Вы находитесь на странице: 1из 15

Encyclopedia of Polymer Sceince and Technology Copyright c 2005 John Wiley & Sons, Inc. All rights reserved.

IONIC LIQUIDS, POLYMERIZATION IN


Introduction
The rst room temperature ionic liquid (RTIL), [EtNH3 ][NO3 ] (mp 12 C), was reported in 1914 (1), and since then a great deal of research effort has been exerted to exploit possible applications of these compounds. Initially, RTILs were used mainly as electrolytes in batteries or for metal electrodeposition; however, nowadays they are nding an ever increasing range of applications. Ionic liquids may be viewed as molten salts and are liquids containing only cations and anions at ambient or near ambient (<100 C) temperature. They exhibit a relatively wide electrochemical window, good electronic and ionic conductivities, a broad range of room temperature liquid compositions, and may have negligible vapor pressure, and excellent chemical, thermal, air and moisture stability (26). They are composed of weakly coordinating anions, for example, BF4 and PF6 and, hence, are highly polar yet noncoordinating solvents. Most imidazolium and pyridinium ionic liquids have polarities similar to those of small molecule alcohols (7,8). Their hydrophilicity/lipophilicity is tunable by varying the combination of cations and anions. Thus, RTILs have been referred to as designer solvents (9), and millions of different RTILs could potentially be synthesized and optimized for specic applications. Depending on their structures, RTILs are able to dissolve a variety of organic, inorganic, and organometallic compounds. Their ease of handling, low vapor pressures, and potential for recycling make them promising potentially environmentally benign reaction media to replace volatile molecular solvents in both the chemical industry and in academic research. Furthermore, RTILs have limited miscibility with some common organic solvents but high compatibility with transition metals. As a consequence, a biphasic or phase-separable (organic/RTIL biphasic systems) catalysis concept can be developed in which a homogeneous catalyst is immobilized in one liquid phase (RTIL) and the reactants and/or products reside largely in another liquid phase (organic) (10), thus enabling easy product and catalyst separation with the retention of the transition metal catalyst in an ionic liquid phase. Various common organic reactions employing metal catalysts have been conducted by employing ionic liquids as alternative reaction media, including Diels-Alder reactions (11), Friedel-Crafts reactions (12), hydrogenations (13), hydroformylations (14), alkylations (15), dimerizations (16), Heck reactions (17), Suzuki couplings (18), Sonogashira couplings (19), ring-closing olen metathesis reactions (20), alcohol oxidations (21), and nucleophilic substitutions (22).
1

IONIC LIQUIDS, POLYMERIZATION IN

Fig. 1. Matrix of RTIL cations and anions.

Some typical RTILs and how structure affects properties are shown in Figure 1. Representative common cations and anions used to prepare RTILs are shown in Figure 2. RTILs are commercially available from ACROS Organic (Fisher Scientic, UK), Chemada Fine Chemicals (Israel), Covalent Associates (Boston, Mass.), C-TRI (Korea), CYTEC (Canada), Sigma-Aldrich (US), BASF (US), Merck KGaA (Germany), QUILL (UK), Sachem (Austin, Tex.), Scionix (UK), and SolventInnovation (Germany).

Ionic Liquids for Polymerization


One of the earliest examples of polymerization in RTILs was reported in 1990 in a Lewis acidic ionic liquid: AlCl3 -1-ethyl-3-methylimidazolium chloride [EMIM]Cl

Fig. 2. Representative common cations and anions used to prepare room temperature ionic liquids.

IONIC LIQUIDS, POLYMERIZATION IN

(23). It was found that the combination of TiCl4 and AlEthylCl2 in AlCl3 /[EMIM]Cl can be catalytically active for ethylene polymerization, although the yield of polyethylene (PE) was low. This work opened a new route for making polymers using RTILs as alternative reaction media. Later, replacing TiCl4 with Cp2 TiCl2 was shown to give higher yields of PE using the same RTIL (24). More examples of related work on using ionic liquids for polymerization or oligomerization of olens can be found in the literature and in patents (25,26). It was believed that the active sites are cationic alkylmetal complex generated by the interaction of added transition metal catalyst with alkylaluminum moieties present in the ionic liquids via halogen abstraction and alkylation. The electrochemical polymerization of benzene in various ionic liquids has been described as a method for synthesis of poly(p-phenylene) (27). In these studies, ionic liquids were used mainly as convenient electrolytes. Carlin and Osteryoung produced a new electroactive material by electrochemical oxidation of triphenylsilyl chloride (Ph3 SiCl) in the acidic ionic liquid (AlCl3 -[EMIM]Cl) (28). The lm formed exhibited reversible redox behavior and was electronically conducting in the oxidized state by incorporation of the cation of the ionic liquid into the lm. However, most of the RTILs mentioned above are the 1st generation ionic liquids, which are haloaluminate-based ionic liquids and thus easily undergo hydrolysis to form undesirable HCl by-products.

Free Radical Polymerization


Free radical polymerization is extensively used for the synthesis of a variety of polymeric materials due to its versatility, synthetic ease, and compatibility with a wide variety of functional groups, which is also coupled with its tolerance to water and protic media. However, conventional free radical polymerization has a signicant drawback, which is related to the reactivity of the propagating free radical chain ends and their propensity to undergo a variety of termination and chain transfer reactions. The materials obtained are therefore polydisperse with limited control over molecular weight and architecture. Ionic polymerizations (anionic or cationic) were for many years the only living polymerization techniques available that achieved efcient control over the structure and architecture of vinyl polymers. Although these techniques can generate polymers with low polydispersity, controlled molecular weight, and control over chain ends and macromolecular architectures, they are not suitable for the polymerization and copolymerization of many vinylic monomers. This limitation is due to the incompatibility of the growing polymer chain ends (carbanion or carbocation) with numerous functional groups and certain monomer families. In addition, stringent reaction conditions, including the use of ultrapure reagents and the need for total or near-total exclusion of water and oxygen, limit the widespread commercial use of these techniques. These challenges spurred polymer chemists to develop new concepts which allow a living or controlled polymerization using special free radical polymerization techniques. As a consequence, nitroxide-mediated polymerization (NMP), atom transfer radical polymerization (ATRP), and radical addition-fragmentation and transfer (RAFT) were developed. The general strategy behind these processes is the reversible termination of the growing polymeric chain with a mediating radical to reduce the overall concentration of the propagating radical chain ends.

