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.
Polystyrene was crosslinked in either 1,2-dichloroethane or carbon tetrachloride in the presence of aluminum chloride. Apparently, the reactions involve Friedel–Crafts substitution of phenyl ring with CH2CH2Cl or CCl3 groups, which than participate in crosslinking, giving CH2CH2 or CCl2, CCl, and C bridges. In the first stage a charge-transfer complex is formed between AlCl3, polystyrene and the solvent. After heating this complex above 35–40°C a rapid formation of HCl occurs and a crosslinked polymer is formed. This final product is insoluble, infusible, and inflammable. It decomposes at 400°C without melting.  相似文献   

2.
A process for the chemical modification of polybutadienes and natural rubber by various metallocene compounds is described. Soluble products of up to 43% ferrocene content were obtained. The effect of substrate, metallocene, and reaction conditions on the course and extent of substitution was investigated. The glass transition temperature Tg was found to increase considerably with the degree of substitution, e.g., cis-polybutadiene substituted with ferrocene (18 mole-%) has a Tg of 30°C, as compared with ?91°C for the unsubstituted polymer.  相似文献   

3.
The polymerization of the complex of methyl methacrylate with stannic chloride, aluminum trichloride, or boron trifluoride was carried out in toluene solution at several temperatures in the range of 60° to ?78°C by initiation of α,α′-azobisisobutyronicrile or by irradiation with ultraviolet rays. The tacticities of the resulting polymers were determined by NMR spectroscopy. Both the 1:1 and the 2:1 methyl methacrylate–SnCl4 complexes gave polymers with similar tacticities at the polymerization temperatures above ?60°C. With decreasing temperature below ?60°C, the isotacticity was more favored for the 2:1 complex, whereas the tacticities did not change for the 1:1 complex. On the ESR spectroscopy of the polymerization solution under the irradiation of ultraviolet rays at ?120°C, the 1:1 SnCl4 complex gave a quintet, while the 2:1 SnCl4 complex gave both a quintet and a sextet. The sextet became weaker with increasing temperature and disappeared at ?60°C. This behavior of the sextet corresponds to the change of the tacticities of polymer for the 2:1 SnCl4 complex. An intra–intercomplex addition was suggested for the polymerization of the 2:1 complex, which took a cis-configuration on the basis of its infrared spectra. The sextet can be ascribed to the radical formed by the intracomplex addition reaction, while the quintet can correspond to that formed by the intercomplex addition reaction. The proportion of the intracomplex reaction was estimated to be about 0.25 at ?75°C, and the calculated value of the probability of isotactic diad addition of the intracomplex reaction was found to be almost unity.  相似文献   

4.
Chelate Complexes of Rhenium Tetrachloride. The Crystal Structures of ReCl4(DME) and ReCl4(DPPE) · Tolan Bright green crystals of ReCl4(DME) have been prepared by the reaction of rhenium pentachloride with dimethoxyethane (DME) in dichloromethane. ReCl4(DPPE) · tolan was obtained in form of red crystals by the reaction of the alkyne complex [ReCl4(Ph? C?C? Ph)(POCl3)] with bis(diphenylphosphino)ethane (DPPE) in dichloromethane. The complexes were characterized by X-ray structure determinations. ReCl4(DME): Space group I4 2d, Z = 8, 829 observed unique reflexions, R = 0.022. Lattice dimensions at 19.5°C: a = b = 960.60(6), c = 2337.2(6) pm. The complex forms monomeric molecules with DME as chelating ligand; the Re? O bond lengths are 213.1 pm. The chlorine atoms, arranged in trans position to the chelating ligand, have slightly shorter Re? Cl bonds than the chlorine atoms in cis position (232,1 pm). ReCl4(DPPE) · tolan: Space group P21/n, Z = 4,4313 observed unique reflexions, R = 0.040. Lattice dimensions at ?80°C: a = 1095.7(1), b = 1764.2(2), c = 1898.0(2) pm, β = 99.229(8)°. The compound consists in form of monomeric molecules [ReCl4(DPPE)] and diphenylacetylene molecules, which are incorporated in the lattice. The two phenyl rings of the tolan molecules are twisted towards each other along the C? C axis with a dihedral angle of 21°. The DPPE molecules are bonded to the rhenium atom in a chelating fashion with medium Re? P lengths of 250.4 pm. The chlorine atoms, arranged in trans position to this ligand, with Re? Cl bond lengths of 234.5 pm are slightly longer than the Re? Cl bonds in cis position with 232.3 pm.  相似文献   

