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 polymerization of methacrylic acid initiated by permanganate / oxalic acid redox pair has been studied in aqueous media at 30 ± 0.2°C in nitrogen atmosphere. The initial rate of polymerization has been found to be proportional to nearly the first power of the catalyst (KMnO4) concentration ((0.8-6.0)·10?4 mole/l.) at low monomer concentration (5.89·10?2 mole/l.) but the catalyst exponent varies from 1.2 to 0.2 at comparatively high monomer concentration (17.68·10?2 mole/l.) and falls at comparatively higher concentrations of the activator ((COOH)2) at constant concentration of catalyst (1.6·10?4 mole/l.) and monomer (5.89·10?2 mole/l.). The rate is proportional to the first power of the monomer concentration within the range (2.36–14.1)·10?2 mole/l. and catalyst (2.8·10?4 mole/l.). The initial rate increases with increase in polymerization temperature up to 45°C. The overall activation energy has been found to be 9.87 kcal/mole within the temperature range 30-45°C. Organic solvents (water miscible only) and salts (KCl, Na2SO4 and Na2C2O4) depress the rate considerably but the manganous salt (MnSO4) is found to increase the initial rate. A complexing agent, NaF, decreases the rate as well as the maximum conversion. Introduction of new catalyst at intermediate stages of polymerization increases the rate and the maximum conversion.  相似文献   

2.
The reaction mechanism of the redox initiation system persulfate/N,N,N′,N′-tetramethyl-ethylenediamine (TMEDA) was studied by spin trapping technique and ESR spectra. The free radical (CH3)2NCH2CH2(CH3)NCH (1) is one of the initial free radicals responsible for the initiation of the vinyl polymerization in addition to the sulfate free radical (HO3SO˙). The result of amino end group analysis confirmed the fact that the free radical 1 initiates vinyl polymerization and becomes the end group of the resulting polymers. The effect of TMEDA concentration on the final percentage of conversion of acrylamide polymerization and on the molecular weights of the resulting polymer were also studied. An initiation mechanism was proposed.  相似文献   

3.
The synthesis and the “living” cationic polymerization of a new saccharidic monomer namely the 1,2 : 3,4‐di‐O‐isopropylidene‐6‐O‐(2‐vinyloxy ethyl)‐D ‐galactopyranose have been investigated. The monomer structure has been characterized by 1H and 13C NMR, elemental analysis and mass spectrometry. It has been polymerized with acetaldehyde diethyl acetal/trimethylsilyl iodide as initiating system in presence of ZnCl2 as co‐initiator. Macromolecules of controlled chain length (from 2 500 g·mol–1 to 7 500 g·mol–1) were obtained, bearing protected saccharidic moieties as side‐groups, and aldehyde as ω‐end‐groups which could be turned into amine ω‐end‐groups. Finally, the saccharidic residues were deprotected in order to perform the binding of DNA probes (oligonucleotides) onto the hydrophilic polymer chains through the anomeric site of the galactose moieties.  相似文献   

4.
Polymerization of acrylic acid in aqueous solution initiated by permanganate-oxalic acid redox pair has been studied at 32 ± 0.2°C in nitrogen atmosphere. The rate of polymerization has been found to be nearly independent of oxalic acid concentration within the range 1.87 to 9.33 · 10?3 mole/l. and decreases only at higher concentrations of the oxalic acid. The rate has also been found to vary with the first power of the monomer concentration (within the range 1.44 to 5.76 · 10?2 mole/l.) and the first power of the catalyst concentration (8.0 to 28.0 · 10?5 Mole/l.). It is, however, proportional to half power at relatively high catalyst concentration, at fixed concentrations of oxalic acid (1.03 · 10?2 mole/l.) and the monomer (5.76 · 10?2 mole/l.). At higher concentration of monomer the catalyst exponent has been found to be nearly unity for both the higher and lower concentrations of the catalyst. The initial rate of polymerization increases with increase in temperature. The overall energy of activation has been found to be 19.56 kcal/mole within the temperature range 30 – 45°C. Organic solvents and salts (KCI, Na2SO4, and Na2C2O4) depress the rate but MnSO4 4H2O has been found to increase the initial rate but depress the maximum conversion.  相似文献   

