Startseite Synthesis of random copolymer of isobutylene with p-methylstyrene by cationic polymerization in ionic liquids
Artikel Open Access

Synthesis of random copolymer of isobutylene with p-methylstyrene by cationic polymerization in ionic liquids

  • Xiaoqian Zhang EMAIL logo , Wenli Guo , Yibo Wu EMAIL logo , Yuwei Shang , Shuxin Li und Weihao Xiong
Veröffentlicht/Copyright: 20. Juni 2018
Veröffentlichen auch Sie bei De Gruyter Brill
e-Polymers
Aus der Zeitschrift e-Polymers Band 18 Heft 5

Abstract

Poly(isobutylene-co-p-methylstyrene) (IB/p-MeSt) random copolymer is a new generation of polyisobutylene-based elastomer. The cationic copolymerization of IB with p-MeSt was thoroughly examined by using various initiating systems in [Hmim][NTf2] at −30°C. The effects of initiating systems and monomer feed ratio on the monomer conversion, molecular weight and copolymer composition are discussed. The polymers were characterized by 1H-nuclear magnetic resonance (NMR), Fourier transform infrared (FTIR) spectroscopy and matrix-assisted laser desorption/ionization-time-of-flight-mass spectroscopy (MALDI-TOF-MS) to determine their chemical composition and molecular structure. The results show that high polarity, high viscosity and ionic environment of ionic liquids (ILs) influenced the copolymerization. The corresponding mechanism of cationic copolymerization in ILs is also proposed.

1 Introduction

Polyisobutylene (PIB) is widely applied in various industrial areas because of its attractive properties, namely, extremely low permeability, excellent oxidative stability and chemical resistance (1), (2), (3), (4). p-Methylstyrene (p-MeSt) is an inexpensive and attractive monomer that provides rigid polymers of pendant methyl repeating units with attractive thermal, mechanical, high glass-transition temperatures and low dielectric constants. Poly(isobutylene-co-p-methylstyrene) (IB/p-MeSt) random copolymer is a new generation of PIB-based elastomer that contains several percentages of “reactive” p-MeSt. IB/p-MeSt random copolymer also has many properties, which are superior to the traditional butyl rubber (5), (6). IB/p-MeSt random copolymer can only be synthesized by cationic polymerization. However, the traditional cationic polymerizations usually occur in chlorinated solvents, which are toxic, volatile and corrosive. In addition, in cationic polymerization, the disposal of the conventional Lewis acid catalysts (e.g. SnCl4, TiCl4, AlCl3, AlCl3OBu2 and BF3OEt2) (7), (8), (9), (10), (11) remains an issue. Thus, cationic polymerization in environmentally friendly green solvents is an important development direction for the present PIB industry.

Ionic liquids (ILs), which are composed solely of anions and cations, are a class of chemicals that have recently emerged as alternatives to environmentally damaging volatile organic compounds. ILs are regarded as polar but non-coordinating solvents with high charge density because of their ionic nature (12), (13). Thus, ILs are not simple solvents for cationic polymerizations. As a new style of green solvent, ILs possess numerous peculiar properties, such as non-flammability, negligible vapor pressure, low melting point, high thermal stability and chemical inertia (14), (15), (16), (17), (18); they can be widely used in several synthesis fields (19), (20), (21), (22), (23), (24), (25), (26), (27).

However, to date, studies on the application of ILs in ionic polymerization, especially in cationic copolymerization, are limited. It has been reported that ILs could also be used as efficient catalysts of cationic polymerizations (28). In addition, several water- and air-stable ILs, such as N-butyl-N-methyl pyrrolidinium bis(trifluoromethanesulfonyl)amide ([C4mpyr][NTf2]) (29), trihexyltetradecylphosphonium bis(trifluoromethanesulfonyl)amide ([P6,6,6,14][NTf2]) (30), 1-butyl-3-methylimidazolium hexafluorophospate ([Bmim][PF6]) (31), (32), 1-octyl-3-methylimidazolium tetrafluoroborate ([Omim][BF4]) (33), 1-octyl-3-methylimidazolium bis(trifluoromethanesulfonyl)amide ([Omim][NTf2]), 1-octyl-3-methylimidazolium bromide ([Omim][Br]), and 1-octyl-3-methylimidazolium hexafluorophospate ([Omim][PF6]) (34), have been successfully applied as cationic polymerization solvents. To date, cationic copolymerization in IL solvents has not been reported yet, and the mechanism is still vague.

