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 93 毫秒
1.
Polymerization of MMA was carried out in near-bulk and in fairly dilute solutions in DMF at 40° with (1) CPB-BZ2O2 combination as the redox initiator in the dark and (2) CPB as the sole photoinitiator in visible light. Rp was proportional to ([CPB] [BZ2O2])0.18 for the redox polymerization and to [CPB]0.35 for the photopolymerization both in near-bulk (10% DMF, v/v) and in high dilution (50% DMF, v/v). Polymerization was inhibited by hydroquinone in each case indicating a radical mechanism. Effect of several solvents/additives on the polymerization revealed that dimethyl formamide, acetonitrile and pyridine acted as rate enhancing solvents in the redox polymerization while in the photopolymerization they acted as normal diluents; benzene, methanol, acetone and chloroform exhibited inert nature while formamide and acetamide behaved as retarding additives in both types of polymerization. Initiator transfer of the degradative type reasonably explains kinetic nonidealities.  相似文献   

2.
Kinetics of solution polymerization of styrene was studied using pyridine as solvent and BZ2O2 and azobisisobutyronitrile (AIBN) as initiators at 60°C. Normal kinetic features (Rp ∝ [AIBN]0.5 · [styrene]1.0) were observed for the AIBN-initiated polymerization, with pyridine playing the role of an inert diluent; but in the BZ2O2-initiated polymerization, the monomer exponent was found to vary from a low value of 0.45 at a relatively low initiator concentration (1 × 10?2 mole/liter) to a value higher than the usual value of unity (1.18) at a much higher concentration of the initiator (16 × 10?2 mole/liter). The initiator exponent value was found to be 0.5 (usual) up to 20% v/v dilution with pyridine, but it showed a tendency to decrease with increase in pyridine content beyond 20% v/v. The k/kt value for each initiator system, however, was found to remain constant over the whole concentration range of pyridine. The unusual kinetic features were explained on the basis of predominance of one or the other of two competitive reactions in BZ2O2-initiated system: (a) higher rate of decomposition of BZ2O2 in pyridine and (b) primary radical depletion by reaction with pyridine, depending upon the concentration of BZ2O2 and pyridine.  相似文献   

3.
Photopolymerization of MMA was studied kinetically at 35° using TEA-BZ2O2 redox system as initiator. The initiator exponent is 0.34 but the monomer exponent depends on the solvent. Solvents (acetonitrile, pyridine and bromobenzene) giving negative or fractional monomer exponent show a rate enhancing effect through actively influencing the initiation step; benzene and chloroform give first order dependence of rate on [monomer] and behave as normal (inert) diluents. Initiation of polymerization takes place through radicals generated by photodecomposition of TEA-BZ2O2 complex formed in situ, the radical generation step being solvent or monomer dependent. Kinetic non-idealities are interpreted in terms of significant initiator dependent termination via degradative chain transfer.  相似文献   

4.
Solution polymerization of MMA, with pyridine as the solvent and BZ2O2 and AIBN as thermal initiators, was studied kinetically at 60°C. The monomer exponent varied from 0.45 to 0.91 as [BZ2O2] was increased from 1 × 10?2 to 30 × 10?2 mole/liter in a concentration range of 8.3-4.6 mole/liter for MMA. For AIBN-initiated polymerization the monomer exponent remained constant at 0.69 as [AIBN] varied from 0.4 × 10?2 to 1.0 × 10?2 mole/liter in the same concentration range for MMA. The k2p/kt Value increased in both cases with an increase in pyridine concentration in the system. This was explained in terms of an increase in the kp value, which was due presumably to the increased reactivity of the chain radicals by donor-acceptor interaction between the molecules of solvent pyridine and propagating PMMA radicals and in terms of lowering the kt value for the diffusion-controlled termination reaction due to an increase in the medium viscosity and pyridine content.  相似文献   

5.
Abstract

The kinetics of phase-transfer-agent-assisted free-radical polymerization of methyl methacrylate using K2S2O8 as the water-soluble initiator and triethylbenzylammonium chloride (TEBA) as the phase-transfer catalyst (PTC) was investigated in toluene-water biphase media at 60°C. The effect of varying [MMA], [K2S2O8], [TEBA], [H+], the ionic strength of the medium, and the temperature on the rate of polymerization (R p) was studied. R p was found to be proportional to [MMA]2, [K2S2O8]1, and [TEBA]0.5. Based on the kinetic results, a mechanism involving initiation of polymerization by phase-transferred S2O8 2- and termination by Q+ (quaternary ammonium ion) is proposed.  相似文献   

