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.
Using 1H-NMR spectra of radical copolymers of methyl acrylate (M1) and 4-vinyl pyridine (M2), the terminal model of copolymerization has been tested. Reactivity ratios (r1 = 0.18 r2 = 1.77) were found. The probability of “coisotactic” alternating addition (σ = 0.98) has been obtained as the best fit of experimental methoxy signal fractions and the theoretical curves.  相似文献   

2.
Radical-initiated copolymerization of N-isopropylacrylamide (NIPA) with maleic (MA) and citraconic (CA) anhydrides was carried out in the presence of 2,2-azobisisobutyronitrile (AIBN) as an initiator in 1,4-dioxane at 65 °C under nitrogen atmosphere. Structure and monomer unit compositon of the copolymers obtained from a wide range of monomer feed were determined by elemental analysis (content of N for NIPA units), Fourier transform infrared and 1H NMR spectroscopy. Monomer reactivity ratios for NIPA (M1)-MA (M2) and NIPA (M1)-CA (M2) pairs were determined by Kelen-Tüdõs (KT) and non-linear regression (NLR) methods using elemental and 1H NMR spectroscopy analyses data. They are r1=0.45 and r2=0.08 (KT, N analysis), r1=0.44 and r2=0.10 (KT, 1H NMR), r1=0.45 and r2=0.078 (NLR) for NIPA-MA monomer pair and r1=0.52 and r2=0.02, r1=0.44 and r2=0.04, r1=0.51 and r2=0.014 for NIPA-CA monomer pair, respectively. Observed tendency towards alternating copolymerization at ?50 mol% NIPA concentration in monomer feed and relatively high activity of NIPA growing radical was explained by H-bond formation between CO (anhydride) and NH (amide) fragments during chain growth reactions. Intrinsic viscosity, molecular weight and thermal behaviour of the synthesized copolymers were found to depend on the type of comonomer and the amount of NIPA units in the copolymers. These functional amphiphilic copolymers containing anion- and cation-active groups show both temperature and pH sensitivity and can be used for biological purposes as physiologically active macromolecular systems.  相似文献   

3.
Copolymers of ethyl acrylate (M1) and n-butyl methacrylate (M2) were prepared by benzoyl peroxide initiation in solution at 60° and copolymer compositions estimated by [13C]-NMR spectroscopy. The kinetic behaviour is approximately ideal with r1 = 1/r2 = 0.47. Relaxation times T1 were determined for six of the carbons in M2 units and one in the M1 units; they range from 0.07 to 1.65 sec.  相似文献   

4.
2-Trimethylsilyloxy-1,3-butadiene (TMSBD), the silyl enol ether of methyl vinyl ketone, was homopolymerized with a radical initiator to afford polymers with a molecular weight of ca. 104. Radical copolymerizations of TMSBD with styrene (ST) and acrylonitrile (AN) in bulk or dioxane at 60°C gave the following monomer reactivity ratios: r1 = 0.64 and r2 = 1.20 for the ST (M1)–TMSBD (M2) system and r1 = 0.036 and r2 = 0.065 for the AN (M1)–TMSBD (M2) system. The Q and e values of TMSBD determined from the reactivity ratios for the former copolymerization system were 2.34 and ?1.31, respectively. The resulting polymer and copolymers were readily desilylated with hydrochloric acid or tetrabutylammonium fluoride as catalyst to yield analogous polymers having carbonyl groups in the polymer chains.  相似文献   

5.
The copolymers prepared in this study by free radical copolymerization of N-vinylpyrrolidone (M 2) with 4-vinylbenzylchloride (M 1) using 2,2′-azobisisobutyronotrile (AIBN) initiator in 1,4-dioxane solvent at 70°C were characterized by FTIR, 1H-NMR and 13C-NMR techniques. Polymer solubility was tested in both polar and nonpolar solvents. The thermal properties were studied by thermogravimetric analysis (TGA) and differential scanning calorimeter (DSC). Copolymer compositions were established by H1-NMR spectra, while reactivity ratios of the monomers were computed using the linearization methods viz., Fineman-Ross (FR) (r 1 = 1.67 and r 2 = 0.67), Kelen-Tudos (KT) (r 1 = 1.77 and r 2 = 0.65) and extended Kelen-Tudos (EK-T) (r 1 = 1.72 and r 2 = 0.63) methods at lower conversion. Furthermore, reactivity ratios in nonlinear error-in-variables method (RREVM) also compute the reactivity ratios (r 1 = 1.76 and r 2 = 0.66); these are found to be in good agreement with each other. The distribution of monomer sequence along the copolymer chain was calculated using a statistical method based on the calculated reactivity ratios.  相似文献   