5.
The polymerization of α-methylstyrene with CF3SO3H and H2SO4 as initiators was studied at 30°C in dichloroethane by the stopped-flow/rapid-scan spectroscopic technique. The propagating cation shows λmax at 336–340 nm. The initiation process and the early stage of propagation were analyzed kinetically on the basis of the cation formation and monomer consumption by taking into account the equilibrium monomer concentration. The rate of initiation with CF3SO3H was found to be proportional to the concentrations of CF3SO3H and monomer, but the initiation with H2SO4 is not proportional to the H2SO4 concentration. The rate constant of initiation was estimated to be (2 ± 1)·103 1.mol?1.s?1 with CF3SO3H, and similar values were found with H2SO4. The rate constant of propagation is 3.104 1.mol?1.s?1 with CF3SO3H as initiator and 106 ? 107 1.mol?1.s?1 with H2SO4 as initiator. These kp values are close to that obtained previously in the radiation polymerization. Finally, λmax values of the propagating cation, obtained by the stopped-flow technique, were collected.  相似文献   

6.
The polymerization of methacrylamide initiated by the redox system K2S2O8/ascorbic acid has been studied at 35 ± 0,2°C under the influence of atmospheric oxygen. The molecular oxygen autocatalyses the polymerization rate. The rate of polymerization is independent of the activator (ascorbic acid) concentration within the range 2,83 · 10?3 to 11,3 · 10?3 mol · l?1, this does not hold below or above the given concentration range. The rate varies linearly at low monomer concentration (up to 17,76 · 10?2 mol · l?1). The catalyst exponent decreases from nearly unity (0,94) to 0,57 with increasing concentration of the catalyst probably due to participation of primary radicals in the termination of growing chain. The initial rate is not changed by the addition of a strong acid (H2SO4) within the range 15 · 10?5 to 30 · 10?5 mol · l?1. With the activator (ascorbic acid) alone, an optimum is observed within the pH range 2,9–3,5. The initial rate and the limiting conversion increases with increasing polymerization temperature. The overall energy of activation as calculated from the ARRHENIUS plot has been found to be 16,0 kcal.mol?1 within the temperature range 30–45°C. Organic solvents (water miscible only) and small amounts of neutral salts, (KCl, Na2SO4) depress the initial rate and the maximum conversion. The addition of small amounts of catalysts like Cu2⊕, Mn2⊕, Agoplus; increases the initial rate but no appreciable increase in the limiting conversion is observed.  相似文献   

7.
Polymerization of 2-hydroxyethyl methacrylate (HEMA) initiated by the redox system manganese triacetate (Mn)(III)/cyanoacetic acid (CAA) was investigated kinetically in DMF and in HOAc/water (80 : 20 v/v). Monomer (M) conversion was followed gravimetrically. Polymerization was carried out at 31°C and at 41°C. The rate of polymerization Rp in DMF is expressed by the equation, RP, ∝ [M]1,4 [Mn(III)]0,4 [CAA]0,4. In HOAc/water, the polymerization exhibited a strong dependence on the concentration of Mn(III). That is, at low concentrations of Mn(III), the rate variations are [M]1,0, [CAA]0,5 and [Mn(III)]0,5, while at high Mn(III) concentration the polymerization rates are independent of [Mn(III)] and vary as a function of [M]1,5 and [CAA]0,5. Enhanced oxidation of the primary radicals by Mn(III) at higher concentrations has been elicited to explain the observed orders. The degree of polymerization was found to increase with increase in concentrations of monomer and decrease in concentrations of CAA and Mn(III). The polymerization rates as well as the molecular weights of the polymers are higher in HOAc/water than in DMF. The enhanced polymerization rates have been attributed to the increased propagation and the reduced termination rates.  相似文献   