6.
Photopolymerization of MMA with the use of H2O2 as the photoinitiator under visible light at 30°C was studied. Kinetic features in bulk monomer and in the presence of different diluents differ significantly. Usual free radical kinetics with square-root dependence of rate on initiator, indicating bimolecular termination of chain radicals, were observed for bulk polymerization. On dilution with various solvents polymerization was found to be retarded to different (usual and more than usual) extents, the observed monomer exponent value being much higher than unity in many cases. This deviation from normal kinetics has been interpreted in terms of the predominance of degradative initiator transfer in the diluted systems.  相似文献   

7.
The polymerization of acrylonitrile (AN) initiated by the system of tetramethyl tetrazene (TMT) and bromoacetic acid (BA) in dimethylformamide (DMF) was studied. The TMT–BA system could initiate the polymerization of AN more easily than TMT alone. The polymerization was confirmed to proceed through a radical mechanism. The initial rate of polymerization Rp was expressed by the equation: Rp = [TMT]0.62-[BA]0.5[AN]1.5. The overall activation energy for the polymerization was estimated as 9.4 kcal/mole. In the absence of monomer, the reaction of TMT with BA in DMF was also studied kinetically by measuring the evolution of nitrogen gas. The reaction was first-order in TMT and first-order in BA; the rate data at 49°C were k2 = 9.1 × 10?2l./mole-sec., ΔH? = 17.0 kcal/mole, and ΔS? = ? 6.6 eu. In addition, the treatment of TMT with BA in benzene led to the formation of tetramethylhydrazine radical cation, which was identified by its ESR spectrum. On the other hand, the relatively strong interaction between TMT and DMF was observed by absorption spectrophotometry.  相似文献   

8.
The electro-initiated polymerization of acrylonitrile initiated by the anodic oxidation of an aqueous acid solution (80% HOAc + 20% H2O) containing Mn(OAc)2 · 4H2O/CNCH2COOH has been investigated in the 30–40°C temperature range. The kinetics and mechanism of the process has been investigated as a function of variables and a suitable mechanism proposed. From the experimental observations the rate of polymerization is seen to be proportional to [An]1.5I0.5[Mn+2]0.5 and [CAA]0.5. The rate of polymerization gradually decreases at a higher applied current. The rate was independent of [CAA]0.5. The rate of polymerization gradually decreases at a higher applied current. The rate was independent of CAA at high concentration. The average degrees of polymerization (P n) increases with increasing AN and decreasing [CAA], [Mn+2] and applied current, I. The initiation is due to the anodic oxidation of Mn+2–CNCH2COOH complex. Both the initiation of polymerization by the primary radical, viz., CN? C?? COOH as well as the oxidation of the primary radical at the electrode are equally significant reactions and neither can be neglected in comparison with the other. Predominant mutual termination accounts for all the observed data.  相似文献   

9.
In this original experiment, reverse atom transfer radical polymerization technique using CuCl2/hexamethyl tris[2-(dimethylamino)ethyl]amine (Me6-TREN) as catalyst complex was applied to living radical polymerization of 4-vinylpyridine (4VP) with azobisisobutyronitrile (AIBN) as initiator. N,N-Dimethylformamide was used as solvent to improve the solubility of the reaction system. The polymerization not only showed the best control of molecular weight and its distribution, but also provided a rather rapid reaction rate with the molar ratio of [4VP]:[AIBN]:[CuCl2]:[Me6-TREN] = 400:1:2:2. The rate of polymerization increased with increasing the polymerization temperature and the apparent activation energy was calculated to be 51.5 kJ· mol1. Use of Cl as the halogen in copper halide had many advantages over the use of Br. The resulting poly(4-vinylpyridine) was successfully used as the macroinitiator to proceed the block polymerization of styrene in the presence of CuCl/Me6-TREN catalyst complex via a conventional ATRP process in DMF.  相似文献   

10.
The aqueous polymerization of MMA was studied kinetically at 40° C using low concentrations of Py-SO2 complex as initiator. For [Py-SO2] < 2 × 10?2 mol/L, Rp ∞ [PY-SO2]0.5 [M]1.5, and for [Py-SO2] > 2 × 10?2 mol/L, Rp ∞ [Py-SO2]0,0[M]1.08. Polymerization is considered to proceed by a radical mechanism. The radical generation or the initiation step is believed to proceed through equilibrium complexation between the Py-SO2 complex and monomer molecules. For [Py-SO2] < 2 × 10?2 mol/L, the polymerization is characterized by bimolecular termination. Above this [Py-SO2], chain termination by a degradative initiator transfer process assumes prominence.  相似文献   

