Next Article in Journal
Modified NiFe2O4-Supported Graphene Oxide for Effective Urea Electrochemical Oxidation and Water Splitting Applications
Next Article in Special Issue
Temperature Dependence of the Number of Defect-Structures in Poly(vinylidene fluoride)
Previous Article in Journal
3D-QSAR and Molecular Dynamics Study of Isoxazole Derivatives to Identify the Structural Requirements for Farnesoid X Receptor (FXR) Agonists
Previous Article in Special Issue
Incorporation of Aramids into Polybenzimidazoles to Achieve Ultra-High Thermoresistance and Toughening Effects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nitroxide-Mediated Controlled Radical Copolymerization of α-Trifluoromethylstyrenes with Styrenes

Department of Chemistry, Ochanomizu University, 2-1-1 Otsuka, Bukyo-ku, Tokyo 104-8610, Japan
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(6), 1214; https://doi.org/10.3390/molecules29061214
Submission received: 6 February 2024 / Revised: 5 March 2024 / Accepted: 7 March 2024 / Published: 8 March 2024
(This article belongs to the Special Issue Themed Issue Dedicated to Prof. Bernard Boutevin)

Abstract

:
Fluorinated polymers are important materials in everyday life; however, most monomers of widely used fluoropolymers are gaseous, and their polymerization is difficult in an ordinary laboratory. Therefore, partially fluorinated polymers have recently been reported. As an easy-to-handle fluorine-containing monomer, α-trifluoromethylstyrene (TFMST) can be used to produce partially fluorinated polymers with trifluoromethyl groups in the main chain; however, TFMST does not homopolymerize, and there are limited reports on its copolymerization with styrene (ST). In this study, we applied the controlled radical polymerization method, which is effective for the polymerization of ST, to the copolymerization of TFMST and ST. We also showed that nitroxide-mediated polymerization is effective. The content ratio of TFMST in the TFMST–ST copolymer can be controlled between 10% and 40% by changing its monomer ratio. Additionally, the polymerization of TFMST and ST with substituents was performed to increase structural variations. The thermal stability as well as water and oil repellency of the synthesized polymers with different composition ratios and substituents were also evaluated.

1. Introduction

Fluorinated polymers are important materials in everyday life because of their unique properties, such as heat resistance, water repellency, and low dielectric constant [1,2,3,4]; however, well-established fluoropolymers, such as polytetrafluoroethylene, poly(vinylidenedifluoride), and poly(chlorotrifluoroethylene) are prepared from fluorinated olefin monomers that are gaseous, pose severe safety concerns [5,6], and are difficult to polymerize in an ordinary laboratory. Most of these polymers have very high melting points and poor solubility in common solvents and they lack formability [7,8]. Therefore, (semi)fluorinated polymers using fluorine-containing monomers, which are easy to handle, have recently attracted considerable attention [9]. Semifluorinated polymers include those with a fluoroalkylene in the main chain and those with a fluorine functional group in the polymer side chain. To incorporate fluoroalkylene into the backbone, the use of telechelic functional poly(perfluoroethers) [10] and leveraging difunctional perfluorinated monomers [9,11,12] have been reported. These polymerizations yield fluorine-substituted main chain polymers; however, the polymerization cannot be controlled [13]. In contrast, as examples of semifluorinated polymers with fluorine functional groups on the side chains, polymerization of fluorinated acrylate [14,15] or styrene (ST) [16,17] has been reported. Although it is easy to handle the monomer, the weakness of the main chain is a problem; therefore, strength—one of the advantages of fluoropolymers—cannot be realized. Here, when α-(trifluoromethyl)acrylate or α-trifluoromethylstyrene (TFMST) is used as a monomer, the tertiary hydrogen in the polymer main chain is replaced by a trifluoromethyl group. This displacement improves chemical resistance. Nevertheless, the polymerization of α-(trifluoromethyl)acrylate [18,19] and TFMST [20] is hindered by the presence of a bulky, electron-withdrawing trifluoromethyl group at the α-position. For instance, for TFMST, the kinetic study of radical copolymerization with ST was reported by Ito [20], and the monomer reactivity ratios of TFMST and ST revealed rTFMST = 0.00 and rST = 0.60; therefore, TFMST did not homopolymerize. Although TFMST retards the polymerization rates, its copolymerization with ST proceeds. Ameduri reported controlled radical polymerization using iodine transfer polymerization (ITP) [21,22]. Despite the potential of the TFMST–ST copolymer, there was only one report on the controlled copolymerization of TFMST and ST; therefore, further synthetic studies were conducted. Three typical controlled radical polymerizations effective for ST, atom transfer radical polymerization (ATRP) [23,24,25], reversible addition-fragmentation chain transfer (RAFT) [26,27,28], and nitroxide-mediated polymerization (NMP) [29,30,31] were applied to the TFMST–ST copolymerization. Copolymerization of TFMST and ST substituted with trifluoromethyl groups has also been investigated. Although the copolymerization of TFMST–ST limits the number of trifluoromethyl groups introduced, the fluorine content of the polymer can be improved by introducing additional fluorine functional groups into the monomer. The thermophysical properties as well as water and oil repellency of the resulting polymers have also been evaluated.