In this study, first we report the application of IL as a solvent in the cationic copolymerization of IB with p-MeSt, initiated by various initiating systems. The effects of monomer feed ratios on monomer conversion, molecular weight and copolymer composition were investigated. In addition, the copolymerization mechanism of IB/p-MeSt in ILs is proposed based on the results. This study expands the application of ILs and enriches and develops the theory of cationic polymerization.

2 Experimental section

2.1 Materials

1-Butyl-3-methylimidazolium bis(trifluoromethanesulfonyl)imide ([Bmim][NTf2]) and 1-hexyl-3-methylimidazolium bis(trifluoromethanesulfonyl)imide ([Hmim][NTf2]) were purchased from Sigma Aldrich (USA), and the purities were more than 99.0%. The ILs were dried and degassed at 70°C for a few days. When the H2O content of IL was reduced to less than 30 ppm, it can be used (35). p-Methylstyrene (p-MeSt, TCI Shanghai, China, purity >98%) was dried by introduction of CaH2 before use. Isobutylene (IB, Beijing Yansan Petroleum Chemical Corp., Beijing, China, purity >99.9%) was cooled in the cold bath of a glovebox before use. Dichloromethane (CH2Cl2, Beijing Chemical Co., Beijing, China, 99.9%) and n-hexane (Hex, Beijing Chemical Co., Beijing, China, >98%) were distilled twice over CaH2 under reduced pressure before use. Commercially available titanium tetrachloride (TiCl4, Aldrich, USA, 99.9%), boron trifluoride etherate complex (BF3OEt2, Aldrich, USA, purity >98%) and anhydrous methanol (Beijing Chemical Co., Beijing, China, purity >99.9%) were used as received. Cumyl chloride (CumCl, J&K Scientific Ltd., Shanghai, China, purity >97.0%) was also used with further purification by double distillation from CaH2 under reduced pressure. The 2-chloro-2,4,4-trimethylpentane (abbreviated as TMPCl) was synthesized according to the literature (36). The clean and quantitative formation of TMPCl adduct was confirmed by 1H nuclear magnetic resonance (NMR) spectroscopy.

2.2 Polymerization copolymer of IB with p-MeSt

The procedures and cationic polymerizations were carried out under a dry nitrogen atmosphere ([H2O]<0.5 ppm; [O2]<10 ppm) in the MBraun 150-M glovebox. Fifty milliliter screw-cap vials with an IKA-MS3 vortex stirrer were used as polymerization reactors. A representative procedure is described as follows: Into a 50 ml screw-cap vial 4.85 ml of [Hmim][NTF2], 1.94 ml of p-MeSt and 0.34 ml of CumCl stock solution in dichloromethane (0.4 mol/l) were added and mixed thoroughly, and then the mixture were cooled to −30°C. Then 1.29 ml of IB was added to the cold mixture. IB should be added into the system after the addition of p-MeSt resulting from its limited solubility in IL. The polymerization was started by the addition of 0.31 ml of BF3OEt2 at −30°C. The polymer precipitated over the course of the polymerization. After a predetermined time, the polymerization was terminated by the induction of excess prechilled methanol. ILs and initiator could be dissolved in methanol, and the mixture was handled under decompress filter with 0.25 μm utra-filtration membrane. After that the products were washed several times with fresh methanol and then dried in a vacuum oven at 40°C to a constant weight in a few days. The IL was reused according to our previous report (37). Monomer conversions were determined by gravimetric analysis.

2.3 Measurements

The number-average molecular weight (Mn) and molecular weight distribution (MWD; i.e. Mw/Mn) of the polymers were detected by gel permeation chromatography (GPC, waters GPC, USA) using tetrahydrofuran (THF) as an eluent at a flow rate of 1.0 ml/min at room temperature. The GPC system was equipped with four Waters styragel columns connected in the following series: 500, 103, 104 and 105 at 30°C. The columns were calibrated against standard polystyrene samples.

NMR spectroscopy (Siemens, Germany) of the polymers was performed on a Bruker AV600 MHz spectrometer using CDCl3 as a solvent at 25°C. 1H-NMR spectra of solutions in CDCl3 were calibrated to tetramethylsilane as an internal standard (δH=0.00).

Matrix-assisted laser desorption ionization time-of-flight mass spectrometry (MALDI-TOF-MS) analyses were performed on a Ultraflex (AB Sciex, USA) TOF-MS, equipped with a 337 nm, 50 Hz N2 laser, a reflector and a delayed extraction. The apparatus was operated at an accelerating potential of 20 kV in the reflected mode. The matrix solution [DHB (2,5-dihydroxybenzoic acid), 10 μl of 20 g l−1 in THF] was mixed with 1 μl of polymer solution (10 g l−1 in THF). The spectra were recorded in the reflector mode in the absence of any cationizing agent. The final solution (1 μl) was deposited onto the target and dried in air at room temperature before irradiation. The mass spectra represent averages over 250 consecutive laser shots. External calibrations were performed with peptide calibration standard (Bruker Daltonic, Brenem, Germany).

