Polyethylene-Based Block Copolymers for Anion Exchange

Sep 10, 2015 - US Army Research Laboratory, Aberdeen Proving Ground, Maryland 21005, United ... S. Piril Ertem , Tsung-Han Tsai , Soenke Seifert , Fre...
0 downloads 0 Views 9MB Size
Article pubs.acs.org/Macromolecules

Polyethylene-Based Block Copolymers for Anion Exchange Membranes Yifan Li,† Ye Liu,‡ Alice M. Savage,§ Frederick L. Beyer,§ Soenke Seifert,∥ Andrew M. Herring,‡ and Daniel M. Knauss*,†

Downloaded by UNIV OF MASSACHUSETTS AMHERST on September 14, 2015 | http://pubs.acs.org Publication Date (Web): September 10, 2015 | doi: 10.1021/acs.macromol.5b01457



Department of Chemistry and Geochemistry and ‡Department of Chemical and Biological Engineering, Colorado School of Mines, Golden, Colorado 80401, United States § US Army Research Laboratory, Aberdeen Proving Ground, Maryland 21005, United States ∥ X-ray Science Division, Argonne National Laboratory, Argonne, Illinois 60439, United States S Supporting Information *

ABSTRACT: Block copolymer membranes with a semicrystalline polyethylene component were prepared by anionic polymerization and postpolymerization functionalization reactions. Polybutadiene-b-poly(4-methylstyrene) (PB-b-P4MS) precursors with four different block compositions and 92−95% 1,4-content in the polybutadiene block were produced by living anionic polymerization in a nonpolar solvent. The polybutadiene block was subsequently hydrogenated to prepare a polyethylene block, and the hydrogenated block copolymers were then brominated at the arylmethyl group of the P4MS block. Subsequent quaternization reaction with trimethylamine led to the anion exchange membranes. The degree of crystallinity in the polyethylene block was determined by differential scanning calorimetry to be approximately 24−27%. The postpolymerization modification reactions were examined by 1H NMR and IR spectroscopy. The amount of quaternary ammonium groups was quantified by ion exchange capacity (IEC) measurements. Membranes with IEC’s ranging from 1.17 to 1.92 mmol/g were prepared. The IEC was varied by changing the relative amount of P4MS in the precursor block copolymer, and water uptake and ionic conductivity were found to increase with increasing IEC. Small-angle Xray scattering (SAXS) experiments and transmission electron microscopy (TEM) showed phase-separated, bicontinuous structures at all compositions. Materials with higher IEC show improved ionic conductivity as well as lower activation energy of ion conduction compared to less functionalized membranes. The hydroxide conductivity of the block copolymer membrane with an IEC of 1.92 mmol/g reached 73 mS/cm at 60 °C in water. Tensile measurements indicated excellent mechanical properties of the semicrystalline membranes for potential use as alkaline fuel cell membrane materials.



INTRODUCTION

mechanical properties and ionic conductivity under the operating conditions of an alkaline fuel cell. Polyethylene has been investigated as a component of blended membrane materials for PEMFCs to provide mechanical properties and to control membrane swelling and methanol permeability.13,14 AFC membranes containing polyethylene as membrane material have also been recently reported. Cationic functionalized cyclooctene has been randomly copolymerized with cyclooctene by ring-opening metathesis polymerization and subsequently hydrogenated to form solvent processable polyethylene AEMs with good mechanical and ion conductive properties.12,15 Another strategy to incorporate polyethylene in AEMs investigated the radiation grafting reaction of vinylbenzyl chloride monomer onto commercially available polyethylene and subsequent conversion

Alkaline fuel cells (AFCs) have been successfully implemented by NASA on the space shuttles using pure hydrogen and precious metal catalysts. However, the real promise of AFCs are their potential advantages over proton exchange membrane fuel cells (PEMFCs), which include their facile electrochemical kinetics for complex fuels and the possibility of utilizing nonprecious-metal catalysts, i.e., nickel1 or silver.2−4 The use of liquid electrolytes such as aqueous potassium hydroxide in traditional AFCs results in carbonate or bicarbonate precipitates from the reaction of potassium hydroxide with carbon dioxide in the air, resulting in reduced performance.5 Recently, research has focused on the development of anion exchange membranes (AEMs) with various tethered cationic functional groups, including quaternary ammonium6−10 or phosphonium,11,12 as solid polymer electrolytes for AFCs to eliminate the formation of precipitates. However, it remains a major challenge to develop a chemically stable AEM with sufficient © XXXX American Chemical Society