2. Results and Discussion

First, copolymerization of ST and TFMST (90:10 mixture) was conducted using three different polymerization methods with high-ST-control capability (Table 1). All methods produced polymers with almost 9:1 incorporation of ST and TFMST. The composition ratio of ST to TFMST in the synthesized polymers was determined by comparing the integrated values of 19F NMR of CF3 of benzotrifluoride (BTF: internal standard) and TFMST (Figure S1 and Table S1).
When ATRP using (1-bromoethyl)benzene (3b) as an initiator and Cu-tetramethylethylenediamine (TMEDA) complex as a catalyst was investigated [32], the reaction proceeded to give a product with an Mn of 10,000, and a dispersity (Đ = Mw/Mn; Mw represents the weight-average molecular weight) of 1.30 was obtained (entry one). Then, RAFT polymerization was performed (entry two) [33]. The reaction was performed at 75 °C for 48 h using 2-cyano-2-propyl dodecyl trithiocarbonate (3b) as a chain transfer reagent and azobis(isobutyronitrile) (AIBN) as a radical initiator, yielding a product with an Mn of 2200. Because of the low molecular weight at 48 h, the reaction time was extended to 96 h but only an Mn of 4500 products was obtained in low yield (entry three). Then, we tried the NMP. The reaction was performed using Hawker’s universal initiator, N-tert-butyl-N-(2-methyl-1-phenylpropyl)-O-(1-pheylethyl)hydroxylamine (3c) [34], and polymerization proceeded but with low yield and Mn (entry four). We also performed the reaction using 2-methyl-2-[1-(diethoxyphosphinyl)-2,2-dimethylpropyl](tert-butyl)aminooxy]propanoic acid (BlocBuilder MA® (3d)) [35,36] and found that the copolymer, which contains ST–TFMST = 91:9, was obtained in 56% yield with an Mn of 7300 and dispersity of 1.14 (entry five). Therefore, NMP with 3d as the initiator was chosen for future studies.
NMP using different monomer ratios of ST and TFMST or α-methylstylene (MST) was investigated (Table 2). First, the polymerization of ST was performed to obtain polystyrene with an Mn of 8300 and a dispersity of 1.07 in 65% yield (entry 1). The polymerization of TFMST alone was also performed. However, the reaction did not proceed, confirming that it did not homopolymerize. Copolymerization with different ratios of ST and TFMST resulted in lower yields and increased dispersion as the ratio of TFMST increased (entries 2–5). The content of TFMST in the copolymer was at the same ratio when the ratio of ST–TFMST was 90/10. Nevertheless, as the content of TFMST increased, its content increased up to 40% but below the monomer content (Figure S2). Furthermore, the reaction rate and yield decreased as the ratio of TFMST increased. When equal amounts of ST and TFMST were used, an approximately 2:1 copolymer was produced. This result is consistent with the report on free radical polymerization by Ueda et al. [20]. These newly synthesized copolymers were soluble in organic solvents such as chloroform, THF, and toluene. As a control experiment, ST–MST copolymerization was also examined but MST was not introduced into the polymer chain at a 90:10 ratio (entry five). Copolymers with ST and MST contents of 7:3 were obtained when the monomer ratio was 50:50 (entry six).
The thermal stability of the resulting copolymers was then assessed by differential scanning calorimetry (DSC) and thermal gravimetric analysis. The 5% weight loss decomposition temperature (Td5) was found to be approximately 270 °C for all polymers. The glass transition temperature (Tg) was lower for the ST–MST copolymer than for polystyrene but higher for polymers containing TFMST, with lower TFMST content. Although the polymer chain length synthesized here is relatively short, the copolymer of ST and TFMST tends to have higher Tg than polystyrene, while Td5 does not change significantly.
Thermogravimetric coupled with gas chromatography–mass spectrometry (TG-GC-MS) measurements were performed on copolymers with high TFMST content (40% TFMST content; Table 1, entry five) to investigate their degradation behavior (Figure S5). TG at a temperature between 30 and 600 °C was measured and the degradation products were observed by GC–MS. Based on the molecular weight, the degradation products derived from ST and TFMST were mainly observed, with less desorption of trifluoromethyl-derived degradation products. This result indicates that the depolymerization to ST and TFMST occurs without the elimination of the trifluoromethyl group. In addition, decomposition occurs in two stages, at around 280 and 410 °C. Since the ceiling temperature of polystyrene is 395 °C, the living end of the polymer is considered to be decomposed in the first peak.
Next, we followed the polymerization reaction over time (Figure 1, Tables S2 and S3). The Mn (blue dot), percentage of TFMST in the remaining monomer (mol%: red dot), and degree of Đ (green dot) were plotted against reaction times. When the monomer ratio of ST to TFMST was 90:10 (Figure 1a), the molecular weight increased over time, and the remaining percentage of TFMST remained constant. Đ was large in the early stages of the reaction but finally reached 1.2. When the monomer ratio is 68:32 (Figure 1b), the percentage of TFMST remaining in the monomer increases with time. This means that ST is preferentially consumed, and the monomer ratio changes during the reaction. As the ratio of TFMST increases during the reaction, the reaction rate decreases, and more TFMST is incorporated in the later stages of the reaction than in the earlier stages, which may result in a gradient of TFMST contents in the polymer chain.
The copolymerization of TFMST and the substituted ST were investigated (Table 3). Substituted ST and TFMST reacted in a ratio of 50:50 for 6 h, and a copolymer was obtained in all cases. The p-methoxystyrene-TFMST copolymer had a broad dispersity (entry one), whereas the p-trifluoromethylstyrene copolymer had a narrow dispersity, although the yield was low (entry two). When pentafluorostyrene was used, polymers with higher Td5 and Tg values than polystyrene were obtained (entry three).
Then, the copolymerization of substituted TFMST and ST was investigated (Table 4). TFMSTs with methyl, methoxy, and trifluoromethyl at the para position were synthesized according to protocols in the literature [37,38] (Scheme S1). As a result, copolymers of both monomers with electron-donating (entries one and two) and electron-withdrawing substituents (entry three) were obtained. Compared with the copolymers of unsubstituted TFMST and ST, Td5 did not change substantially but showed a higher Tg in all cases, especially for the copolymer with the pentafluoro group 5da.
Finally, we measured the water and oil contact angles of the synthesized copolymers (Figure 2, Tables S4–S6). A copolymer ST and TFMST (66:34) had a water contact angle of 97°. This value is smaller than that reported by Ameduri [13]; however, the difference may be because their polymer was synthesized by ITP and has a perfluorohexyl group at the polymer end. The oil repellency measured using dodecane as oil was 8°. Copolymers with different polymer ends (Table S4) and different ratios of ST to TFMST (Table S5) had no substantial effect on water and oil repellency, even with increasing fluorine content. The substituted copolymers were also measured (Table S6, Figure 2). The results for 5aa, 5da with a fluorine functional group on the ST monomer, and 5ad with a trifluoromethyl group on the TFMST monomer were also shown (Figure 2). As a result, a slight improvement in water and oil repellency was observed when pentafluorophenyl groups were present [39].