8.
The bimodal molecular weight distribution (MWD) of the polymers and the polymerization rate in the cationic polymerization of styrene by CF3SO3H were studied under a variety of conditions. Both the decrease of the dielectric constant of the reaction mixture and the addition of a common-ion salt (Bu4N+SO3CF) reduced the polymerization rate and the formation of the higher molecular weight portion of the polymers (the high polymer). It appears that of the two propagating species the one which forms the high polymer is more ionically dissociated and more reactive in propagation. Salt effects indicate that in nitrobenzene the propagating species is a solvent-separated ion pair and/or a free ion. The effects of monomer concentration on the bimodal MWD have shown that there are different chain-breaking reactions for the two propagating species. The possibility that the monomer is complexed with one of the growing species is also discussed.  相似文献   

9.
Equilibrium constants were determined for the formation of cyclic oligomers of the general formula $ {\hbox{-\hskip-5pt-\hskip-5pt-}[({\rm CH}_3})_2 {\rm Si \hbox{---} Si(CH_3})_2\hbox{---} {\rm O\hbox{-\hskip-6.5pt-}]}_n , n = 2 - 6$, in the polymerization of the n = 2 homologue, octamethyl-1,4-dioxa-2,3,5,6-tetrasilacyclohexane ( 1 ). The formation of a linear polymer is thermodynamically favoured in bulk and in concentrated solutions. The equilibrium concentration of cyclics is little dependent on temperature. The six-membered ring, n = 2, used here as a monomer, is the most abundant cyclic in the equilibrium mixture. Kinetics of the conversion of 1 during the polymerization initiated with CF3SO3H, CH3SO3H and CF3COOH was studied in methylene dichloride solution. In its first stage, the reaction follows first order, but it slows down at higher monomer conversion. The order of the reaction with respect to the acid is two. Some kinetic features of the polymerization initated with CF3SO3H were similar to those exhibited by the polymerization of octamethylcyclotetrasiloxane (D4). In particular small amounts of additional water have rather small effect on the rate, but change to a considerable extent the entropy and enthalpy of activation. The concentration of cyclic oligohomologues passes through a maximum during monomer conversion, which indicates the kinetic enhancement of the cyclisation process. This observation is somewhat similar to those in the polymerization of hexamethylcyclotrisiloxane (D3). However, in contrast to the polymerization of D3 and similarly to the polymerization of D4, the excessive formation of oligomers being multiples of the monomer is not observed.  相似文献   

10.
The polymerization of methacrylamide initiated by potassium permanganate/oxalic acid redox system has been studied at 35 ± 0.2°C in a nitrogen atmosphere. The rate of polymerization is independent of the activator (oxalic acid) concentration within the range 5·10?3 to 10·10?3 mole·1?1, except at very high (above 10·10?3 mole·1?1) or at very low (below 5·10?3 mole·1?1) concentrations of the activator. The rate varies linearly at low monomer concentration (up to 5.87·10?2 mole·1?1). The catalyst exponent decreases from nearly unity (0.91) to 0.66 with increase in concentration of the catalyst (KMnO4) probably due to participation of primary radicals in the termination of the growing chain. The addition of strong acid (H2SO4) within the range 5 · 10?4–15 · 10?4 mole · l?1, shows a constant pH of 2.7 resulting in no change in initial rate. With activator alone (H2C2O4 · 2H2O), within the pH range 2.7 to 2.9 an optimum is observed. The initial rate increases with increase in polymerization temperature. The overall energy of activation as calculated from the ARRHENIUS plot has been found to be 15.1 kcal · mole?1 within the temperature range 30–50°C. Organic solvents (water miscible only) depress the initial rate and the maximum conversion. Small amounts of neutral salts (KCl and Na2SO4), however, show no appreciable effect on the polymerization rate, but small amounts of manganous salts (MnSO4) can increase the initial rate to a considerable extent. High concentrations of MnSO4 result in termination of the growing chain. A complexing agent, Na2F2, decreases the initial rate but increases the maximum conversion. Introduction of fresh catalyst at intermediate stages of polymerization increases both the rate of the reaction and the conversion.  相似文献   

