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 39 毫秒
1.
An absorbance probe method was used for the investigation of photolysis of cationic photoinitiators. The rates of the photolysis of diphenyliodonium hexafluorophosphate (DPIH), diphenyliodonium tetrafluoroborate (DPIB), di(tert-butylphenyl)iodonium tetrafluoroborate (DTIB), di(tert-butylphenyl)iodonium bromate (DTIBr), triphenylsulfonium hexafluorophosphate (TPS) and cyclopropyldiphenylsulfonium tetrafluoroborate (CPS) were studied in the presence of acid indicator quinaldine red (QR) in acetonitrile. Diphenyliodonium hexafluorophosphate and triphenylsulfonium hexafluorophosphate showed the highest photolysis rate. Photopolymerization of 1,3-di(9-carbazolyl)-2-propanol glycidyl ether (DCPGE) initiated with the iodonium and sulfonium salts in bulk and in solution was studied. It was established that the highest initial rate of polymerization is characteristic of DCPGE photopolymerization initiated with DPIH and TPS in bulk. The oligomers of DCPGE of number average molecular weight ( [`(Mn)]\overline{M_n} ) ranging from 710 to 1220 were obtained in these reactions in bulk and those with [`(Mn)]\overline{M_n} ranging from 1300 to 1600 were obtained in solution.  相似文献   

2.
Summary. The complexes [CTA][Mn(II)(SQ)3] were isolated in the solid state and purified. SQ is the o-semiquinone of L-dopa or dopamine and CTA is the cetyltrimethylammonium cation. These complexes were characterized by Raman, infrared, EPR and thermogravimetry (TG) techniques. The EPR spectra of the solids presented an intense signal characteristic of the o-semiquinone radical anion with g=2.0062 and g=2.0063 for L-dopa and dopamine, respectively. Six characteristic lines around the organic radical signal confirm the presence of the Mn2+ ion. The most intense Raman bands were observed at for dopamine and at 1356 cm–1 for L-dopa and assigned to a C–O stretching with major C1–C2 character. The absence of an intense Raman band at ca. , characterizes the ligands as an o-semiquinone radical anion. Broad bands in the region can be assigned to deformations associated with the five-member ring chelate including the manganese ion, the oxygens, and the C1–C2 bonds. The more intense IR bands for the dopamine and the L-dopa-derived ligands at are assigned to CO. Mass loss mechanisms for the two complexes, based on the TG results, were proposed and confirm the formula proposed.  相似文献   

3.
A novel micellization induced by photolysis was attained using a poly(4-tert-butoxystyrene)-block-polystyrene diblock copolymer (PBSt-b-PSt). BSt-b-PSt showed no self-assembly in dichloromethane and existed as isolated copolymers. Dynamic light scattering demonstrated that the copolymer produced spherical micelles in dichloromethane by the irradiation with a high-pressure mercury lamp in the presence of photoacid generators, such as bis(alkylphenyl)iodonium hexafluorophosphate (BAI), diphenyliodonium hexafluorophosphate (DPI), and triphenylsulfonium triflate (TPS). The irradiation time to promote the micellization increased in the order of BAI < DPI < TPS, depending on the UV absorption intensity of the photoacid generators. The efficiency to promote the micellization was also dependent on the block length of the copolymer. Under an identical PBSt block length, the copolymer with the shorter PSt block length more easily formed micelles. The 1H NMR analysis confirmed that the PBSt-b-PSt copolymer was converted into poly(4-vinyl phenol)-block-PSt, resulting in micelles by self-assembly.  相似文献   

4.
Polyisoprene (PIP) was found to react with trifluoroacetic acid (TFA) to give an adduct. Saponification of the ester gave a new alternating copolymer of ethylene and α-methyl vinyl alcohol. TFA did not react with polybutadiene (PBD) under these conditions, thus providing a way to produce amphiphilic block copolymers from PBD-b-PIP. TFA addition to the PIP block took place cleanly at an ambient temperature with 2 equiv of the acid in toluene to give block copolymer of PBD and trifluoroacetated PIP. This polymer is very soluble in toluene regardless of molecular weight. Methanolysis with NaOMe cleaved the ester to give PDB-b-(ethylene-alt-α-methyl vinyl alcohol) (PIPOH) in a MeOH/toluene mixture. Low molecular weight hydroxylated diblock copolymer is a viscous liquid when the ratio of PIP/PBD is 0.1 ( $\overline {M_n }$ = 4100, D = 1.3), but a solid with the ratio 0.5 ($\overline {M_n }$ = 7170, D = 1.6). High molecular weight polymer ($\overline {M_n }$ = 114,000, D = 1.4) with PIP/PBD ratio 0.1 is a hazy rubbery material. Block copolymers of PIPOH and poly(methacrylic acid) was also obtained from copolymers of PIP and poly(t-butyl methacrylate). The hydroxylated copolymers showed surface activity by monolayer formation on a Langmuir–Blodgett trough. The transfer of the monolayer on a silicon wafer gave z-type deposition, with the average ellipsometer thickness of the layer being about 40 Å thick per monolayer for $\overline {M_n }$ = 4100 copolymer. © 1994 John Wiley & Sons, Inc.  相似文献   