3. Materials and Methods

3.1. General Information

3.1.1. Materials

ST (99%) was purchased from Yoneyama Yakuhin Kogyo Co., Ltd. (Osaka, Japan). The chemicals 4-methoxystyrene (>98%), 4-(trifluoromethyl)styrene (>98%), 2-bromo-3,3,3-trifluoropropene (>97%), phenylbronic acid (>97%), 4-methylphenylbronic acid (>97%), 4-methoxyphenylbronic acid (>97%), 4-(trifluoromethyl)phenylbronic acid (>97%), and benzotrifluoride (>98%) were purchased from TCI (Mountain View, CA, USA). Chemicals such as 2,3,4,5,6-pentafluorostyrene (+97%) and AIBN were purchased from Wako (Richmond, VA, USA), while α-methylstyrene (98%) was purchased from Nacalai tesque (Kyoto, Iapan). N-tert-butyl-N-(2-methyl-1-phenylpropyl)-O-(1-phenylethyl)hydroxylamine (>98%) was purchased from Santa Cruz Biotechnology, Inc. (Dallas, TX, USA), and stored in the dark at −28 °C. BlocBuilder MA was supplied from Arkema (Colombes, France) and kept in the dark at 0 °C. Potassium carbonate (K2CO3) was purchased from Takahashi Pure Chemical Co. (Yokohama, Japan). Bis(triphenylphosphine)palladium(II) dichloride (PdCl2(PPh3)2) (97%) and ammonium chloride (NH4Cl) (99.5%) were purchased from Kanto Chemical Co., Inc. (Tokyo, Japan). The styrene monomers of 4-methoxystyrene, 4-(trifluoromethyl)styrene, and 2,3,4,5,6-pentafluorostyrene were used after removal of the polymerization inhibitor.

3.1.2. General Measurements

1H NMR and 19F NMR were measured on a JEOL GSX-400 spectrometer (JEOL, Akishima, Japan, 400 MHz for 1H, 101 MHz for 13C, and 376 MHz for 19F) and a JEOL ECX-500 spectrometer (JEOL, Akishima, Japan, 500 MHz for 1H, 126 MHz for 13C, and 471 MHz for 19F). All chemical shifts are reported as n parts-per-million downfield of the standards. Tetramethylsilane 1H 0.0 was used as the internal standard for chloroform-d solutions. The values of the number-average molecular weight (Mn) and polydispersity (Mw/Mn) were estimated by gel permeation chromatography in THF calibrated with polystyrene as standard, operating at 1.0 mL/min with an L-2490 refractive index detector, two Showdex KF-803L columns (8.0 × 300 mm), and KF-806L (8.0 × 300 mm) at 40 °C. THF was purchased from Kanto Chemical Co., Inc.

