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.
Summary: Propagation rate coefficients, kp, for free‐radical polymerization of butyl acrylate (BA) previously reported by several groups are critically evaluated. All data were determined by the combination of pulsed‐laser polymerization (PLP) and subsequent polymer analysis by size exclusion (SEC) chromatography. The PLP‐SEC technique has been recommended as the method of choice for the determination of kp by the IUPAC Working Party on Modeling of Polymerization Kinetics and Processes. Application of the technique to acrylates has proven to be very difficult and, along with other experimental evidence, has led to the conclusion that acrylate chain‐growth kinetics are complicated by intramolecular transfer (backbiting) events to form a mid‐chain radical structure of lower reactivity. These mechanisms have a significant effect on acrylate polymerization rate even at low temperatures, and have limited the PLP‐SEC determination of kp of chain‐end radicals to low temperatures (<20 °C) using high pulse repetition rates. Nonetheless, the values for BA from six different laboratories, determined at ambient pressure in the temperature range of ?65 to 20 °C mostly for bulk monomer with few data in solution, fulfill consistency criteria and show excellent agreement, and are therefore combined together into a benchmark data set. The data are fitted well by an Arrhenius relation resulting in a pre‐exponential factor of 2.21 × 107 L · mol?1 · s?1 and an activation energy of 17.9 kJ · mol?1. It must be emphasized that these PLP‐determined kp values are for monomer addition to a chain‐end radical and that, even at low temperatures, it is necessary to consider the presence of two radical structures that have very different reactivity. Studies for other alkyl acrylates do not provide sufficient results to construct benchmark data sets, but indicate that the family behavior previously documented for alkyl methacrylates also holds true within the alkyl acrylate family of monomers.

Arrhenius plot of propagation rate coefficients, kp, for BA as measured by PLP‐SEC.  相似文献   


2.
Summary: The termination kinetics of dibutyl itaconate (DBI) bulk polymerization was studied via SP–PLP–ESR single pulse–pulsed laser polymerization with time‐resolved detection of free‐radical concentration by electron‐spin resonance, at temperatures between 0 and 60 °C. As is characteristic of PLP experiments, termination rate coefficients, kt(i,i), are measured for radicals of (almost) identical chain length (CL) i. CL‐averaged 〈kt〉, for chain lengths up to 200 monomer units, and also kequation/tex2gif-stack-1.gif referring to termination of very small‐size radicals are directly deduced from measured DBI radical concentration vs time traces. At 45 °C, 〈kt〉 is (3.4 ± 0.6) · 105 L · mol?1 · s?1 and kequation/tex2gif-stack-2.gif is (7.2 ± 1.0) · 105 L · mol?1 · s?1. Both rate coefficients are independent of monomer conversion up to the highest experimental conversion of 18%. The associated activation energies are EA(〈kt〉) = 23.0 ± 3.2 kJ · mol?1 and EA(kequation/tex2gif-stack-3.gif) = 27.6 ± 2.8 kJ · mol?1, respectively. “Model‐dependent” and “model‐free” analyses of radical concentration vs time profiles indicate a pronounced CL dependence of kt(i,i) for DBI radicals of moderate size, 5 < i < 50. The lowering of kt(i,i) with CL corresponds to an exponent α close to 0.5 in a power‐law expression kt(i,i) = kequation/tex2gif-stack-4.gif · i?a. At higher chain lengths, the variation of kt(i,i) with CL becomes weaker and may be represented by an α value of 0.16 or even below. These results are consistent with models according to which α varies to a larger extent at low CL and to a smaller extent at high CL with the crossover region between the two ranges being located somewhere around i = 100.

Conversion‐dependence of 〈kt〉 and kequation/tex2gif-stack-5.gif from laser‐induced photopolymerization of DBI.  相似文献   


3.
Propagation rate coefficients, kp, which have been previously reported by several groups for free‐radical bulk polymerizations of cyclohexyl methacrylate (CHMA), glycidyl methacrylate (GMA), benzyl methacrylate (BzMA), and isobornyl methacrylate (iBoMA) are critically evaluated. All data were determined by the combination of pulsed‐laser polymerization (PLP) and subsequent polymer analysis by size‐exclusion chromatography (SEC). This so‐called PLP‐SEC technique has been recommended as the method of choice for the determination of kp by the IUPAC Working Party on Modeling of Polymerisation Kinetics and Processes. The present data fulfill consistency criteria and the agreement among the data from different laboratories is remarkable. The values for CHMA, GMA, and BzMA are therefore recommended as constituting benchmark data sets for each monomer. The data for iBoMA are also considered reliable, but since SEC calibration was established only by a single group, the data are not considered as a benchmark data set. All kp data for each monomer are best fitted by the following Arrhenius relations: CHMA: , GMA: , BzMA: , iBoMA: . Rather remarkably, for the methacrylates under investigation, the kp values are all very similar. Thus, all data can be fitted well by a single Arrhenius relation resulting in a pre‐exponential factor of 4.24 × 106 L · mol?1 · s?1 and an activation energy of 21.9 kJ · mol?1. All activation parameters refer to bulk polymerizations at ambient pressure and temperatures below 100 °C. Joint confidence intervals are also provided, enabling values and uncertainties for kp to be estimated at any temperature.