The functional groups of poly(p-MeSt) and [Bmim][NTf2] were investigated using a Fourier transform infrared spectrometer (FTIR, Nicolet-Nexus 670, China).

The glass transition temperature (Tg) was recorded by TA DSC Q2000 (differential scanning calorimetry, USA) and the instrument was calibrated using indium as a standard (sample size 6−10 mg, heating rate 10°C min-1, nitrogen atmosphere).

3 Results and discussion

3.1 Copolymerization of IB with p-MeSt using various initiating systems

ILs are regarded as polar but non-coordinating solvents with high charge density; thus, they do not behave as simple solvents for ionic polymerizations. We comprehensively compared the cationic copolymerization in IL with those in organic solvents by employing a series of initiating systems to understand the ionic environment and its effect on cationic polymerizations.

As shown in Table 1, the cationic copolymerization of IB with p-MeSt was conducted in [Hmim][NTf2] at −30°C using various initiating systems. Cationic copolymerization in the traditional molecular solvent n-Hex/CH2Cl2 was also compared, as shown in Table 1. Generally, initiator/Lewis acid combinations, called initiating systems, play an important role in cationic copolymerization (38). The selected initiators include (a) H2O; (b) TMPCl, selected as a model of PIB living end; and (c) CumCl, selected as a similar model of poly(p-MeSt) living end. Lewis acids, which include TiCl4 and BF3OEt2, are commonly used in cationic polymerizations. From Table 1, the oligomers were obtained in the copolymerization of an equimolar mixture of IB and p-MeSt in [Hmim][NTf2]. In comparison with the products obtained from n-Hex/CH2Cl2 solvent, those obtained from [Hmim][NTf2] exhibited lower Mn and Mw/Mn, whereas the IB content in the product polymer changed. Notably, the product obtained from IL using BF3OEt2 Lewis acid resulted in a higher yield, regardless of the initiator used. Here, we selected the CumCl/BF3OEt2 initiating system to study the cationic copolymerization of IB with p-MeSt further.

Table 1:

Results for the cationic copolymerization of IB with p-MeSt using various initiating systems.a

RunInitiating systemSolventMn (GPC)Mw/MnYield %bIB content (%)c
1TMPCl/TiCl4[Hmim][NTf2]13111.207.1
2CumCl/TiCl412271.1913.6
3H2O/TiCl419351.987.9
4TMPCl/BF3OEt211461.2182.636.3
5CumCl/BF3OEt211721.2186.236.1
6H2O/BF3OEt211021.2561.633.1
7TMPCl/TiCl4V(Hex)/V(CH2Cl2)=50:5025882.4193.8
8CumCl/TiCl431122.2595.0
9H2O/TiCl425042.6292.3
10TMPCl/BF3OEt252751.9960.033.6
11CumCl/BF3OEt251282.1467.732.0
12H2O/BF3OEt244352.1668.336.1
  1. aConditions: [p-MeSt]o=[IB]o=1.768 mol l−1; [TMPCl]o=[CumCl]o=[H2O]o=6.162 mmol l−1; [BF3OEt2]o=[TiCl4]o=0.111 mol l−1; Mn(theor)=50,000 g mol−1; at −30°C for 60 min.

  2. bTotal monomer conversion.

  3. cIB content in product polymer; Determined by 1H-NMR; “–“ stands for no test because of their low yield.

3.2 Effects of comonomer feed ratios

The effects of monomer ratios in the feed on the copolymer composition was studied. The IB monomer is almost insoluble in pure [Hmim][NTf2], whereas it can be dissolved in the mixture of [Hmim][NTf2] and p-MeSt. The feed ratio of [IB]0 to [p-MeSt]0 was changed from 0:100 (mol:mol) to 60:40 (mol:mol). When the feed ratio was larger than 60:40, IB was insoluble in the polymerization system. The effect of the comonomer composition on the copolymerization is shown in Table 2. From the table, the comonomer composition significantly influenced Mn, Mw/Mn, and the yield. Under identical reaction conditions, a high concentration of IB in the comonomers resulted in narrow Mw/Mn and low Mn due to the rapid cross-transfer relative to propagation. Moreover, yield had a maximum value as the concentration of IB increases under the same reaction time. In addition, the polymerizations that generated oligomers only indicated that chain transfer occurred in [Hmim][NTf2].

