Amide group-containing polar solvents as ligands for iron-catalyzed atom transfer radical polymerization of methyl methacrylate

Jun Zhoua, Jirong Wanga, Jianyu Hana, Dan Heb, Danfeng Yanga, Zhigang Xue*a, Yonggui Liaoa and Xiaolin Xie*a
aKey Laboratory for Large-Format Battery Materials and Systems, Ministry of Education, School of Chemistry and Chemical Engineering, Huazhong University of Science and Technology, Wuhan 430074, China. E-mail: zgxue@mail.hust.edu.cn; xlxie@mail.hust.edu.cn; Fax: +86 27 87543632; Tel: +86 27 87793241
bKey Laboratory of Optoelectronic Chemical Materials and Devices of Ministry of Education, School of Chemical and Environmental Engineering, Jianghan University, Wuhan 430056, China

Received 27th March 2015 , Accepted 17th April 2015

First published on 17th April 2015


Abstract

A series of amide group-containing polar solvents, formamide (Fo), N-methylformamide (MFo), N,N-dimethylformamide (DMF), acetamide (Ac), N-methylacetamide (MAc), N,N-dimethylacetamide (DMAc), urea, tetramethyl urea (TMU), 2-pyrrolidone (2-Py), N-methyl-2-pyrrolidone (NMP) and 5-methyl-2-pyrrolidone (MPy), were used as both solvents and ligands for iron(II)-catalyzed atom transfer radical polymerizations (ATRPs) of methyl methacrylate (MMA), with ethyl 2-bromo-2-phenylacetate (EBPA) as the initiator. Most of the polymerizations were well-controlled in character, and the structures of the polar solvents greatly affected the catalytic activity. In addition, the living features of the systems remained in the presence of limited amounts of polar solvents. Some of the polar solvents (MFo, TMU and 2-Py) were also employed for iron(III)-catalyzed activators generated by electron transfer (AGET) ATRPs of MMA, and the results were as good as those of the ATRPs.


Introduction

Metal-catalyzed controlled radical polymerization (CRP), also known as atom transfer radical polymerization (ATRP), has been a versatile tool for synthesizing well-defined polymers1–15 with controlled molecular weights (Mn) and narrow molecular weight distributions (MWDs, Mw/Mn) since it was first presented in 1995.16–18 The catalyst used in ATRP is considered to be the most essential and important issue for controlling a polymerization,6,7,14,15 and a great deal of effort has thus been made to research the effects of various complexes on polymerizations, such as those based on copper,10,11,19–27 ruthenium2,6,28–35 and iron.14,15,36–57

On account of the environmental and sustainability aspects, the iron-based catalysts are one of the most promising catalysts for ATRP among those catalysts. Firstly, iron exists widely in the Earth’s crust as an extremely abundant metal, which can be obtained easily at a low price. Secondly, iron has low toxicity and is biocompatible, properties which are significant for industrial applications.2,6,7 What’s more, the oxidation states that usually exist, iron(II) and iron(III), have empty orbitals to coordinate to many ligands, and form active complexes used in CRP on a large scale. Compared with copper-based catalysts, iron complexes used in ATRP have generally been less active. However, the development of environmentally benign and sustainable iron-based systems is becoming a trend of “green” chemistry or “green” reactions in the applications of organic chemistry,58–61 polymer chemistry6,7,14,15,62 and electrochemistry.63–68 Various ligands have been applied for complexing with iron salts to form iron-based catalysts,14,15 such as the traditional phosphine (P), nitrogen (N) or P–N ligands. Notably, some organic acid- and salt-containing ligands have also been developed for iron-catalyzed ATRP.15 Although these catalysts are active for ATRP, more attention needs to be paid to the cost and separation of these iron complexes for potential practical applications of the ATRP technique.

Recently, FeBr2-catalyzed ATRPs of methyl methacrylate (MMA) in polar solvents were reported by Matyjaszewski et al.69,70 The resultant polymers had molecular weights which agreed with the theoretical values, and the MWDs remained low (Mw/Mn < 1.3) when conducted in N-methyl-2-pyrrolidone (NMP), N,N-dimethylformamide (DMF) and acetonitrile (MeCN), but the polymers obtained in dimethyl sulfoxide (DMSO) had poor controllability (Mw/Mn > 2.0). The experimental results suggested that certain polar solvents act as ligands for iron species due to the coordination ability of the solvents with FeBr2. Actually, polar solvents which have lone pairs of electrons (nitrogen, oxygen, or phosphorus atoms) can also coordinate with organoiron complexes, and their coordination and catalytic properties have been studied in our previous work.71,72 To study the function of polar solvents as ligands for conducting CRP, Bulgakova et al.73 and Xue et al.74 reported the activators generated by electron transfer (AGET) ATRP of MMA using polar solvents (DMF, NMP or MeCN) in the absence of additional ligands. Most of the polymerizations showed the typical characteristics of “living”/controlled radical polymerization. In addition, we have also used alcohols, such as methanol, ethanol, ethylene glycol (EG) and glycerol, as reducing agents for the iron-catalyzed AGET ATRP of MMA in the presence of polar solvents as ligands.75 Very recently, Zhu and coworkers76 also reported the iron-catalyzed AGET ATRP of MMA using polyethylene glycol 400 (PEG-400) as both a solvent and a ligand.

The polymerizations were well controlled even when the concentration of PEG-400 was reduced to a catalytic amount. This system is much more environmentally friendly when compared with most traditional iron-catalyzed ATRP systems, due to the “green” nature of PEG-400. Compared to the conventional iron-based catalytic systems, it is particularly significant that the polymerizations were found to be more environmentally friendly ATRPs in the presence of polar solvents as ligands.