IONIC LIQUIDS, POLYMERIZATION IN

In the absence of other reactions leading to initiation of new polymer chains (ie, no reaction of the mediating radical with the vinylic monomer), the concentration of reactive chain ends is extremely low, minimizing irreversible termination reactions, such as combination or disproportionation. All chains would thus be initiated only from the desired initiating species, and growth should occur in a living fashion, allowing a high degree of control over the entire polymerization process with well-dened polymers being obtained (29). Although polymerization in moisture-sensitive ionic liquids has certain advantages when compared to traditional solvents, little polymerization work was done in RTILs until the introduction of the 2nd generation of air, moisture, and thermally stable, neutral ionic liquids, composed of dialkylimidazolium cations and PF6 or BF4 anions. In spite of the fact that these ionic liquids are anticipated to be potentially useful solvents for ionic polymerization due to their highly polar nature, the rst report of polymerization in the new nonhygroscopic ionic liquids was based on a free radical polymerization processatom transfer radical polymerization (ATRP). Carmichael and co-workers (30) observed rst-order kinetic behavior over a wide range of reaction temperatures using ethyl-2-bromoisobutyrate initiator and CuBr/N-propyl-2-pyridylmethanimine as the catalyst pair in a 50/50 (V/V) MMA/[BMIM]PF6 system. The reactions are faster as compared to those in common solvents, while the polydispersities (M w /M n , where M w is weightaverage molecular weight and M n is number-average molecular weight) are narrow: 1.301.43. By extraction with toluene, the polymer can be separated from the solution while copper catalyst remains in the ionic liquid phase, which facilitates catalyst re-use by adding fresh monomer. The use of the RTIL medium also eliminated the need for postpurication to eliminate toxic copper salts, which entails passing the polymer solution through purication columns. Carmichael and co-workers attributed this polymerization behavior to the increased polarity of ionic liquids, because a similar increase in the rate has been observed with other polar/co-ordinating solvents. In a recent communication (31), Hong and coworkers reported large increases in both the rate of polymerization and the molecular weights that were obtained in the polymerizations of methyl methacrylate (MMA) in [BMIM]PF6 as compared to identical polymerizations carried out in conventional organic solvents. These dramatic increases were believed to be at least in part because of the high viscosity of the polymerization media. Thus, a diffusioncontrolled termination mechanism was proposed to explain the diminished chain terminations in such a viscous system, which would account for a simultaneous increase in both rate of polymerization and degree of polymerization. Very recently, Benton and Brazel (32) reported similar behavior, high molecular weights and rapid polymerization rates, for MMA in [BMIM]PF6 . Harrison and co-workers (33) used a pulse laser polymerization (PLP) technique to polymerize MMA in [BMIM]PF6 , which is the IUPAC-recommended standard procedure for measuring the rate constant of propagation (kp ). They found that both the propagation and termination rates were signicantly affected by [BMIM]PF6 . They attributed the increase of the propagation rate to the increased polarity of the ionic liquid solution, which lowers the activation energy of propagation through charge-transfer interactions. The termination rate is decreased simply because of the increased viscosity of the polymerization medium. Both an increase of kp and decrease of kt contribute a roughly ten-fold increase of overall rate of polymerization. Quite

IONIC LIQUIDS, POLYMERIZATION IN

recently, Cheng and co-workers (34) reported free radical polymerization of acrylonitrile in [BMIM]BF4 using AIBN as initiator. These workers also note that ionic liquids are excellent media for achieving high molecular weight polymers. They attribute this result to low chain transfer constants for RTILs and the ability of these solvents to stabilize the growing radical chain ends. Clearly most of the early studies on polymerization in RTILs used [BMIM]PF6 as the medium. This was largely due to the commercial availability and easy synthesis of this material. However, [BMIM]PF6 can readily hydrolyze to release toxic HF as a by-product, and there is an ever increasing focus on the use of more benign RTILs in polymer systems. Zhang and co-workers (35) measured the polarity and viscosity of over a dozen RTILs based upon a wide range of cations and anions, and attempted to correlate these physical data with the low conversion conventional free radical polymerization behavior of styrene and MMA in these RTILs. This study revealed no obvious trends between polarity or viscosity of RTILs and polymerization rate or molecular weight of the formed polymer. At around the same time, Strehmel and co-workers (36) reported that the higher the viscosity of the RTIL, the higher the yields and molecular weights of the polymers. However, unlike in the work by Zhang and co-workers, these workers carried their polymerizations to high (near quantitative) conversions, where the Trommsdorff effect strongly impacts polymerization behavior. Clearly, the higher viscosities will affect polymerization behavior in this regime and make it difcult to ascertain polymerization behavior in the non-Trommsdorf regime. The RTILs employed in this latter study were composed of cations and anions of similar structures except the alkyl length and positions on the cation, while either cations or anions are structurally different in the former case. In our opinion, it is still too early to make a conclusion regarding the effects of polarity and viscosity of the RTIL on polymerization behavior. It would be very interesting to repeat the experiments of Strehmel and co-workers (36) but keep conversions low. The polarity of molecular solvents is a complex of many interactions including: H-bonding, -interaction, or van der Waals forces. In RTILs, the situation is even more complicated since both cations and anions may have their own distinct interactions. Zhang and co-workers used solvatochromic dye (Nile red) to measure the polarities of RTILs and the results had similar trend as those in Reference (37). Since Nile red is a positively solvatochromic, it preferentially interacts with anions of RTILs. The apparent polarities deduced may be not reect the real values, thus further experiments need to be carried out to address this problem. In addition, the polymers synthesized in RTILs all have similar glass transition temperatures and microstructures as compared to those obtained in benzene or in bulk, based upon thermal analysis and 13 C NMR experiments (38). Biedron and co-workers (39) reported heterogeneous ATRP in [BMIM]PF6 . Alkyl acrylates (methyl, butyl, hexyl, and dodecyl) are either soluble, partly soluble, or completely insoluble in this RTIL depending on the length of the alkyl substituent. For the heterogeneous systems, the alkyl acrylate formed an upper monomer phase while the CuBr/pentamethyldiethylenetriamine (PMDETA) catalyst remained in the lower RTIL phase. Methyl acrylate (MA) and poly(methyl acrylate) (PMA) are miscible with [BMIM]PF6 and form a homogeneous polymerization system, therefore, all reactions proceed in one phase. For the three other acrylates, the growing macromolecular chains react with the monomer at the