Table 2:

Effect of the initial feed ratios.a

Run[IB]0:[p-MeSt]0IB content (%)bMn (calcd.)Mn (GPC)Mw/MnYield %c
10:100012,58051482.1062.9
210:9013.017,54020851.5787.7
320:8020.618,17018981.4690.7
430:7025.718,92014161.3994.6
540:6031.317,94011571.2989.7
650:5033.117,18010571.2585.9
760:4038.782009501.2141.0
  1. aConditions: [CumCl]o=15.405 mol l−1; [BF3OEt2]o=0.277 mol l−1; Mn(theor)=20,000 g mol−1; at −30°C for 60 min.

  2. bIB content (mol %) in product polymer; determined by 1H-NMR.

  3. cTotal monomer conversion.

The GPC curves of copolymers with different compositions are shown in Figure 1A. The curves were unimodal, which suggested the successful formation of the corresponding copolymer. In addition, a clear lateral shift of maxima toward a high molecular weight region was observed as IB:p-MeSt ratios decreased. Therefore, the Mn of the product evidently increased with the IB monomer content increasing in the initial feed ratios (Figure 1). IB could easily participate in the polymerization when dissolved, although dissolving IB in the polymerization system was difficult. This indicated that IB was more “active” in the chain transfer reaction.

Figure 1: GPC curves of copolymers and refractive index () and UV (256 nm) GPC traces of IB/p-MeSt copolymer are shown in Figure 1B.(A) GPC traces of the synthesized copolymers; (B) the UV/RI ratio, RI and UV traces of IB/p-MeSt copolymer obtained in [Hmim][NTf2] (monomer feed ratio of: p-MeSt-IB=1:1).
Figure 1:

GPC curves of copolymers and refractive index () and UV (256 nm) GPC traces of IB/p-MeSt copolymer are shown in Figure 1B.

(A) GPC traces of the synthesized copolymers; (B) the UV/RI ratio, RI and UV traces of IB/p-MeSt copolymer obtained in [Hmim][NTf2] (monomer feed ratio of: p-MeSt-IB=1:1).

The refractive index (RI) and UV (256 nm) GPC traces of IB/p-MeSt copolymer are shown in Figure 1B. The RI signal was proportional to the total mass of the polymer chain. The UV signal was only proportional to the number of aromatic monomer units incorporated into the chain because the UV absorption of the IB unit was negligible at 256 nm. The RI and UV traces of the IB/p-MeSt random copolymers obtained in [Hmim][NTf2] were close to perfect overlapping. Thus, the IB units were introduced into the copolymer homogeneously, and the product was the random copolymer. The UV/RI ratio was proportional to the p-MeSt content of the given molecular weight fraction and was constant over the entire molecular weight range. The Tg of IB/p-MeSt copolymer was strongly dependent on its microstructure. Figure 2 shows the DSC curves for copolymers with different IB:p-MeSt ratios. All samples provided only one Tg value, which was only between that of IB (−66.7°C) and p-MeSt (110°C). The Tg value decreased and moved to that of IB with the increase in IB:p-MeSt ratios. This result indicated that the copolymer without phase separation was a random copolymer.

Figure 2: DSC curves of the synthesized copolymers.
Figure 2:

DSC curves of the synthesized copolymers.

3.3 Microstructure of the copolymer

The structure of the copolymer in a feed ratio of 50:50 was characterized by IR, 1H-NMR, and MALDI-TOF-MS. Figure 3 shows the FTIR spectra of the obtained polymers. In Figure 3A, the band at 3018 cm−1 was attributed to the C-H in the phenyl ring, and the band at 812 cm−1 was the out-of-plane CH bending of two adjacent aromatic hydrogens of p-MeSt. However, the band at 1363 cm−1 could only be caused by the twisted vibration absorption peak of tertiary butyl of IB. Figure 4 shows the 1H-NMR spectrum of the polymers with the signal assignment to the possible structures. Figure 4A shows the aromatic ring protons at δ 6.3–7.3 ppm and the methyl protons of IB at δ 1.0 ppm (37), (38). The IR and 1H-NMR data suggested that the copolymer was successfully obtained. Figure 4A also shows three possible end groups of the copolymers, which include the methoxy terminal group (-OCH3, 3.1 ppm) that resulted from the quenching of the polymerization with methanol (i) and indanyl rings (ii) and (iii) formed by the Friedel-Crafts reactions. The characteristic resonance at 4.0 ppm was assigned to the indanyl ring (ii), and that at 1.68 ppm was assigned to the indanyl ring (iii) (39), (40). Therefore, the main chain transfer reaction was a Friedel-Crafts reaction.

