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.
Although zinc? cobalt (III) double metal cyanide complex (Zn? Co (III) DMCC) catalyst is a highly active and selective catalyst for carbon dioxide (CO2)/cyclohexene oxide (CHO) copolymerization, the structure of the resultant copolymer is poorly understood and the catalytic mechanism is still unclear. Combining the results of kinetic study and electrospray ionization‐mass spectrometry (ESI‐MS) spectra for CO2/CHO copolymerization catalyzed by Zn? Co (III) DMCC catalyst, we disclosed that (1) the short ether units were mainly generated at the early stage of the copolymerization, and were hence in the “head” of the copolymer and (2) all resultant PCHCs presented two end hydroxyl (? OH) groups. One end ? OH group came from the initiation of zinc? hydroxide (Zn? OH) bond and the other end ? OH group was produced by the chain transfer reaction of propagating chain to H2O (or free copolymer). Adding t‐BuOH (CHO: t‐BuOH = 2:1, v/v) to the reaction system led to the production of fully alternating PCHCs and new active site of Zn? Ot‐Bu, which was proved by the observation of PCHCs with one end ? Ot‐Bu (and ? OCOOt‐Bu) group from ESI‐MS and 13C NMR spectra. Moreover, Zn?OH bond in Zn? Co (III) DMCC catalyst was also characterized by the combined results from FT‐IR, TGA and elemental analysis. This work provided new evidences that CO2/CHO copolymerization was initiated by metal? OH bond. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

2.
Anionic copolymerization of β-cyanopropionaldehyde (M1) with methyl isocyanate (MeI, M2) was studied with use of benzophenone–dilithium complex as initiator at ?78°C. The values of monomer reactivity ratio were determined to be r1 = 8.3 ± 0.3 and r2 = 0.01 ± 0.01. The structure of resulting copolymer was investigated by means of NMR analysis. The MeI unit is presumed to enter the copolymer chain through its C?N opening.  相似文献   

3.
This article deals with the latent reactive polymers having isocyanate moiety obtained from the radical copolymerization of 2‐propenyl isocyanate ( 2PI ) with styrene, 2PI with methyl methacrylate ( MMA ), and 2‐methacryloyloxyethyl isocyanate ( MOI ) with styrene. The radical copolymerization was carried out in benzene (5.00 M by total monomer) in the presence of AIBN (3.00 mol % of total monomer) at 60 °C for 24 h. The isocyanate moiety in each copolymer was stable at room temperature for more than 6 months under aerobic atmosphere, because no change of the infrared absorption based on isocyanate group of the resulting copolymer at around 2250 cm?1 was observed. Isocyanate moiety of obtained copolymer (poly( 2PI ‐co‐ St )) reacted with excess diamines or diols at 80 °C in THF solution to afford the crosslinked polymer quantitatively. These results could demonstrate that isocyanate moiety in the copolymers showed thermal and reactive latency. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 2448–2453, 2006  相似文献   

4.
4‐Methoxycarbonyloxybenzoic acid (MCOBA) and 6‐methoxycarbonyloxy‐2‐naphthoic acid (MCONA) were synthesized as new monomers to replace 4‐acetoxybenzoic acid (ABA) and 6‐acetoxy‐2‐naphthoic acid (ANA) in the synthesis of liquid crystal polymers. MCOBA and MCONA (73 : 27, mol : mol) were reacted at temperatures ranging from 220 to 325°C in bulk. The copolymer (Mw = 14,200) has a Tg (90°C) and a Tm (249°C). The MCOBA/MCONA copolymer is lighter in color than the ABA/ANA copolymer. During the copolymerization, six by‐products were collected, isolated, and analyzed, and their formation was investigated. The copolymerization rate was studied by the measurement of evolved carbon dioxide. The polymerization of MCOBA and MCONA is cleaner and faster than the polymerization of ABA and ANA. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 1703–1707, 1999  相似文献   