IONIC LIQUIDS, POLYMERIZATION IN

interface but reside predominantly in the monomer phase. In all cases, the reactions were living radical polymerizations. The advantages of heterogeneous ATRP are easy separation of polyacrylate from the RTIL phase after reaction, with less copper catalyst contamination and reduction of side reactions due to the absence of catalyst in the upper monomer phase, as compared to bulk ATRP. It was found that using a chiral RTIL (1-(R-(+)-2 -methylbutyl)-3-methylimidazolium hexauorophosphate) could impact the stereoregularity of polymer produced by ATRP (40). Sarbu and co-workers (41) used different catalyst systems (iron or copper halides) to successfully carry out ATRP in a range of 1-butyl-3-methylimidazolium ionic liquids. In iron-mediated ATRP, no additional ligand was required to achieve a controlled polymerization of MMA although both initiation rates and rates of reaction were low. Systems without organic ligands were effective only in a phosphonate ionic liquid for copper-mediated ATRP of MMA, while a ligand was required in ionic liquids with halide or carbonate anions. ATRP in ionic liquids proceeds with low initiation efciency. This has been attributed to the high concentration of the catalyst in the RTIL phase, into which the small initiators can easily diffuse and thus generate high concentration of free radicals. The free radical can undergo irreversible termination and cause the low initiation efciency. However, initiator efciency can be improved by employing macroinitiators, which have little tendency to diffuse into the RTIL phase. In addition, the catalyst can be regenerated after removal of polymer and unreacted monomer. Reversible atom transfer radical polymerization of MMA in [BMIM]PF6 , [BMIM]BF4 , and [DMIM]BF4 (1-dodecyl-3-methylimidazolium tetrauoroborate) has been described (42). In [BMIM]PF6 , the reverse ATRP of MMA was achieved with 2,2 -azobisisobutyronitrile (AIBN)/CuCl2 /2,2 -bipyridine initiation system, in which the system is homogeneous throughout the reaction. Due to a cage effect with ionic liquids molecules, the termination of the primary radicals through decomposition of AIBN might occur before they can initiate polymerization, which accounts for the low initiation efciency of AIBN. The Cu catalyst is also soluble in the ionic liquid. As a consequence of both factors, less catalyst is needed to effectively mediate the polymerization process in ionic liquids than in other reverse ATRPs. The ionic liquids and catalyst can be recovered and reused. Zhang and co-workers (43) used both a bimolecular initiation system (benzoyl peroxide (BPO) + 2,2,6,6-tetramethylpiperidine 1-oxyl (TEMPO)) and a universal initiator system (2,2,5-trimethyl-3-(1-phenylethoxy)-4-phenyl-3-azahexane (TMPPAH)) for the attempted nitroxide-mediated polymerization of MMA and St. Polymers were produced but polymerizations were not living/controlled for either monomer. Possible reasons are the low diffusion rates of mediating radicals and slow degradation of free TEMPO at elevated temperatures in the presence of [BMIM]PF6 . Ryan and co-workers (44) demonstrated a controlled/living nitroxidemediated polymerization of MA in 50% V/V of [HMIM]PF6 initiated by the initiator pair AIBN + 4-oxo-2,2,6,6-tetramethyl-1-piperidinyl-N-oxyl (4-oxo-TEMPO) at 140 C155 C. The reaction rates under these conditions were greater than for similar reactions conducted in anisole. In both of these studies, self-polymerization (thermal polymerization) was observed because of the high temperatures used. Other kinds of living free radical polymerization, such as RAFT and charge transfer polymerization, have also been reported in RTILs (45).

IONIC LIQUIDS, POLYMERIZATION IN

Fig. 3. SEC traces of (a) PS block from BPO in [BMIM]PF6, (b) PS-b-PMMA before extraction, and (c) PS-b-PMMA after extracting with cyclohexane and acetonitrile.