Figure 3: FTIR spectra of the synthesized copolymers IB/p-MeSt copolymer (A); PIB (B); poly(p-MeSt) (C).
Figure 3:

FTIR spectra of the synthesized copolymers IB/p-MeSt copolymer (A); PIB (B); poly(p-MeSt) (C).

Figure 4: Representative 1H NMR spectrum of the synthesized IB/p-MeSt copolymer (A); PIB (B); poly(p-MeSt) (C).
Figure 4:

Representative 1H NMR spectrum of the synthesized IB/p-MeSt copolymer (A); PIB (B); poly(p-MeSt) (C).

The copolymer was also analyzed via MALDI-TOF-MS. Figure 5A shows the spectra of the copolymer, revealing two series of periodic groups of peaks. Each series of groups was separated by 56 m/z, which corresponded to the addition of an isobutyl group; whereas each isotopic distribution in the cluster was separated by 6 m/z. This result could be accounted for by the mass difference between one p-MeSt unit (118 u) and two isobutyl units (2×56=112 u) (41). Thus, when moving from left to right in an oligomer group, a gain of 6 m/z corresponded to one more p-MeSt monomer and two less isobutyl monomers. When moving from one grouping of oligomers to another, one isobutyl group was gained or lost. In addition, a series of peaks could be attributed to the copolymer chains that bear a methoxy terminal group (structure a) at the ω-end caused by the quenching of the polymerization with methanol (Mn=926.6, 932.6, and 939.5 g mol−1; Figure 5B). The other peak was 22 Da above the previous value (Mn=939.5 g mol−1) and could be attributed to the copolymer with indanyl ring structures (structures b, Mn=961.2 g mol−1), which were the results of Friedel-Crafts alkylations. These results suggested that the IB/p-MeSt random copolymer could be obtained in [Hmim][NTf2] via cationic copolymerization. In addition, the chain transfer reaction resulted in a lower Mn relative to the theoretical value.

Figure 5: Spectra of the copolymer.(A) MALDI-TOF-MS for cationic copolymerization (in IL) of p-MeSt and IB (feed ratio of IB: p-MeSt-IB=1:1); (B) expanded view of spectrum in Figure 5A.
Figure 5:

Spectra of the copolymer.

(A) MALDI-TOF-MS for cationic copolymerization (in IL) of p-MeSt and IB (feed ratio of IB: p-MeSt-IB=1:1); (B) expanded view of spectrum in Figure 5A.

3.4 Effects of solvents on copolymerization

Although NTf2−1 can stabilize the propagating carbocationic species through the moderation or delocalization of the positive charges in homopolymerization of p-MeSt (37), the factors for copolymerization are highly complex. In [Hmim][NTf2], the copolymerization only generated oligomers. Thus, we attempted to identify the reason for this low molecular weight. We investigated the different polarity of solvents, including ILs and the traditional solvents for cationic polymerizations. Table 3 shows the results of the different polarities of solvents on copolymerization. The results show that a high polarity of solvents yielded low Mn and narrow Mw/Mn. Therefore, the high polarity of ILs resulted in the low Mn and narrow Mw/Mn, although they could stabilize carbocation (36). In addition, the high viscosity of ILs that resulted in the low releasing rate of reaction heat possibly caused the low molecular weight (32).

Table 3:

Effects of different type of solvents and their polarity on copolymerization of IB with p-MeSt.a

RunSolventPolarityMn (GPC)Mw/MnYield %b
1V(n-Hex)/V(CH2Cl2)=60:40
33642.7585.6
2CH2Cl226502.3092.4
3[Bmim][NTF2]12471.2580.8
4[Hmim][NTF2]10571.2585.9
  1. aConditions: [p-MeSt]o=[IB]o=1.768 mol·l−1; [CumCl]o=15.405 mmol·l−1, [BF3OEt2]o=0.277 mol·l−1; Mn(theor)=20,000 g/mol; at −30°C.

  2. bTotal monomer conversion.

3.5 Copolymerization mechanism in ILs

From the aforementioned results, we proposed the corresponding mechanism of cationic copolymerization of IB with p-MeSt in ILs system, as shown in Scheme 1. BF3OEt2 activated the C-Cl bond of the cumyl chloride to form the initiating cationic species. The copolymerization in ILs occurred by chain breaking via the predominant Friedel-Crafts alkylations, which created a new polymer chain via protic reinitiation and resulted in a lower Mn relative to the theoretical value. The high polarity, high viscosity, and ionic environment of ILs influenced the copolymerization, although ILs did not participate in the initiation or termination reactions in the entire polymerization process.