Received: July 2, 2015 Revised: September 2, 2015

A

DOI: 10.1021/acs.macromol.5b01457 Macromolecules XXXX, XXX, XXX−XXX

Article

Downloaded by UNIV OF MASSACHUSETTS AMHERST on September 14, 2015 | http://pubs.acs.org Publication Date (Web): September 10, 2015 | doi: 10.1021/acs.macromol.5b01457

Macromolecules

(Sigma-Aldrich, 99.9%), and trimethylamine (Sigma-Aldrich, 25 wt % aqueous solution) were used without further purification. All polymerization glassware, glass syringes, needles, and stir bars were dried in an oven at 180 °C overnight, and the glassware was further flame-dried under an argon purge after being equipped with a stir bar and sealed with a rubber septum. Polybutadiene-b-poly(4-methylstyrene) (PB-b-P4MS) by Anionic Polymerization. Polymerizations were conducted at room temperature. A series of block copolymers (A, B, C, and D) were synthesized following the same polymerization procedure while varying the relative composition of butadiene and 4-methylstyrene blocks. A typical procedure (copolymer A) is as follows. Dry cyclohexane (50 mL) from a column solvent purifier was added to a rubber septum-sealed, flame-dried, and argon-purged 100 mL round-bottom flask equipped with a stir bar. sec-BuLi (100 μL, 0.14 mmol) was then injected with a 250 μL syringe. The solution turned to a pale yellow and was allowed to stir for 15 min. An argon-purged pressure vessel containing butadiene (7.22 g, 0.133 mol) over calcium hydride was connected to the round-bottom flask with a cannula. The pressure vessel was gently warmed, and the butadiene was transferred from the pressure vessel to the reaction flask over 15−20 min. The cannula was then removed, and the reaction was left to stir for 24 h. An aliquot (1 mL) of reaction solution was removed by syringe after 24 h and added to argon-purged methanol to terminate a sample of polybutadiene. 4-Methylstyrene (5.10 mL, 39.5 mmol) was introduced into the reaction flask using a 5 mL syringe. A bright orange color was immediately observed, and the reaction solution was left to stir for 12 h. After the reaction time, a small amount of argon-purged methanol was introduced by syringe to terminate the living polymer chain ends as observed by the disappearance of the orange color. The polymer solution was precipitated into excess methanol containing 2,6-di-tertbutyl-p-cresol (0.2 g, about 0.02 wt % of polymer) as radical inhibitor, and the polymer was isolated by filtration, washed with more methanol, and dried in a vacuum oven at 40 °C overnight. 1H NMR (500 MHz, CDCl3, δ): 1.34−1.65 (m, 2H; −CH2CH(Ar)), 1.66−2.06 (s, 1H; −CH(Ar)), 2.04 (s, 2H; trans −CH2−), 2.09 (s, 2H; cis −CH2−), 2.07−2.51 (s, 3H; −CH3); 4.91−5.02 (m, 2H; −CH2− CH(CHCH2)−), 5.32−5.49 (m, 2H; −CH2−CHCH−CH2−), 5.53−5.63 (m, 1H; −CH2−CH(CHCH2)−), 6.15−7.16 (m, 4H; ArH). Hydrogenation of PB-b-P4MS Block Copolymer. Hydrogenation of the polybutadiene block of the copolymer was accomplished using p-toluenesulfonyl hydrazide to form polyethylene-b-poly(4-methylstyrene) (PE-b-P4MS). The general procedure (described for copolymer A) is as follows: PB-b-P4MS (1.012 g, 0.015 mol butadiene repeat units) was weighed into a 250 mL threeneck round-bottom flask equipped with a magnetic stirbar, a condenser, and an argon bubbler. Xylenes (80 mL) was added into the reaction flask, and after dissolution, the reaction flask was immersed in an oil bath and heated to 120 °C. p-Toluenesulfonyl hydrazide (22.3 g, 0.12 mmol) was added, and the solution was allowed to stir for 12 h. The hot solution was precipitated in an excess of methanol and filtered, and the white powder was washed with hot deionized water to remove byproducts. The final polymer was recovered by filtration and dried in a vacuum oven at 80 °C overnight. 1 H NMR (500 MHz, CDCl3, δ): 1.2−1.65 (m, 2H; −CH2CH(Ar)), 1.54 (s 2H; −CH2−); 1.66−2.06 (s, 1H; −CH(Ar)); 2.07−2.51 (s, 3H; −CH3), 6.15−7.16 (m, 4H; ArH). Bromination of PE-b-P4MS Block Copolymer. Bromination of the 4-methyl groups of the poly(4-methylstyrene) block in PE-b-P4MS was done using NBS using a modified procedure from that described in the literature.28−30 In a typical reaction (described for copolymer A), PE-b-P4MS (1.38 g, 2.70 mmol of 4-methylstyrene repeat units) was dissolved in 50 mL of carbon tetrachloride in a two-neck 150 mL round-bottom flask. The solution was heated with an oil bath to 72 °C with stirring under an argon atmosphere. AIBN (6.5 mg, 0.040 mmol) was first added to the solution followed by NBS (0.432 g, 2.43 mmol, 90% to the amount of 4-methylstyrene) under an argon atmosphere. The solution was refluxed and allowed to stir for 24 h. The solution turned yellow during the bromination. The hot solution was