5.
The solution and bulk copolymerization of dicyclopentadiene (DCP) and maleic anhydride (MAH) occurs over the temperature range 80–240°C, upon the addition of a free-radical catalyst which has a short half-life at the reaction temperature. An unsaturated 1/1 MAH/DCP copolymer, derived from the copolymerization of MAH with the norbornene double bond, followed by a Wagner-Meerwein rearrangement, is obtained in the presence of a large excess of DCP at 80° C, while a saturated 2/1 MAH/ DCP copolymer, derived from the cyclocopolymerization of the residual cyclopentene unsaturation, is obtained at higher temperatures or in the presence of excess MAH. The copolymers prepared under other conditions with intermediate MAH/DCP mole ratios contain both 1/1 and 2/1 repeating units. The copolymer obtained from bulk copolymerization above 170° C contains units derived from cyclopentadiene-MAH cyclocopolymerization as well as DCP-MAH copolymerization.  相似文献   

6.
Diphenyl-, phenylethyl, and phenylmethylketene have been copolymerized with phenyl isocyanate by use of sodium naphthalene in dimethylformamide (DMF) at ?45°C. Reactivity ratios of phenyl isocyanate (r2) with diphenylketene (r1) were r1 = 0.10, r2 = 0.29; with phenylethylketene (r1) were r1 = 1.6, r2 = 0.10; and with phenyl methyl ketene (r1) were r1 = 4.8, r2 = 0.02. The same initiator and solvent system were used for homopolymerization of phenylethylketene and copolymerization with m-chloro-, p-chloro-, p-methoxy-, and m-methoxyphenyl isocyanate as well as with phenyl isocyanate. Molecular weights ranged from 1740 to 4000. The effect of substituents on the order of isocyanate incorporation into the copolymer was m-Cl = p-Cl > m-MeO > H > p-MeO. Phenylethylketene was also copolymerized with m-methoxyphenyl, p-methoxyphenyl, and p-tolyl isocyanate in tetrahydrofuran (THF) at ?78°C. Molecular weights ranged from 2800 to 10,500. The least reactive isocyanate was incorporated into the copolymer to a greater extent in this solvent than in the more polar DMF. DTA showed the presence of crystallinity only in polymers of high isocyanate content. The ketenes copolymerized less readily with alkyl isocyanates, such as ethyl, n-butyl, n-hexyl, and cyclohexyl isocyanate, than with the aromatic isocyanates when sodium naphthalene was used in either DMF or THF.  相似文献   

7.
To study living anionic polymerization, 3‐(triethylsilyl)propyl isocyanate (TEtSPI) monomer was synthesized by hydrosilylation of allylamine with triethylsilane and treatment of the resulting amine with triphosgene. The polymerization of TEtSPI was performed with sodium naphthalenide (Na‐Naph) as an initiator and in the absence and presence of sodium tetraphenylborate (NaBPh4) as an additive in tetrahydrofuran (THF) at ?78 and at ?98 °C. A highly stabilized amidate anion for living polymerization of isocyanates was generated for the first time with the combined effect of the bulky substituent and the shielding action of the additive NaBPh4, extending the living character at least up to 120 min at ?98 °C. Even the anion could exist at ?78 °C for 10 min. A block copolymer, poly(n‐hexyl isocyanate)‐b‐poly[(3‐triethylsilyl)propyl isocyanate]‐b‐poly(n‐hexyl isocyanate), was synthesized with quantitative yields and controlled molecular weights via living anionic polymerization in THF at ?78 °C for TEtSPI and ?98 °C for n‐hexyl isocyanate, respectively, with Na‐Naph with three times of NaBPh4 as a common ion salt. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 933–940, 2004  相似文献   