Combined with CuBr/PMDETA catalyst, dendritic polyarylether 2bromoisobutyrates (Gn-Br, n=13) can be used as effective macroinitiators for ATRP of N-hexylmaleimide and styrene in [BMIM]PF6 at lower temperature than in anisole. The afforded polymers had well-dened molecular weights and low polydispersities (1.18 < PDI < 1.36) (46). The two monomers showed a stronger tendency to form alternating structures in RTILs as compared to polymerizations conducted in anisole. Limited solubility of some monomers and polymers in ionic liquids presents opportunities to make diblock copolymers by sequential addition polymerization. Zhang and co-workers (47) demonstrated the formation of PSt-b-PMMA by sequential addition in [BMIM]PF6 through conventional free radical polymerization using benzoyl peroxide (BPO) as initiator. St was polymerized rst and the polymer gradually precipitated out when the conversion reached around 50% due to the insolubility of PSt in [BMIM]PF6 . The chain coils wrapped the macroradicals inside resulting in prolonged lifetimes because of diminished termination. Unreacted St monomer was pumped out on a high vacuum line. After adding the second monomer (MMA), diblock copolymers were formed at room temperature although the re-initiation was not 100% (see Figure 3). Reversal of the polymerization sequence only afforded homopolymer, PMMA. Kubisa and co-workers used a similar strategy to produce PBA-b-PMA and PMA-b-PBA block copolymers (PBA is poly(butyl acrylate) by different sequences of addition of monomers using the ATRP method in the same ionic liquid (48). They found that when conversion of

IONIC LIQUIDS, POLYMERIZATION IN

MA, used as rst monomer, exceeded 70%, a signicant percentage of dead PMA chains contaminated the nal diblock copolymer. Conversely, when MA was added to the solution of PBA in [BMIM]PF6 , clean diblock was formed, essentially free of homopolymer, and BA can be polymerized to complete conversion. Ma and coworkers (42,49) also described successful synthesis of block copolymers, where St was polymerized by chlorine-end-capped PMMA as macroinitiator through reverse ATRP in [BMIM]PF6 , [BMIM]BF4 , and [C12 MIM]BF4 . The effect of solvents on reactivity ratios and sequence length distributions for free radical polymerization has been extensively studied. The calculated reactivity ratios of St and MMA (rSt = 0.381 0.02 and rMMA = 0.464 0.02) in [BMIM]PF6 by nonlinear method (CONTOUR computer program (50)) are significantly different from those (rSt = 0.54 0.04 and rMMA = 0.50 0.04) in benzene at 60 C (51). The boot-strap model (52), polarity of the solvents, interaction between solvent and monomers (eg solventmonomer complex), viscosity and system heterogeneity all possibly contribute to the different reactivity ratios in RTILs and in benzene. [BMIM]PF6 was found to be an efcient plasticizer for PMMA, prepared by in situ radical polymerization in this ionic liquid (53). The polymers have physical characteristics comparable with those containing traditional plasticizers (phthalates, adipates, and trimellitates) and retain greater thermal stability. RTILs are normally used as inert solvents for polymerization. However, Mays and co-workers discovered that polymerization of MMA can be carried out at ambient temperature in the ionic liquid trihexyl-tetradecyl-phosphonium bis(2,4,4trimethylpentyl)phosphinate ([H3 TDP] [(PM3 )2 P]) via a redox-initiated polymerization mechanism (54). The cation of [H3 TDP][(PM3 )2 P] apparently reacts with BPO in a redox reaction, with the cation as reductant and BPO as oxidant. The generated radical efciently initiated polymerization at room temperature to produce PMMA with high yield and high molecular weight.

Ionic Polymerization
Living anionic polymerization usually gives narrower molecular weight distribution polymers as compared to those obtained by living free radical polymerizations. These polymerizations can proceed to complete conversion with far fewer side reactions such as termination and chain transfer. The ability to introduce terminal functional groups by using selective termination agents is an additional advantage of anionic polymerization. On the other hand, the stringent purication requirements for living anionic polymerization put some limits on the commercial synthesis of polymers using this method. Even though RTILs are in some ways ideal potential solvents for ionic polymerizations, because the RTIL can possibly help to stabilize the growing carbanions or carbocations due to their ionic nature, there are no reports in the literature on anionic polymerization in ionic liquids. This is primarily because it is very difcult to adequately purify ionic liquids to the standards required for anionic polymerizations. Our preliminary data (not yet published) indicate that termination occurs due to impurities in the RTIL. Vijayaraghavan and co-workers (55) reported the rst case of cationic polymerization of styrene in N-butyl-N-methylpyrrolidinium bis(triuoromethanesulfonyl)amide

IONIC LIQUIDS, POLYMERIZATION IN

([P14 ][Tf2 N]). The cationic initiator used was a Brnsted acid: bis(oxalato)boric acid (HBOB). The polymerization proceeded in a living manner but the reaction temperature was high (60 C). PSt produced had low molecular weights (1300 1700) and moderate polydispersities (1.31.5). The author suggested that RTILs might promote the dissociation of the acid that can efciently initiate polymerization. Biedron and Kubisa (56) conducted cationic polymerization of styrene in [BMIM]PF6 with a 1-phenetyl chloride/TiCl4 initiating system. Polymers were obtained but chain transfer was signicant, resulting in a lack of control over molecular weight and molecular weight distribution.