Scheme 1: The mechanism of cationic copolymerization of IB with p-MeSt in ILs.
Scheme 1:

The mechanism of cationic copolymerization of IB with p-MeSt in ILs.

4 Conclusions

The IB/p-MeSt random copolymer with high conversion, average molecular weight (Mn) ranging from 800 to 2100, and narrow molecular weight distribution (Mw/Mnca. 1.1–1.5) could be prepared via cationic copolymerization in an IL medium. The different compositions of IB from 13 mol% to 39 mol% in IB/p-MeSt random copolymers could be mediated by changing the monomer ratio in the feed. The GPC curves were unimodal throughout the copolymerization. The polymerizations that generated oligomers only indicated that chain transfer occurred in ILs. The high polarity, high viscosity, and ionic environment of ILs were the possible reasons for the chain transfer of copolymerization.

Award Identifier / Grant number: 51373026

Award Identifier / Grant number: 51573020

Award Identifier / Grant number: 51503019

Funding statement: This work was supported by the Hebei Province Science Foundation for Youths (E2018508091), the Fundamental Research Funds for the Central Universities (3142017097), the Langfang Technology Support Project (2017011040), the National Natural Science Foundation of China (No. 51373026, No. 51573020, and No. 51503019), and Beijing Natural Science Foundation (No. 2172022 and No. 2162014).

References

1. Kang J, Erdodi G, Kennedy JP. Polyisobutylene‐based polyurethanes with unprecedented properties and how they came about. J Polym Sci Part A Polym Chem. 2011;49:3891–904.10.1002/pola.24839Suche in Google Scholar

2. Alvarez Albarran A. Modular surface functionalization of polyisobutylene-based biomaterials. Diss Theses Gradworks. 2014;50(12):273–90.Suche in Google Scholar

3. Pratap G, Mustafa S, Hollis W, Heller J. Living carbocationic polymerization of isobutylene by tert‐amyl alcohol/BCl3/1‐methyl‐2‐pyrrolidinone initiating system. J Appl Polym Sci. 1992;44(6):1069–74.10.1002/app.1992.070440616Suche in Google Scholar

4. Ren P, Wu YB, Guo WI, Li SX, Mao J, Xiao F, Li K. Synthesis, characterization and haemocompatibility of poly(styrene-b-isobutylene-b-styrene) triblock copolymers. Polym Korea. 2011;35:40–6.10.7317/pk.2011.35.1.40Suche in Google Scholar

5. Xie Y, Chang JJ, Wu YB, Yang D, Wang H, Zhang T, Li SX, Guo WL. Synthesis and properties of bromide‐functionalized poly(isobutylene‐co‐p‐methylstyrene) random copolymer. Polym Int. 2016;66:468–76.10.1002/pi.5286Suche in Google Scholar

6. Tsou AH, Favis BD, Hara Y, Bhadane PA, Kirino Y. Reactive compatibilization in brominated poly(isobutylene‐co‐p‐methylstyrene) and polyamide blends. Macromol Chem Phys. 2009;210(5):340–8.10.1002/macp.200800465Suche in Google Scholar

7. Vasilenko IV, Shiman DI, Kostjuk SV. Highly reactive polyisobutylenes via AlCl3OBu2‐coinitiated cationic polymerization of isobutylene: effect of solvent polarity, temperature, and initiator. J Polym Sci Part A Polym Chem. 2012;50(4):750–8.10.1002/pola.25830Suche in Google Scholar

8. De P, Faust R. Determination of the absolute rate constant of propagation for ion pairs in the cationic polymerization of p-methylstyrene. Macromolecules 2005;38(13):5498–505.10.1021/ma050449iSuche in Google Scholar

9. Jun A, Hidehiro Y, Motomasa Y, Shokyoku K, Sadahito A. Living cationic polymerization of styrene derivatives using SnCl4/EtAlCl2 in the presence of a Lewis base. Polym Preprints. 2009;50(1):156.Suche in Google Scholar

10. Satoh K, Nakashima J, Kamigaito M, Sawamoto M. Novel BF3OEt2/R−OH initiating system for controlled cationic polymerization of styrene in the presence of water. Macromolecules 2001;34(3):396–401.10.1021/ma0006070Suche in Google Scholar