8.
We found that phenethylcarbamates that bear ortho‐salicylate as an ether group (carbamoyl salicylates) dramatically accelerate O?C bond dissociation in strong acid to facilitate generation of isocyanate cation (N‐protonated isocyanates), which undergo subsequent intramolecular aromatic electrophilic cyclization to give dihydroisoquinolones. To generate isocyanate cations from carbamates in acidic media as electrophiles for aromatic substitution, protonation at the ether oxygen, the least basic heteroatom, is essential to promote C?O bond cleavage. However, the carbonyl oxygen of carbamates, the most basic site, is protonated exclusively in strong acids. We found that the protonation site can be shifted to an alternative basic atom by linking methyl salicylate to the ether oxygen of carbamate. The methyl ester oxygen ortho to the phenolic (ether) oxygen of salicylate is as basic as the carbamate carbonyl oxygen, and we found that monoprotonation at the methyl ester oxygen in strong acid resulted in the formation of an intramolecular cationic hydrogen bond (>C?O+?H???O<) with the phenolic ether oxygen. This facilitates O?C bond dissociation of phenethylcarbamates, thereby promoting isocyanate cation formation. In contrast, superacid‐mediated diprotonation at the methyl ester oxygen of the salicylate and the carbonyl oxygen of the carbamate afforded a rather stable dication, which did not readily undergo C?O bond dissociation. This is an unprecedented and unknown case in which the monocation has greater reactivity than the dication.  相似文献   

9.
This article presents a route to a novel polyester having sequentially ordered two orthogonal reactive groups. The polyester was given by the imidazole‐initiated alternating copolymerization of allyl glycidyl ether (AGE) and a bislactone 1 . This copolymerization system is characterized by the following three reaction behaviors: (1) the selective participation of only one of the two lactone moieties of 1 to the copolymerization to give a linear polyester, and the consequent introduction of the second lactone into the side chain of the polyester, (2) the participation of the epoxy moiety in AGE to the copolymerization, and the consequent introduction of the carbon–carbon double bond into the side chain of the polyester, and (3) arrangement of the sequentially ordered two orthogonal reactive groups according to the alternating manner. The introduction of the two reactive groups to the side chain of the alternating copolymer allowed two routes of sequential chemoselective reactions: (A) The ring‐opening reaction of the lactone moiety with n‐propylamine and the following Pt‐catalyzed hydrosilylation of the carbon–carbon double bond with dimethylphenylsilane and (B) the sequential reactions of the reverse order. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2009  相似文献   

10.
New copolymers of amino acid and urethane (PAU), in which a polyurethane segment was combined with poly(γ‐methyl‐L‐glutamate) (PMLG) of various contents, were synthesized by the copolymerization of the polyurethane prepolymer (UPP) having isocyanate groups at both terminals of the chain and γ‐methyl‐L ‐glutamate‐N‐carboxyanhydride (NCA) to improve the elastic recovery and adhesion of PMLG for application of PMLG to synthetic leather. The copolymerization of the UPP with NCA was carried out by applying the reactivity of isocyanate and the polymerization mechanism of NCA using the primary amine and tertiary amine as initiators. Infrared (IR) and 13C‐NMR spectra of these PAUs as well as the chemical analysis of the PAU intermediates showed that the PAUs would have a multiblock–triblock structure: namely, the PAUs consisted of the block copolymer segments of urethane and a small amount of PMLG at the center of the copolymer chain and most of the PMLG at both terminals of the copolymer chain. The elastic recovery and adhesion of these PAUs were significantly larger than those of the PMLG with the maintenance of a good sense of touch, which was a unique asset of PMLG. Furthermore, it was found that the PAUs had intermediary moisture permeability between that of PMLG and polyurethane. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 383–389, 1999  相似文献   

11.
The living anionic copolymerization of styrene with 1,2‐bis(4′‐ethenylphenyl)ethane (1) or p‐divinylbenzene (PDVB) with sec‐butyllithium in benzene was carried out. The copolymerizations of styrene with more than 20 mol % of 1 gave insoluble polymers in quantitative yields, whereas the yield showed the maximum (97%) for PDVB at 15 mol %. The content of unreacted double bonds of the network polymer formed by the copolymerization with PDVB was four times as large as that formed with 1. Gas chromatographic analyses of the copolymerization suggested close reactivities of the double bonds between styrene and 1, whereas a rapid consumption of PDVB compared with styrene was observed in their copolymerization. The r1, r2,and r1r2 values for the copolymerization of styrene with 1 were determined to be 1.00, 1.09, and 1.09, respectively, which suggests that a more homogeneous network structure can be attained with 1. The living chain end of the produced living gel initiated the polymerization of tert‐butyl methacrylate to give an insoluble block copolymer in a good yield. The hydrolysis of the ester group of the block copolymer led to an amphiphilic copolymer that exhibited a characteristic property of a hydrogel. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 2543–2547, 2000  相似文献   