6.
The radical copolymerization of N,N-diallyl-N,N-dimethylammonium chloride (AMAC) (M1) with ethylene glycol vinyl ether (M2) in an aqueous medium proceeds at a high rate to afford random copolymers. The reactivity ratios equal to r 1 = 2.18 and r 2 = 0.01 indicate that AMAC is a more active comonomer. The overall reaction order in comonomers is 2.4, and the effective activation energy is 97.4 ± 2 kJ/mol. The monomer M1 enters into copolymerization by both of the double bonds with the formation of pyrrolidinium structures in the chain through the cyclization stage.  相似文献   

7.
Copolymerizations of tri-n-butyltin acrylate (M1) with (a) methyl methacrylate (M2), (b) propyl methacrylate (M3), (c) butyl methacrylate (M4) and (d) acrylonitrile (M5) in solution at 70 using AIBN as initiator led to monomer reactivity ratios as follows: (a) r1 = 0.401 and r2 = 2.199, (b) 0.323 and 1.713, (c) 0.196 and 1.65, and (d) 0.243 and 1.008. The variation of the average copolymer composition with conversion for two copolymers from M1 with M2 and M4 were calculated and verified experimentally. Four terpolymer compositions involving M1 and M5 with M2 or M4 were prepared and the terpolymer compositions were calculated on the basis of tin and nitrogen analyses. The variations of instantaneous and average terpolymer composition with conversion fit satisfactorily the experimental results over a wide range of conversion.  相似文献   

8.
The copolymerization of 2-hydroxyethyl acrylate (HEA, M_1) and methylmethacrylate (MMA, M_2) in cyclohexanone was studied. The multiple experiments ofsolution copolymerization with low conversion were carried out at two sensitive compositionfeed points at 60, 80, 100, 120 and 140℃, respectively. The composition of the copolymerswas analyzed by ~1H-NMR. The reactivity ratios which were estimated by the Error-in-Variable Method (EVM) of Mayo-Lewis equation were found to be r_1 = 0.328, r_2 = 1.781for 60℃; 0.375, 1.709 for 80℃; 0.406, 1.654 for 100℃; 0.439, 1.540 for 120℃ and 0.455,1.400 for 140℃, and the 95% joint confidence intervals of the reactivity ratios were alsodetermined. According to r_1 and r_2, Arrhenius relations and the activity energy differencebetween the homo- and cross-propagation were calculated.  相似文献   

9.
The homopolymerization and copolymerization of butadiene-1-carboxylic acid (Bu-1-Acid) (M1) were studied in tetrahydrofuran at 50°C with azobisisobutyronitrile as an initiator. The initial rate of polymerization was proportional to [AIBN]1/2 and [Bu-1-Acid]1. The overall activation energy for the polymerization was 22.87 kcal/mole. For copolymerization with styrene (M2) and acrylonitrile (M2), the monomer reactivity ratios r1, r2 were determined by the Fineman-Ross method, as follows; r1 = 5.55, r2 = 0.08 (M2 = styrene); r1 = 11.0, r2 = 0.03 (M2 = acrylonitrile). Alfrey-Price Q-e values calculated from these values were 6.0 and +0.11, respectively. The Bu-1-Acid unit in the copolymer as well as the homopolymer was found from infrared and NMR spectral analyses to be composed of a trans-1,4 bond. The hydrogen-transfer polymerization of Bu-1-Acid leading to polyester was attempted with triphenylphosphine as initiator, but did not occur.  相似文献   