11.
The mechanism of the polymerization of acrylamide in aqueous medium initiated by the glycerol (R)/Ce(IV) redox system was studied. The rate of monomer disappearance was found to be directly proportional to [M]3/2, and R1/2 at lower concentrations of Ce(IV). The rate of the disappearance of ceric ions was found to be inversely proportional to [M] and directly proportional to [Ce(IV)] and [R]. Consistent with the findings of earlier investigations, a complex formation between monomer, acrylamide, and ceric ion is indicated. The experimental results show the termination to be mutual. On the basis of these and other kinetic results, an appropriate mechanism is proposed. At higher concentration of Ce(IV), a different mechanism seems to operate. The rates of disappearance of both the monomer and ceric ions are retarded on addition of anions like HSO4?, SO4? ?, or CIO4?, but they are accelerated on the addition of Mn(II) ions. Interpretations of the above observations are furnished.  相似文献   

12.
This article reports the synthesis of novel amphiphilic triblock copolymers with a semi‐branched PLURONIC®7R structure by atom transfer radical polymerization (ATRP) in aqueous media. Poly(ethylene oxide)s (PEOs) with molecular weights 10 000 and 16 000 were end‐functionalized and used as bifunctional macroinitiators for the polymerization of oligo(propylene oxide) monomethacrylate by ATRP in a 1/3 v/v water/methanol mixture and in a 1/1 v/v water/1‐propanol mixture. Deviations from first‐order kinetics with respect to the monomer concentration were observed indicating that termination reactions were taking place. However, linear plots were obtained, when ln[M]0/[M] was plotted against time2/3 as suggested by Fischer. The effect on the control of the polymerization by adding Cu(II)Br2 to the polymerization medium has been investigated. When 10 mol‐% of Cu(II)Br2 was substituted for Cu(I)Br, normal first‐order kinetics were observed. A large reduction in the rate of polymerization was observed for the polymerization initiated by bifunctional PEO10 000 initiator, but almost no reduction in the rate of polymerization was observed, when the bifunctional PEO16 000 initiator was used. When the polymerizations were conducted in 1/1 v/v water/1‐propanol, unexpectedly high rates of polymerization were observed.

Synthesis of amphiphilic block copolymers with a semi‐branched PLURONIC®R architecture using ATRP.  相似文献   


13.
Sodium dodecyl 2-hydroxy-3-methacryloyloxypropyl phosphate as an amphiphilic vinyl monomer was found to polymerize spontaneously without initiator in water. The polymerization takes place at a monomer concentration higher than the critical micelle concentration via a free-radical mechanism, and complete conversion is achieved even at a very low initial monomer concentration (3 mmol/L) and a rather low temperature (35°C). In benzene, capable of forming inverse micelles, the polymerization also occurs spontaneously. In methanol, which gives an isotropic solution, the spontaneous polymerization proceeds only at high concentration (near 300 mmol/L). The 1H spin-spin relaxation time (T*2) of the monomer changes discontinuously in a range of monomer concentration from 250 to 300 mmol/L in CD3OD, suggesting that some kind of monomer mutual interaction takes place in methanol too. Thus, it can be concluded that the micellar aggregation monomer molecules triggers the polymerization.  相似文献   