In considering the role of polar solvents, which act as ligands in the iron-catalyzed ATRPs mentioned above, they have electron donor groups, such as the oxygen atom in DMF or NMP, and all of them are able to coordinate to iron(II) ([Ar]3d64s0) or iron(III) ([Ar]3d54s0) which have empty orbitals. We expected that the coordination ability of polar solvents to iron salts and their universality would enable them to be applied in iron-catalyzed ATRP, and to be appropriate substitutes for traditional P ligands or N ligands. In this article, combining the various advantages of iron catalysts and polar solvents, the iron-catalyzed ATRP of MMA (Scheme 1) was investigated using polar solvents with amide groups as ligands. These amide group-containing ligands can be classified by their structures as polar solvents based on formamide, polar solvents based on urea or polar solvents based on pyrrolidone (see Scheme 2). They all have the ability to coordinate with iron salts which depend on their structures. The majority of them have been used for ATRP for the first time in this work. As expected, some of them resulted in good controllability of molecular weights and MWDs. In addition, the different structures of the polar solvents greatly affected their catalytic activities, which is significant for the optimization of polar solvents for ATRP applications. In order to ease the process of separating the solvents from the polymerization system for the purpose of environmental protection, we also reduced the concentrations of the polar solvents to catalytic amounts, and ATRP was still feasible. These amide group-containing polar solvents were also employed in iron(III)-catalyzed AGET ATRP (see Scheme 1), and most of the systems showed controlled features.


image file: c5ra05460e-s1.tif
Scheme 1 Mechanism of iron-catalyzed ATRP using amide group-containing polar solvents as ligands.

image file: c5ra05460e-s2.tif
Scheme 2 Structures of amide group-containing polar solvents.

Experimental

Materials

Methyl methacrylate (MMA, 98+%, Sinpharm) was passed through a column filled with neutral alumina, dried over calcium hydride (CaH2), distilled under reduced pressure and stored in a freezer under argon. Formamide (Fo, 99+%, Sinpharm), N,N-dimethylformamide (DMF, 99.5+%, Sinpharm), N-methylacetamide (MAc, 99+%, Sinpharm), N,N-dimethylacetamide (DMAc, 99+%, Sinpharm), and N-methyl-2-pyrrolidone (NMP, 95+%, Sinpharm) were dried over CaH2 and distilled under reduced pressure. N-Methylformamide (MFo, 99+%, Acros), acetamide (Ac, 98+%, Sinpharm), 2-pyrrolidone (2-Py, 99+%, Xiya), 5-methyl-2-pyrrolidone (MPy, 98+%, TCI), urea (99+%, Sinpharm), tetramethyl urea (TMU, 98+%, TCI), ethyl 2-bromo-2-phenylacetate (EBPA, 95%, Alfa Aesar), iron(II) bromide (FeBr2, 98+%, Alfa Aesar), iron(III) bromide (FeBr3, 98+%, Alfa Aesar), vitamin C (VC, 99+%, Sinpharm), sodium carbonate (Na2CO3, Sinpharm) and ethylene glycol (EG, 99+%, Sinpharm) were used without further purification.

Polymerization procedures

A typical polymerization procedure with the molar ratio of [MMA]0/[EBPA]0/[FeBr2]0/[solvent]0/[Na2CO3]0 = 200[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]20[thin space (1/6-em)]:[thin space (1/6-em)]2 is as follows. A Schlenk flask (25 mL) was charged with FeBr2 (61 mg, 0.28 mmol) and Na2CO3 (60 mg, 0.57 mmol). The flask was sealed with a rubber septum and was cycled three times between vacuum and argon (Ar) to remove oxygen. Degassed polar solvents and MMA (6 mL, 56.3 mmol) were then added to the flask through degassed syringes. The solution was stirred for 30 min at room temperature. After three additional freeze–pump–thaw cycles, the initiator, EBPA (49.5 μL, 28.2 mmol) was added, and the flask was immersed in a thermostated oil bath at 60 °C. At timed intervals, samples were withdrawn from the flask with a degassed syringe. The monomer conversion was determined gravimetrically after removal of the unconverted monomer and solvent under reduced pressure, and the resulting residue was diluted with tetrahydrofuran (THF) and then filtered through a column filled with neutral aluminum oxide to remove the iron catalyst. The poly(methyl methacrylate) (PMMA) solution was then precipitated using an excess of n-hexane, and these polymers were dried under vacuum overnight at 80 °C for gel permeation chromatography (GPC) characterization. The same experimental procedures were carried out for the iron(III)-catalyzed AGET ATRP.

Chain extension experiment

A predetermined quantity of PMMA macroinitiator (PMMA–Br) obtained by ATRP of MMA with a molar ratio of [MMA]0/[PMMA]0/[FeBr2]0/[TMU]0/[Na2CO3]0 = 500[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]50[thin space (1/6-em)]:[thin space (1/6-em)]5 was added to a Schlenk flask, and then a predetermined quantity of MMA, FeBr2, TMU and Na2CO3 was added. The rest of the procedure was the same as that described above. The chain extension polymerization was carried out under stirring at 60 °C.

Measurements

1H NMR spectroscopy was performed using a Bruker AV400 NMR spectrometer, with deuterated chloroform (CDCl3) as the solvent and tetramethylsilane (TMS) as the standard. The number-average molecular weight (Mn,GPC) and Mw/Mn of the polymers were determined with GPC using an Agilent 1100 gel permeation chromatograph, with a PLgel 79911GP-104 (7.5 × 300 mm, 10 μm bead size) column. THF was used as the eluent at a flow rate of 1 mL min−1 at 35 °C. Linear polystyrene standards were used for calibration.

Results and discussion

Some previous works74–76 have demonstrated that EBPA is the optimal initiator for the iron-catalyzed ATRP of MMA, because both the phenyl and ester groups in EBPA contribute to the stabilization of the generated radicals. Therefore, all polymerizations were conducted with EBPA as the initiator in this study. General amide group-containing polar solvents were chosen for our study as far as possible, and we classified them as polar solvents based on formamide, polar solvents based on urea or as polar solvents based on pyrrolidone by their structures.