5.
<正> (Ph2Ppy)2(μ-Cl)2Cu2Cl2 (C34H28Cl4Cu2N2P2): Mr=795.5, triclinic, Pl, a=13.891(2), b= 13.196(3), c= 20.158(4) A, d= 90.28(1)°, β=110.05(2)°, r= 90.13(1)°, Z=4, V= 3471.04A3, Dx=1.53 gcm-3, R=0.066, Rw= 0.076 for 5486 observed unique reflections. The complex was prepared by the reaction of (Ph2Ppy)2NiCl2 with CuC≡CPh in CH2Cl2 solution.  相似文献   

6.
Conditions of formation of the Ti2NiH3.3 hydride phase in the reaction of the Ti2Ni intermetallic compound with ammonia and hydrogen have been determined. The products of the reaction of the intermetallide with ammonia in the presence of the NH4Cl activator in the temperature range 100–500°C have been identified. It has been shown that the use of ammonia at temperatures >400°C leads to the formation of titanium nitride and nickel.  相似文献   

7.
Chain-chlorinated polystyrene samples have been prepared and characterised, containing from 2·5 to 34·3% chlorine by weight (from 0·1 to 1·3 chlorine atoms per styrene unit). Chlorination has been found to involve mainly substitution of the tertiary hydrogen atoms, followed by methylene substitution at reactant concentrations of more than 1 mole Cl2 per styrene unit. Reaction with chlorine is quantitative in CCl4 in the absence of air and the amount of ring-chlorination as a side reaction is very small.Chlorination in the backbone destabilises the polymer, which first loses hydrogen chloride on heating. The double bonds formed provide points of weakness for chain scission, which occurs at lower temperatures than for PS. Under programmed heating, the dehydrochlorination and chain scission reactions overlap to some extent in temperature range, but highly conjugated partially degraded polymer can be made from extensively chain-chlorinated PS. At more than 1 Cl per styrene unit, HCl is almost the sole volatile product, although the conjugated polymer breaks up to chain fragments above 300°C.  相似文献   

8.
The First Fluoro-trithiatetraphospha-heptanes Fluorinated α- and β-tetraphosphorus-trisulfur molecules were prepared by the reaction of α- or β-P4S3X2 (X = Cl, Br, I) with (n-butyl)3SnF. The substitution reaction of the α-isomers yields under retention of the configuration at the phosphorus atoms α-P4S3XF and α-P4S3F2. In the reaction of the β-isomers more products were observed, because the configuration of the phosphorus atom can be retained or inversed in the first step of the substitution which yields β-P4S3XexoFexo or β-P4S3XexoFendo. The mass relation of the products depends on the halogen ligand. In the second substitution β-P4S3(Fexo)2 or β-P4S3FexoFendo are formed. β-P4S3(Fendo)2 was not observed. By the reaction of β-P4S3I2 with BiX3 (X = Cl, Br, I) we also were able to prepare small amounts of β-P4S3XendoXexo-molecules (X = I, Cl, Br) with an inversed configuration at one phosphorus atom. The 31P- and 19F-NMR parameter of all compounds are discussed.  相似文献   

9.
The photo-oxidation of PVC has been studied over the temperature range 30–150°C. Initiation with ultraviolet (2537A) radiation has been correlated with the presence of minute amounts of ozone. The contribution of atomic oxygen and singlet oxygen (1Δg) molecules to the initiation mechanism is discussed. The β-chloroketones probably formed in the photo-oxidation of PVC, decomposed according to a Norrish type I reaction without loss of chlorine atoms. The gaseous products of the photo-oxidation of PVC at 30°C were carbon dioxide, carbon monoxide, hydrogen, and methane. Hydrogen chloride was obtained only when PVC was heated at high temperatures. When PVC was photo-oxidized and then heated at high temperature, benzene was obtained in addition to hydrogen chloride. The gaseous products from the photo-oxidations of model compounds, such as 4-chloro-2-butanone and 2,4-dichloropentane, were also compared with those from PVC. Hydrogen chloride was detected only after photo-oxidation at temperatures of 25°C or higher. Therefore, it was concluded that hydrogen chloride is mainly a product of thermal decomposition. Since unsaturation was not observed in photo-oxidized PVC films, the cause of discoloration is unclear. When PVC was modified by stabilizers or additives, the oxidative degradation was further complicated by side reactions with the additives.  相似文献   