5.
Living cationic polymerization of p-methylstyrene has been investigated with use of hydrogen iodide/zinc halide initiating systems (HI/ZnX2; X = Cl, I) in toluene and methylene chloride solvents. The best results were obtained with HI/ZnCl2 below 0°C where the zinc salt was employed in excess over hydrogen iodide (HI/ZnCl2 = 1/5 molar ratio). Under these conditions, the number-average molecular weights $ \left( {\overline {M_n } } \right)$ of the produced polymers increased proportionally to monomer conversion, further increased upon addition of fresh feeds of p-methylstyrene into completely polymerized reaction mixtures, and were in good agreement with the calculated values assuming that one living polymer chain forms per molecule of hydrogen iodide; the molecular weight distributions (MWDs) of the polymers were fairly narrow $ \left( {{{\overline {M_w } } \mathord{\left/ {\vphantom {{\overline {M_w } } {\overline {M_n } }}} \right. \kern-\nulldelimiterspace} {\overline {M_n } }} = 1.1 - 1.3} \right) $ throughout these processes. In contrast, when equimolar mixtures of hydrogen iodide and zinc chloride or iodide were employed, the polymerizations were rather slow even in methylene chloride at +25°C, and the product polymers exhibited bimodal MWDs, the lower-molecular weight fraction of which was mediated by long-lived growing species. Addition of tetra-n-butylammonium iodide as a common ion salt (nBu4NI/HI = 1/200 molar ratio) led to unimodal MWDs consisting of the long-lived lower polymer fraction alone.  相似文献   

6.
Summary. The synthesis and reaction with two oxidation agents is described for N-phenyl-1-(2-oxo-1-azacycloalkyl)methanesulfonamides. Their oxidation was carried out using radicals and 3-chloroperbenzoic acid. In both cases, the EPR spectra of corresponding aminoxyl radicals were recorded. Their simulation confirmed that the –SO2– group in the neighbourhood of the – – fragment does not prevent the interaction of the unpaired electron with the methylene protons and the nitrogen atom of the heterocyclic ring.  相似文献   

7.
Four novel onium salts (onium‐polyoxometalate) have been synthesized and characterized. They contain a diphenyliodonium or a thianthrenium (TH) moiety and a polyoxomolybdate or a polyoxotungstate as new counter anions. Outstandingly, these counter anions are photochemically active and can sensitize the decomposition of the iodonium or TH moiety through an intramolecular electron transfer. The phenyl radicals generated upon UV light irradiation (Xe–Hg lamp) are very efficient to initiate the radical polymerization of acrylates. Cations are also generated for the cationic polymerization of epoxides. Remarkably, these novel iodonium and TH salts are characterized by a higher reactivity compared with that of the diphenyliodonium hexafluorophosphate and the commercial TH salt, respectively. Interpenetrating polymer networks can also be obtained under air through a concomitant cationic/radical photopolymerization of an epoxy/acrylate blend (monomer conversions > 65%). The photochemical mechanisms are studied by steady‐state photolysis, cyclic voltammetry, and electron spin resonance techniques. © 2015 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2015 , 53, 981–989  相似文献   

8.
Laser flash photolysis of diphenyliodonium salts produces phenyliodinium radical cation (PhI), which was also generated independently by flash-induced electron transfer from iodobenzene to a phenanthrolinium salt. Apparent second-order rate constants were determined for reaction of the transient (PhI) with nucleophiles, including iodobenzene and cyclohexene oxide. Quantum yields of formation of acid from stationary photolysis of diphenyliodonium hexafluoroarsenate were found to be significantly higher than yields of iodobenzene. These results may be explained by facile reaction of PhI with PhI to yield a new iodonium salt together with a proton. High reactivity of PhI with cyclohexene oxide suggests that the transient may directly initiate cationic polymerization of epoxides.  相似文献   

9.
The photoinitiated ring‐opening cationic polymerization of a monofunctional benzoxazine, 3‐phenyl‐3,4‐dihydro‐2H‐1,3‐benzoxazine, with onium salts such as diphenyliodonium hexafluorophosphate and triphenylsulfonium hexafluorophosphate as initiators was examined. The structures of the polymers thus formed were complex and related to the ring‐opening process of the protonated monomer either at the oxygen or nitrogen atoms. The phenolic mechanism also contributed, but its influence decreased with decreasing monomer concentration. Thermal properties of the polymers were also investigated by differential scanning calorimetry and thermogravimetric analysis. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 3320–3328, 2003  相似文献   