of benzyl chloride groups to quaternary ammonium cations.16,17 AEMs based on polyethylene have also been reported by blending anion exchange particles with linear polyethylene and a water-soluble additive to prepare heterogeneous membranes.18 Although cationic functionalized polyethylene has been designed as a component of AEMs, the studies are based on polymers with either randomly distributed cationic functional groups or randomly grafted cationic chains. Other research on AEMs has been directed toward preparing well-defined block copolymers having a hydrophobic block that forms the matrix for mechanical strength and a hydrophilic (cationic) block for conductivity.19−28 Amphiphilic block copolymers for AEMs can provide phase-separated materials with controlled morphology for ion conduction by controlling the block composition. A notable strategy in the synthesis of well-defined block copolymers is to use living anionic polymerization of vinyl monomers. Anionic polymerization offers various benefits to prepare block copolymers as the reactive site in a living anionic polymerization persists throughout the reaction such that no termination or chain transfer occurs on the time scale of the reaction. Because there is no termination, monomers can be sequentially polymerized to yield well-defined block copolymers. Through the use of living anionic polymerization, polymers can be synthesized with controlled molecular weight, composition, and functional groups. Anionic polymerization can be used to control the microstructure of butadiene polymerizations, and high 1,4-structured polybutadiene can be obtained by conducting anionic polymerization of butadiene in nonpolar solvents. Subsequent hydrogenation of high 1,4content polybutadiene leads to a polyethylene-like structure that is semicrystalline. Through this process, polyethylenebased block copolymers as AEMs can be prepared by the sequence of living anionic polymerization followed by hydrogenation and postpolymerization functionalization. In this study, polybutadiene-b-poly(4-methylstyrene) (PB-bP4MS) block copolymers were synthesized by living anionic polymerization in a nonpolar solvent. The high 1,4-content polybutadiene block was hydrogenated to a polyethylene-like block, and the poly(4-methylstyrene) block was ultimately converted to an ion conductive block. The following research investigates in detail the block copolymerization, hydrogenation, and cationic group functionalization reactions. The polyethylene block copolymer membranes are prepared to study their properties, including water uptake, hydration number, ionic conductivity, and morphology.