10.
Styrene (M1) has been copolymerized with o-, m- and p-halostyrenes (M2) at temperatures between 40 and 110°C, using azoisobutyronitrile as initiator; the halostyrenes were labelled with 14C in the β-position. The compositions of the copolymers were determined by liquid scintillation counting. Since [M1] ? [M2], a simplified form of the copolymer composition equation was used to determine reactivity ratios r1: account was taken of the isotope effect resulting from the labelling in the β-position. Arrhenius parameters of r1 were found; they show that polar effects predominate in determining the magnitude of r1. Steric effects, which counteract the polar effects, are small.  相似文献   

11.
A new approach to obtaining thermoset organotin polymers, which permits control of crosslinking site distribution and, through it, a better control of properties of organotin antifouling polymers, is reported. Tri-n-butyltin acrylate and tri-n-butyltin methacrylate monomers were prepared and copolymerized, by the solution polymerization method with the use of free-radical initiators, with several vinyl monomers containing either an epoxy or a hydroxyl functional group. The reactivity ratios were determined for six pairs of monomers by using the analytical YBR method to solve the differential form of the copolymer equation. For copolymerization of tri-n-butyltin acrylate (M1) with glycidyl acrylate (M2), these reactivity ratios were n = 0.295 ± 0.053, r2 = 1.409 ± 0.103; with glycidyl methacrylate (M2) they were r1 = 0.344 ± 0.201, r2 = 4.290 ± 0.273; and with N-methylolacrylamide (M2) they were r1 = 0.977 ± 0.087, r2 = 1.258 ± 0.038. Similarly, for the copolymerization of tri-n-butyltin methacrylate (Mi) with glycidyl aery late (M2) these reactivity ratios were r1 = 1.356 ± 0.157, r2 = 0.367 ± 0.086; with glycidyl methacrylate (M2) they were r1 = 0.754 ± 0.128, r2 = 0.794 ± 0.135; and with N-methylolacrylamide (M2) they were r1 ?4.230 ± 0.658, r2 = 0.381 ± 0.074. Even though the magnitude of error in determination of reactivity ratios was small, it was not found possible to assign consistent Q,e values to either of the organotin monomers for all of its copolymerizations. Therefore, Q,e values were obtained by averaging all Q,e values found for the particular monomer, and these were Q = 0.852, e = 0.197 for the tri-n-butyltin methacrylate monomer; and Q = 0.235, e = 0.401 for the tri-n-butyltin acrylate monomer. Since the reactivity ratios indicate the distribution of the units of a particular monomer in the polymer chain, the measured values are discussed in relation to the selection of a suitable copolymer which, when cross-linked with appropriate crosslinking agents through functional groups, would give thermoset organotin coatings with an optimal balance of mechanical and antifouling properties.  相似文献   

12.
The copolymerization of 4-cyclopentene-1,3-dione (M2) with p-chlorostyrene and vinylidene chloride is reported. The copolymers were prepared in sealed tubes under nitrogen with azobisisobutyronitrile initiator. Infrared absorption bands at 1580 cm.?1 revealed the presence of a highly enolic β-diketone and indicated that copolymerization had occurred. The copolymer compositions were determined from the chlorine analyses and the reactivity ratios were evaluated. The copolymerization with p-chlorostyrene (M1) was highly alternating and provided the reactivity ratios r1 = 0.32 ± 0.06, r2 = 0.02 ± 0.01. Copolymerization with vinylidene chloride (M1) afforded the reactivity ratios r1 = 2.4 ± 0.6, r2 = 0.15 ± 0.05. The Q and e values for the dione (Q = 0.13, e = 1.37), as evaluated from the results of the vinylidene chloride case, agree closely with the previously reported results of copolymerization with methyl methacrylate and acrylonitrile and confirm the general low reactivity of 4-cyclopentene-1,3-dione in nonalternating systems.  相似文献   