95% joint confidence intervals for Arrhenius parameters A and EA for cyclohexyl (CHMA), glycidyl (GMA), benzyl (BzMA), and isobornyl (iBoMA) methacrylate; for details see text.  相似文献   


4.
Summary: The telomerization of 10‐undecenol with alkyl hydrogenphosphonate was studied in order to synthesize telomers of different molecular weights. The study showed that telomers from 10‐undecenol could be obtained despite the fact that the double bond has a low reactivity. The kinetic constant Kp2/KTe was determined to be 7 × 10?4 l · mol?1 · s?1 at 135 °C and the transfer constant CT was 0.057. These values are normal for a low activity telogen such as hydrogenphosphonate and for slightly reactive monomers like 10‐undecenol.

SEC chromatogram of telomers obtained in the reaction of 10‐undecenol addition.  相似文献   


5.
The propagation rate coefficient (kp) of the vinyl pivalate (VPi) has been determined from an EPR quantification of the concentration of propagating radicals. The polymerizations followed first‐order kinetics with respect to the monomer, and the concentration of propagating radicals did not vary with time. The chain lengths were estimated to be sufficiently large to not affect kp. The resulting Arrhenius parameters for kp of VPi in heptane are Ap = 1.39 × 107 dm3 · mol?1 · s?1 and Ep = 20.5 kJ · mol?1. They are very similar to those of vinyl acetate as previously determined by the PLP method. The present results thus indicate that EPR gives consistent kp values for vinyl esters.

  相似文献   


6.
The copolymerization of p-tert-butoxystyrene ( 1 ) (M1) and dibutyl fumarate ( 2 ) (M2) initiated with dimethyl 2,2′-azobisisobutyrate ( 3 ) was studied in benzene at 60°C kinetically and by means of electron paramagnetic resonance (EPR) spectroscopy. The monomer reactivity ratios were determined to be r1 = 0,18 and r2 = 0,01, indicating that homopropagation of M2 is almost negligible in the copolymerization. The copolymerization system was revealed to involve EPR-observable propagating polymer radicals under practical copolymerization conditions. The apparent rate constants of propagation (kp) and termination (kt) determined by EPR show a rapid increase in the range from 0,9 to 1,0 of feed composition (f1 = {[M1]/([M1] + [M2])}) of M1. From the relationship between kp and f1 based on Fukuda's penultimate model, the rate constants of propagation of copolymerization were evaluated; k111 = 140 L · mol?1 · s?1, k211 = 4,3 L · mol?1 · s?1, k112 = 778 L · mol?1 · s?1, k212 = 24 L · mol?1 s?1 and k121 = 19 L · mol?1 · s?1, suggesting a pronounced penultimate effect.  相似文献   

7.
Pulsed laser polymerization (PLP) with subsequent analysis of molecular mass distribution (MMD) is used to determine the rate coefficient of chain transfer to an agent A, ktrA, by varying pulse repetition rate such that the contributions of PLP‐induced and chain‐transfer‐induced peaks to the MMD change to a significant extent. It is shown by simulation that the relative heights of these peaks may be used to estimate ktrA. The method is applied to evaluation of the rate coefficient of chain transfer to dodecyl mercaptan with butyl methacrylate polymerizations at ?11, 0, 20 and 40 °C. The Arrhenius parameters for this coefficient are determined to be: A(ktrA) = (2.2 ± 0.6) × 106 L · mol?1 · s?1 and Ea(ktrA) = (22.1 ± 0.7) kJ · mol?1.