10.
Diphenylphenacylsulfonium tetrafluoroborate (DPPS+BF4–) salt possessing both phenacyl and sulfonium structural units was synthesized and characterized. DPPS+BF4– absorbs light at relatively higher wavelengths. The direct and sensitized initiation activity of the salt in both cationic and free radical photopolymerizations was investigated and compared with that of its analogue triphenylsulfonium tetrafluoroborate (TPS+BF4–). Differential scanning photocalorimetry and conventional gravimetric studies revealed that DPPS+BF4– showed higher efficiency for direct and sensitized photopolymerizations of most of the monomers investigated. Although, principally both homolytic and/or heterolytic cleavage is possible, theoretical studies suggested that homolytic pathway is more favored for the generation of reactive initiating species. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2018 , 56, 451–457  相似文献   

11.
Reaction of guaiazulene (1) with thiophene-2,5-dicarbaldehyde (2) in methanol in the presence of hexafluorophosphoric acid at 25 °C for 3 h gives as high as 90% isolated yield of the delocalized dicarbenium-ion compound, 2,5-thienylenebis(3-guaiazulenylmethylium) bis(hexafluorophosphate) (3). Similarly, reaction of 1 with furan-2,5-dicarbaldehyde (4) under the same conditions as the above reaction affords the corresponding dicarbenium-ion compound, 2,5-furylenebis(3-guaiazulenylmethylium) bis(hexafluorophosphate) (5), in 84% isolated yield. Along with a facile preparation and the spectroscopic and electrochemical properties of 3 and 5, comparative studies on the 1H and 13C NMR spectral and chemical properties of 3 and 5 with those of the delocalized mono- and dicarbenium-ion compounds [i.e., (3-guaiazulenyl)(2-thienyl)methylium hexafluorophosphate (7), (2-furyl)(3-guaiazulenyl)methylium hexafluorophosphate (9), α,α′-bis(3-guaiazulenylmethylium) bis(tetrafluoroborate) (10), 1,2-phenylenebis(3-guaiazulenylmethylium) bis(hexafluorophosphate) (11), and 1,4-phenylenebis(3-guaiazulenylmethylium) bis(tetrafluoroborate) (12)] are reported. Moreover, referring to the results of the X-ray crystallographic analyses of 7, 9, 11, and 12, the optimized 2,5-thienylenebis(3-guaiazulenylmethylium)- and 2,5-furylenebis(3-guaiazulenylmethylium)-ion structures for 3 and 5, calculated by a WinMOPAC (version 3.0) program using PM3 as a semiempirical Hamiltonian, are described.  相似文献   

12.
13.
The products of photolysis of N-substituted salicylic acid amides, viz., 2-hydroxy-3-tert-butyl-5-ethylbenzoic acid N-(4-hydroxy-3,5-di-tert-butylphenyl)amide (1) and 2-hydroxybenzoic acid N-[3-(4-hydroxy-3,5-di-tert-butylphenyl)prop-1-yl]amide (2), in heptane were studied by optical spectroscopy and stationary and nanosecond laser photolysis (Nd: YAG laser, 355 nm). It was shown by the method of partial deuteration of amides 1 and 2 that they exist in both the unbound state and as complexes with intraand intermolecular hydrogen bond. Amides 1 and 2 are subjected to photolysis, which results in the formation of a triplet state and phenoxyl radicals RO? presumably due to the absorption of the second photon by the excited singlet state. The formation of radical products due to N–H bond ionization was not observed. The main channel of decay of the triplet state and radicals RO? is triplet–triplet annihilation and recombination (k r ≈ 2.3?108 L mol–1 s–1), respectively. The UV irradiation of compounds 1 and 2 leads to the excitation of the amide groups, and no formation of radical products due to N–H bond ionization was observed.  相似文献   

14.
Summary In nitromethane, the methanolysis of aryl acetates is catalysed by the tetrafluoroborate and hexafluorophosphate salts of the ( 5-ethyltetramethylcyclopentadienyl) ( 6-benzen)rhodium(III) cation. Under the conditions of the methanolysis, the anion of the latter salt reacts with methanol to give dimethyl phosphorofluoridate. The hydrogen fluoride formed also in this reaction is thought to be responsible for the greater efficiency of the hexafluorophosphate salt as a catalyst for the methanolysis.  相似文献   