11. Kostjuk SV, Dubovik AY, Vasilenkol IV, Mardykin VP, Gaponik LV, Kaputsky FN, Antipin LM. Novel initiating system based on AlCl3 etherate for quasiliving cationic polymerization of styrene. Polym Bull. 2004;52(3):227–34.10.1007/s00289-004-0280-2Suche in Google Scholar

12. Reichardt C. Polarity of ionic liquids determined empirically by means of solvatochromic pyridinium N-phenolate betaine dyes. Green Chem. 2005;7(5):339–51.10.1039/b500106bSuche in Google Scholar

13. Earle MJ, Engel BS, Seddon KR. Keto–Enol tautomerism as a polarity indicator in ionic liquids. Aust J Chem. 2004;57(2):149–50.10.1071/CH03259Suche in Google Scholar

14. Kowsari MH, Fakhraee M. Influence of butyl side chain elimination, tail amine functional addition, and C2 methylation on the dynamics and transport properties of imidazolium-based [Tf2N] ionic liquids from molecular dynamics simulations. J Chem Eng Data. 2015;60(3):551–60.10.1021/je500618wSuche in Google Scholar

15. Figoli A, Marino T, Simone S, Di Nicolo E, Li XM, He T, Tornaghi S, Drioli E. Towards non-toxic solvents for membrane preparation: a review. Green Chem. 2014;16(9):4034–59.10.1039/C4GC00613ESuche in Google Scholar

16. Cevasco G, Chiappe C. Are ionic liquids a proper solution to current environmental challenges? Green Chem. 2014;16(5):2375–85.10.1039/c3gc42096eSuche in Google Scholar

17. Losetty V. Recent advances and thermophysical properties of acetate-based protic ionic liquids. Chem Sci. 2016;7(2):1–7.10.4172/2150-3494.1000128Suche in Google Scholar

18. Patel R, Kumari M. Interaction between pyrrolidinium based ionic liquid and bovine serum albumin: a spectroscopic and molecular docking insight. Biochem. & B Anal. Biochem. 2016;5(2). DOI: 10.4172/2161-1009.1000265.Suche in Google Scholar

19. Lozinskaya EI, Shaplov AS, Vygodskii YS. Direct polycondensation in ionic liquids. Eur Polym J. 2004;40(9):2065–75.10.1016/j.eurpolymj.2004.05.010Suche in Google Scholar

20. Zhang S, Dias Goncalves L, Lefebvre H, Tessier M, Rousseau B, Fradet A. Direct poly(β-alanine) synthesis via polycondensation in ionic liquids. ACS Macro Lett. 2012;1(8):1079–82.10.1021/mz300264vSuche in Google Scholar

21. Vygodskii YS, Lozinskaya EI, Shaplov AS. Ionic liquids as novel reaction media for the synthesis of condensation polymers. Macromol Rapid Commun. 2002;23(12):676–80.10.1002/1521-3927(20020801)23:12<676::AID-MARC676>3.0.CO;2-2Suche in Google Scholar

22. Brusseau SGN, Boyron O, Schikaneder C, Santini CC, Charleux B. Nitroxide-mediated controlled/living radical copolymerization of methyl methacrylate with a low amount of styrene in ionic liquid. Macromolecules 2011;44(2):215–20.10.1021/ma102321zSuche in Google Scholar

23. Jeličić A, Beuermann S, García N. Influence of ionic liquid structure on the propagation kinetics of methyl methacrylate. Macromolecules 2009;42(14):5062–72.10.1021/ma900774eSuche in Google Scholar

24. Muginova SV, Galimova AZ, Polyakov AE, Shekhovtsova TN. Ionic liquids in enzymatic catalysis and biochemical methods of analysis: capabilities and prospects. J Anal Chem. 2010;65(4):331–51.10.1134/S1061934810040027Suche in Google Scholar

25. Li Y, Qiang Q, Zheng X, Wang Z. Controllable electrochemical synthesis of Ag nanoparticles in ionic liquid microemulsions. Electrochem Commun. 2015;58:41–5.10.1016/j.elecom.2015.05.020Suche in Google Scholar

26. Biedroń T, Kubisa P. Atom‐transfer radical polymerization of acrylates in an ionic liquid. Macromol Rapid Commun. 2001;22(15):1237–42.10.1002/1521-3927(20011001)22:15<1237::AID-MARC1237>3.0.CO;2-ESuche in Google Scholar

27. Zhang H, Zhang Y, Liu W, Wang H. Kinetic study of atom transfer radical polymerization of methyl methacrylate in ionic liquids. J Appl Polym Sci. 2008;110(1):244–52.10.1002/app.28643Suche in Google Scholar