12.
N-Aryltetramethylguanidines catalyze the formation of polyurethane foams by a novel exchange reaction involving C?N double bond of the guanidines and the isocyanate groups of the polyisocyanates. The catalyst sites are transferred to the growing polymer network, achieving a rapid catalytic build-up of the foam.  相似文献   

13.
The investigations presented deal with the experimental results of the copolymerization of maleic anhydride (MAn) with alkenes. The course of the reaction is explained by the overall rate of the copolymerization (v Br), which correlates with the solution viscosity of the copolymer, and the dependence of the v Br maximum on the mole ratio of the monomers at constant total monomer concentration. The use of solvents with increasing donor power leads to increased complexing of the free MAn molecules and of the MAn radical chain ends. The results demonstrate that, for low 1-alkenes, the addition of the MAn chain radical is the rate-determining step of the copolymerization. As the substituents on the olefinic double bond become larger or the double bond shifts to the 1,2-position, the addition of MAn to the hydrocarbon radical becomes more and more the rate-determining step. On the other hand, an increase of the CT complexation of the MAn polymer radical by use of donor solvents decreases the alkene addition rate.  相似文献   

14.
Monomer reactivity ratios and copolymerization parameters were determined for n-octadecyl acrylate and N-n-octadecylacrylamide with several monomers not previously reported. Values of Q and e for the long-chain acrylate now agreed more closely than before with the average of values for the lower acrylate homologs. The average polarity parameter for N-n-octadecylacrylamide still showed more electron withdrawal from the double bond (e = 0.42) than was expressed by the average literature value (e = ?0.10) for N-n-octylacrylamide. Because penultimate effects were absent in this work, the reason for the discrepancy seems to reside in the copolymer analysis. Reactivity ratios for oleyl and octadecyl acrylate with methyl methacrylate were similar. Coefficients for the relation between overall rates of copolymerization and composition were obtained on some systems by curve fitting with a computer. They generally showed a slight minimum.  相似文献   

15.
The γ-radiation-induced free-radical copolymerization of ethylene and CO has been investigated over a wide range of pressure, initial gas composition, radiation intensity, and temperature. At 20°C., concentrations of CO up to 1% retard the polymerization of ethylene. Above this concentration the rate reaches a maximum between 27.5 and 39.2% CO and then decreases. The copolymer composition increases only from 40 to 50% CO when the gas mixture is varied from 5 to 90% CO. A relatively constant reactivity ratio is obtained at 20°C., indicating that CO adds 23.6 times as fast as an ethylene monomer to an ethylene free-radical chain end. For a 50% CO gas mixture, the above value of 23.6 and the copolymerization rate decrease with increasing temperature to 200°C. The kinetic data indicate a temperature-dependent depropagation reaction. Infrared examination of copolymers indicates a polyketone structure containing ? CH2? CH2? and ? CO? units. The crystalline melting point increases rapidly from 111 to 242°C., as the CO concentration in the copolymer increases from 27 to 50%. Molecular weight of copolymer formed at 20°C. increased with increasing CO, indicating M?n values >20,000. Increasing reaction temperature results in decreasing molecular weight. Onset of decomposition for a 50% CO copolymer was measured at ≈250°C.  相似文献   

16.
The goal of this research is to synthesize biodegradable polymers that would have nitroxyl radical biological functions. Linear aliphatic polyesters were chosen as the starting materials. The hydroxyl‐terminated polylactide/?‐caprolactones (PBLC‐OHs) were first synthesized by melt ring‐opening copolymerization in the presence of benzyl alcohol and stannous octoate. PBLC‐OHs were used as the precursor for the synthesis of double bond‐functionalized polylactide/?‐caprolactones (PBLC‐Mas) by reacting the hydroxyl end groups of PBLC‐OH with maleic anhydride in melt at 130 °C. Acrylic acid/lactide/?‐caprolactone graft copolymers (PBLCAs) were then successfully carried out by the radical copolymerization of acrylic acid and PBLC‐Ma initiated by azobisisobutyronitrile. Finally, nitroxyl radicals [4‐amino‐2,2,6,6‐tetramethylpiperidine‐1‐oxy (TAM)] were incorporated into the carboxylic acid sites of the acrylic acid/lactide/?‐caprolactone copolymer (TAM‐PBLCA) by reacting TAM with PBLCA in the presence of N,N′‐carbonyl diimidazole. A high content of TAM was incorporated into the PBLCA copolymer. The polymers synthesized were characterized by 1H and 13C NMR, Fourier transform infrared spectroscopy, and electron paramagnetic resonance spectra. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 39: 4214–4226, 2001  相似文献   