11.
The polymerization of acrylonitrile (AN) initiated by tetramethyl-2-tetrazene (TMT) and p-toluenesulfonic acid (TSA) in dimethylformamide (DMF) was studied. The polymerization was confirmed to proceed through a radical mechanism. The initial rate of polymerization R was expressed by the equation: Rp = k[TMT]0.6 [TSA]0.46 [AN]1.35. The overall activation energy for the polymerization was estimated as 20.7 kcal/mole. In the absence of monomer, the reaction of TMT with TSA was also studied kinetically by measuring the evolution of nitrogen. From these results and ESR measurement of the TMT/TSA system, a possible initiation mechanism is proposed.  相似文献   

12.
The overall reaction rate kinetics of polymerization of diethyleneglycol dimethacrylate and copolymerization of it with styrene in bulk and in the presence of inert diluents were investigated. Theresults indicated that these reactions can be treated as free radical polymerization with highly diffu-sion controlled termination reaction in which the termination rate constant is an empirically derivedfunction of monomer conversion: K_t=K_(to)(1-c ln[M]/ [M_0])~(-1) in which K_(to) is the initial terminationrate constant and c is a factor related to the magnitude of diffusion co?re The following equationof monomer conversion as a function of time could then be derived: U=1-exp {1/c [1-(1+ckt/2)~2]}in which k=K_P(R_i/2K_(to))~(1/2) and t is the time of reaction. Excellent agreement between the theoreticaland experimental overall reaction kinetic curves was obtained. The equation is valid for crosslinkingand noncrosslinking free radical polymerizations in which the self-acceleration effcct is effective fromthe very beginning of the reaction. The equation can be expressed in a more generally applicableform: U=1--exp{1/e[1--(1+?t/n)~n] in which n≥0.  相似文献   

13.
Polymerization of methyl acrylate in HClO4 and HNO3 was studied in the temperature range 10–15°C. The kinetics of the polymerization were found to be very simple, involving initiation and termination by cobaltic ions. Kinetic studies on polymerization of acrylonitrile in HClO4 and HNO3 revealed that water oxidation, and monomer oxidation were side reactions as in the case of methyl methacrylate. Experimental evidence favored the simultaneous initiation by Co3+ and CoOH2+ species. In H2SO4, certain unusual features were encountered. At low [Co3+], linear termination as well as termination by mutual combination occurred. Another interesting aspect was that CoSO4+ initiated at low [Co3+]. This was unlike the case of other monomers in H2SO4. The rates of polymerization and rates of cobaltic ion disappearance were measured with respect to changes in [M], [Co3+], [H+], temperature, etc. The various rate constants were evaluated.  相似文献   

14.
The kinetics of redox -initiated polymerization of acrylic acid (AA) by the systerm Mn3+-isobutyric acid (IBA) in sulfuric acid was studied in the temperature range of 35–50°C. The overall rates of polymerization (Rp), disappearance of manganic ion (?Rm), and degree of polymerization (X n), were measured with variation in [monomer], [Mn3+], [IBA], H+, μ, [Mn2+], and temperature. The polymerization is initiated by the organic free radical that develops from the Mn3+-isobutyric acid oxidation reaction. Two types of termination reactions, one by the metal ion (Mn3+) and the other by the MN3+-isobutyric acid complex are proposed to explain the kinetic results. The various rate parameters were evaluated an discussed.  相似文献   

15.
The photolysis of formamide vapor at 2062 Å has been studied in a flow system with results essentially similar to those obtained previously under static conditions and higher conversions. The rotating-sector technique has been applied to the radical-chain decomposition of formamide under conditions (305°C, 11.5 torr) such that decomposition of the carbamyl (NH2CO) radical was rate controlling, so that [NH2CO] ? [NH2]. A rate constant of (3.1 ± 1.0) × 1010(M·sec)?1 was obtained for bimolecular chain termination by carbamyl radicals. A concurrent first-order radical loss, probably at the surface, was taken into account by the treatment described by Shepp. Both oxamide and HNCO were tentatively identified as termination products, suggesting the occurrence of both combination and disproportionation, but quantitative estimates of the relative rates were not possible. From the rate constant for chain termination, and relative rate constants obtained previously, Arrhenius parameters A = (5.9 ± 2.0) × 1012 sec?1 and A0 = (1.04 ± 0.35) × 1014 (M·sec)?1 were estimated for the unimolecular decomposition of carbamyl radicals in the high and low pressure limits.  相似文献   

16.
The kinetics of oxidation of N,N‐dimethylformamide by chromium(VI) has been studied spectrophotometrically in aqueous perchloric acid media at 20°C. The rate showed a first‐order dependence on both [Cr(VI)] and [DMF], and increased markedly with increasing [H+]. The order with respect to [HClO4] was found to lie between 1 and 2. The rate was found to be independent of ionic strength as well as of any inhibition effect of Mn(II). The formation of superoxochromium(III) ion was detected in an aerated solution of chromium(VI), DMF and HClO4. The proposed mechanism, involving two reaction pathways, leads to the rate law, rate = Ka1 [HCrO4] [DMF] (kI Ka2 [H+]²+kII[H+]). The first pathway, with rate constant kI, involves the formation of chromium(V) and a free radical. The second pathway, with rate constant kII, involves the formation of Cr(IV), CO2 and dimethylamine. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 409–415, 1999  相似文献   