10.
Living cationic polymerization of a vinyl ether with a naphthyl group [2‐(2‐naphthoxy)ethyl vinyl ether, βNpOVE] was achieved using base‐assisting initiating systems with a Lewis acid. The Et1.5AlCl1.5/1,4‐dioxane or ethyl acetate system induced the living cationic polymerization of βNpOVE in toluene at 0 °C. The living nature of this reaction was confirmed by a monomer addition experiment, followed by 1H NMR and matrix‐assisted laser desorption ionization time‐of‐flight mass spectrometry (MALDI‐TOF‐MS) analyses. In contrast, the polymerization of αNpOVE was not fully controlled; under similar conditions, it produced polymers with broad molecular weight distributions. The 1H NMR and MALDI‐TOF‐MS spectra of the resultant poly(αNpOVE) revealed that the products had undesirable structures derived from Friedel–Crafts alkylation. The higher reactivity of αNpOVE in electrophilic substitution reactions, such as the Friedel–Crafts reaction, was attributable to the greater electron density of the naphthyl ring, which was calculated based on frontier orbital theory. The naphthyl groups significantly affected the properties of the resultant polymer. For example, the glass transition temperatures (Tg) of poly(NpOVE)s are higher by approximately 40 °C than that of poly(2‐phenoxyethyl vinyl ether). © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

11.
The rate of polymerization of thiophene, at concentrations of catalyst (SnCl4), and thiophene of the same order as was subsequently used in studying the reaction between thiophene and di(chloromethyl)benzene, is of the order of 10-2%/hr at 30°C. There is no significant self-condensation of DCMB under the same conditions. Since the reaction between thiophene and DCMB is complete at 30°C in minutes rather than hours, it is assumed that self-condensation of thiophene or DCMB during the reaction between them will be negligible and should not influence the course of the reaction or the structure of the resulting polymer. Reaction at 30°C is much too fast for convenient study. A temperature of 0°C is more appropriate and was used in subsequent kinetic work. The first two products of the condensation of p-di(chloromethyl)benzene (DCMB) with thiophene have been identified by a combination of mass, infrared, and nuclear magnetic resonance spectroscopy as thenylchloromethylbenzene (TCMB) and dithenylbenzene (DTB). DCMB, TCMB, and DTB have been estimated quantitatively during the course of the reaction by gas-liquid chromatography (GLC), and it has been established that the rates of each of the two reaction steps is first-order with respect to the chloro compound (DCMB and TCMB respectively), thiophene, and SnCl4. Rate constants for these two consecutive reactions were calculated to be k1 = 2.79 × 10-4l.2/mole2-sec, k2 = 6.37 × 10-3l.2/mole2-sec; the corresponding energies of activation are E1 = 7.93 kcal/mole, E2 = 7°67 kcal/mole. These rate constants are appreciably higher than values previously obtained for the corresponding DCMB–benzene reactions.  相似文献   

12.
Zn pack coating formation takes place in three steps as differential scanning calorimetry shows. The initial step (at 193.9°C) is endothermic and involves the transformation of α-NH4Cl to β-NH4Cl and the NH4Cl decomposition to NH3 and HCl. During the second step (at 248.6°C), which is exothermic, Zn2+ salts are formed and most probably ZnCl2. Finally at 264.1°C (endothermic reaction) it seems that ZnCl2 is decomposed to form Zn that is deposited on the ferrous substrate. The as-cast Zn diffuses in the iron substrate forming the gamma and delta phase of the Fe–Zn phase diagram. Al2O3 is not involved in the above-mentioned mechanism and acts only as filler.  相似文献   

13.
Samples of β-Co2(OH)3Cl and Zn5(OH)8Cl2 · H2O have been prepared and their thermal decomposition studied in air and N2 by DTA and TG up to 1000°C. X-Ray diffraction analysis of the thermal treatment products in air at various temperatures from 100 to 100°C was also carried out. The results obtained made it possible to establish the steps through which the pyrolysis of both compounds proceeds.  相似文献   