17.
Synthesis and anionic polymerization of the isocyanate of L-leucine methyl ester (LeuMI) were carried out. LeuMI was prepared by the reaction of L-leucine methyl ester hydrochloride with phosgene. Anionic polymerization of LeuMI was carried out at –50°C in dimethyl-formamide (DMF) using sodium methoxide, potassium tert-butoxide, methyllithium, or sodium cyanide under a nitrogen atmosphere. The obtained polymers were insoluble in common organic solvents such as DMF, tetrahydrofuran, chloroform, and dimethyl sulfoxide, but were soluble in trifluoroacetic acid. Further, anionic copolymerization of LeuMI with n-butyl isocyanate (BI) was also carried out to observe that the smaller the content of LeuMI unit in the copolymer was, the larger the specific rotation of the polymer. From the circular dichroic spectral analyses of the polymers it was confirmed that the capability of helix formation of poly (LeuMI) was smaller than poly(BI). The relaxation of the helical structure of the polymer in trifluoroacetic acid solution was observed, and the smaller the content of LeuMI unit in the copolymer was, the faster the relaxation of the helical structure. © 1995 John Wiley & Sons, Inc.  相似文献   

18.
Copolymers of 3,4-dihydro-2H-pyran-2-carboxyaldehyde (acrolein dimer) with phenyl isocyanate were obtained under several conditions. Infrared and NMR analyses showed that the isocyanate always reacted with acrolein dimer forming urethane linkages, not block units of isocyanate. An alternating copolymer was obtained from the copolymerization in the presence of anionic catalysts such as butyllithium at room temperature, irrespective of the monomer ratios employed. The isocyanate content in the copolymer prepared with an Al(C2H5)2Cl catalyst was increased by elevating polymerization temperature. The copolymerizability of aldehydes with the isocyanate depends upon the polarity of aldehyde group.  相似文献   

19.
The radiation induced copolymerization of chlorotrifluoro ethylene (CTFE) with various butenes was studied at temperatures between ?20°C and +40°C using 60Co-γ rays. In the case of isobutene (IB) an almost alternating crystalline copolymer is formed in a heterogeneous reaction. At high IB-concentrations a cationic homopolymerization of this olefin occurs simultaneously to the radical copolymerization. The copolymerization rate increases with increasing temperature and degree of conversion. The highest rates are obtained for monomer mixtures with about 80 to 90 mole % CTFE. The decrease in rate for monomer mixtures with still higher CTFE concentrations is assumed to be partly due to the low IB-concentration and partly to degradative chain transfer by the isobutene. In support of this assumption molecular weights and melting points of the copolymer have been determined. Similar results were obtained for butene-1 but in this case, no cationic homopolymerization was observed and the reaction proceeded homogeneously. Cis- and trans-butene-2 only acted as polymerization inhibitors.  相似文献   

20.
Di(mesityl)cyclohexenylphosphine undergoes hydroboration with Piers' borane [HB(C6F5)2] to yield the cyclohexylene‐anellated frustrated Lewis pair 5 . This P/B pair splits H2 with the formation of the product 4 and adds to the C?O double bond of phenyl isocyanate to yield 6 . In the crystal, compound 5 features a puckered four‐membered heterocyclic core structure with a long P? B bond (av. 2.197(5) Å). The activation energy of the P? B cleavage of the frustrated Lewis pair 5 was determined by dynamic 19F NMR spectroscopy at ΔG(298 K)=12.1±0.3 kcal mol?1.  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