Polar solvents based on formamide as ligands

As mentioned in the Introduction section, polar solvents (NMP, DMF, MeCN and DMSO) can coordinate to iron salts, and can be used as ligands for ATRP. As far as we know, the oxygen atom in the amide group can act as an electron donor to coordinate to an iron ion.77,78 To probe the role of the amide group as a ligand, various polar solvents based on formamide that have linear structures (Fo, MFo, DMF, Ac, MAc, and DMAc) were used as ligands for the ATRP of MMA at 60 °C, with a molar ratio of [MMA]0/[EBPA]0/[FeBr2]0 = 200[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 and [MMA]0/[polar solvent]0 = 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v) without additional ligands. As shown in Table 1, no reaction occurred in the absence of Na2CO3 for almost all of the solvents (entries 1, 3, 6, 9, and 11), except for DMF (entry 13), which was consistent with work previously reported.69 Zhu’s work48,79–81 suggested that a catalytic amount of base can significantly accelerate iron-catalyzed ATRP, and that the controllability of the polymerizations can be maintained as well. They considered that an increased pH could reduce the redox potential (E1/2) of polymerizations. The lower redox potential induced a faster reversible cleavage (activation) of a carbon–halogen terminal and, in turn, generated more radical species, which resulted in the enhancement of rate and controllability of the polymerization. Moreover, a base (Na2CO3 or NaOH) was also applied for the iron(III)-catalyzed AGET ATRP of MMA with FeX3 (X = Br or Cl)/polar solvent catalytic systems using different alcohols as reducing agents in our previous work,75 and the results revealed the exciting rate-enhancement effect of base on polymerization. Likewise, Bai et al.82 reported Cu-catalyzed AGET ATRP using alcohol as a reducing agent. Cu(II) was reduced to Cu(I), and the alcohol was oxidized to an aldehyde or ketone in the presence of base. The polar solvents studied in this article have lower coordination abilities than the traditional P or N ligands. The reactions may proceed very slowly. No polymerization conversion was even obtained at all after a long time in the absence of base (entries 1, 3, 6, 9 and 11 in Table 1). In light of the above results, base was added to our systems. As expected, most of them showed good controllability and relatively fast polymerization rates when the ratio of Na2CO3 to FeBr2 was 2. What interested us was that the resulting PMMA obtained from the DMAc system (entry 7) had a number-average molecular weight (Mn) consistent with the predicted value and a low MWD value (Mw/Mn = 1.23). In the case of the MAc system, the Mn matched the theoretical value well (entry 4). However, the MWD was a little bit high. When the polymerization was carried out in Ac or Fo (entries 2 and 10), the molecular weights were greatly bigger than the theoretical values and the polydispersities (PDIs) were high. A possible reason was the poor dissolubility of Ac or Fo in MMA.
Table 1 FeBr2-catalyzed ATRPs of MMA using polar solvents based on formamide as ligandsa
Entry Solvent [M]0/[solvent]0b Base Time (h) Conv. (%) Mn,thc (g mol−1) Mn,GPC (g mol−1) Mw/Mn
a [MMA]0/[EBPA]0/[FeBr2]0/[Na2CO3]0 = 200[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2 or 0, 60 °C.b The ratio of [M]0 to [solvent]0 is the molar ratio, except where marked with “(v/v)”, which stands for the volume ratio.c Mn,th = ([MMA]0/[initiator]0) × MMMA × conversion + MEBPA, MMMA and MEBPA represent the molecular weights of MMA and EBPA, respectively.
1 Ac (solid) 4[thin space (1/6-em)]:[thin space (1/6-em)]1   30 0 NA NA NA
2 Ac (solid) 100[thin space (1/6-em)]:[thin space (1/6-em)]1 Na2CO3 17 70 14[thin space (1/6-em)]250 23[thin space (1/6-em)]600 1.87
3 MAc 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v)   68 0 NA NA NA
4 MAc 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v) Na2CO3 29 34 7050 6900 2.04
5 MAc 10[thin space (1/6-em)]:[thin space (1/6-em)]1 Na2CO3 1.5 56 11[thin space (1/6-em)]500 12[thin space (1/6-em)]600 1.28
6 DMAc 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v)   25 0 NA NA NA
7 DMAc 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v) Na2CO3 3.5 57 11[thin space (1/6-em)]600 13[thin space (1/6-em)]600 1.23
8 DMAc 10[thin space (1/6-em)]:[thin space (1/6-em)]1 Na2CO3 1.5 42 8700 9360 1.31
9 Fo 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v)   30 0 NA NA NA
10 Fo 10[thin space (1/6-em)]:[thin space (1/6-em)]1 Na2CO3 1.3 25 5250 120[thin space (1/6-em)]000 3.5
11 MFo 10[thin space (1/6-em)]:[thin space (1/6-em)]1   43 0 NA NA NA
12 MFo 10[thin space (1/6-em)]:[thin space (1/6-em)]1 Na2CO3 3.2 50 10[thin space (1/6-em)]250 14[thin space (1/6-em)]200 1.43
13 DMF 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v)   7 15 3250 2300 1.27
14 DMF 10[thin space (1/6-em)]:[thin space (1/6-em)]1 Na2CO3 25 40 8250 10[thin space (1/6-em)]800 1.56


In addition, when the amount of solvent was reduced such that the molar ratio of [MMA]0/[solvent]0 was 10[thin space (1/6-em)]:[thin space (1/6-em)]1, almost all of the reactions still expressed controlled characters (entries 5, 8, 12 and 14). This finding is very significant for the application of ATRP in environmentally-friendly industrial processes. Kinetic plots for the ATRP of MMA with a molar ratio of [MMA]0/[EBPA]0/[FeBr2]0/[solvent]0/[Na2CO3]0 = 200[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]20[thin space (1/6-em)]:[thin space (1/6-em)]2 at 60 °C are depicted in Fig. 1a. The polymerizations proceeded with approximately first-order kinetics in all cases, indicating a constant concentration of growing radicals during polymerizations. The dependence of Mn and Mw/Mn on the monomer conversion within different polar solvents based on formamide is shown in Fig. 1b. The molecular weights of all PMMA increased linearly with the conversion. It is noted that all the experimental molecular weights were slightly higher than the corresponding theoretical ones, which may be attributed to the low initiator efficiency. Other explanations need further study. The MWD values from polymerizations with DMAc and MAc were relatively low (Mw/Mn = 1.18–1.28), but those for MFo were a little higher (Mw/Mn = 1.42–1.52). Even so, this could still be regarded as a “living”/controlled radical polymerization.


image file: c5ra05460e-f1.tif
Fig. 1 (a) Kinetic plots of ln([M]0/[M]) versus time and (b) plots of Mn (filled symbols) and Mw/Mn (open symbols) values versus conversion for FeBr2-catalyzed ATRP of MMA using polar solvents based formamide as ligands. [MMA]0/[EBPA]0/[FeBr2]0/[solvent]0/[Na2CO3]0 = 200[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]20[thin space (1/6-em)]:[thin space (1/6-em)]2, 60 °C. ■ = MAc; ▲ = DMAc; ● = MFo.