17.
《印度化学会志》2023,100(8):101058
The kinetics of Cu(II) accelerated L-valine (Val) oxidation by hexacyanoferrate(III) in CTAB micellar medium were investigated by measuring the decline in absorbance at 420 nm. By adjusting one variable at a time, the progression of the reaction has been inspected as a function of [OH], ionic strength, [CTAB], [Cu(II)], [Val], [Fe(CN)63−], and temperature using the pseudo-first-order condition. The results show that [CTAB] is the critical parameter with a discernible influence on reaction rate. [Fe(CN)6]3- interacts with Val in a 2:1 ratio, and this reaction exhibits first-order dependency with regard to [Fe(CN)63−]. In the investigated concentration ranges of Cu(II), [OH], and [Val], the reaction demonstrates fractional-first-order kinetics. The linear increase in reaction rate with added electrolyte is indicative of a positive salt effect. CTAB significantly catalyzes the process, and once at a maximum, the rate remains almost constant as [CTAB] increases. Reduced repulsion between surfactant molecules' positive charge heads brought on by the negatively charged [Fe(CN)6]3-, OH, and [Cu(OH)4]2- molecules may be responsible for the observed drop in CMC of CTAB.  相似文献   

18.
Anodic electroinitiated polymerization of acrylamide has been studied in DMF and DMSO in the presence of Co(NO3)2 or Co(ClO4)2 in the temperature range 25–40°C. The kinetics and mechanism of the process has been investigated as a function of variables and a suitable mechanism proposed. From the experimental observations, the rate of polymerization is seen to be proportional to [AM]1.5, [I]0.5, and [Co2+]1. Current densities exceeding 15 mA/cm2 have no effect on the rates. The average degrees of polymerisation (P?n) increase with increasing [AM] and decreasing [Co2+] and applied current, I. It has been shown that a monomer-metal ion complex is oxidized at the anode, generating radical species. The polymerization and termination are confined to the anode compartment. The process is very efficient compared to the NO mediated reaction.  相似文献   

19.
The influence of stereoregular poly(methyl methacrylate) (PMMA) as a polymer matrix on the initial rate of radical polymerization of methyl methacrylate (MMA) has been measured between ?11 and +60°C using a dilatometric technique. Under proper conditions an increase in the relative initial rate of template polymerization with respect to a blank polymerization was observed. Viscometric studies showed that the observed effect could be related to the extent of complex formation between the polymer matrix and the growing chain radical. The initial rate was dependent on tacticity and molecular weight of the matrix polymer, solvent type and polymerization temperature. The accelerating effect was most pronounced (a fivefold increase in rate) at the lowest polymerization temperature with the highest molecular weight isotactic PMMA as a matrix in a solvent like dimethylformamide (DMF), which is known to be a good medium for complex formation between isotactic and syndiotactic PMMA. The acceleration of the polymerization below 25°C appeared to be accompanied by a large decrease in the overall energy and entropy of activation. It is suggested that the observed template effects are mainly due to the stereoselection in the propagation step (lower activation entropy Δ Sp?) and the hindrance of segmental diffusion in the termination step (higher activation energy Δ Et?) of complexed growing chain radicals.  相似文献   

20.
Photopolymerization of methyl methacrylate in bulk and in solution at 40 °C using monochloroacetic acid –dimethyl aniline (MCAA–DMA) combination as photoinitiator was studied kinetically. The apparent activation energy was found to be 4.39 kcal/mol (18.37 kJ/mol) while the kinetic parameter kp2/kt was 1.27 × 10−2 1/mol/sec. The kinetic data indicated that polymerization followed a radical mechanism. The initiator order was found to be 0.25, indicating significant deviation from the square root dependence for normal free radical kinetics. The non‐ideality in the kinetics can be explained on the basis of significant initiator‐dependent termination through primary radicals or degradative initiator transfer. The observed monomer order was significantly less than unity (i.e. nonideal behavior) for use of carbon tetrachloride, chloroform, methylethyl ketone and acetic acid as diluents, but it was unity (i.e. ideal behavior) for use of benzene as the diluent. Solvents other than benzene contributed to enhancement of rate of polymerization by influencing the radical generation step. End‐group analysis indicated the incorporation of DMA and MCAA moieties as end‐groups in the polymers. Copyright © 1999 John Wiley & Sons, Ltd.  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