3.1.3. Thermal Measurements

Thermogravimetric analysis of the polymers was performed on a BRUKER TG-DTA2000SA (Bruker, Billerica, MA, USA) under nitrogen conditions at 10 °C min−1 from 0 °C to 500 °C. DSC was measured on a BRUKER DSC100SA under nitrogen conditions at 10 °C min−1 from −30 °C to 250 °C. The procedure of thermogravimetric (TG) (NETZSCH STA2500 Regulus, Netzsch, Selb, Germany) coupled with gas chromatography–mass spectrometry (GC/MS) was used to carry out pyrolysis/combustion experiments. In total, a 1.3 mg sample was pyrolyzed/combusted in the TG at temperatures between 30 °C and 600 °C with a heating rate of 10 °C/min. N2 gas with a flow rate of 100 mL/min was used as the carrier gas to create an inert atmosphere in the pyrolysis process. The evolved gases at a target temperature (300 °C for an inert atmosphere that corresponds to the maximum mass loss rate) were sampled via an autoinjector system that was connected to the TG system. The remainder of the gas was purged to the atmosphere. All transfer lines of the autoinjector were maintained at 300 °C to avoid any condensation of volatile gases whose boiling temperature was less than 300 °C. Gas analysis was carried out with gas chromatography–mass spectrometry (GC/MS) (Agilent 8890 GC system JMS-Q1600GC (Agilent, Santa Clara, CA, USA) for GC, JMS-Q1500GC Master-Quad GC/MS (JEOL, Akishima, Japan) for MS). The GC instrument was equipped with HP-5MS (5%-phenyl)-methylpolysiloxane nonpolar (15 m length, 0.25 mm I.D., and 0.25 μm film) and an HP-MOLESIEVE (30 m length, 0.53 mm I.D) column. The initial oven temperature of 40 °C was kept isothermal for 3 min, then heated to 600 °C at a rate of 10 °C/min, and held at 600 °C for 5 min. The temperatures of the MS source and MS quad were 300 and 150 °C, respectively. The detector consists of a mass selective detector, and electron impact mass spectra were acquired with 70 eV ionizing energy with a scanning range from 10 to 1000 Da and with a scan rate of 1 scan s−1. The MS transfer line and the ion source temperature were maintained at 300 °C.

3.1.4. Water and Oil Repellency Measurements

The polymer film obtained for the contact angle assessments was formed using a spin coater (Mikasa Opticoat MS-A100 model, Tokyo, Japan). A solution of 100 mg of the gained polymer in 270 µL of toluene (30 wt%) was placed on a 70 × 40 mm slide glass. The substrate surface was covered with the dissolved polymer using a pipette, followed by spinning at 100–400 rpm for 20 s to spread and form a uniform thin film. The film sample was then heated at 110 °C for 6 min and left overnight. The contact angle of the film sample was measured using a Nick LSE-ME5 (Nick, Saitama, Japan) contact angle measurement device. Measurements were performed at least three different sites, and the final value was averaged. A drop of water (5 μL) or dodecane (6 μL) was deposited using a microsyringe. The angles were determined using the Laplace–Young equation approximately 1 s after the drop deposition.

3.2. General Procedure for Synthesis of Substituted TFMST

To a 100 mL flask equipped with a magnetic stir bar, arylbronic acid (8 mmol), 2-bromo-3,3,3-trifluoropropene (1.65 mL, 16 mmol), dry THF (24 mL), K2CO3 (2.0 M, 16 mL), and PdCl2(PPh3)2 (168 mg, 0.24 mmol) was added. The mixture was frozen–degassed with liquid nitrogen and vacuum evacuation (×3). After dissolution, the reaction vessel was injected with argon. The resulting solution was stirred at 65 °C for 24 h. After the reaction mixture was cooled to room temperature, the mixture was quenched with saturated aqueous NH4Cl, extracted with Et2O, dried over Na2SO4, filtered, and concentrated under pressure. After removing the solvent, the crude product was purified by distillation under reduced pressure (pressure: 20 mbar, bath temperature: 90 °C) to obtain a colorless oil. Compounds were identified according to the literature [36,37].

3.3. General Procedure for Copolymerization

3.3.1. ATRP Polymerization

To a 10 mL flask equipped with a magnetic stir bar, 3a (18.5 mg, 0.1 mmol), Cu-TMEDA (46.4 mg, 0.1 mmol), and monomers (styrenes and TFMSTs, 10 mmol) were added. The mixture was frozen–degassed with liquid nitrogen and vacuum evacuation (×3). After dissolution, the reaction vessel was injected with Ar. The resulting solution was stirred at 110 °C for 12 h. After cooling the reaction mixture to room temperature, the mixture was dissolved in a small amount of CH2Cl2 and precipitated in cold hexane. The resulting polymer was dried under a vacuum.

3.3.2. RAFT Polymerization

To a 10 mL flask equipped with a magnetic stir bar, 3b (34.6 mg, 0.1 mmol), AIBN (4.1 mg, 0.025 mmol) in 1,4-dioxane (2.5 mL), and monomers (styrenes and TFMSTs, 10 mmol) were added. The mixture was frozen–degassed with liquid nitrogen and vacuum evacuation (×3). After dissolution, the reaction vessel was injected with Ar. The resulting solution was stirred at 75 °C. After cooling the reaction mixture to room temperature, the mixture was dissolved in a small amount of CH2Cl2 and precipitated in cold hexane. The resulting polymer was dried under a vacuum.