14.
Nucleophilic substitution of PVC with sodium thiophenate was carried out in cyclohexanone solution at 5, 25, 40, 60, and 70°C. The initial rate obeys an Arrhenius law from 25 to 60°C, with an activation energy of 70 kJ/mol. Conversion limits are observed which strongly depend on the temperature. The stereoselectivity of the reaction with respect to the configurational triads does not depend on the temperature: the distribution of configurations is only dependent on the conversion. Assuming an SN2 substitution mechanism governed by steric factors, the Monte Carlo simulation procedure described in a prior study is shown to give a good account for all temperatures above 40°C assuming for the mm, mr or rm, and rr triads a reactivity such as Rmm = 2 Rmr and Rrr nil at low temperature and very low at temperatures ≥ 40°C. The low conversion limits observed at 5 and 25°C cannot be explained by a limited accessibility of a part of the polymer. Finally, it is shown that the elimination reaction, which remains limited, does not interfere with the substitution process.  相似文献   

15.
The branching reaction in the radical polymerization of vinyl acetate was studied kinetically. Branching occurs by polymer transfer as well as terminal double-bond copolymerization. The chain-transfer constants to the main chain (Cp,2) and to the acetoxy methyl group (Cp,1) on the polymer were calculated on the basis of the experimental data described in the preceding paper giving Cp,2 = 3.03 × 10?4, Cp,1 = 1.27 × 10?4 at 60°C, and Cp,2 = 2.48 × 10?4, Cp,1 = 0.52 × 10?4 at 0°C. Chain transfer to monomer is important with respect to the formation of the terminal double bond. The total values of transfer constants to the α- or β-position in the vinyl group and the acetoxymethyl group in vinyl acetate was determined to be 2.15 × 10?4 at 60°C. The transfer constant to the acetyl group in the monomer (Cm,1) was also evaluated to be 2.26 × 10?4 at 60°C from the quantitative determination of the carboxyl terminals in PVA. These facts suggest that the chain-transfer constant to the α- or β-position in the monomer (Cm,2) is nearly equal to zero within experimental error. Copolymerization reactivity parameters of the terminal double bond were also estimated. In conclusion, it has become clear that the formation of nonhydrolyzable branching by the terminal double-bond reaction can be almost neglected, and hence that the long branching in PVA is formed only by the polymer transfer mechanism. On the other hand, a large number of hydrolyzable branches in PVAc are prepared by the terminal double-bond reaction rather than by polymer transfer.  相似文献   

16.
Mixtures of reagent grade CaHPO4, MgO, and CaCO3 were sintered at 1000°C both in air and in dry carbon dioxide to investigate the stability field of β-Ca3(PO4)2 in the ternary CaO? P2O5? MgO system along the CaO? P2O5 and the β-Ca3(PO4)2? Mg3(PO4)2 join. The extent of substitution of Mg for Ca in β-Ca3(PO4)2 was determined to be near 14.7% Mg of the cat-ions, both in air and in dry carbon dioxide. Lattice parameters of products with various degrees of substitution are presented. There is a decrease of the c-axis up to 10% substitution after which an increase of the c-axis occurs on further substitution of Ca by Mg. This is probably related to substitution in two different sublattices. At 1000°C the carbonate ion could not be incorporated in the β-Ca3(PO4)2 structure, and along the CaO? P2O5 join the Ca/P range of solid solution for β-Ca3(PO4)2 was found to be immeasurably small both in air and in dry carbon-dioxide.  相似文献   

17.
The reaction of of 4‐amino‐5‐ethyl‐2H‐1,2,4‐triazole‐3(4H)‐thione (AETT, L ) with furfural in methanol led to the corresponding Schiff‐Base ( L1 ). The reaction of L1 with [Cu(PPh3)2]Cl in methanol gave to the neutral compound [( L1 )Cu(PPh3)2Cl] ( 1 ). By recrystallization of 1 from CH3CN the complex [( L1 )Cu(PPh3)2Cl]·CH3CN ( 1a ) was obtained. All compounds were characterized by infrared spectroscopy, elemental analyses as well as by X‐ray diffraction studies. Crystal data for L1 at ?80 °C: space group with a = 788.4(1), b = 830.3(2), c = 928.8(2) pm, α = 84.53(1)°, β = 65.93(1)°, γ = 72.02(1)°, Z = 2, R1 = 0.0323; for 1 at ?100 °C: space group with a = 1166.3(1), b = 1423.8(2), c = 1489.1(2) pm, α = 62.15(1)°, β = 72.04(1)°, γ = 88.82(1)°, Z = 2, R1 = 0.0338 and for 1a at ?100 °C: space group P21/c with a = 1294.1(1), b = 1019.8(2), c = 3316.9(4) pm, β = 94.73(1)°, Z = 4, R1 = 0.0435.  相似文献   