EXPERIMENTAL SECTION

Materials. 4-Methylstyrene (Sigma-Aldrich, 96%) was dried over calcium hydride and distilled under reduced pressure two times before use. 1,3-Butadiene (Sigma-Aldrich, ≥99%) was transferred from a gas cylinder to an argon purged pressure vessel containing calcium hydride and a stir bar while cooling the vessel to −78 °C with an acetone/dry ice bath. The butadiene stirred over the calcium hydride for approximately 20 min right before use. sec-Butyllithium (sec-BuLi) (1.4 M in cyclohexane) solution and di-n-butylmagnesium (1.0 M in heptane) solution were obtained from Sigma-Aldrich and used as received. 2,2′-Azobis(isobutyronitrile) (AIBN) (Sigma-Aldrich, 98%) was purified by recrystallization in methanol, and N-bromosuccinimide (NBS) was recrystallized from deionized water. Cyclohexane (HPLC grade, Fisher) was passed through columns on a Pure Solv solvent purification system (Innovative Technology). 2,6-Di-tert-butyl-p-cresol (BHT, Eastman), xylenes (Mall, ACS reagent grade), p-toluenesulfonyl hydrazide (TSH) (Sigma-Aldrich, 97%), carbon tetrachloride B

DOI: 10.1021/acs.macromol.5b01457 Macromolecules XXXX, XXX, XXX−XXX

Article

Downloaded by UNIV OF MASSACHUSETTS AMHERST on September 14, 2015 | http://pubs.acs.org Publication Date (Web): September 10, 2015 | doi: 10.1021/acs.macromol.5b01457

Macromolecules precipitated in an excess of methanol and filtered, and the yellow powder was redissolved and precipitated two more times to remove byproducts. The polyethylene-b-poly(vinylbenzyl bromide) (PE-bPVBBr) was isolated by filtration and dried in a vacuum oven at 80 °C overnight. 1H NMR (500 MHz, CDCl3, δ): 1.2−1.65 (m, 2H; −CH2CH(Ar)); 1.54 (s, 2H; −CH2−); 1.66−2.06 (s, 1H; −CH(Ar)); 2.07−2.51 (s, 3H; −CH3); 4.20−4.76 (s, 2H; −CH2Br); 6.04−7.62 (m, 4H; ArH). PE-b-PVBBr Block Copolymer Membrane Preparation. PE-bPVBBr block copolymer powder was dissolved in xylenes at 90 °C with a concentration of 3 g/100 mL. The solution was drop-cast onto an 80 °C preheated glass slide, and the solvent was allowed to evaporate. The resulting membrane was placed in a vacuum oven at 80 °C overnight to remove remaining xylenes. Amination of PE-b-PVBBr Block Copolymer. The PE-b-PVBBr membrane was soaked in a 25 wt % aqueous trimethylamine solution for 48 h to convert benzyl bromide to benzyltrimethylammonium bromide ([VBTMA][Br]). The PE-b-P[VBTMA][Br] membrane was then repeatedly immersed in fresh deionized water multiple times, rinsed thoroughly with deionized water, and then dried in a vacuum oven at 60 °C overnight. Ion Exchange of Hydroxide Membrane. The quaternary ammonium functionalized block copolymer membranes were soaked in 1 M KOH solution for 48 h to exchange bromide to hydroxide. The membranes were subsequently immersed in deionized water for a few hours with the deionized water changed several times during the soaking process. Characterization. The molecular weights and molecular weight distributions for polybutadiene and PB-b-P4MS were determined by gel permeation chromatography (GPC) using a Waters 600 HPLC pump equipped with a Wyatt Technology Optilab RI detector and a Wyatt Technology miniDAWN multiangle laser light scattering (MALLS) detector. Elutions were carried out with two Polymer Laboratories PLgel 5 μm Mixed-D columns at room temperature with THF at a flow rate of 1.0 mL/min. Samples were dissolved in THF at 5 mg/mL, and molecular weights were measured using Astra software supplied by Wyatt Technology. 1H NMR spectra of copolymers were obtained using a JEOL ECA 500 MHz spectrometer. All polymer samples were dissolved in CDCl3 using tetramethylsilane (TMS) or CHCl3 as standard. IR measurements were performed by attenuated total reflectance (ATR)−Fourier transform infrared (FTIR) spectroscopy using a Thermo-Electron Nicolet 4700 spectrometer with a Smart Orbit attachment in a spectral range of 500−4000 cm−1. Glass transition temperatures (Tg) and crystallinity of membranes were determined by differential scanning calorimetry (DSC) on a TA DSC20 running TA software with a heating rate of 10 °C/min under a nitrogen purge. The degree of crystallinity was calculated by comparing the enthalpy of the DSC tested sample with the enthalpy value for 100% crystalline polyethylene.31,32 Theoretical ionic exchange capacities (IECs) were calculated based on the complete conversion of benzyl bromide to quaternary ammonium. Titrated IEC was determined for samples in the bromide form by the Mohr’s titration method.33,34 The membranes were dried in a vacuum oven at 60 °C for 24 h, and their dry weights were quickly measured. About 0.10 g of membrane was accurately weighed and soaked in 50 mL of 0.2 M NaNO3 solution for 8 h. After 8 h, the film was transferred into fresh 0.2 M NaNO3 solution for another 8 h. This process was repeated a total of three times in 24 h to exchange the bromide ion to nitrate. The NaNO3 solution was titrated to determine the amount of free bromide ions with 0.0992 M AgNO3 solution by using K2CrO4 as color indicator. The bromide titration was repeated three times, and the titrated IEC was calculated with the equation IEC (mmol/g) =