  相似文献   


8.
The absolute rate constants of propagation kp and of termination kt of ethyl α-cyanoacrylate (ECNA) were determined in bulk at 30°C by means of the rotating sector method under conditions to suppress anionic polymerization; kp = 1 622 1 · mol?1 · s?1 and kt = 4,11 · 108 1 · mol?1 · s?1 for the polymerization in the presence of acetic acid, and kp = 1610 1 · mol?1 · s?1 and kt = 4,04 · 108 l · mol?1 · s?1 for the polymerization in the presence of 1,3-propanesultone. The magnitude of k/kt determined was 6,39 · 10?3 l · mol?1 · s?1. The absolute rate constants for cross-propagation in ECNA copolymerizations were also evaluated. Quantitative comparison of the rate constants with those of common monomers and polymer radicals shows that the strong electron-withdrawing power of the ethoxycarbonyl and cyano groups enable the poly(ECNA) radical to add to monomers as fast as the other polymer radicals. The relatively high reactivity of ECNA, regardless of the type of attacking polymer radical, is interpreted by a transition state greatly stabilized by both the ethoxycarbonyl and the cyano groups.  相似文献   

9.
Electron paramagnetic resonance (EPR) spectroscopy is used for measuring rate coefficients of addition, kad, and fragmentation, kβ, together with the associated equilibrium constants, Keq, for butyl acrylate polymerizations mediated by S‐ethyl propan‐2‐ylonate‐S’‐propyl trithiocarbonate (EPPT) and by SS’‐bis(methyl‐2‐propionate) trithiocarbonate (BMPT). Experiments at ?40 °C yield kad = (3.4 ± 0.3) × 106 L mol?1 s?1, kβ = (1.4 ± 0.4) × 102 s?1, and Keq = (2.6 ± 0.8) × 104 L mol?1 for EPPT and kad = (4.1 ± 0.9) × 106 L mol?1 s?1, kβ = (4.5 ± 0.5) × 101 s?1, and Keq = (8 ± 4) × 104 L mol?1 for BMPT. The Keq values are in satisfactory agreement with data from ab initio calculations.

  相似文献   


10.
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   

11.
The accurate characterization of molar‐mass distributions of poly(acrylic acid) (PAA) and poly(methacrylic acid) (PMAA) by size‐exclusion chromatography (SEC) is addressed. Two methods are employed: direct aqueous‐phase SEC on P(M)AA and THF‐based SEC after esterification of P(M)AA to the associated methyl esters, P(M)MA. P(M)AA calibration standards, P(M)AA samples prepared by pulsed‐laser polymerization (PLP), and PAA samples prepared by reversible addition‐fragmentation chain transfer (RAFT) are characterized in a joint initiative of seven laboratories, with satisfactory agreement achieved between the institutions. Both SEC methods provide reliable results for PMAA. In the case of PAA, close agreement between the two SEC methods is only observed for samples prepared by RAFT polymerization with weight‐average molar mass between 80 000 and 145 000 g mol?1 and for standards with peak molar masses below 20 000 g mol?1. For standards with higher molar masses and for PLP‐prepared PAA, the values from THF‐based SEC are as much as 40% below the molar masses determined by aqueous‐phase SEC. This discrepancy may be due to branching or degradation of branched PAA during methylation. While both SEC methods can be recommended for PMAA, aqueous‐phase SEC should be used for molar‐mass analysis of PAA unless the sample is not branched.

  相似文献   


12.
The telomerization of acrylic acid (AA) with thioglycolic acid (TGA) initiated with 2,2′‐azoisobutyronitrile (AIBN) was first investigated in organic medium (THF, 65°C). The kinetic study of this telomerization led to the determination of the TGA transfer constant (CT = 3.2) and to the ratio kp/√kte equal to 0.48 L1/2·mol–1/2·s–1/2. Then, the same study was performed both in aqueous medium and in water/THF mixture in order to investigate the solvent effect on the transfer constant. From these works it is emphasised that the nature of the solvent plays an important role on the kinetics parameters. First, this research underlined an increase of the kp/√kte value by raising the water proportion in water/THF mixtures. Then, the kinetic study showed the highest value for the kp/√kte constant, equal to 2.48 L1/2·mol–1/2·s–1/2 when the telomerization proceeded in water. Consequently, the value of CT, which is directly influenced by the kp/√kte constant, presented a decrease from CT = 3.2 in THF to a value equal to 0.5 in water. By this way, the « ideal » case of telomerization (CT = 1) was reached for a mixture of solvents; 80% water/20% THF (v/v).  相似文献   