18.
Abstract

Two copper(I) complexes, [Cu(H2net)2Cl] · CH2Cl2 (1) and [Cu(H2nmt)2Cl]2 · (CHCl3)2 (2), were synthesized by the reaction of CuCl2 · 2H2O with N-(p-nitrophenyl)-N′-(ethoxycarbonyl)-thiourea (H2net) and N-(p-nitrophenyl)-N′-(methoxycarbonyl)-thiourea(H2nmt), respectively. Both complexes crystallize in the monoclinic space group C2/c. For complex 1, a = 29.52(2), b=13.920(6), c = 14.873(3)Å; β= 101.75(2)°, V = 5984(4) Å3, Z = 8 and R = 0.053; for complex 2, a = 30.68(1), b = 13.369(4), c = 14.226(7) Å, β = 99.52(4)°. V = 5754(4) Å3, Z = 4 and R = 0.063. In complex 1, two H2net molecules are bonded to Cu(I) atom through two S atoms forming a mononuclear complex with trigonal geometry for the Cu(I) ion [Cl(1)-Cu-S(1)=118.54(7), Cl(1)-Cu-S(2)=119.70(7), S(1)-Cu-S(2)=112.17(8)°, Cu-S(1) = 2.251(2), Cu-S(2) = 2.255(2), Cu-Cl(1) = 2.263(2) Å]. Complex 2 is a dimer formed by long Cu-S interactions [Cu-S* = 2.607(3) Å] from adjacent twc H2nmt molecules; the Cu(I) ion has distorted tetrahedral coordination [Cl(1)-Cu-S(1) = 119.8(1), Cl(1)-Cu-S(2)=120.0(1), S(1)-Cu-S(2)=108.85(9)°] with unequal Cu-S [2.268(2), 2.247(2)Å] and Cu-Cl(1) [2.255(2)Å] bonds.  相似文献   

19.
Heteroatom-functionalized Methylgold Complexes: Synthesis and Structure of Chloromethyl(triphenylphosphine)- and Phenylthiomethyl(trimethylphosphine)gold [AuCl(PPh3)] reacts with Mg(CH2Cl)Cl, prepared in situ from CH2ClI and iPrMgCl, in ether at –65 °C to give [Au(CH2Cl)(PPh3)] ( 1 a ). 1 a reacts with LiI, NaOMe and PPh3 to give [Au(CH2I)(PPh3)] ( 2 ), [Au(CH2OMe)(PPh3)] ( 3 ) and [Au(CH2PPh3)(PPh3)]Cl ( 4 ), respectively. 2 decomposes rapidly at room temperature, yielding ethylene and [AuI(PPh3)]. The reaction of [AuCl(PMe3)] with LiCH2SPh in THF affords [Au(CH2SPh)(PMe3)] ( 5 ). The chloromethyl and the phenylthiomethyl complex 1 a and 5 were isolated and characterized by NMR (1H, 13C, 31P) spectroscopy as well as by single-crystal X-ray structure analysis. In the solid state discrete molecules of 1 a and 5 are found with linear C–Au–P units [C–Au–P 179,8(4)° ( 1 a ), 179,1(1)° ( 5 )]. The angle Au–C–Cl (115,4(6)°) in 1 a is slightly greater than the tetrahedral angle.  相似文献   

20.
1-Chloro-2-β-naphthylacetylene (ClβNA) polymerized in good yields in the presence of MoCl5-based catalysts. The highest weight-average molecular weight of poly(ClβNA) reached about 3 × 105. The polymer was a yellow solid (absorption cutoff in CHCl3 450 nm). It was soluble in toluene, chloroform, etc., and provided a tough film by the solvent casting method. The polymer retained its weight up to 300°C in air; it was thermally less stable than poly(1-chloro-2-phenylacetylene) but more stable than poly(β-naphthylacetylene). The oxygen permeability coefficient (PO2) of this polymer was 19 barrers (25°C), which is fairly small for a substituted polyacetylene. © 1996 John Wiley & Sons, Inc.  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