28. Vasilenko IV, Berezianko IA, Shiman DI, Kostjuk SV. New catalysts for the synthesis of highly reactive polyisobutylene: chloroaluminate imidazole-based ionic liquids in the presence of diisopropyl ether. Polym Chem. 2016;7(36):5615–9.10.1039/C6PY01325BSuche in Google Scholar

29. Vijayaraghavan R, MacFarlane DR. Organoborate acids as initiators for cationic polymerization of styrene in an ionic liquid medium. Macromolecules 2007;40(18):6515–20.10.1021/ma070668zSuche in Google Scholar

30. Vijayaraghavan R, Macfarlane DR. Novel acid initiators for the rapid cationic polymerization of styrene in room temperature ionic liquids. Sci China Chem. 2012;55(8):1671–76.10.1007/s11426-012-4658-ySuche in Google Scholar

31. Biedro T, Kubisa P. Cationic polymerization of styrene in a neutral ionic liquid. J Appl Polym Sci. 2004;42(13):3230–5.10.1002/pola.20158Suche in Google Scholar

32. Han L, Wu YB, Dan Y, Wang H, Zhang XQ, Wei XL, Guo WL, Li SX. Characteristics and mechanism of styrene cationic polymerization in 1-butyl-3-methylimidazolium hexafluorophosphate ionic liquid. RSC Adv. 2016;6(107):105322–30.10.1039/C6RA22284FSuche in Google Scholar

33. Wu Y, Han L, Zhang X, Mao J, Gong L, Guo W, Gu K, Li S. Cationic polymerization of isobutyl vinyl ether in an imidazole-based ionic liquid: characteristics and mechanism. Polym Chem. 2015;6(13):2560–8.10.1039/C4PY01784FSuche in Google Scholar

34. Yoshimitsu H, Kanazawa A, Kanaoka S, Aoshima S. Cationic polymerization of vinyl ethers with alkyl or ionic side groups in ionic liquids. J Polym Sci Part A Polym Chem. 2016;54(12):1774–84.10.1002/pola.28039Suche in Google Scholar

35. Huddleston JG, Visser AE, Reichert WM, Willauer HD, Broker GA, Rogers RD. Characterization and comparison of hydrophilic and hydrophobic room temperature ionic liquids incorporating the imidazolium cation. Green Chem. 2001;3(4):156–64.10.1039/b103275pSuche in Google Scholar

36. Hadjikyriacou S, Acar M, Faust R. Living and controlled polymerization of isobutylene with alkylaluminum halides as coinitiators. Macromolecules 2004;37(20):7543–7.10.1021/ma049082sSuche in Google Scholar

37. Zhang X, Guo W, Wu Y, Gong L, Li W, Li X, Li S, Shang Y, Yang D, Wang H. Cationic polymerization of p-methylstyrene in selected ionic liquids and polymerization mechanism. Polym Chem. 2016;7(32):5099–112.10.1039/C6PY00796ASuche in Google Scholar

38. De P, Faust R. Living carbocationic polymerization of p-methoxystyrene using p-methoxystyrene hydrochloride/SnBr4 initiating system: determination of the absolute rate constant of propagation for ion pairs. Macromolecules 2004;37(21):7930–7.10.1021/ma0490451Suche in Google Scholar

39. Ashbaugh JR, Ruff CR, Shaffer TD. Characterization of signals in the solution 1H NMR spectrum of poly(isobutylene‐co‐p‐methylstyrene) and their utility. J Polym Sci Part A Polym Chem. 2000;38(9):1680–6.10.1002/(SICI)1099-0518(20000501)38:9<1680::AID-POLA34>3.0.CO;2-YSuche in Google Scholar

40. Lubnin AV, Országh I, Kennedy JP. The microstructure of poly(isobutylene-co-p-methylstyrene) by NMR spectroscopy. J Macromol Sci. Part A Pure Appl Chem. 1995;32(11):1809–30.10.1080/10601329508009363Suche in Google Scholar

41. Cox FJ, Johnston MV, Qian K, Peiffer DG. Compositional analysis of isobutylene/p-methylstyrene copolymers by matrix-assisted laser desorption/ionization mass spectrometry. J Am Soc Mass Spectrom. 2004;15(5):681–8.10.1016/j.jasms.2003.12.017Suche in Google Scholar

Received: 2018-01-25
Accepted: 2018-04-24
Published Online: 2018-06-20
Published in Print: 2018-09-25

©2018 Walter de Gruyter GmbH, Berlin/Boston

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Heruntergeladen am 29.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/epoly-2018-0017/html
Button zum nach oben scrollen