropynylbenzene with branched alkoxy and alkyl groups [CH3C≡CRC6H2RC≡CCH3, R = 2-methylpropoxy (1a), 3-methylbutoxy (1b), 4-methylpentoxy (1c), cyclohexylmethoxy (1d), 2-ethylhexoxy (1e), 2-octoxy (1f), 2-ethylhexyl (1g), and 2-octyl (1h)] were polymerized with Mo(CO)6 in the presence of 4-(trifluoromethyl)phenyl to afford poly(2,5-di(alkoxy or alkyl)-p-phenyleneethynylene)s (2ah). Polymer 2a was insoluble in any solvents, but the other polymers (2bh) were soluble in common organic solvents. The polymers with relatively long side chains (2eh) had high molecular weight over 1.6 × 104 and gave free-standing membranes by solution-casting method. The densities of membranes of 2eh were 0.914–0.998, and their fractional-free volume values were relatively large (0.094–0.158). The oxygen permeability coefficients of membranes of 2eh were 18.4, 12.7, 4.85, and 19.3 barrers, respectively. It was found that poly(p-phenyleneethynylene) with 2-octyl side groups, which have the branch at the nearest position from main chain, exhibited the highest gas permeability.  相似文献   

5.
Dipropynylbenzenes with alkyl groups (CH3C ≡ CRC6H2RC≡CCH3, R=n-C6H13, n-C8H17, n-C10H21, 1ac, respectively) were polymerized with Mo(CO)6 to afford solvent-soluble poly(2,5-dialkyl-p-phenyleneethynylene)s (2ac). The polymers (2ac) had high molecular weight over 3×104, and gave free-standing membranes by solution casting method. According to thermogravimetric analysis (TGA), these poly(p-phenyleneethynylene)s showed high thermal stability (T0 ≥380 °C). The densities of membranes of poly(2,5-dialkyl-p-phenyleneethynylene)s (2ac) were 0.936–0.965, and their fractional free volume (FFV) were relatively large (ca. 0.14–0.15). The oxygen permeability coefficients (PO2) of membranes of 2ac were 4.88, 7.06, and 16.6 barrers, respectively.  相似文献   

6.
Summary Methylated (β-cyclodextrin (me-β-CD) was used to complex the hydrophobic monomers n-butyl acrylate (1), n-hexyl acrylate (2) and cyclohexyl acrylate (3) yielding the corresponding water soluble host/guest complexes 1a–3a. The complexes were copolymerized in water by free radical mechanism and the reactivity ratios were determined by measuring the monomer consumption by HPLC. The following reactivity ratios were found: copolymerization of 1a and 2a: r1= 1.01 ± 0.01; r2= 1.04 ± 0.01; copolymerization of 3a and 2a: r1= 0.74; r2= 1.28; copolymerization of 3a and 1a: r1= 0.75 ± 0.04; r2= 1.13 ± 0.01. In contrast to that, the copolymerization of the uncomplexed monomers 1–3 in organic medium (DMF/H2O) leads to nearly ideal statistical copolymers in all cases. Received: 28 November 2000/Accepted: 12 January 2001  相似文献   

7.
Adsorption equilibrium constants for methyl oleate and methyl linoleate in vapor phase on supported copper and nickel catalysts have been determined using the technique of pulse gas chromatography. The results are discussed in relation to selectivity in fat hydrogenation. Notation: A, column cross-section, m2 ; an,bn, nth Fourier coefficients; c, concentration of adsorbate in bulk flow, mol/m3 ; c* = c/ ∫ 0 cdt, normalized concentration of adsorbate in bulk flow; Ci, concentration of adsorbate in catalyst pores, mol/m3 ; ca, concentration of adsorbate on catalyst surface, mol/kg; cTOT, active area of catalyst as measured by hydrogen adsorption, mol/kg; De, effective diffusion coefficient of adsorbate in catalyst, m2/s; Dea, axial dispersion coefficient based on void cross-section, m2/s; hn, nth coefficient in Hermite polynomial expansion; Hn nth Hermite polynomial; ΔHA, adsorption enthalpy, kJ/mol; ΔHvap , heat of vaporization, kJ/moll; ka, adsorption rate constant, m3/kgs;KA, adsorption equilibrium constant, m3/kg; K0 , preexponential factor defined in Eqn. 8, m3/kg; kf, mass transfer coefficient, m/s; L, bed length, m; q, flow rate, m3/s; R, particle radius, m; Rg, gas constant; t, time, s; T, temperature, K; TF, period of Fourier expansion, s; u = q/A, linear velocity, m/s; z, length coordinate in packed column, m. Greek symbols: δ(t), Dirac delta function; ∈B, void fraction of bed; ∈-p, particle void fraction, ρrp, particle density, kg/m3 ; ξ, radial coordinate in particle, m; μ1, first absolute moment, μ2, second central moment.  相似文献   