Polar solvents based on urea as ligands

Urea is generally used as a nitrogenous fertilizer in agricultural production and exists extensively in nature. It contains the amide group, like the polar solvents based on formamide discussed above. Moreover, urea has one more amidogen than the latter. This may endow the urea with a stronger coordination ability to form catalyst complexes with iron salts. However, there was no reaction when urea was added into the system with a ratio of [MMA]0/[urea]0 = 4[thin space (1/6-em)]:[thin space (1/6-em)]1, as shown in entry 1 of Table 2. In comparison with the results obtained with polar solvents based on formamide, it took a long time for the urea to coordinate with FeBr2 to express activity. Thus, Na2CO3 was used to accelerate the polymerization. It can be seen from Table 2 that the resulting polymer had a high experimental molecular weight and polydispersity (entry 2). The result indicated that the base’s rate-enhancing effect had occurred. Another reason for the poor controllability may be the weak solubility of urea in MMA. The viscosity of the polymerization system increased along with the reaction, and lots of catalyst complex formed at the bottom of the flask, resulting in nonuniformity of the system.
Table 2 FeBr2-catalyzed ATRPs of MMA using polar solvents based on urea as ligandsa
Entry Solvent [M]0/[solvent]0b Base Time (h) Conv. (%) Mn,th (g mol−1) Mn,GPC (g mol−1) Mw/Mn
a [MMA]0/[EBPA]0/[FeBr2]0/[Na2CO3]0 = 200[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2 or 0, 60 °C.b The molecular weight was too small and out of the range of the instrument.
1 Urea (solid) 4[thin space (1/6-em)]:[thin space (1/6-em)]1   30 0 NA NA NA
2 Urea (solid) 100[thin space (1/6-em)]:[thin space (1/6-em)]1 Na2CO3 5 35.2 7290 92[thin space (1/6-em)]100 2.7
3 TMU 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v)   56 11 2400 NAb NAb
4 TMU 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v) Na2CO3 2 64 13[thin space (1/6-em)]000 12[thin space (1/6-em)]000 1.24
5 TMU 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v) Na2CO3 3 78 16[thin space (1/6-em)]000 14[thin space (1/6-em)]000 1.24
6 TMU 10[thin space (1/6-em)]:[thin space (1/6-em)]1 Na2CO3 3 41 8450 7400 1.27


The results from TMU-containing systems were quite different from those with urea (entries 3–6 in Table 2). When the initial volume ratio of MMA to TMU was 2, the reaction could be carried out in the absence of base (entry 3). However, the polymerization reached a low conversion of 11% in 56 h, which is extremely low compared with those for the polymerizations shown in entries 4 and 5 ([catalyst]0/[Na2CO3]0 = 1[thin space (1/6-em)]:[thin space (1/6-em)]2). It can be seen that the polymerization reached a very high conversion of 64%, with good values of molecular weight and MWD (Mw/Mn = 1.24) of the resultant polymers, in just 2 h when using Na2CO3. The catalytic amount of base not only improved the rate of polymerization, but also kept the “living” feature of the system (Mw/Mn = 1.24–1.27). Some differences can be seen when comparing the results for TMU with those for polar solvents based on formamide, as shown in Table 1. The polymerizations had higher rates and better controllability when using TMU as the ligand, especially when the ratio of [MMA]0/[solvent]0 was 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v). For example, when MAc or DMAc were used as ligands, the conversions of monomers reached 34% and 57% in 29 h and 3.5 h, respectively. But for TMU, the conversion reached 78% in just 3 h. Bearing in mind the structures of these solvents, the four methyl groups on TMU could enhance its coordination ability because the methyl group is a better electron donor (electron-donating substituent) than the hydrogen group, and can improve the electron cloud density of nitrogen or oxygen.83 Therefore, the FeBr2/TMU complex showed a better catalytic activity. The polymerization was still well controlled when the molar ratio of the [TMU]0 to [MMA]0 was reduced to 1[thin space (1/6-em)]:[thin space (1/6-em)]10, although the polymerization rate decreased a little (entry 6). The linear increase observed for the TMU in Fig. 2a indicated that the concentration of growing radicals remained constant during the polymerization process. A relatively long induction period existed. The evolution of the Mn and MWD values with monomer conversion is shown in Fig. 2b. The molecular weights of all PMMA increased linearly with the conversion, and deviated little from the theoretical molecular weights. In addition, the values of Mw/Mn remained relatively small (<1.3). All of these observations suggest that TMU is suitable for carrying out the ATRP of MMA, acting as a ligand to form an active complex with FeBr2.


image file: c5ra05460e-f2.tif
Fig. 2 (a) Kinetic plots of ln([M]0/[M]) versus time and (b) plots of Mn (filled symbols) and Mw/Mn (open symbols) values versus conversion for the FeBr2/TMU-catalyzed ATRP of MMA. [MMA]0/[EBPA]0/[FeBr2]0/[TMU]0/[Na2CO3]0 = 200[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]20[thin space (1/6-em)]:[thin space (1/6-em)]2, 60 °C.