In order to determine water uptake of AEMs, membranes were presoaked in deionized water. After 24 h of soaking, the fully hydrated membranes were removed from the water and the extra water on the membrane surface was blotted dry with KimWipes, and the membrane was weighed immediately. The quaternized membranes were immersed again into deionized water for another 15 min to let them rehydrate. This method was repeated five times to obtain the average mass of hydrated membranes. The membranes were finally dried under vacuum at 60 °C for 24 h. The water uptake (WU) was determined by WU (%) =

mdry

× 100%

where mhydrated is hydrated mass of membrane after removing all surface water and mdry is dry mass of membrane after drying in the vacuum oven. Synchrotron small-angle X-ray scattering (SAXS) experiments were performed at the Advanced Photon Source at Argonne National Laboratory on beamline 12-ID-B under conditions described previously.35 An environmental chamber was used to control the temperature and humidity.36,37 The humidity was stepped to 25%, 50%, 75%, and 95% relative humidity (RH) over 20, 20, 40, and 90 min, respectively, in order to hydrate the membrane. Transmission electron microscopy was performed on unstained sections prepared by cryo-ultramicrotomy of the membranes in the bromide counterion form. Each membrane was microtomed using a Leica UC7 ultramicrotome equipped with a Leica FC7 cryochamber, generating sections ranging in thickness between 50 and 70 nm. A JEOL JEM-2100F TEM and a Gatan 806 high-angle annular dark field scanning TEM (HAADF STEM) detector were used to collect dark field images from each sample. The TEM was operated at 200 kV, with a 40 μm condenser aperture, a HAADF STEM collection angle of 48− 168 mrad, and 0.5 nm spot size. A Gatan digital micrograph was used to collect and analyze the data.38 The ionic conductivity of quaternized membranes was measured by electrochemical impedance spectroscopy (EIS) using a Bio-Logic VMP3 potentiostat under controlled temperature, and data were collected using EC Laboratories software. Membranes were held in four electrode cells with platinum electrodes. In this work, EIS was performed in deionized ultrapure water (18 MΩ) under a nitrogen atmosphere by changing temperature from 50 to 90 °C. Significant contamination from carbon dioxide and the formation of carbonate and bicarbonate were considered negligible during the measurement. Activation energy (Ea) of ionic conductivity can be determined by using an Arrhenius relationship between the conductivity and temperature, which can be expressed as