8.
The title compound, [Sm2(C3H2O4)3(H2O)6], was investigated by X-ray diffraction. It crystallizes in the monoclinic space group C2/c with cell parameters a = 17.1650(8) ?, b = 12.3010(5) ?, c = 11.1420(4) ?, β = 127.5161(10)°, Z = 4 and V = 1866.04(14) ?3. The Sm atom lies on a two-fold axis and has nine-coordination with six oxygen atoms from carboxylate groups and three water molecules. The compound forms a layer-type polymeric structure. The layers are formed by samarium and one independent malonate group to give a three-dimensional framework. The extensive network of hydrogen bonds and bridge bonds observed in this structure enhances the structural stability. The thermal dehydration of the compound was investigated by thermogravimetric analysis.  相似文献   

9.
Hydrate effects on the conformations of ethylene oxide oligomers (EO-x, x = 1–8 mers) were examined using quantum chemical calculations (QCC). Conformational analyses were carried out by RHF/6-31G. The models were constructed by locating a water molecule to each ether–oxygen in the structures optimized for non-hydrate oligomers. Hydrate ratio, h (h = H2Omol/Omol in oligomer), was set from 0 to 1.0. The six type conformations with repeated units of O–C, C–C and C–O bonds were examined. Conformational energy, E c (HF), was calculated as difference between the energy of oligomer with water molecules and that of non-hydrogen and/or hydrogen bonding water molecules. Hydrate energies for each conformer, ∆μ h (kcal/m.u., based on E c in non-hydrate state), were negative and linearly decreased with the increase of h values, and such effects with the increase of h values were weaken with increasing x values. These results were consistent with our previous results calculated using the permittivity, ε (ε = 0–80.1), by QCC. In non-hydrate (h = 0), the (ttt) x conformers were the most stable independent of x. However, in hydrate states (h = 0.44–0.67), the (tg+t) x conformers were the most stable independent of x values, and in h = 1, the (tg+t)8 conformer (8-mer) was most stable [∆E c(g) = −1.3 kcal/m.u., ∆E c(g): energy difference between a given oligomer and the (ttt) x oligomer]. These results supported the experimental those based on NMR analyses using dimethoxyethane and triglyme solutions. Molecular lengths (l) of (tg+t) x , (tg+g) x and (g+g+g+) x conformers having higher x values significantly decreased with increasing h values. Such contraction with hydration, however, was independent of ΔE c(g) values of each conformer.  相似文献   

10.
The structural phase transition occurring in Friedel's salt, the chlorinated compound 3CaO·Al2O3·CaCl2·10H2O (AFm phase), was studied by synchrotron and standard X-ray powder diffraction. The compound transforms at 35 °C from a rhombohedral (rh) high-temperature (HT) phase [R−3c; a=5.744(2) Å, c=46.890(3) Å] to a monoclinic (m) low-temperature (LT) phase [C2/c; a=9.960(4) Å, b=5.7320(2) Å, c=16.268(7) Å, β=104.471(2)°]. The LT and HT phases were refined with the Rietveld method from synchrotron data recorded at 20 and 40 °C. Variations of the lattice parameters as a function of temperature are reported between 8 and 48 °C. The rh→m transition is characterized by a unit cell volume expansion of 1% and a movement of the interlayer species: a shift of 0.45 Å of the Cl anions along [010]h and a shift of 0.25 Å of the water molecules along [210]h of the hexagonal cell. The m phase distortion is due to an ordering of the hydrogen bonds between chloride anions and H-atoms of the water molecules.  相似文献   

11.
Static measurements have been used to predict the dynamic response of ldquo;unboundedrdquo; open-cell noninked (dry) and inked foam materials. Creep, ec(t), and recovery, er(t), were determined in compression from static and dynamic modes. Force measurements, f(t), and strain decay, e(t), were used to determine the change in creep, δec(t). The change in creep represents the plastic strain, ep1(t=th), and is uniquely defined by the recovery function, er(t=th), where th is the hold time. Creep and recovery results of various classes of foam materials and nonfoam materials were found to fit a master curve of the form Fr(t) =exp[–kr(th)t] = [er(t) - e00( th)]/[e0 (t = 0) - e00(th)] at a reduced time of kr(th) t [kr(th)] C0/(th)a (where Co depends on the material's “dry” or “wet” state), a is a function of the type of material, and em is the permanent set]. These empirical results are applicable to printing ink transfer and print quality. Other important factors of concern are diffusion processes within the polymer matrix and the nature of the polymer (e. g., chemical constitution, porosity, molecular weight, and solubility). © 1995 John Wiley & Sons, Inc.  相似文献   