Polar solvents based on pyrrolidone as ligands

NMP has been reported to coordinate well to iron salts in general, and the generated catalyst complex has expressed dramatic living characteristics when used in CRP, both in ATRP69 and in AGET ATRP.74,75 To identify the activity for the ATRP of MMA catalyzed by FeBr2/NMP, some experiments were run under the same conditions as above. From entries 9 and 10 in Table 3, it can be seen that the conversion was up to 33% in 7 h when the ratio of [MMA]0/[NMP]0 was 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v). What’s more, the polymerizations were well-behaved (Mw/Mn = 1.39). This result confirmed the excellent activity of NMP in ATRP of MMA as a ligand, especially compared with the other amide group-containing polar solvents mentioned in this article. Polymerizations also proceeded in some other polar solvents in the absence of base, such as DMF, TMU and MPy, with conversions of monomers reaching 15%, 11% and 19% in 7 h, 56 h and 6 h, respectively; these polymerizations were relatively slower than when NMP was used as a ligand. In order to further probe the ability of NMP as a ligand in ATRP, the same experiments were conducted in the presence of a catalytic amount of NMP ([MMA]0/[NMP]0 = 10[thin space (1/6-em)]:[thin space (1/6-em)]1), and base was also added into the reaction to compare with the other systems. From Table 3, it can be seen that the resulting polymer reached a conversion of 33% in just 3 h, with a value of Mn,GPC which was pretty similar to Mn,th, and also had a narrow MWD (Mw/Mn = 1.21) (entry 11). Thereafter, the conversion reached 68% in 6.7 h, with a well-controlled result (entry 12).
Table 3 FeBr2-catalyzed ATRPs of MMA using polar solvents based on pyrrolidone as ligandsa
Entry Solvent [M]0/[solvent]0 Base Time (h) Conv. (%) Mn,th (g mol−1) Mn,GPC (g mol−1) Mw/Mn
a [MMA]0/[EBPA]0/[FeBr2]0/[Na2CO3]0 = 200[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2 or 0, 60 °C.
1 2-Py 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v)   27 0 NA NA NA
2 2-Py 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v) Na2CO3 1.5 48 9800 19[thin space (1/6-em)]500 1.38
3 2-Py 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v) Na2CO3 2 74 15[thin space (1/6-em)]000 20[thin space (1/6-em)]900 1.35
4 2-Py 10[thin space (1/6-em)]:[thin space (1/6-em)]1 Na2CO3 4.4 17 3650 4650 2.23
5 MPy 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v)   6 19 4050 7200 1.33
6 MPy 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v)   27 67 13[thin space (1/6-em)]650 16[thin space (1/6-em)]800 1.54
7 MPy 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v) Na2CO3 2.3 54 11[thin space (1/6-em)]000 10[thin space (1/6-em)]500 1.93
8 MPy 10[thin space (1/6-em)]:[thin space (1/6-em)]1 Na2CO3 2 40 8250 10[thin space (1/6-em)]700 1.8
9 NMP 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v)   7 33 6850 4300 1.39
10 NMP 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v)   23 45 9250 5700 1.42
11 NMP 10[thin space (1/6-em)]:[thin space (1/6-em)]1 Na2CO3 3 33 6850 6700 1.21
12 NMP 10[thin space (1/6-em)]:[thin space (1/6-em)]1 Na2CO3 6.7 68 13[thin space (1/6-em)]850 12[thin space (1/6-em)]300 1.22


2-Py and MPy contain an amide group (see Scheme 2), and have ringed structures, similar to NMP. They were chosen to test the activity as ligands as well. As seen from entries 1–8 in Table 3, when the polymerization was carried out with a ratio of [MMA]0/[polar solvent]0 of 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (v/v) in the absence of Na2CO3, there was no polymerization in 27 h for 2-Py, but for MPy, the conversion reached up to 19% in 6 h. A possible explanation for this is that the methyl group of MPy increases the ability of the amide group to coordinate to the iron salt, and makes it express higher “living” features. This explanation is also applicable to NMP. The polymerizations using 2-Py showed poor controllability, presumably caused by its poor dissolution in MMA. In the absence of base, the resulting polymers when MPy was used showed higher Mn,GPC values than the theoretical molecular weights, which may have arisen from low initiator efficiency. When the base was added into the polymerizations, the experimental molecular weights behaved well, but the values of Mw/Mn were a little higher. A suitable amount of base needs to be chosen for the ATRP of MMA using MPy as a ligand.

The obtained results above demonstrated the potential of using amide group-containing polar solvents for Fe(II)-catalyzed ATRP of MMA. The addition of base was essential for most of the systems. The role of the base has been stated above. The selected polar solvents act as ligands that increase the dissolubility of FeBr2 in MMA, and also adjust the catalytic activity, which results in the good living features of the polymerizations. It’s particularly interesting that polymerization can be carried out in the absence of base when using DMF, TMU, NMP and MPy. As mentioned above, all of them have an electron donor (methyl group) near or on the amide group, and this can increase the activity of the solvents by enhancing the solvents’ coordination to iron salts. For TMU in particular, the four methyl groups around the oxygen atom greatly heighten the activity. As a result, the polymerizations have higher rates and the resultant polymers are well-controlled in the presence of TMU. This conclusion has potential application in the choice of suitable polar solvents for CRP. Moreover, the configurations of the different solvent molecules may also influence their activities, which needs further investigation.

Iron(III)-catalyzed AGET ATRP

AGET ATRP is easier to operate than normal ATRP on account of using a high oxidation state catalyst, and has similar “living” characteristics to ATRP when suitable reducing agents (RAs) are added. The iron(III)-catalyzed AGET ATRP of MMA using different reducing agents in the presence of some polar solvents (DMF, NMP and MeCN) has been reported in our previous studies,74,75 and most of the polymerizations showed well-controlled features. To identify the activity of some other polar solvents based on the amide group from this article, three solvents (MFo, TMU and 2-Py) were selected for the FeBr3-catalyzed AGET ATRP of MMA, using ethylene glycol (EG) or vitamin C (VC) as the reducing agent. It can be seen from Table 4 that all of the polymerizations were conducted successfully with an additional limited amount of base. The polymerization rate with VC was obviously faster than that with EG, because of the stronger reducing ability of VC compared with EG. Some of the resulting polymers showed well-controlled features, such as those in entries 2, 3 and 4 with MWD values of 1.38, 1.23 and 1.37, respectively, and with little deviation of Mn,GPC with respect to Mn,th at the same time. As for the other instances shown in Table 4 (entries 1, 5 and 6), the differences between the molecular weights from experiment and from theory were a little big. All of them had the phenomenon that some precipitates formed during these reactions, which caused the nonuniformity of the systems. In spite of this, the chosen amide group-containing polar solvents have potential for conducting iron(III)-catalyzed AGET ATRP. The results here further express the universality of the selected polar solvents used for CRP.
Table 4 FeBr3-catalyzed AGET ATRPs of MMA using amide group-containing polar solvents as ligandsa
Entry Solvent RAb Time (h) Conv. (%) Mn,th (g mol−1) Mn,GPC (g mol−1) Mw/Mn
a [MMA]0/[EBPA]0/[FeBr3]0/[Na2CO3]0 = 200[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2, 60 °C.b The ratio of RA to catalyst is 3[thin space (1/6-em)]:[thin space (1/6-em)]1 for EG, and 1[thin space (1/6-em)]:[thin space (1/6-em)]1 for VC.
1 MFo EG 6.1 80 16[thin space (1/6-em)]300 27[thin space (1/6-em)]580 1.59
2 MFo VC 2.7 70 14[thin space (1/6-em)]250 16[thin space (1/6-em)]500 1.38
3 TMU VC 15 53 10[thin space (1/6-em)]850 8600 1.23
4 2-Py EG 8.5 17 3650 3550 1.37
5 2-Py EG 12 77 15[thin space (1/6-em)]650 21[thin space (1/6-em)]850 1.46
6 2-Py VC 1.1 62 12[thin space (1/6-em)]650 24[thin space (1/6-em)]250 1.55