Ring-Opening Polymerization
Ring-opening metathesis polymerization of norbornene using cationic allenylidene precatalys, [(p-cymene)RuCl(Pcy3 )( C C CPh2 )][OTf], in a biphasic medium, (1-butyl-2,3-dimethylimidazolium hexauorophosphate ([BDMIM]PF6 )/toluene), was carried out (57). The catalyst was in the ionic liquid phase, and the upper toluene phase dissolved the polymer that was formed. Both the ionic liquid and catalyst could be reused up to six times without signicant loss of catalytic activity and with quantitative yields of polymer. After reloading of catalyst, the catalyst system can still be re-used up to ve times. The enhanced recycling capabilities as compared to other alkylidene catalysts were attributed to the ionic character of the cationic allenylidene complex, while the other systems gradually lost their catalytic abilities within three successive runs. Biedron and co-workers (58) presented the results of the cationic ring opening polymerization of 3-ethyl-3-hydroxymethyloxetane in the most common hydrophilic ionic liquid, [BMIM]BF4 , using the BF3 -Et2 O initiation system. The advantage of carrying out this reaction in ionic liquids over solution or bulk polymerizations is the high reaction temperature that can be used (up to 180 C because intermolecular hydrogen bonding that leads to the formation of aggregates is reduced at the elevated temperature). The intramolecular H-bonding, facilitating intramolecular chain transfer, is not signicantly affected. The multihydroxyl, branched structure was preserved, and the molecular weights of the polymers were in the same range as those made in organic solvents or in bulk. On the contrary, the application of ionic liquids can also help to reduce the reaction temperature. At higher temperatures, side reactions may be signicant. Polycarbonate made by ring opening polymerization of ethylene carbonate experiences decarboxylation when the reaction is carried out at 180200 C (59). In acidic ionic liquids such as [BMIM]Cl-AlCl3 or [BMIM]Cl-SnCl2 , the reaction can take place at temperatures of 100120 C depending on the ionic liquid used. The drawback of these new reaction systems is that decarboxylation was still not negligible, and low levels of ethylene carbonate polymers were formed.

Enzyme-Catalyzed Polymerization
Candida Antarctica lipase (lipase CA) catalyzed formation of biodegradable polyesters in [BMIM]PF6 and [BMIM]BF4 has been reported (60). Ring-opening

10

IONIC LIQUIDS, POLYMERIZATION IN

polymerization of -caprolactone and polycondensation of dicarboxylic acid diesters with 1,4-butanediol were examined. Higher molecular weight products and improved conversions suggest the potential for green polymer chemistry by the combination of nontoxic enzyme catalyst, mild reaction conditions with potentially environmentally benign solvents.

Condensation Polymerization
Polyimides and polyamides are important materials due to their attractive thermal, mechanical, and electrical properties. Usually, the synthesis of high molecular weight polyimides in organic solvents requires high temperatures and the presence of an acidic catalyst. Vygodskii and co-workers (61) studied the condensation polymerization behavior in RTILs composed of different anions and cations. High molecular weight polyimides and polyamides were obtained using RTILs as novel reaction media, without addition of catalyst.

Transition-Metal Catalyzed Polymerization


Poly(phenyl-acetylene) (PPA) is a conjugated polymer with interesting photoconductivity, photoluminescence, nonlinear optical, and membrane properties. It can be obtained by different polymerization methods including radical, cationic, metathesis catalyst or ZieglerNatta polymerizations. High molecular weight PPA can also be obtained by Rh(I) catalyzed phenylacetylene polymerization in ionic liquids such as n-butylpyridinim tetrauoroborate ([bupy]BF4 ) or [BMIM]PF4 (62). The catalyst used was either (diene)Rh(acac) or [(diene)RhCl2 ]2 , and the cocatalyst used was triethylamine. The polymer was separated from the ionic liquid using either extraction with toluene or ltration by adding methanol into ionic liquids to form a suspension of PPA in the solvent mixture methanol/RTIL. Extraction usually facilitates catalyst recycling, but cannot permit the maximum product recovery. While more polymers can be recovered through the ltration method, the Rh(I) complex solution after ltration shows no catalytic activity. The molecular weights ranged from 55,000 to 200,000. Rogers and co-workers (63) used 1-hexylpyridinium bis(triuoromethanesulfonyl) imide ([C6 Pyr][NTf2 ])/methanol as a solvent pair for palladium-catalyzed alternating copolymerization of styrene and carbon monoxide, in which the palladium catalyst was LPd(OAc)2 (L=2,2 -bipyridine and 1,10-phenanthroline). [C6 Pyr][NTf2 ] and catalyst can be recycled, and the yields and molecular weights were higher as compared to those obtained when the polymerization was carried out in methanol alone. Furthermore, the catalyst stability and propagation rate were improved due to the inhibited chain transfer and catalyst decomposition in [C6 Pyr][NTf2 ].

Electrochemical Polymerization
Electrochemical polymerization is a major method to synthesize conducting polymer for potential application in energy storage devices, electrochromic devices,

IONIC LIQUIDS, POLYMERIZATION IN

11

and light-emitting diodes. This process shows some advantages over chemical synthesis including faster reactions, simple procedures, generation of the polymer in the doped state, and easy control of the lm thickness. Naudin and coworkers (64) have used ionic liquids for the electrochemical polymerization of poly(3-(4-uorophenyl)thiophene) (PFPT). The electrochemical behavior was similar to that in common nonaqueous electrolyte. However, X-ray photoluminescence spectroscopy revealed the presence of some ionic liquid residue in the formed lm. Other kinds of conjugated polymers such as polypyrrole, polyaniline, and unsubstituted polythiophene were also synthesized in ionic liquid (65).

Other Applications of Ionic Liquids in Polymers Systems


Polyaniline nanoparticles with diameters from about 3080 nm can be fabricated by interfacial polymerization at the interface between aqueous media and RTILs (66). Electronic devices such as electrochemical mechanic actuators have improved cycle life and larger generated strain using [BMIM]PF6 , [BMIM]BF4 , or [BMIM][NTf2 ] as electrolytes as compared to traditional organic or aqueous solvents, because ionic liquids have high ion conductivity, large electrochemical windows, and fast ion mobility (67). Single-ion conductive membrane material, made out of polymerizable ionic liquids by incorporating some ethylene oxide spacer cross-linkers, has high ionic conductivity (1.3710 4 S cm 1 ), and the lm formed is transparent and free standing (68). In order to overcome migration of ions of RTILs along with potential gradient with target ions in membrane application, poly(VdF-co-HFP) containing zwitterionic ionic liquids was reported as a new polymer gel electrolyte which only transports target ions (69). RTILs have been used as porogens in cross-linked polymer systems formed by free radical polymerization of cross-linking monomers (70). RTILs are also effective solvents for regenerating cellulose via the aid of microwave irradiation (71). Forming hydrogen-bonding between hydroxyl functions on the polymer chains and anions of RTILs such as Cl , Br , and SCN , which are hydrogen bond acceptors, can break the intermolecular hydrogen bonding between linear glucose polymer chains. A recent review on application of ionic liquids in polymerization is available in the literature (72).