14.
The kinetics of radical polymerization of 4-vinylphenyloxirane ( 1 ), 4-vinylbenzyloxirane ( 2 ), and 2-(4-vinylphenyl)oxetane ( 3 ) initiated by 2,2′-azoisobutyronitrile (AIBN) was studied. In the case of 1 the initial rate of polymerization was found to depend on the initiator and monomer concentrations as (rp)total ∝? [AIBN]0,5 · [M1]1,36. The higher order of the polymerization rate with respect to 1 was interpreted as due to a concurrent thermally initiated polymerization; the rate of the latter was found to depend on the monomer concentration squared. The value of the ratio propagation rate constant over square root of termination rate constant kp/kt1/2 = 2,8 · 10?2 dm3/2 · mol?1/2 · s?1/2 was determined from the measured dependences (rp)total = f[AIBN] and (rp)total = f([M1]), corrected for the rate of thermally initiated polymerization of 1 . On the other hand, the kinetics of radical polymerization of 2 and 3 did not deviate from the standard scheme valid for radical polymerization; in both cases the observed reaction order with respect to initiator and monomer was 0,5 and 1, respectively. Radical copolymerization of 4-vinylphenyloxirane (M1) with styrene (M2) was characterized by monomer reactivity ratios r1 = 1,06 and r2 = 0,78, respectively, corresponding to the Q, e-scheme values Q = 0,9 and e = ?0,36 for monomer 1 .  相似文献   

15.
Living/controlled radical polymerization of dimethyl(methacryloyloxy)methyl phosphonate (MAPC1) has been attempted using degenerative transfer to produce block copolymers. RAFT polymerization of this monomer is sensitive to very low level of oxygen and in any case limited to low monomer conversion. Reverse iodine transfer polymerization (RITP) leads to higher monomer conversion with a limited amount of living polymer (55% by 1H NMR), precluding an efficient synthesis of block copolymers. A PMMA‐b‐PMAPC1 diblock copolymer was therefore synthesized by iodine transfer polymerization (ITP) of MAPC1 from a PMMA‐I macro‐chain transfer agent prepared by RITP. The diblock copolymer, purified by selective precipitation, contains an average of five MAPC1 units and has potential an adhesive and as a corrosion inhibitor.

  相似文献   


16.
A newly designed automatically controlled stirred reactor suitable for kinetic measurements of reactions with half lives ≥2s has been applied to follow the anionic polymerization of methyl methacrylate in THF with Na+ as a counter ion in the presence of an excess of NaB(C6H5)4. As initiators were used: benzylsodium reacted with α-methylstyrene (I), fluorenylsodium (II), and 9-methylfluorenylsodium (III). With I the initiation is fast as compared with the polymerization reaction which is first order in monomer concentration. Within the range of ?50°C to ?100°C an almost unperturbed “living” polymerization is observed. The Arrhenius plot of the rate constants results in a straight line with activation energy Ea = 4,4kcal·mol?1 (= 18kJ·mol?1) and frequency exponent A = 7,0.II and III are slow initiators, II giving rise to side reactions because of the “acidic” proton in 9-position after initiation, III exhibiting a rate constant of initiation ki = 1l·mol?1·s?1 at ?72°C. The termination reaction is becoming increasingly important with increasing temperature and seems to be a unimolecular reaction with Ea,t = 11,5kcal·mol?1 (= 48 kJ·mol?1) and At = 10. Since the basic feature of the reactor is the possibility of drawing samples, polymers from each state of the reaction were available to be investigated also with respect to their tacticity. The monomer addition was shown to follow Bernoullian statistics. A structure of the “living” end being in harmony with the results observed is discussed.  相似文献   

17.
2-Methyl-1.3-dioxepane was polymerized in 1.2-dichloroethane with triethyl oxonium tetrafluoroborate and boron trifluoride as initiators. This polymerization involves a monomer-polymer equilibrium and the polymer consists of regularly alternating tetrahydrofuran and acetaldehyde units. The thermodynamic parameters for the polymerization were determined from the temperature dependence of the equilibrium monomer concentration as follows: ΔHSS = ?2.1 ± 0.3 kcal/mole and ΔS0SS = ?8.9 ± 1.3 cal/mole deg. The ceiling temperature for the 1 mole/1. solution is ?37°C. The attempted polymerization of 2.2-dimethyl-1.3-dioxepane gave only a crystalline cyclic dimer.  相似文献   