13.
Radical propagation kinetics of the bulk homopolymerizations of vinyl pivalate (VPi) and vinyl benzoate (VBz) have been studied using pulsed‐laser polymerization (PLP) combined with size exclusion chromatography (SEC). As part of the study, the Mark–Houwink para­meters of poly(VPi) and poly(VBz) in tetrahydrofuran are determined using a triple detector SEC. The observed significant increase (by ≈ 20%) of the bulk VPi propagation rate coefficient (kp) as pulse repetition rate is increased from 200 to 500 Hz is similar to that reported for vinyl acetate (VAc). Data collected in the temperature range of 25–85 °C for VPi is well fit by the Arrhenius relation ln(kp/L mol?1 s?1) = 15.73?2093(T/K). The activation energy is similar to that found for vinyl acetate (VAc), with kp values higher by ≈ 50%. PLP studies in ethyl acetate and in heptane find no substantial solvent effect on VPi or VAc kp values. Attempts to measure the propagation kinetics of VBz by PLP are not successful, suggesting that significant radical stabilization occurs for the system. Small‐scale batch poly­merization experiments demonstrate relative polymerization rates of these vinyl ester monomers that are consistent with the PLP results.

  相似文献   


14.
To improve the knowledge of emulsion copolymerization of monomers both swelling their copolymers, but which are of quite different polarity (water solubility), a series of styrene (S)/methyl acrylate (MeA) copolymerizations was carried out in batch at 50°C with potassium persulfate as initiator. The overall rates of copolymerization increase with the amount of MeA in the monomer feed. Copolymer composition follows the usual copolymerization equation if bulk/solution reactivity ratios (rij) and monomer partition between aqueous and organic phase are taken into account (simulation). However, accurate kinetic data at low conversion (gas chromatography) put in evidence an enhanced polymerization of the more hydrophilic monomer (MeA), which can be attributed to polymerization in the water phase. Particle sizes increase with conversion and tend to a limiting value, the higher the MeA content is. Particle number (Np), which is practically constant with conversion of S homopolymerization, tends to increase with MeA content as polymerization proceeds. This trend is enhanced if the emulsifier (sodium dodecanesulfonate, SDS) concentration is increased. Overall propagation rate constants were estimated as function of the experimental conditions and monomer concentration within the particles. From kinetic data (rate of polymerization) and Np, it was found that the average number of radicals per particle, ñ, remains close to 0,5. It was then possible considering S(kp = 125 1 · mol?1 · s?1) as a standard monomer, to estimate the polymerization rate constant for MeA (335 1 · mol?1 · s?1). Since adsorption of emulsifier was shown to be closely related to particle surface composition, the specific area As of SDS was measured on latices at various conversions and initial monomer feeds. As conversions increases, the particle surface appears to be richer and richer in MeA, which corresponds to a particle structuration. Strong and weak acid group titration is also in quite good agreement with the colloidal behaviour.  相似文献   

15.
Summary: The free‐radical polymerization kinetics of 4‐acetoxystyrene (4‐AcOS) is studied over a wide temperature range. Pulsed‐laser polymerization, in combination with dual detector size‐exclusion chromatography, is used to measure kp, the propagation rate coefficient, between 20 and 110 °C. Values are roughly 50% higher than those of styrene, while the activation energy of 28.7 kJ · mol−1 is lower than that of styrene by 3–4 kJ · mol−1. With known kp, conversion and molecular weight data from 4‐AcOS thermal polymerizations conducted at 100, 140, and 170 °C are used to estimate termination and thermal initiation kinetics. The behavior is similar to that previously observed for styrene, with an activation energy of 90.4 kJ · mol−1 estimated for the third‐order thermal initiation mechanism.

Joint confidence (95%) ellipsoids for the frequency factor A and the activation energy Ea from non‐linear fitting of kp data for 4‐AcOS (black) and styrene (grey).  相似文献   


16.
The aqueous-phase polymerization of N,N′-methylenebis(acrylamide) initiated by potassium peroxodisulfate in the absence and in the presence of the anionic emulsifier sodium dodecylsulfate was kinetically investigated at 50°C by conventional gravimetric and dilatometric methods. The rate of polymerization is found to be proportional to the 0,75 and 0,24 oder with respect to potassium peroxodisulfate and N,N′-methylenebis(acrylamide) concentrations, respectively. On the other hand, it is independent of the concentration of sodium dodecylsulfate. This agrees with the polymerization of a monomer soluble in water. Therefore, the equations for a homogeneous polymerization were applied to evaluate the experimental results. The calculated ratio kp/kt0,5 of the rate constants of propagation kp and termination kt for the N,N′-methylenebis(acrylamide) polymerization at zero conversion in the absence of emulsifier are scattered in the interval between 3,1 and 3,4 dm1,5 · mol?0,5 · s?0,5 and in the presence of emulsifier in the interval between 2,4 and 3,5 dm1,5 · mol?0,5 · s?0,5. They are close to those obtained for the homogeneous polymerization of acrylamide in the aqueous phase. The lower values of kp/kt0,5 ≈ 0,3–0,6 dm1,5 · mol?0,5 · s?0,5 determined for the polymerization of N,N′-methylenebis(acrylamide) for conversions between 30 and 60% follow from the hindered termination reaction within the polymer particles. The polymer dispersions formed are unstable. The growth of the polymer particles proceeds predominantly by coalescence. This suggests a kinetics which does not follow the Smith-Ewart theory but is characterized by a continuous particle nucleation and agglomeration. The interval 1 occurs at the beginning of the dispersion polymerization when polymer particles are being formed. Interval 2 follows, once the number of polymer particles has been fixed.  相似文献   