σT = σ0e−Ea / RT where σ is the ionic conductivity, σ0 is the Arrhenius constant, R is gas constant, and Ea is the activation energy of ionic conductivity under certain relative humidity. Tensile properties of block copolymer membranes were measured at dry and hydrated conditions on an Instru-Met frame tensile tester (MTS Model: A30-33) using a 1000 lb load cell (with a resolution of less than 0.1 lb) and a cross-head speed of 5 mm/min at room temperature. The dry samples were prepared by drying membranes in a vacuum oven at 60 °C for 24 h. Samples were cut into strips with dimensions of approximately 60 mm × 8 mm, and membrane thicknesses of different samples were measured in a range of 50−80 μm. Stress was calculated from the cross-sectional area of the film, and the strain was obtained by measuring the increase of the film gauge length. The Young’s modulus was determined by the initial slope of the stress vs strain curve at very low strains.

c AgNO VAgNO 3

mhydrated − mdry



3

mdry membrane

RESULTS AND DISCUSSION Studies have demonstrated random polyethylene copolymers for anion exchange membranes, with the semicrystalline polyethylene providing good mechanical properties, low

where c is the concentration of AgNO3 solution, V is the volume of the AgNO3 solution consumed during titration, and mdry membrane is the dry mass of the membrane. C

DOI: 10.1021/acs.macromol.5b01457 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

Downloaded by UNIV OF MASSACHUSETTS AMHERST on September 14, 2015 | http://pubs.acs.org Publication Date (Web): September 10, 2015 | doi: 10.1021/acs.macromol.5b01457

Scheme 1. Sequential Anionic Polymerization to Prepare PB-b-P4MS Block Copolymers

Table 1. Characterization of PB and PB-b-P4MS Copolymers PB block

a

PB-b-P4MS copolymer

samples

[Bd]/[sec-BuLi]

Mna (g/mol)

Mw/Mna

1,4-PBb (%)

[4MS]/[sec-BuLi]

Mna (g/mol)

Mw/Mna

PB:P4MSb (wt %)

A B C D

950 750 898 590

50 100 39 800 47 300 32 700

1.08 1.06 1.09 1.05

92 92 95 95

282 370 526 646

64 800 61 000 75 400 68 900

1.16 1.08 1.13 1.14

77:23 68:32 60:40 51:49

Determined by light scattering (GPC). bMeasured by 1H NMR spectroscopy.

dimensional swelling, and good chemical stability.13,14,39 However, well-defined diblock copolymers containing polyethylene as the hydrophobic block for alkaline fuel cell membranes have not been reported. In this study, a polyethylene block copolymer with quaternary ammonium functionality is prepared through a series of reactions. The reaction sequence is designed to exploit living anionic polymerization to form polybutadiene (PB) with high 1,4 content as a precursor to polyethylene and the subsequent polymerization of poly(4-methylstyrene) as a precursor to a benzyl bromide functionality and, ultimately, a benzyltrimethylammonium functionality. Some comparison can be made to AEMs prepared by chloromethylation and subsequent quaternization with trimethylamine of triblock copolymers of polystyrene-b-poly(ethylene-ran-butylene)-b-polystyrene (SEBS),25−27 which are also produced by anionic polymerization of butadiene; however, the route described here yields high 1,4-content PB which upon hydrogenation provides semicrystalline polyethylene rather than the rubbery poly(ethylene-ran-butylene). Furthermore, the bromination of the P4MS block avoids the potentially dangerous chloromethylation step. PB-b-P4MS block copolymer was synthesized through the sequential anionic polymerization of butadiene and 4methylstyrene initiated by sec-BuLi in cyclohexane (Scheme 1). Living anionic polymerization was selected since this technique offers significant advantages over other polymerization methods to synthesize a well-defined PB block with high 1,4 enchainment of monomer. While the anionic polymerization of butadiene has been widely studied, few references to the anionic polymerization of 4-methylstyrene exist.29,40−44 A block copolymer of butadiene and 4-methylstyrene can be produced through either reacting butadiene or 4-methylstyrene as the first block. However, the chain transfer to toluene in butyllithium initiated anionic polymerization of styrene is well studied,45 and chain transfer to the para-methyl group is possible during the polymerization. 4-Methylstyrene homopol-