18.
The polymerization of N-(2,6-dimethylphenyl)itaconimide (1) with azoisobutyronitrile (2) was studied in tetrahydrofuran (THF) kinetically and spectroscopically with the electron spin resonance (ESR) method. The polymerization rate (Rp) at 50°C is given by the equation: Rp = K [2] 0,5 · [1] 2,1. The overall activation energy of the polymerization was calculated to be 91 kJ/mol. The number-average molecular weight of poly (1) was in the range of 3500–6500. From an ESR study, the polymerization system was found to involve ESR-observable propagating polymer radicals of 1 under the actual polymerization conditions. Using the polymer radical concentration, the rate constants of propagation (kp) and termination (kt) were determined at 50°C. kp (24–27 L · mol?1 · s?1) is almost independent of monomer concentration. On the other hand, kt (3,8 · 104–2,0 · 105 L · mol?1 · s?1) increases with decreasing monomer concentration, which seems mainly responsible for the high dependence of Rp on monomer concentration. Thermogravimetric results showed that thermal degradation of poly (1) occurs rapidly at temperatures higher than 360°C and the residue at 500°C was 12% of the initial polymer. For the copolymerization of 1 (M1) with styrene (M2) at 50°C in THF the following copolymerization parameters were found; r1 = 0,29, r2 = 0,08, Q1 = 2,6, and e1 = +1,1.  相似文献   

19.
The polymerization of N-octadecylmaleimide ( 1 ) initiated with azodiisobutyronitrile ( 2 ) was investigated kinetically in benzene. The overall activation energy of the polymerization was calculated to be 94,2 kJ·mol?1. The polymerization rate (Rp) at 50°C is expressed by the equation, Rp = k[ 2 ]0,6[ 1 ]1,7. The homogeneous polymerization system involves ESR-detectable propagating polymer radicals. Using Rp and the polymer radical concentration determined by ESR, the rate constants of propagation (kp) and termination (kt) were evaluated at 50°C. kp (33 L · mol?1 · s?1 on the average) is substantially independent of the monomer concentration. On the other hand, kt (0,3 · 104 – 1,0 · 104 L · mol?1 · s?1) is fairly dependent on the monomer concentration, which is ascribable to a high dependence of kt on the chain length of rigid poly( 1 ). This is the predominant factor for the high order with respect to the monomer concentration in the rate equation. In the copolymerization of 1 (M1) and St (M2) with 2 in benzene at 50°C, the following copolymerization parameters were obtained: r1 = 0,11, r2 = 0,09, Q1 = 2,1, and e1 = +1,4.  相似文献   

20.
Homopolymerization and copolymerization of N‐(2‐hydroxyethylmethyl)acrylamide ( 1 ) with N,N′‐diethyl‐1,3‐bis(acrylamido)propane ( 2 ) have been studied in ethanol/water solvents or in bulk. The polymerization rate exponent depends on the polymerization temperature and increased from 0.99 at 50 °C to 1.21 at 75 °C. The initial polymerization rate dependence on temperature between 50 and 75 °C in Arrhenius coordinates was linear. However, the slope increased with an increase in monomer concentration. The overall activation energy Ea varied from 64 kJ · mol?1 at 0.5 mol · L?1 to 76 kJ · mol?1 at 2.0 mol · L?1 monomer concentration. In addition, the reaction order with respect to the initiator AIBN deviated from the standard free radical polymerization kinetics. The exponent 0.42 indicated a significant participation of primary radicals in termination reactions. The polymerization reaction order declined from ideality as well, and the Ea and polymerization enthalpy changed with the batch composition variation, which could be explained through monomer/monomer, monomer/solvent, or monomer/polymer complexation. The chain transfer to polymer was considered as an origin for the polymer network formation in monomer 1 homopolymerization if its concentration in the batch was 1 mol · L?1 or higher. The more intensive polymerization acceleration with the monomer dilution and the greater heat released reduction with decrease in the monomer 1 to 2 ratios were evidenced in copolymerization. Nevertheless, the deviations from conventional reaction kinetics were not as pronounced as with acrylic or methacrylic acid polymerizations.

  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