Chain end analysis and chain extension experiment

The chain end of the PMMA prepared from the FeBr2-catalyzed ATRP in the presence of a limited amount of TMU was analyzed by 1H NMR spectroscopy, as shown in Fig. 3. The signal labelled ‘a’ (3.77 ppm) was attributed to the methyl ester group at the chain end, and the signal ‘b’ (3.60 ppm) came from the other methyl ester groups in PMMA. The signals ‘c’ (3.38 ppm), ‘d’ (7.17–7.37 ppm) and ‘e’ (3.97–4.17 ppm) corresponded to the protons derived from EBPA in methine, phenyl and methylene groups. The molecular weight (Mn,NMR) can be determined by the integrals in the 1H NMR spectrum based on eqn (1):
 
Mn,NMR (g mol−1) = (Ia,b/3) × 100.12/(Ie/2) + 243.1 (1)

image file: c5ra05460e-f3.tif
Fig. 3 1H NMR spectrum of PMMA (Mn,GPC = 7200, Mw/Mn = 1.45) with CDCl3 as the solvent. [MMA]0/[EBPA]0/[FeBr2]0/[TMU]0/[Na2CO3]0 = 200[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]20[thin space (1/6-em)]:[thin space (1/6-em)]2, 60 °C.

The calculated molecular weight of PMMA from the 1H NMR spectrum (Mn,NMR = 6300) is in agreement with the GPC result (Mn,GPC = 7200). The result suggested that the end of the obtained PMMA was end-capped by the EBPA moieties.

In order to further confirm the mechanism of ATRP carried out using polar solvents as ligands, a chain extension experiment was done using one of the resulting polymers as a macroinitiator. The macroinitiator (PMMA–Br, Mn,GPC = 10[thin space (1/6-em)]400, Mw/Mn = 1.29) came from the ATRP with a ratio of [MMA]0/[EBPA]0/[FeBr2]0/[NMP]0/[Na2CO3]0 = 200[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]20[thin space (1/6-em)]:[thin space (1/6-em)]2 at 60 °C with a conversion of 61%, and the chain-extended polymer (PMMA-b-PMMA–Br) was obtained from an ATRP with a ratio of [MMA]0/[PMMA–Br]0/[FeBr2]0/[NMP]0/[Na2CO3]0 = 500[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]50[thin space (1/6-em)]:[thin space (1/6-em)]5 at 60 °C in 1 h. As shown in Fig. 4, a peak shift can be seen from the macroinitiator to the chain-extended PMMA with Mn,GPC = 30[thin space (1/6-em)]100 and Mw/Mn = 1.40. The increased value of Mw/Mn may be due to a little of the polymer being out of activity in the macroinitiator. The successful chain extension reaction confirms the controlled features of the polymerizations.


image file: c5ra05460e-f4.tif
Fig. 4 GPC curves for the chain extension experiment.

Conclusions

In summary, iron(II)-catalyzed ATRPs of MMA were carried out in a series of amide group-containing polar solvents without additional ligands. Most of the systems showed characteristics of “living”/controlled radical polymerization in the presence of a catalytic amount of base. The results here confirmed the coordination potential of polar solvents containing the amide group to iron salts, and the formed catalyst complexes showed good activity. The structures of the polar solvents had a great effect on their coordination abilities, which is very important for the polymerization activity. The addition of an electron donor (methyl group) in the structures improved the solvents’ activity significantly, which can be seen from the results when DMF, TMU, MPy and NMP were used. It is particularly interesting that the polymerizations remained controlled even when the polar solvents were reduced to a limited amount (the initial molar ratio of catalyst to solvent was 1[thin space (1/6-em)]:[thin space (1/6-em)]20), which is of great significance for green applications. Some polar solvents were also employed for iron(III)-catalyzed AGET ATRP of MMA; as expected, they behaved well, similarly as for ATRP. More details about the mechanism of the coordination between polar solvents and the metal salt, and work expanding the scope of polar solvents applied for CRP, still need to be investigated in future studies.

Acknowledgements

We are grateful to the National Natural Science Foundation of China (21304036, 51210004, 51433002, and 51473056), and the Specialized Research Fund for the Doctoral Program of Higher Education (20120142120024) for support of this work.