Conclusions
Room temperature ionic liquids have proved over recent years to be useful and unique reaction media for a variety of chemical reactions. Some potential benets of working with RTILs as solvents for polymerization are wide and varied (Table 1): Faster reaction rates and better selectivities are often observed in RTILs as compared to those in common organic solvents. In addition to these advantages, the nonvolatility, nonammability, and recycling potential also help RTILs become a preferred alternative reaction medium to meet environmental and other requirements. Although polymerization in 1st generation RTILs was not very attractive when compared to similar reactions carried out in organic solvents, the 2nd generation RTILs are showing much greater promise. This is evidenced by

12

IONIC LIQUIDS, POLYMERIZATION IN

Table 1. Potential Benets of Polymerization in RTILs Ability to rapidly generate high molecular weight polymers with low residual monomer Simplied methods to synthesize complex molecular architectures such as block- and graft copolymers, and copolymers with new monomer sequences, as compared with more complicated techniques such as anionic, cationic and living radical polymerization Ability to design reactions to be carried out at much higher temperatures, as RTIL volatility is much lower and thermal stability is much higher than traditional solvent Improvement in polymer physical properties, such as tensile strength, elastic modulus and impact strength, based on the reaction mechanism in RTILs Ability to work with biological components in a nondenaturing environment during reactions that incorporate enzymes, proteins, and other biologically active agents into polymer networks.

the increasing number of papers published in this area every year, and interesting and important discoveries. When conventional free radical polymerizations are carried out in RTILs, there is an increase in the propagation rate due to the high polarity of the RTIL and a decrease in the termination rate because of the high viscosity of the RTILs, and thus the overall reaction rate is greatly increased. For living/controlled polymerization such as ATRP, the reactions in RTILs are still controllable with faster reaction rates and easy separation of catalysts and/or ligands from the resulting polymers. Several other polymerization processes have been conducted in RTILs, with the notable exception of anionic polymerization, which needs careful work in order to nd the right RTILs, purication techniques, and polymerization conditions.

BIBLIOGRAPHY
1. P. Walden, Bull. Acad. Imper. Sci. (St. Petersburg) 1800 (1914). 2. C. L. Hussey, Pure Appl. Chem. 60, 17631772 (1988). 3. K. R. Seddon, in G. Mamantov and R. Marassi, eds., Molten Salt Chemistry, Reidel Publishing Co., Dordrecht, The Netherlands, 1987, p. 365. 4. J. S. Wilkes, J. A. Levisky, R. A. Wilson, and C. L. Hussey, Inorg. Chem. 21, 12631264 (1982). 5. C. L. Hussey, in G. Mamantov and C. Mamantov, eds., Advances in Molten Salts Chemistry, Vol. 5, Elsevier Science Publishing Co., New York, 1983, pp. 185230. 6. (a) M. K. Dieter, C. J. Dymek Jr., N. E. Heimer, J. W. Rovang, and J. S. Wilkes, J. Am. Chem. Soc. 110, 2722 (1988); (b) J. S. Wilkes and M. J. Zaworotko, J. Chem. Soc., Chem. Commun. 965967 (1992); (c) H. L. Ngo, K. LeCompte, L. Hargens, and A. B. McEwen, Thermochim. Acta 357358, 97102 (2000). 7. S. N. V. K. Aki, J. F. Brennecke, and A. Samanta, Chem. Commun. 413 (2001). 8. H. Zhang, K. Hong, and J. W. Mays, in Abstract of Papers, ACS 227th National Meeting, Anaheim, CA, USA, 2004. 9. M. Freemantle, Chem. Eng. News 32 (Mar. 30, 1998); 37 (May 15, 2000); 21 (Jan. 1, 2001); 26 (May 3, 2004). 10. B. Cornils and W. A. Herrmann, in B. Cornils and W. A. Herrmann, eds., Applied Homogeneous Catalysis with Organometallic Compounds, Weinheim, New York, 1996, Chap. 4.1, p. 1167.

IONIC LIQUIDS, POLYMERIZATION IN


11. 12. 13. 14. 15. 16.

13

17. 18.

19. 20. 21. 22. 23. 24. 25.

26.

27.

28. 29.

30. 31. 32. 33. 34. 35. 36.