13.
Copolymers of 2-sulfoethyl methacrylate, (SEM) were prepared with ethyl methacrylate, ethyl acrylate, vinylidene chloride, and styrene in 1,2-dimethoxyethane solution with N,N′-azobisisobutyronitrile as initiator. The monomer reactivity ratios with SEM (M1) were: vinylidene chloride, r1 = 3.6 ± 0.5, r2 = 0.22 ± 0.03; ethyl acrylate, r1 = 3.2 ± 0.6, r2 = 0.30 ± 0.05; ethyl methacrylate, r1 = 2.0 ± 0.4, r2 = 1.0 ± 0.1; styrene, r1 = 0.6 ± 0.2, r2 = 0.37 ± 0.03. The values of the copolymerization parameters calculated from the monomer reactivity ratios were e = +0.6 and Q = 1.4. Comparison of the monomer reactivities indicates that SEM is similar to ethyl methacrylate with regard to copolymerization reactivity in 1,2-dimethoxyethane solution. The sodium salt of 2-sulfoethyl methacrylate, SEM?Na, was copolymerized with 2-hydroxyethyl methacrylate (M2) in water solution. Reactivity ratios of r1 = 0.7 ± 0.1 and r2 = 1.6 ± 0.1 were obtained, indicating a lower reactivity of SEM?Na in water as compared to SEM in 1,2-dimethoxyethane. This decreased reactivity was attributed to greater ionic repulsion between reacting species in the aqueous medium.  相似文献   

14.
Copolymers of the cyclic ketene acetals, 2-methylene-5,5-dimethyl-1,3-dioxane, 3 , (M1) with 2-methylene-1,3-dioxolane, 4 , (M2) or 2-methylene-1,3-dioxane, 5 , (M2), were synthesized by cationic copolymerization. An experimental method was designed to study the reactivity of these very reactive and extremely acid sensitive cyclic ketene acetal monomers. The reactivity ratios, calculated using a computer program based on a nonlinear minimization algorithm, were r1 = 6.36 and r2 = 1.25 for the copolymerization of 3 with 4 , and r1 = 1.56 and r2 = 1.42 for the copolymerization of 3 with 5. FTIR and 1H-NMR spectra when combined with the values of r1 and r2 showed that these copolymers were formed by a cationic 1,2-polymerization (ring-retained) route. Furthermore the tendency existed to form very short blocks of M1 or M2 within the copolymers. Cationic copolymerization of cyclic ketene acetals have the potential to be used for synthesis of novel polymers. © 1996 John Wiley & Sons, Inc.  相似文献   

15.
Abstract

Radical homopolymerization of N-[4-N′-(α-methylbenzyl)-aminocarbonylphenyl]maleimide ((S)-MBCP) was carried out at 50 and 70°C for 24 h to give optically active polymers ([α]25 D = 159.8 to 163.4°). Radical copolymerizations of (S)-MBCP (M1) were performed with styrene (ST, M2, methyl methacrylate (MMA, M2) in THF at 50°C. The monomer reactivity ratios (r 1, r 2) and the Alfrey-Price Q, e values were determined as follows: r 1 = 0.32, r 2= 0.14, Q 1 = 1.74, e 1 = 0.96 in the (S)-MBCP-ST system; r 1 = 0.54, r 2 = 0.93, Q 1 = 1.11, e 1 = 1.23 in the (S)-MBCP-MMA system. Chiroptical properties of the polymers and the copolymers were also investigated, and asymmetric induction into the copolymer main chain is discussed.  相似文献   

16.
4-Phenyl-2-butene (4Ph2B) undergoes monomer-isomerization copolymerization with 4-methyl-2-pentene (4M2P) and 2-and 3-heptene (2H and 3H) with TiCl3–(C2H5)3Al catalyst at 80°C to produce copolymer consisting exclusively of 1-olefin units. For comparison the copolymerization of 4-phenyl-1-butene (4Ph1B) with 4-methyl-1-pentene (4M1P) and 1-heptene (1H) was carried out under similar conditions. The composition of the copolymers obtained from these copolymerizations was determined from the ratios of optical densities D1380 and D1600 of infrared (IR) spectra of their thin films. The apparent monomer reactivity ratios for the monomer-isomerization copolymerization of 4Ph2B with 4M2P, 2H, and 3H in which the concentration of olefin monomer in the feed was used as internal olefin and those for the copolymerization of 4Ph1B with 4M1P and 1H were determined as follows: 4Ph2B(M1)-4M2P(M2); r1 = 0.90, r2 = 0.20, 4Ph1B(M1)-4M1P (M2); r1 = 0.40, r2 = 0.70, 4Ph2B(M1)-2H(M2); r1, = 0.45, r2 = 1.85, 4Ph2B(M1)-3H(M2); r1 = 0.50, r2 = 1.20, 4Ph1B(M1)-1H(M2); r1 = 0.55, r2 = 0.75. The difference in monomer reactivity ratios seemed to originate from the rate of isomerization from 2- or 3-olefins to 1-oletins in these monomer-isomerization copolymerizations.  相似文献   