3.3.3. NMP Polymerization

BlocBuilder MA (3d) (38.1 mg, 0.1 mmol) and monomers (ST and TFMSTs, 10 mmol) were added to a 10 mL flask equipped with a magnetic stir bar. The mixture was frozen–degassed with liquid nitrogen and vacuum evacuation (×3). After dissolution, the reaction vessel was injected with Ar. The resulting solution was stirred at 110 °C for 6 h. After cooling the reaction mixture to room temperature, the mixture was dissolved in a small amount of CH2Cl2 and precipitated in cold hexane. The resulting polymer was dried under a vacuum.

4. Conclusions

In this study, controlled copolymerization of TFMST and ST was investigated. We found that NMP using BlocBuilder MA (3d) as the initiator yielded the TFMST–ST copolymer in low dispersity (Đ). The content ratio of TFMST can be controlled between 10% and 40% by changing its monomer ratio. Additionally, TFMST and ST with their substituents were polymerized to obtain novel copolymers. Thermal measurements of the synthesized copolymers were performed. The resulting copolymers exhibited lower Td5 than polystyrene, except for those containing 10% TFMST. All copolymers exhibited a higher Tg than polystyrene, especially for the copolymer with the pentafluoro group 5da. Moreover, a slight improvement in the water and oil contact angle was observed with the copolymer with pentafluoro groups of 5da.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules29061214/s1, Scheme S1: synthesis of monomers; Figure S1: example of 19F NMR spectra (CDCl3) of synthesized polymer; Table S1: the data used to determine the molar composition of the polymers; Figure S2: TFMST ratio in monomer vs. polymer; Table S2: time course of polymerization using 90 mol% of ST; table S3: time course of polymerization using 68 mol% of ST; Figure S3: TG curves with different TFMST content; Figure S4: TG curves with different substituent; Figure S5: TG-GC-MS in mass spectrum; Table S4: static contact angle measurements of polymers with different polymerization methods; Table S5: static contact angle measurements of polymers with different ratios of TFMST; Table S6: Static contact angle measurements of polymers with different Ar1 or Ar2. Refs. [40,41,42] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, T.K. and T.Y.; methodology, validation and investigation, Y.I. and A.Y.; writing—original draft preparation, T.Y.; writing—review and editing, T.Y.; visualization, T.Y.; supervision, T.Y.; project administration, T.Y.; funding acquisition, T.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a Grant-in-Aid for Transformative Research Areas (A), MEXT, Japan, grant number JP21H05219.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the Supplementary Materials.