Notes and references

  1. V. Coessens, T. Pintauer and K. Matyjaszewski, Prog. Polym. Sci., 2001, 26, 337–377 CrossRef CAS.
  2. M. Kamigaito, T. Ando and M. Sawamoto, Chem. Rev., 2001, 101, 3689–3746 CrossRef CAS PubMed.
  3. K. Matyjaszewski and J. Xia, Chem. Rev., 2001, 101, 2921–2990 CrossRef CAS PubMed.
  4. W. A. Braunecker and K. Matyjaszewski, Prog. Polym. Sci., 2007, 32, 93–146 CrossRef CAS PubMed.
  5. N. V. Tsarevsky and K. Matyjaszewski, Chem. Rev., 2007, 107, 2270–2299 CrossRef CAS PubMed.
  6. M. Ouchi, T. Terashima and M. Sawamoto, Chem. Rev., 2009, 109, 4963–5050 CrossRef CAS PubMed.
  7. F. di Lena and K. Matyjaszewski, Prog. Polym. Sci., 2010, 35, 959–1021 CrossRef CAS PubMed.
  8. M. Kamigaito, Polym. J., 2011, 43, 105–120 CrossRef CAS PubMed.
  9. L. Bai, L. Zhang, Z. Cheng and X. Zhu, Polym. Chem., 2012, 3, 2685–2697 RSC.
  10. K. Matyjaszewski, Macromolecules, 2012, 45, 4015–4039 CrossRef CAS.
  11. K. Matyjaszewski, Isr. J. Chem., 2012, 52, 206–220 CrossRef CAS PubMed.
  12. W. He, H. Jiang, L. Zhang, Z. Cheng and X. Zhu, Polym. Chem., 2013, 4, 2919–2938 RSC.
  13. K. Matyjaszewski and N. V. Tsarevsky, J. Am. Chem. Soc., 2014, 136, 6513–6533 CrossRef CAS PubMed.
  14. R. Poli, L. E. N. Allan and M. P. Shaver, Prog. Polym. Sci., 2014, 39, 1827–1845 CrossRef CAS PubMed.
  15. Z. Xue, D. He and X. Xie, Polym. Chem., 2015, 6, 1660–1687 RSC.
  16. M. Kato, M. Kamigaito, M. Sawamoto and T. Higashimura, Macromolecules, 1995, 28, 1721–1723 CrossRef CAS.
  17. J.-S. Wang and K. Matyjaszewski, J. Am. Chem. Soc., 1995, 117, 5614–5615 CrossRef CAS.
  18. T. E. Patten, J. Xia, T. Abernathy and K. Matyjaszewski, Science, 1996, 272, 866–868 CAS.
  19. T. E. Patten and K. Matyjaszewski, Acc. Chem. Res., 1999, 32, 895–903 CrossRef CAS.
  20. N. Chan, M. F. Cunningham and R. A. Hutchinson, Polym. Chem., 2012, 3, 1322–1333 RSC.
  21. S. Jana, A. Parthiban and F. M. Choo, Chem. Commun., 2012, 48, 4256–4258 RSC.
  22. S. A. Turner, Z. D. Remillard, D. T. Gijima, E. Gao, R. D. Pike and C. Goh, Inorg. Chem., 2012, 51, 10762–10773 CrossRef CAS PubMed.
  23. Q. Zhang, P. Wilson, Z. Li, R. McHale, J. Godfrey, A. Anastasaki, C. Waldron and D. M. Haddleton, J. Am. Chem. Soc., 2013, 135, 7355–7363 CrossRef CAS PubMed.
  24. A. B. Dwyer, P. Chambon, A. Town, T. He, A. Owen and S. P. Rannard, Polym. Chem., 2014, 5, 3608–3616 RSC.
  25. J. Gao, W. Song, P. Wang and G. Zhai, J. Polym. Sci., Part A: Polym. Chem., 2014, 52, 2562–2578 CrossRef CAS PubMed.
  26. H. Jiang, L. Zhang, X. Jiang, X. Bao, Z. Cheng and X. Zhu, Macromol. Rapid Commun., 2014, 35, 1332–1339 CrossRef CAS PubMed.
  27. M. Y. Khan, Z. Xue, D. He, S. K. Noh and W. S. Lyoo, Polymer, 2010, 51, 69–74 CrossRef CAS PubMed.
  28. T. Opstal and F. Verpoort, Angew. Chem., Int. Ed., 2003, 42, 2876–2879 CrossRef CAS PubMed.
  29. Y. Fukuzaki, Y. Tomita, T. Terashima, M. Ouchi and M. Sawamoto, Macromolecules, 2010, 43, 5989–5995 CrossRef CAS.
  30. K. Nakatani, T. Terashima, M. Ouchi and M. Sawamoto, Macromolecules, 2010, 43, 8910–8916 CrossRef CAS.
  31. H. Yoda, K. Nakatani, T. Terashima, M. Ouchi and M. Sawamoto, Macromolecules, 2010, 43, 5595–5601 CrossRef CAS.
  32. D. He, S. K. Noh and W. S. Lyoo, J. Polym. Sci., Part A: Polym. Chem., 2011, 49, 4594–4602 CrossRef CAS PubMed.
  33. T. Terashima, A. Nomura, M. Ito, M. Ouchi and M. Sawamoto, Angew. Chem., Int. Ed., 2011, 50, 7892–7895 CrossRef CAS PubMed.
  34. K. Nakatani, Y. Ogura, Y. Koda, T. Terashima and M. Sawamoto, J. Am. Chem. Soc., 2012, 134, 4373–4383 CrossRef CAS PubMed.
  35. T. Terashima, S. Nishioka, Y. Koda, M. Takenaka and M. Sawamoto, J. Am. Chem. Soc., 2014, 136, 10254–10257 CrossRef CAS PubMed.
  36. T. Ando, M. Kamigaito and M. Sawamoto, Macromolecules, 1997, 30, 4507–4510 CrossRef CAS.
  37. S. Zhu and D. Yan, Macromol. Rapid Commun., 2000, 21, 1209–1213 CrossRef CAS.
  38. V. C. Gibson, R. K. O’Reilly, W. Reed, D. F. Wass, A. J. P. White and D. J. Williams, Chem. Commun., 2002, 1850–1851 RSC.
  39. R. Ferro, S. Milione, V. Bertolasi, C. Capacchione and A. Grassi, Macromolecules, 2007, 40, 8544–8546 CrossRef CAS.
  40. R. K. O’Reilly, M. P. Shaver, V. C. Gibson and A. J. P. White, Macromolecules, 2007, 40, 7441–7452 CrossRef.
  41. Z. Xue, B. W. Lee, S. K. Noh and W. S. Lyoo, Polymer, 2007, 48, 4704–4714 CrossRef CAS PubMed.
  42. Z. Xue, S. Noh and W. Lyoo, Macromol. Res., 2007, 15, 302–307 CrossRef CAS.
  43. L. Zhang, Z. Cheng, F. Tang, Q. Li and X. Zhu, Macromol. Chem. Phys., 2008, 209, 1705–1713 CrossRef CAS PubMed.
  44. Z. Xue, N. T. B. Linh, S. K. Noh and W. S. Lyoo, Angew. Chem., Int. Ed., 2008, 47, 6426–6429 CrossRef CAS PubMed.