17.
A newer approach to obtaining acrylic thermoset polymers with adequate hydrophilicity required for various specific end uses is reported. Glycidyl methacrylate (GMA) was copolymerized with n-butyl acrylate (n-BA), isobutyl acrylate (i-BA), and 2-ethylhexyl acrylate (2-EHA) in bulk at 60°C. with benzoyl peroxide as free radical initiator. The copolymer composition was determined from the estimation of epoxy group. Reactivity ratios were calculated by the Yezrielev, Brokhina, and Roskin method. For copolymerization of GMA (M1) with n-BA (M2) the reactivity ratios were r1 = 2.15 ± 0.14, r2 = 0.12 ± 0.03; with i-BA (M2) they were r1 = 1.27 ± 0.06, r2 = 0.33 ± 0.031; and with 2-EHA (M2) they were r1 = 2.32 ± 0.14, r2 = 0.13 ± 0.009. The reactivity ratios were the measure of distribution of monomer units in a copolymer chain; the values obtained are compared and discussed.  相似文献   

18.
2-Hydroxypropyl methacrylate (2-HPMA) has been copolym-erized with ethyl methacrylate (EMA), n-butyl methacrylate (BMA), and 2-ethylhexyl methacrylate (EHMA) in bulk at 60°C using benzoyl peroxide as initiator. The copolymer composition has been determined from the hydroxyl content. The reactivity ratios have been calculated by the Yezrielev, Brokhina, and Raskin method. For copolymerization of 2-HPMA (M1) with EMA (M2), the reactivity ratios are r1 = 1.807 ± 0.032 and r2 = 0.245 ± 0.021; with BMA (M2) they are n = 2.378 ± 0.001 and r2 = 0.19 ± 0.01; and with EHMA the values are r1 = 4.370 ± 0.048 and r2 = 0.103 ± 0.006. Since reactivity ratios are the measure of distribution of monomer units in copolymer chain, the values obtained are compared and discussed. This enables us to choose a suitable copolymer for synthesizing thermoset acrylic polymers, which are obtained from cross-linking of hydroxy functional groups of HPMA units, for specific end-uses.  相似文献   

19.
《European Polymer Journal》1986,22(3):195-197
Styrene (M1) and 14C-labelled α-methylstyrene (M2) have been copolymerised in the temperature range 40 to 105°C, under conditions where [M1] ⪢ [M2]. Under these circumstances, the effect of the low ceiling temperature of M2 was circumvented in deriving a value of the reactivity ratio r1. The Arrhenius plot of r1 was linear. Values of r1 determined under more conventional conditions, where [M1] ≈ [M2], have been found to deviate further from this Arrhenius line as the temperature is increased; this illustrates the effect of ceiling temperature under these latter conditions and the unreliability of such values of r1 when comparing reactivities.  相似文献   

20.
Glycidylmethacrylate/methacrylonitrile (G/M) copolymers of different compositions were prepared and a copolymer composition was obtained from quantitative 13C NMR spectroscopy. Reactivity ratios for comonomers were calculated using the Kelen–Tudos (KT) and non linear error in variable (EVM) methods. The reactivity ratios obtained from KT and EVM are rG=1.14±0.1, rM=0.76±0.06 and rG=1.12, rM=0.75, respectively. Complete spectral assignment of 13C- and 1H-NMR spectra were done with the help of Distortionless Enhancement by Polarization Transfer (DEPT) and 2D 13C–1H heteronuclear single quantum coherence (HSQC).  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