15.
Densities, ??, and viscosities, ??, of binary mixtures of 2-methyl-2-propanol with acetone (AC), ethyl methyl ketone (EMK) and acetophenone (AP), including those of the pure liquids, were measured over the entire composition range at 298.15, 303.15 and 308.15?K. From these experimental data, the excess molar volume $V_{\mathrm{m}}^{\mathrm{E}}$ , deviation in viscosity ????, partial and apparent molar volumes ( $\overline{V}_{\mathrm{m},1}^{\,\circ }$ , $\overline{V}_{\mathrm{m},2}^{\,\circ }$ , $\overline{V}_{\phi ,1}^{\,\circ}$ and $\overline{V}_{\phi,2}^{\,\circ} $ ), and their excess values ( $\overline{V}_{\mathrm{m},1}^{\,\circ \mathrm{E}}$ , $\overline{V}_{\mathrm{m,2}}^{\,\circ \mathrm{ E}}$ , $\overline {V}_{\phi \mathrm{,1}}^{\,\circ \mathrm{ E}}$ and $\overline{V}_{\phi \mathrm{,2}}^{\,\circ \mathrm{ E}}$ ) of the components at infinite dilution were calculated. The interaction between the component molecules follows the order of AP > AC > EMK.  相似文献   

16.
17.
Zusammenfassung Das Ziel der vorliegenden Untersuchung ist vor allem die Ermittlung des kristallinen und amorphen Anteils der untersuchten Proben aus dem mittleren Schwankungsquadrat der Elektronendichte, die aus der Absolutintensität der Röntgenkleinwinkelstreuung hergeleitet werden kann. Es sollen ferner die mittleren Dimensionen der kristallinen Anteile aus der Intensität des Auslaufes der Streukurven berechnet werden.
The mean square electron density fluctuations of three Nylon-6 samples (untreated, stretched, stretched and tempered) were obtained from the scattering curves. The degree of crystallinity of each sample was calculated from and from the densities of the crystalline and amorphous phases,d c =1.23 andd a =1.10, which were available from the literature. Crystallinities thus obtained were found to be as much as 50% less than those calculated from measured sample densitiesd. The discrepancy was ascribed to a small uncertainty ind a , the amorphous density: the calculated crystallinities are very sensitive to small changes in this parameter. If one sets the crystalline volume fractions obtained from equal to those obtained fromd, an equation ford a in terms of andd results, and is solved to yieldd a -values only about 1% higher than the literature value and of course consistent with both X-ray and density determinations. The volume fractions of the crystalline phases range from 20 to 30%. The possibility that the samples contained small amounts of air-filled holes may be excluded, since thed a -values (based on andd) calculated on the assumption of a volume fraction of as little as 0.3% for such holes are already much too high.The specific inner surface and an average dimension of the regions of the two phases were also determined. The crystalline regions are roughly 30 Å in size, the amorphous ones three to four times as large, and largest for the tempered sample.


Mit 3 Abbildungen  相似文献   

18.
利用基于进化算法的晶体结构预测软件USPEX,并结合第一性原理方法对HfB的稳定晶体结构进行全局搜索。在基态条件下,新发现2个HfB晶体结构(空间群:P6m2和R3m)。其中,P6m2结构比已报道的Hf B晶体结构(空间群:Pnma、Cmcm、I41/amd和Fm3m)具有最低的基态能量。这些结构中,B原子分别以二维类石墨烯(P6m2和R3m结构),zig-zag链(Pnma、Cmcm和41/amd结构)和孤立原子(Fm3m结构)3种形式存在,从而导致它们具有显著的化学键合特征、高温稳定性和强韧特性差异。  相似文献   

19.
利用基于进化算法的晶体结构预测软件USPEX,并结合第一性原理方法对HfB的稳定晶体结构进行全局搜索。在基态条件下,新发现2个HfB晶体结构(空间群:P6m2和R3m)。其中,P6m2结构比已报道的HfB晶体结构(空间群:Pnma、Cmcm、I41/amdFm3m)具有更低的基态能量。这些结构中,B原子分别以二维类石墨烯(P6m2和R3m结构),zig-zag链(Pnma、Cmcm和41/amd结构)和孤立原子(Fm3m结构)3种形式存在,从而导致它们具有显著的化学键合特征、高温稳定性和强韧特性差异。  相似文献   

20.
Zusammenfassung Die Molekulargewichtsverteilung von neun Polystyrolproben mit enger Verteilung (anionische Polymerisation) wurde mit Hilfe der Säulenfraktionierungsmethode vonBaker undWilliams bestimmt. Die ursprünglichen Polymeren besitzen alle einen höheren niedrigmolekularen Anteil, als einer einheitlichen Verteilungsfunktion entsprechen würde. Die Verteilungsbreite des Hauptanteils (82 bis 90%) entspricht -Werten von 1,03 bis 1,005, die der Gesamtpolymeren liegt bedeutend höher, mit einem unteren Grenzwert von 1,08 bis 1,21. Durch eine einfache einstufige Fraktionierung kann eine merkliche Verschärfung der Verteilung erreicht werden.  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