12.
One novel dysprosium coordination polymer [Dy(PIC)3(H2O)2]n (HPIC = pyridine-4-carboxylic acid) has been synthesized. X-ray analysis reveals that it forms the chain-like molecular structure through the bridged oxygen atoms of the carbonyl groups. The title coordination polymer crystallizes in the monoclinic system, space group C2/c, with lattice parameters: a=20.243(9) Å, b=11.576(5) Å, c=9.834(4) Å, β=110.601(2)°, V=2078(2) Å3, Dc=1.805 mg/m3, Z=4, F(000)=1100, GOF = 1.11, R1=0.0404. The photophysical property has been studied with ultraviolet absorption spectrum, excitation and emission spectrum. The luminescence spectra show the stronger blue emission than yellow emission.  相似文献   

13.
Dense Bi2Te2W3O16 ceramics were prepared by the conventional solid‐state reaction route. X‐ray diffraction data show the room‐temperature (RT) crystal symmetry of Bi2Te2W3O16 to be well described by the centrosymmetric monoclinic C2/c space group [a = 21.280(5) Å, b = 5.5663(16) Å, c = 12.831(3) Å and β = 124.014(19)° and Z = 4]. Raman spectroscopy analyses are in broad agreement with space group assignment, but also revealed the presence of Bi2W2O9 as a secondary phase. This phase is present as plate‐like grains embedded on a fine‐grained equiaxed matrix, as revealed by scanning electron microscopy. From the fitting of infrared reflectivity data the relative permittivity, εr, was estimated as 34.2, and the intrinsic quality factor, Qu × f as 57 500 GHz. At RT and microwave frequencies, Bi2Te2W3O16 ceramics sintered at 720°C for 6 h exhibit εr ~ 34.5, Qu × f = 3173 GHz (at 7.5 GHz), and temperature coefficient of resonant frequency, τf = ?92 ppm/°C. This shows a good agreement between the estimated and measured εr values, but also shows that, in principle, the dielectric losses of the ceramics are of extrinsic origin.  相似文献   

14.
Cubic niobium nitrides δ-NbNx with different x = 0.87–1.015 were prepared by the self-propagating high-temperature synthesis (SHS) under nitrogen pressures of P(N2) = 48–230 MPa.Nitrogen composition x as a function of P(N2) was determined in weight gain experiments and compared with that determined by chemical (Kjeldahl) analysis. For powders with different x, the superconducting transition temperature Tc was measured as a function of lattice parameter a. The Tc values were found to grow linearly with increasing a. A maximum value of Tc = 15.0 K for cubic niobium nitride corresponded to a maximum value of a = 4.3901 Å.  相似文献   

15.
The direct polycondensations of cubane-1,4-dicarboxylic acid with 1,4-phenylenediamine (2 a), 4,4′-oxydianiline (2 b), 4,4′-sulfonyldianiline (2 c), and 9,9′-bis(4-aminophenyl)florene (2 d) were carried out in N-methyl-2-pyrrolidone/pyridine containing triphenylphosphite and lithium chloride at 110 °C for 9 h. Polyamide 3 a obtained from 2 a was scarcely soluble in organic solvent even during heating, and was soluble only in conc-H2SO4, whereas 3 c and 3 d derived from 2 c and 2 d, respectively, were readily soluble in N-methyl-2-pyrrolidone, N,N-dimethylacetamide, N,N-dimethylformamide, and dimethylsulfoxide. After treating polyamide 3 d with the rhodium complex catalyst in NMP, cubane units were quantitatively converted into cyclooctatetraenes. Received: 3 March 1997/Accepted: 1 April 1997  相似文献   

16.
In this paper, we analyze the concentration multiplicity and dynamic behavior for an autocatalytical reaction, A + R → (n + 1)R + products with an overall rate expression given by – γa= kcapcrr(p > 0 and r > 0) in a imperfectly mixed (Cholette's model) CSTR. We proved that non‐ideal mixing had an effect on the number of steady states and dynamic behavior for the reaction orders r > 1 and r = 1. However, the above‐mentioned effect does not happen for the reaction order r < 1. Furthermore, a simulated example was used to demonstrate our results.  相似文献   