Acknowledgments

We are also grateful to Arkema inc. for their generous supply of BlocBuilder MA.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Teng, H. Overview of the development of the fluoropolymer industry. Appl. Sci. 2012, 2, 496–512. [Google Scholar] [CrossRef]
  2. Ameduri, B. Frontispiece: Fluoropolymers: The right material for the right applications. Chem. Eur. J. 2018, 23, 18830–18841. [Google Scholar] [CrossRef] [PubMed]
  3. Puts, G.J.; Crouse, P.; Ameduri, B. Polytetrafluoroethylene: Synthesis and characterization of the original extreme polymer. Chem. Rev. 2019, 119, 1763–1805. [Google Scholar] [CrossRef] [PubMed]
  4. Ameduri, B.; Boutevin, B. Well-Architectured Fluoropolymers: Synthesis, Properties and Applications, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2004. [Google Scholar]
  5. Hercules, D.A.; Parrish, C.A.; Sayler, T.S.; Tice, K.T.; Williams, S.M.; Lowery, L.E.; Brady, M.E.; Coward, R.B.; Murphy, J.A.; Hey, T.A.; et al. Preparation of tetrafluoroethylene from the pyrolysis of pentafluoropropionate salts. J. Fluor. Chem. 2017, 196, 107–116. [Google Scholar] [CrossRef]
  6. Ferrero, F.; Meyer, R.; Kluge, M.; Schroeder, V.; Spoormaker, T. Study of the spontaneous ignition of stoichiometric tetrafluoroethylene-air mixtures at elevated pressures. J. Loss Prev. Proc. Ind. 2013, 26, 759–765. [Google Scholar] [CrossRef]
  7. Tervoort, T.A.; Visjager, J.F.; Smith, P. Melt-processable poly(tetrafluoroethylene)—Compounding, fillers and dyes. J. Fluor. Chem. 2002, 114, 133–137. [Google Scholar] [CrossRef]
  8. Li, M.; Zhang, W.; Wang, C.; Wang, H. Melt processability of polytetrafluoroethylene: Effect of melt treatment on tensile deformation mechanism. J. Appl. Polym. Sci. 2012, 123, 1667–1674. [Google Scholar] [CrossRef]
  9. Jaye, J.A.; Sletten, E.M. Recent advances in the preparation of semifluorinated polymers. Polym. Chem. 2021, 12, 6515–6526. [Google Scholar] [CrossRef]
  10. Friesen, C.M.; Ameduri, B. Outstanding telechelic perfluoropolyalkylethers and applications therefrom. Prog. Polym. Sci. 2018, 81, 238–280. [Google Scholar] [CrossRef]
  11. He, E.; Tum, K.; Cheng, J.; Wang, Y.; Lu, H.; Zhang, L.; Cheng, Z. Synthesis and phase behavior of (semifluorinated alkane)-based side-chain liquid crystalline copolymers. Macromol. Rapid Commun. 2022, 43, 2200266. [Google Scholar] [CrossRef]
  12. Kanbara, T.; Arase, M.; Tanaka, M.; Yamaguchi, A.; Tagami, K.; Yajima, T. Amine-catalyzed synthesis of fluorine-containing polymers through halogen bonding. Chem. Asian J. 2023, 8, e202300035. [Google Scholar] [CrossRef] [PubMed]
  13. Ameduri, B. Controlled radical (co)polymerization of fluoromonomers. Macromolecules 2010, 43, 10163–10184. [Google Scholar] [CrossRef]
  14. Gong, H.; Gu, Y.; Chen, M. Controlled/living radical polymerization of semifluorinated (meth)acrylates. Synlett 2018, 29, 1543–1551. [Google Scholar]
  15. Yao, W.; Li, Y.; Huang, X. Fluorinated poly(meth)acrylate: Synthesis and properties. Polymer 2014, 55, 6197–6211. [Google Scholar] [CrossRef]
  16. Borkar, S.; Jankova, K.; Siesler, H.W.; Hvilsted, S. New highly fluorinated styrene-based materials with low surface energy prepared by ATRP. Macromolecules 2004, 37, 788–794. [Google Scholar] [CrossRef]
  17. Ceretta, F.; Zaggia, A.; Conte, L.; Ameduri, B. Conventional radical polymerization and iodine-transfer polymerization of 4′-nonafluorobutyl styrene: Surface and thermal characterizations of the resulting poly(fluorostyrene)s. J. Ploym. Sci. A Polym. Chem. 2013, 51, 3203–3212. [Google Scholar] [CrossRef]
  18. Iwatsuki, S.; Itoh, T.; Iida, T. Absolute rate constants for radical homopolymerization of methyl α-(trifluoromethyl) acrylate. J. Polym. Sci. Part A Poym. Chem. 1994, 32, 1389–1392. [Google Scholar] [CrossRef]
  19. Agliettot, M.; Passaglia, E.; Mirabello, L.M. Synthesis of new polymers containing α-(trifluoromethy1) acrylate units. Macromol. Chem. Phys. 1995, 196, 2843–2853. [Google Scholar] [CrossRef]
  20. Ueda, M.; Ito, H. Radical reactivity of α-trifluoromethylstyrene. Polym. Chem. 1998, 26, 89–98. [Google Scholar] [CrossRef]
  21. Walkowiak-Kulikowska, J.; Szwajca, A.; Boschet, F.; Gouverneur, V.; Ameduri, B. Poly(vinylidene fluoride)-b-poly(styrene) block copolymers by Iodine transfer polymerization (ITP): Synthesis, characterization, and kinetics of ITP. Macromolecules 2014, 47, 8634–8644. [Google Scholar] [CrossRef]