T. Fischer, A. Sethi, T. Welton, and J. Woolf, Tetrahedron Lett. 40, 793 (1999). A. Stark, B. L. MacLean, and R. D. Singer, J. Chem. Soc., Dalton Trans. 63 (1999). Y. Chauvin, L. Mussmann, and H. Olivier, Angew. Chem. Int. Ed. 34, 2698 (1995). C. P. Mehnert, R. A. Cook, N. C. Dispenziere, and M. Afeworki, J. Am. Chem. Soc. 124, 12932 (2002). M. Badri, J. J. Brunet, and R. Perron, Tetrahedron Lett. 33, 4435 (1992). (a) J. E. L. Dullius, P. A. Z. Suarez, S. Einloft, R. F. de Souza, J. Dupont, J. Fischer, and A. De Cian, Organometallics 17, 815 (1988); (b) S. M. P. Silva, A. Z. Suarez, R. F. de Souza, and J. Dupont, Polym. Bull. 40, 401 (1998). D. E. Kaufmann, M. Nouroozian, and H. Henze, Syn. Lett. 1091 (1996). (a) C. J. Mathews, P. J. Smith, and T. Welton, Chem. Commun. 1249 (2000); (b) R. Rajagopal, D. V. Jarikote, and K. V. Srinivasan, Chem. Commun. 616 (2002); (c) J. McNulty, A. Capretta, J. Wilson, J. Dyck, G. Adjabeng, and A. Robertson, Chem. Commun. 1986 (2002). (a) S. B. Park, H. Alper, Chem. Commun. 1306 (2004); (b) T. Fukuyama, M. Shinmen, S. Nishitani, M. Sato, and I. Ryu, Org. Lett. 4, 1691 (2002). (a) Q. Yao and Y. Zhang, Angew. Chem. Int. Ed. 42, 3395 (2003); (b) N. Audic, H. Clavier, M. Mauduit, J. C. Guillemin, J. Am. Chem. Soc. 125, 9248 (2003). I. A. Ansari and R. Gree, Org. Lett. 4, 1507 (2002). D. Kim and D. Chi, Angew. Chem. Int. Ed. 43, 483 (2004). R. T. Carlin, R. A. Osteryoung, J. S. Wilkes, and J. Rovang, Inorg. Chem. 29, 3003 (1990). R. T. Carlin and J. S. Wilkes, J. Mol. Catal. 63, 125 (1990). (a) M. F. Pinheiro, R. S. Mauler, and R. F. de Souza, Macromol. Rapid Commun. 22, 425 (2001); (b) P. Wasserscheid, C. M. Gordon, C. Hilgers, M. J. Muldoon, and I. R. Dunkin, Chem. Commun. 1186 (2001); (c) O. Stenzel, R. Brull, U. M. Wahner, R. D. Sanderson, and H. G. Raubenheimer, J. Mol. Catal. A: Chem. 192, 217 (2003). (a) EP 558187 (1993), P. W. Ambler, P. K. G. Hodgson, and N. J. Stewart; (b) WO 9521871 (1995), A. A. K. Abdul-Sada, P. W. Ambler, P. K. G. Hodgson, K. R. Seddon, and N. J. Stewart; (c) EP 791643 (1997), P. M. Atkins, M. R. Smith, and B. Ellis; (d) WO 0032658 (2000), V. Murphy. (a) V. M. Kobryanskii and S. A. Arnautov, J. Chem. Soc., Chem. Commun. 9, 727 (1992); (b) V. M. Kobryanskii and S. A. Arnautov, Makromol. Chem. 193, 455 (1992); (c) D. C. Trivedi, J. Chem. Soc., Chem. Commun. 544 (1989); (d) S. A. Arnautov, Synth. Met. 84, 295 (1997); (e) L. M. Goldenberg, R. A. Osteryoung, Synth. Met. 64, 63 (1994). R. T. Carlin and R. A. Osteryoung, J Electrochem. Soc. 141, 1709 (1994). (a) C. J. Hawker, A. W. Bosman, and E. Harth, Chem. Rev. 101, 3661 (2001); (b) M. Kamigaito, T. Ando, and M. Sawamoto, Chem. Rev. 101, 3689 (2001); (c) K. Matyjaszewski and J. Xia, Chem. Rev. 101, 2921 (2001). A. J. Carmicheal, D. M. Haddleton, S. A. F. Bon, and K. R. Seddon, Chem. Commun. 1237 (2000). (a) K. Hong, H. Zhang, J. W. Mays, A. N. Visser, C. S. Brazel, J. D. Holbrey, W. M. Reichert, and R. D. Rogers, Chem. Commun. 1368 (2002); (b) M. G. Benton and C. S. Brazel, Polym. Int. 53, 1113 (2004). (a) S. Harrison, S. R. MacKenzie, and D. M. Haddleton, Chem. Commun. 2850 (2002); (b) S. Harrison, S. R. MacKenzie, and D. M. Haddleton, Macromolecule 36, 5072 (2003). A. L. Cheng, Y. Zhang, T. Zhao, and H. Wang, Macromol. Symp. 216, 9 (2004). H. Zhang, K. Hong, and J. W. Mays, in R. Rogers and C. S. Brazel, eds., Ionic Liquids in Polymer System, ACS Symposium Series, Anaheim CA, Spring, 2004, in press. V. Strehmel, H. Kraudelt, H. Wetzel, E. G rnitz, and A. Laschewsky, in R. Rogers and o C. S. Brazel, eds., Ionic Liquids in Polymer System, ACS Symposium Series, Anaheim CA, Spring, 2004, in press.