17.
A novel 2D sheet-like copper(II) complex formulated [Cu6(trans-oxen)33-OH)2(H2O)2]n-(ClO4)4n.2nH2O (1), where H2oxen is N,N’-bis(2-aminoethyl)oxamide, has been synthesized and characterized by elemental analysis, IR and UV–visible spectroscopy, and single-crystal X-ray diffraction. 1 crystallizes in triclinic space group P-1 with crystallographic data: a = 10.299(2) ?, b = 10.833(2) ?, c = 11.781(2) ?, α = 69.95(3)°, β = 81.10(3)°, γ = 62.44(3)°, and Z = 2. In the crystal structure, the cation [Cu6(trans-oxen)33-OH)2(H2O)2]n4n+ exhibits a two-dimensional sheet-like structure for Cu(II) ions bridged both by trans-oxen and hydroxyl groups. Each Cu(II) ion is located in a slightly distorted square-pyramidal environment. The hydroxyl group bridges three [Cu(oxen)Cu]2+ dications in which one oxen2− is μ2-bridge while the other two act as never before reported μ4-bridge. Perchlorate anions and H2O molecules are inserted between the sheets and bridge the 2D cations via N–H···O and O–H···O hydrogen bonds to form an infinite three-dimensional system. The cyclic voltammetric behavior of 1 and the influence of counter ions on structure are preliminarily investigated.  相似文献   

18.
The coordination polymer, [Cu2(sal)2(pyz)(H2O)2] n (H2sal=salicylic acid, pyz=pyrazine), was synthesized and structurally characterized by X-ray crystallography. The empirical formula of the polymer is CuC9NO4H8and the structural parameters are as follows: M r=257.71, monoclinic, P21/c, Z=4, a=10.175(3) Å, b=7.352(4) Å, c=12.757(1) Å, =105.206(2)°, V=920.9(6) Å3, D c=1.859 g/cm3, (MoK)=23.62 cm–1, F(000)=520 and the final R=0.048 for 2179 observable reflections. Each salicylate ligand connects the three copper centers to afford a novel 2-dimensional (2-D) network structure. The CuII-sal framework forms a rhombus type coordination framework. The pyz ligands fill the voids of the sheet and coordinate to the Cu site.  相似文献   

19.
In this work, two new 312 MAX phases of Zr3PbC2 and Hf3PbC2 were successfully synthesized by spark plasma sintering. It is the first discovery of lead-containing 312 MAX phases, which together with M2PbC (M = Ti, Zr, Hf) form the lead-containing MAX phase family. Considering the extremely low electrical conductivity of Hf2PbC, these two new compounds are of great research value. Based on the Rietveld refinement results, their lattice parameters and atomic positions were well determined, as a = 3.3771(5) Å, c = 20.0070(9) Å for Zr3PbC2 and a = 3.3357(1) Å, c = 19.7659(8) Å for Hf3PbC2, where M atoms are located at (0, 0, 0) and (1/3, 2/3, 0.1258(6)[Zr]; 0.1255(2)[Hf]), Pb atoms are located at (0, 0, 1/4), and C atoms are located at (1/3, 2/3, 0.0663(2)[Zr]; 0.0641(3)[Hf]), respectively. Additionally, the typical laminar microstructure of Zr3PbC2 and Hf3PbC2 grains was observed.  相似文献   

20.
The structure and bonding of 7-silanorbornadienes was investigated using X-ray Diffraction (XRD), solid-state NMR spectroscopy and density functional calculations. The solid state structures of four benzo-7-silanorbornadienes (4a, c, d, e) and of one dibenzo-7-silabenzonorbornadiene (5a) are reported and compared with the results of previous structural investigations. The most prominent features of the molecular structures of all 7-silanorbornadienes are very long Si-C(bridgehead) bonds (d(SiC) = 190.6–196.8 pm) and very acute CSiC bond angles α (α(CSiC) = 78.7–83.9°). All newly investigated 7-silanorborndienes show for tetracoordinated silicon nuclei extremely deshielded 29Si NMR resonances (δ29Si = 65.6–31.6). Solid State NMR investigations for 7-silanorbornadienes anti-4a, b reveal highly anisotropic chemical shift tensors of axial or nearly axial symmetry (4a: δ11 = 161, δ22 = δ33 = −11; 4b: δ11 = 113, δ22 = 14, δ33 = −15). The dominating, strongly deshielding δ11 component is oriented almost perpendicular to the plane spanned by the two bridgehead carbon atoms and the bridging silicon atom. The DFT calculations suggest that the origin of the strong deshielding is a small energy difference between the frontier orbitals, which are strongly localized at the silicon atom. In addition the computations reveal that both the long SiC bonds and the strongly deshielded 29Si NMR chemical shift are direct consequences of the bonding situation in 7-silanorbornadienes which are characterized by through space interaction of the C=C double bonds and by hyperconjugation between the SiC σ-bonds and the unoccupied orbitals of the C=C double bonds.  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