  22. Walkowiak-Kulikowska, J.; Szwajca, A.; Gouverneur, V.; Ameduri, B. Synthesis, characterization, and hermal and surface properties of co- and terpolymers based on fluorinated α-methylstyrenes and styrene. Polym. Chem. 2017, 8, 6558–6569. [Google Scholar] [CrossRef]
  23. Kamigaito, M.; Ando, T.; Sawamoto, M. Metal-catalyzed living radical polymerization. Chem. Rev. 2001, 101, 3689–3731. [Google Scholar] [CrossRef] [PubMed]
  24. Matyjaszewski, K.; Xia, J.H. Atom transfer radical polymerization. Chem. Rev. 2001, 101, 2921–2987. [Google Scholar] [CrossRef] [PubMed]
  25. Tsarevsky, N.V.; Matyjaszewski, K. “Green” atom transfer radical polymerization: From process design to preparation of well-defined environmentally friendly polymeric materials. Chem. Rev. 2007, 107, 2270–2299. [Google Scholar] [CrossRef] [PubMed]
  26. Keddie, D.J.; Moad, G.; Rizzardo, E.; Thang, H. RAFT agent design and synthesis. Macromolecules 2012, 45, 5321–5343. [Google Scholar] [CrossRef]
  27. Favier, A.; Charreyre, M.T. Experimental requirements for an efficient control of free-radical polymerizations via the reversible addition-fragmentation chain transfer (RAFT) process. Macromol. Rapid Commun. 2006, 27, 665–689. [Google Scholar] [CrossRef]
  28. Moad, G.; Rizzardo, E.; Thang, H. Living radical polymerization by the RAFT process—Second update. Aust. J. Chem. 2009, 62, 1402–1472. [Google Scholar] [CrossRef]
  29. Nicolas, J.; Guillaneuf, Y.; Lefay, C.; Bertin, D.; Gigmes, D.; Charleux, B. Nitroxide-mediated polymerization. Prog. Polym. Sci. 2013, 38, 63–235. [Google Scholar] [CrossRef]
  30. Hawker, C.J.; Bosman, A.W.; Harth, E. New polymer synthesis by nitroxide-mediated living radical polymerizations. Chem. Rev. 2001, 101, 3661–3698. [Google Scholar] [CrossRef]
  31. Sciannamea, V.; Jerome, C.; Detrembleur, C. In-situ nitroxide-mediated radical polymerization (NMP) processes: Their understanding and optimization. Chem. Rev. 2008, 108, 1104–1126. [Google Scholar] [CrossRef]
  32. Xia, J.; Matyjaszewski, K. Controlled/”living” radical polymerization. Atom transfer radical polymerization using multidentate amine ligands. Macromolecules 1997, 30, 7697–7700. [Google Scholar] [CrossRef]
  33. Barlow, K.J.; Hao, X.; Hughes, T.C.; Hutt, O.E.; Polyzos, A.; Turner, K.A.; Moad, G. Porous, functional, poly(styrene-co-divinylbenzene) monoliths by RAFT polymerization. Polym. Chem. 2014, 5, 722–732. [Google Scholar] [CrossRef]
  34. Benoit, D.; Chaplinski, V.; Braslau, R.; Hawker, C.J. Development of a universal alkoxyamine for “living” free radical polymerizations. J. Am. Chem. Soc. 1999, 121, 3904–3920. [Google Scholar] [CrossRef]
  35. Banerjee, S.; Domenichelli, I.; Ameduri, B. Nitroxide-mediated alternating copolymerization of vinyl acetate with tert-butyl-2-trifluoromethacrylate using a SG1-based alkoxyamine. ACS Macro Lett. 2016, 5, 1232–1236. [Google Scholar] [CrossRef] [PubMed]
  36. Langlais, M.; Coutelier, O.; Destarac, M. Thiolacetone-functional reversible deactivation radical polymerization agents for advanced macromolecular engineering. Macromolecules 2018, 51, 4315–4324. [Google Scholar] [CrossRef]
  37. Liu, Y.; Zhou, Y.; Zhao, Y.; Qu, J. Synthesis of gem-difluoroallylboronates via FeCl2-catalyzed boration/β-fluorine elimination of trifluoromethyl alkenes. Org. Lett. 2017, 19, 946–949. [Google Scholar] [CrossRef] [PubMed]
  38. Walkowiak-Kulikowska, J.; Boschet, F.; Kostov, G.; Gouverneur, V.; Ameduri, B. On the reactivity of α-trifluoromethylstyrene in radical copolymerizations with various fluoroalkenes. Eur. Polym. J. 2016, 84, 612–621. [Google Scholar] [CrossRef]
  39. Lin, J.W.P.; Dudek, L.P.; Majumdar, D. Wetting properties of homopolymers and copolymers of pentafluorostyrene and methacrylate and homopolymer blends. J. Appl. Polym. Sci. 1987, 33, 657–667. [Google Scholar] [CrossRef]
  40. Pan, R.; Liu, X.; Deng, M. A novel and convenient synthetic method for producing α-(trifluoromethyl)styrenes. J. Fluor. Chem. 1999, 95, 167–170. [Google Scholar] [CrossRef]
  41. Miura, T.; Ito, Y.; Murakami, M. Synthesis of gem-difluoroalkenes via β-fluoride elimination of organorhodium(I). Chem. Lett. 2008, 37, 1006–1007. [Google Scholar] [CrossRef]
  42. Nader, B.S.; Cordova, J.A.; Reese, K.E.; Powell, C.L. A novel fluoride mediated olefination of electron-deficient aryl ketones by alkanesulfonyl halides. J. Org. Chem. 1994, 59, 2898–2901. [Google Scholar] [CrossRef]
Figure 1. Plots of Mn (down; blue dot), residual amount of TFMST (up; red dot (mol%)), and Mw/Mn (up; green dot) against reaction times; (a) ST to TFMST ratio of 90:10; (b) ST to TFMST ratio of 68:32.