14

IONIC LIQUIDS, POLYMERIZATION IN

37. (a) J. F. Deye, T. A. Berger, and A. G. Anderson, Anal. Chem. 62, 615 (1990); (b) A. J. Carmicheal and K. R. Seddon, J. Phys. Org. Chem. 13, 591 (2000); (c) S. V. Dzyuba and R. A. Bartsch, Tetrahedron Lett. 43, 4657 (2002). 38. H. Zhang, L. Bu, M. Li, K. Hong, A. E. Visser, R. D. Rogers, and J. W. Mays, in R. D. Rogers and K. R. Seddon, eds., Ionic Liquids: Industrial Application for Green Chemistry, ACS Symposium, 2001. 39. T. Biedron and P. Kubisa, Macromol. Rapid Commun. 22, 1237 (2001). 40. T. Biedron and P. Kubisa, Polym. Int. 52, 1584 (2003). 41. T. Sarbu and K. Matyjaszewski, Macromol. Chem. Phys. 202, 3379 (2001). 42. (a) H. Ma, X. Wan, X. Chen, and Q. Zhou, J. Polym. Sci.. Part A: Polym. Chem. 41, 143 (2003); (b) H. Ma, X. Wan, X. Chen, and Q. Zhou, Polymer 44, 5311 (2003). 43. H. Zhang, K. Hong, and J. W. Mays, Polym. Bull. 52, 9 (2004). 44. J. Ryan, F. Aldabbagh, P. B. Zetterlund, and B. Yamada, Macromol. Rapid Commun. 25, 930 (2004). 45. (a) S. Perrier, T. P. Davis, A. J. Carmichael, and D. M. Haddleton, Chem Commun. 2226 (2002); (b) S. Perrier, T. P. Davis, A. J. Carmichael, and D. M. Haddleton, Euro. Polym. J. 39, 417 (2003); (c) R. Vijayaraghavan and D. R. MacFarlane, Aust. J. Chem. 57, 129 (2004). 46. Y. Zhao, J. Zhang, J. Jiang, C. Chen, and F. Xi, J. Polym. Sci., Part A: Polym. Chem. 40, 3360 (2002). 47. H. Zhang, K. Hong, and J. W. Mays, Macromolecules 35, 5738 (2002). 48. T. Biedron and P. Kubisa, J Polym. Sci., Part A: Polym. Chem. 40, 2799 (2002). 49. (a) H. Ma, X. Wan, X. Chen, and Q. Zhou, J Polym. Sci., Part A: Polym. Chem. 41, 143 (2003); (b) H. Ma, X. Wan, X. Chen, and Q. Zhou, Polymer 44, 5311 (2003). 50. A. M. van Herk, J. Chem. Educ. 72, 138 (1995). 51. H. Zhang, K. Hong, M. Jablonsky, and J. W. Mays, Chem Commun. 1356 (2003). 52. H. Harwood, J. Macromol. Sci. Macromol Symp. 10/11, 331 (1987). 53. M. P. Scott, C. S. Brazel, M. G. Benton, J. W. Mays, J. D. Holbrey, and R. D. Rogers, Chem. Commun. 1370 (2002). 54. H. Zhang and J. W. Mays, manuscript in preparation (2005). 55. R. Vijayaraghavan, and D. R. MacFarlane, Chem. Commun. 700 (2004). 56. T. Biedron and P. Kubisa, J. Polym. Sci., Part A: Polym. Chem. 42, 3230 (2004). 57. S. Csihony, C. Fischmeister, C. Bruneau, I. T. Horvath, and P. H. Dixneuf, New J. Chem. 26, 1667 (2002). 58. T. Biedron, M. Bednarek, and P. Kubisa, Macromol. Rapid Commun. 25, 878 (2004). 59. J. Kadokawa, Y. Iwasaki, and H. Tagaya, Macromol. Rapid Commun. 23, 757 (2002). 60. H. Uyama, T. Takamoto, and S. Kobayashi, Polym. J. 34, 94 (2002). 61. Y. S. Vygodskii, E. I. Lozinskaya, and A. S. Shaplov, Macromol. Rapid Commun. 23, 676 (2002). 62. P. Mastrorilli, C. F. Nobile, V. Gallo, G. P. Suranna, and G. Farinola, J. Mol. Catal. A: Chem. 184, 73 (2002). 63. M. A. Klingshirn, G. A. Broker, J. D. Holbrey, K. H. Shaughnessy, and R. D. Rogers, Chem Commun. 1394 (2002). 64. E. Naudin, H. A. Ho, S. Branchaud, L. Breau, D. B langer, J. Phys. Chem. B 106, 10585 e (2002). 65. (a) K. Sekiguchi, M. Atobe, and T. Fuchigami, J. Electroanal. Chem. 557, 1 (2003); (b) J. M. Pringle, J. Efthimiadis, P. C. Howlett, J. Efthimiadis, D. R. MacFarlane, A. B. Chaplin, S. B. Hall, D. L. Ofcer, G. G. Wallace, and M. Forsyth, Polymer 45, 1447 (2004); (c) J. H. Shi, C. H. Yang, G. Y. Gao, and Y. F. Li, Chin. J. Chem. Phys. 17, 503 (2004). 66. H. Gao, T. Jiang, B. Han, Y. Wang, J. Du, Z. Liu, and J. Zhang, Polymer 45, 3017 (2004).

IONIC LIQUIDS, POLYMERIZATION IN

15

67. (a) J. Ding, D. Zhou, G. Spinks, G. Wallace, S. Forsyth, M. Forsyth, and D. MacFarlane, Chem. Mater. 15, 2392 (2003); (b) W. Lu, A. G. Fadeev, B. Qi, E. Smela, B. R. Mattes, J. Ding, G. Spinks, J. Mazurkiewics, D. Zhou, G. G. Wallace, D. R. MacFarlane, S. A. Forsyth, and M. Forsyth, Science 297, 983 (2002). 68. S. Washiro, M. Yoshizawa, H. Nakajima, and H. Ohno, Polymer 45, 1577 (2004). 69. H. Ohno, M. Yoshizawa, and W. Ogihara, Electrochim. Acta. 48, 2079 (2003). 70. P. Snedden, A. I. Cooper, K. Scott, and N. Winterton, Macromolecule 36, 4549 (2003). 71. R. P. Swatloski, S. K. Spear, J. D. Holbrey, and R. D. Rogers, J. Am. Chem. Soc. 124, 4974 (2002). 72. P. Kubisa, Prog. Polym. Sci. 29, 3 (2004).

HONGWEI ZHANG University of Tennessee JIMMY W. MAYS Oak Ridge National Laboratory

Вам также может понравиться