  45. Z. Xue, S. K. Noh and W. S. Lyoo, J. Polym. Sci., Part A: Polym. Chem., 2008, 46, 2922–2935 CrossRef CAS PubMed.
  46. Z. Xue, H. S. Oh, S. K. Noh and W. S. Lyoo, Macromol. Rapid Commun., 2008, 29, 1887–1894 CrossRef CAS PubMed.
  47. M. Ishio, M. Katsube, M. Ouchi, M. Sawamoto and Y. Inoue, Macromolecules, 2009, 42, 188–193 CrossRef CAS.
  48. L. Bai, L. Zhang, Z. Zhang, Y. Tu, N. Zhou, Z. Cheng and X. Zhu, Macromolecules, 2010, 43, 9283–9290 CrossRef CAS.
  49. Z. Xue, D. He, S. K. Noh and W. S. Lyoo, Macromolecules, 2009, 42, 2949–2957 CrossRef CAS.
  50. D. He, Z. Xue, M. Y. Khan, S. K. Noh and W. S. Lyoo, J. Polym. Sci., Part A: Polym. Chem., 2010, 48, 144–151 CrossRef CAS PubMed.
  51. Y. Wang, Y. Kwak and K. Matyjaszewski, Macromolecules, 2012, 45, 5911–5915 CrossRef CAS.
  52. H. Aoshima, K. Satoh, T. Umemura and M. Kamigaito, Polym. Chem., 2013, 4, 3554–3562 RSC.
  53. K. Nishizawa, M. Ouchi and M. Sawamoto, Macromolecules, 2013, 46, 3342–3349 CrossRef CAS.
  54. A. Simakova, M. Mackenzie, S. E. Averick, S. Park and K. Matyjaszewski, Angew. Chem., Int. Ed., 2013, 52, 12148–12151 CrossRef CAS PubMed.
  55. W. He, L. Cheng, L. Zhang, Z. Liu, Z. Cheng and X. Zhu, Polym. Chem., 2014, 5, 638–645 RSC.
  56. Y. Liu, T. Xu, L. Zhang, Z. Cheng and X. Zhu, Polym. Chem., 2014, 5, 6804–6810 RSC.
  57. S. I. Nakanishi, M. Kawamura, H. Kai, R. H. Jin, Y. Sunada and H. Nagashima, Chem.–Eur. J., 2014, 20, 5802–5814 CrossRef CAS PubMed.
  58. C. Bolm, J. Legros, J. Le Paih and L. Zani, Chem. Rev., 2004, 104, 6217–6254 CrossRef CAS PubMed.
  59. E. B. Bauer, Curr. Org. Chem., 2008, 12, 1341–1369 CrossRef CAS.
  60. B. Plietker, Iron Catalysis in Organic Chemistry: Reactions and Applications, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim, 2008 Search PubMed.
  61. I. Bauer and H.-J. Knölker, Chem. Rev., 2015 DOI:10.1021/cr500425u.
  62. Y.-H. Ng, F. di Lena and C. L. L. Chai, Polym. Chem., 2011, 2, 589–594 RSC.
  63. J. Yang, J. Wang, Y. Tang, D. Wang, B. Xiao, X. Li, R. Li, G. Liang, T.-K. Sham and X. Sun, J. Mater. Chem. A, 2013, 1, 7306–7311 CAS.
  64. J. Liu, M. N. Banis, Q. Sun, A. Lushington, R. Li, T.-K. Sham and X. Sun, Adv. Mater., 2014, 26, 6472–6477 CrossRef CAS PubMed.
  65. J. Mosa, M. Aparicio, A. Duran, C. Laberty-Robert and C. Sanchez, J. Mater. Chem. A, 2014, 2, 3038–3046 CAS.
  66. C. Gong, F. Deng, C.-P. Tsui, Z. Xue, Y. S. Ye, C.-Y. Tang, X. Zhou and X. Xie, J. Mater. Chem. A, 2014, 2, 19315–19323 CAS.
  67. C. Gong, Z. Xue, X. Wang, X.-P. Zhou, X.-L. Xie and Y.-W. Mai, J. Power Sources, 2014, 246, 260–268 CrossRef CAS PubMed.
  68. X. Wang, C. Gong, D. He, Z. Xue, C. Chen, Y. Liao and X. Xie, J. Membr. Sci., 2014, 454, 298–304 CrossRef CAS PubMed.
  69. Y. Wang and K. Matyjaszewski, Macromolecules, 2010, 43, 4003–4005 CrossRef CAS.
  70. H. Schroeder, M. Buback and K. Matyjaszewski, Macromol. Chem. Phys., 2014, 215, 44–53 CrossRef CAS PubMed.
  71. Z. Xue, J.-C. Daran, Y. Champouret and R. Poli, Inorg. Chem., 2011, 50, 11543–11551 CrossRef CAS PubMed.
  72. Z. Xue and R. Poli, J. Polym. Sci., Part A: Polym. Chem., 2013, 51, 3494–3504 CrossRef CAS PubMed.
  73. S. A. Bulgakova, E. S. Volgutova and I. E. Khokhlova, Open J. Polym. Chem., 2012, 2, 99–104 CrossRef CAS.
  74. D. Yang, D. He, Y. Liao, Z. Xue, X. Zhou and X. Xie, J. Polym. Sci., Part A: Polym. Chem., 2014, 52, 1020–1027 CrossRef CAS PubMed.
  75. Z. Xue, J. Zhou, D. He, F. Wu, D. Yang, Y. S. Ye, Y. Liao, X. Zhou and X. Xie, Dalton Trans., 2014, 43, 16528–16533 RSC.
  76. M. Ding, X. Jiang, J. Peng, L. Zhang, Z. Cheng and X. Zhu, Green Chem., 2015, 17, 271–278 RSC.
  77. K. Ding, F. Zannat, J. C. Morris, W. W. Brennessel and P. L. Holland, J. Organomet. Chem., 2009, 694, 4204–4208 CrossRef CAS PubMed.
  78. G. R. Ferrell, C. Moore, A. L. Rheingold and C. J. A. Daley, Acta Crystallogr., Sect. E: Struct. Rep. Online, 2009, 65, m1512–m1513 CAS.
  79. L. Bai, L. Zhang, Z. Zhang, J. Zhu, N. Zhou, Z. Cheng and X. Zhu, J. Polym. Sci., Part A: Polym. Chem., 2011, 49, 3980–3987 CrossRef CAS PubMed.
  80. L. Bai, L. Zhang, Z. Zhang, J. Zhu, N. Zhou, Z. Cheng and X. Zhu, J. Polym. Sci., Part A: Polym. Chem., 2011, 49, 3970–3979 CrossRef CAS PubMed.
  81. T. Guo, L. Zhang, H. Jiang, Z. Zhang, J. Zhu, Z. Cheng and X. Zhu, Polym. Chem., 2011, 2, 2385–2390 RSC.
  82. Y. Wang, X. Li, F. Du, H. Yu, B. Jin and R. Bai, Chem. Commun., 2012, 48, 2800–2802 RSC.
  83. M. P. Shaver, L. E. N. Allan and V. C. Gibson, Organometallics, 2007, 26, 4725–4730 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra05460e

This journal is © The Royal Society of Chemistry 2015