Figure 1. Plots of Mn (down; blue dot), residual amount of TFMST (up; red dot (mol%)), and Mw/Mn (up; green dot) against reaction times; (a) ST to TFMST ratio of 90:10; (b) ST to TFMST ratio of 68:32.
Molecules 29 01214 g001
Figure 2. Water and oil contact angles of copolymers 5aa, 5da, and 5ad.
Figure 2. Water and oil contact angles of copolymers 5aa, 5da, and 5ad.
Molecules 29 01214 g002
Table 1. Reactions using different polymerization methods.
Table 1. Reactions using different polymerization methods.
Molecules 29 01214 i001
EntryReagentsTemp.
[°C]
Time
[h]
Yield (a)
[%]
Ratio of ST (b)
[%]
Ratio of TFMST (b)
[%]
Mn (c)Mw/Mn (c)
1 (e)3a (1 eq.), Cu-TMEDA (1 eq.)110126291910,0001.30
2 (f)3b (1 eq.), AIBN (0.25 eq.)
1,4-dioxane
754828 (d)901022001.27
3 (f)3b (1 eq.), AIBN (0.25 eq.)
1,4-dioxane
759621871345001.10
4 (g)3c (1 eq.)110616901028001.18
5 (g)3d (1 eq.)11065691973001.14
(a) Cold hexane-insoluble part; (b) determined by 19F NMR using BTF as an internal standard after purification; (c) determined by gel permeation chromatography in THF based on linear polystylene as calibration standard; (d) concentrated residue; (e) the mixture of 1a (9 mmol), 2a (1 mmol), 3a (0.1 mmol), and Cu-TMEDA (0.1 mmol) under an Ar atmosphere was stirred at 110 °C for 12 h; (f) the mixture of 1a (9 mmol), 2a (1 mmol), 3b (0.1 mmol), and AIBN (0.025 mmol) in 1,4-dioxane (2.5 mL) under an Ar atmosphere was stirred at 75 °C for 48 or 96 h; (g) the mixture of 1a (9 mmol), 2a (1 mmol), and 3c or 3d (0.1 mmol) under an Ar atmosphere was stirred at 110 °C for 6 h.
Table 2. Reaction using different ratios of ST vs. TFMST or MST (a).
Table 2. Reaction using different ratios of ST vs. TFMST or MST (a).
Molecules 29 01214 i002
EntryST (1a)
[eq.]
2a or 4
[eq.]
Yield (a)
[%]
Ratio of ST
[%]
Ratio of (TF)MST
[%]
Mn (b)Mw/Mn (b)Td5%
[°C]
Tg
[°C]
1100-65100-83001.0727571
2902a (10)56919 (c)73001.14285103
3702a (30)548218 (c)83001.31279107
4502a (50)326634 (c)74001.3026180
5302a (70)206040 (c)61001.4726191
6 (d)904 (10)47100071001.11--
7 (e)504 (50)277129 (f)75001.3427865
(a) Cold hexane-insoluble part; (b) determined by gel permeation chromatography in THF based on linear polystyrene as calibration standard; (c) determined by 19F NMR using BTF as an internal standard after purification; (d) 8 h; (e) 22 h; (f) determined by 1H NMR.
Table 3. Reaction using substituted ST (a).
Table 3. Reaction using substituted ST (a).
Molecules 29 01214 i003
EntrySTYield (b)
[%]
CopolymerRatio of ST (c)
[%]
Ratio of TFMST (c)
[%]
Mn (d)Mw/MnTd5%
[°C]
Tg
[°C]
11b265ba693158001.50252100
21c195ca772347001.1823894
31d355da703093001.49293130
(a) The mixture of 1b, c, or d (5 mmol), 2a (5 mmol), and 3c (0.1 mmol) under an Ar atmosphere was stirred at 110 °C for 6 h; (b) cold hexane-insoluble part; (c) determined by 19F NMR using BTF as an internal standard after purification; (d) determined by gel permeation chromatography in THF based on linear polystyrene as calibration standard.
Table 4. Reaction using substituted MST (a).
Table 4. Reaction using substituted MST (a).
Molecules 29 01214 i004
EntryTFMSTYield (b)
[%]
CopolymerRatio of ST (c)
[%]
Ratio of TFMST (c)
[%]
Mn (d)Mw/Mn (d)Td5%
[°C]
Tg
[°C]
12b205ab732786001.24251107
22c335ac683261001.4725593
32d465ad703010,0001.37278110
(a) The mixture of 1a (5 mmol), 2b, c, or d (5 mmol), and 3c (0.1 mmol) under an Ar atmosphere was stirred at 110 °C for 17 h; (b) cold hexane-insoluble part; (c) determined by 19F NMR using BTF as an internal standard after purification; (d) determined by gel permeation chromatography in THF based on linear PSt as calibration standard.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kanbara, T.; Ito, Y.; Yamaguchi, A.; Yajima, T. Nitroxide-Mediated Controlled Radical Copolymerization of α-Trifluoromethylstyrenes with Styrenes. Molecules 2024, 29, 1214. https://doi.org/10.3390/molecules29061214

AMA Style

Kanbara T, Ito Y, Yamaguchi A, Yajima T. Nitroxide-Mediated Controlled Radical Copolymerization of α-Trifluoromethylstyrenes with Styrenes. Molecules. 2024; 29(6):1214. https://doi.org/10.3390/molecules29061214

Chicago/Turabian Style

Kanbara, Tadashi, Yuriko Ito, Airi Yamaguchi, and Tomoko Yajima. 2024. "Nitroxide-Mediated Controlled Radical Copolymerization of α-Trifluoromethylstyrenes with Styrenes" Molecules 29, no. 6: 1214. https://doi.org/10.3390/molecules29061214

Article Metrics

Back to TopTop