17.
The SP‐PLP‐EPR technique is used to carry out a detailed investigation of the radical termination kinetics of 1H, 1H, 2H, 2H‐tridecafluorooctyl methacrylate (TDFOMA) in bulk at relatively low conversion. Composite‐model behavior for chain‐length‐dependent termination rate coefficients, kti,i, is observed. It is found that for TDFOMA, ic ≈ 60 independent of temperature, and αs ≈ 0.65 and αl ≈ 0.2 at 80 °C and above. However, at lower temperatures the situation is strikingly different, with the significantly higher average values of αs = 0.89 ± 0.15 and αl = 0.32 ± 0.10 being obtained at 50 °C and below. This makes TDFOMA the first monomer to be found that exhibits clearly different exponent values, αs and αl, at lower and higher temperature, and that has both a high αs, like an acrylate, and a high ic, like a methacrylate.  相似文献   

18.
The kinetics of free radical solution (co)polymerization of N‐acryloylmorpholine (NAM) and N‐acryloxysuccinimide (NAS) have been investigated at 60°C in 1,4‐dioxane using 2,2′‐azoisobutyronitrile as the initiator. First, the kp/kt1/2 value for the homopolymerization of NAM monomer was estimated to be 4.97 mol–1/2·L1/2· s–1/2, which is indicative of a high ability of NAM to homopolymerize. Then, the reactivity ratios were determined to be rNAS =  0.63 ± 0.03 and rNAM =  0.75 ± 0.01. This binary comonomer system behaves like a perfectly random one, allowing the synthesis of macromolecules homogeneous in composition. Average copolymer compositions determined by 1H NMR and 13C NMR spectroscopy were in good agreement with the theoretical ones. Finally, monomer sequence distribution was studied by 13C NMR analysis.  相似文献   

19.
Methyl α-(alkoxymethyl)acrylates were prepared from methyl α-(bromomethyl)acrylate in 80–90% yield. The monomers homopolymerize fast to yield low-molecular-weight polymers. The monomers bearing a linear alkoxymethyl group except for the ethoxymethyl group are characterized by a relatively low ceiling temperature. The rate constants for propagation kp and termination kt of methyl α-(butoxymethyl)acrylate were evaluated to be kp = 298 dm3 · mol?1 · s?1 and kt = 8 · 106 dm3 · mol?1 · s?1 at 60°C, respectively. The α-(alkoxymethyl)acrylates are more reactive than methyl methacrylate toward polystyrene radical, except for the α-(dodecyloxymethyl)acrylate which is slightly less reactive, indicating that an increase in the reactivity by the electron-withdrawing character of the alkoxy group prevails over the steric hindrance against addition of the polymer radical, except for the large dodecyloxymethyl group.  相似文献   

20.
The single pulse (SP) – pulsed laser polymerization (PLP) technique has been applied to measure kt /kp, the ratio of termination to propagation rate coefficients, for free‐radical bulk homopolymerizations of methyl acrylate (MA) and dodecyl acrylate (DA) between 10 and 50°C at pressures from 10 to 2 500 bar. kt /kp is obtained from experimental monomer concentration vs. time traces that are determined via time‐resolved (μs) near infrared monitoring of monomer conversion induced by single excimer laser pulses of about 20 ns width. With kp being known from PLP–SEC experiments, chain‐length averaged kt is immediately obtained from kt /kp. For MA, kt remains constant up to about 15% monomer conversion and clearly decreases upon further polymerization. For DA at pressures of 100 bar and above, a plateau value of constant kt is observed up to about 60% monomer conversion whereas at lower pressure, e. g. at 10 bar, kt slightly increases in the very initial conversion region, but also exhibits a plateau kt value at moderate and high conversions. The occurrence of such plateau kt values and their pressure and temperature dependence are consistent with the view that plateau regions of kt are best understood in terms of diffusion control via segmental mobility.  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