ymer has previously been prepared by anionic polymerization, and the research identified the importance of low-temperature polymerization and limiting the conversion to less than 60% in order to avoid chain transfer.42 In order to synthesize welldefined linear block copolymer and minimize any chain transfer to pendent benzylic methyl groups, the block copolymerization was designed to polymerize butadiene first, followed by polymerization of the 4-methylstyrene. In addition, the 4methylstyrene polymerization was terminated prior to high monomer conversion. In the first step, the anionic polymerization of butadiene was carried out using sec-BuLi in cyclohexane at room temperature following a procedure previously reported.46 sec-BuLi-initiated anionic polymerization of butadiene in nonpolar cyclohexane solvent has been demonstrated to result in predominately 1,4structured PB.47 Linear polybutadienyllithium chains were formed by adding butadiene monomer to the solution of secBuLi in cylcohexane, resulting in a light yellow color during butadiene polymerization. Once the butadiene polymerization was completed, a small amount of polymer solution (1 mL) was removed and terminated with degassed methanol for analysis. A series of four PB polymers were produced with molecular weights controlled by the molar ratios of butadiene/sec-BuLi. The molecular weights were targeted to range between 33 000 and 50 000 g/mol. GPC was used to analyze the molecular weights attained for each PB polymerization and found excellent agreement between the actual and targeted molecular weights (Table 1), indicating complete conversion of butadiene monomer to polymer. The molecular weight distributions of the PB samples are relatively narrow and in the range of 1.05− 1.09. The diblock copolymer was prepared by adding 4methylstyrene to the living polybutadienyllithium chains. As the 4-methylstyrene was introduced, the solution color changed from light yellow (representative of the polybutadienyllithium chain end) to bright orange immediately, indicating the successful initiation of 4-methylstyrene. After polymerization D

DOI: 10.1021/acs.macromol.5b01457 Macromolecules XXXX, XXX, XXX−XXX

Article

Downloaded by UNIV OF MASSACHUSETTS AMHERST on September 14, 2015 | http://pubs.acs.org Publication Date (Web): September 10, 2015 | doi: 10.1021/acs.macromol.5b01457

Macromolecules

Figure 1. GPC chromatograms of PB-b-P4MS copolymers (A) copolymer A, (B) copolymer B, (C) copolymer C, and (D) copolymer D.

spectroscopy. Figure 2a shows the spectrum of one PB-b-P4MS copolymer (sample C) consisting of characteristic signals of

of the 4-methylstyrene, the block copolymerization was terminated by the introduction of degassed methanol, and the solution immediately turned to colorless, indicating the termination of the living polymer chains. The polymer solution was precipitated into an excess of methanol to isolate the block copolymer. A series of PB-b-P4MS diblock copolymers (Table 1) were prepared by varying the molar ratio of 4methylstyrene/sec-BuLi. The molecular weight of the P4MS blocks in the different copolymer samples was determined by the difference in the molecular weight of the PB block and the final copolymer measured by GPC. The conversion of 4methylstyrene was calculated to be approximately 43−49% based on the mole ratio value of 4-methylstyrene/sec-BuLi. The progression of the copolymerization is clearly demonstrated by the GPC chromatograms in Figures 1A−D for the four different copolymerizations (copolymers A, B, C, and D). Solid lines represent the PB block, and the dashed lines of the PB-b-P4MS copolymer show effective initiation and growth of the P4MS block from the PB homopolymer. The dashed peaks are shifted to lower elution volume in all chromatograms, and the effective crossover from polybutadienyllithium to polybutadiene-b-poly(4-methylstyrenyl)lithium can be observed by the resulting monomodal peaks. Molecular weight distributions of the block copolymers (Mw/ Mn = 1.08−1.16) become only slightly broader after adding the second block. The slight increase in polydispersity could be due to a slow crossover from polybutadienyllithium to polybutadiene-b-poly(4-methylstyrenyl)lithium or to a small amount of chain transfer to the para-methyl group. Small shoulders in the high molecular weight region are observed in some chromatograms, presumably a result from some polymer chain coupling by oxygen during termination reactions. The microstructure of the PB block and the relative block composition of the copolymers were determined by 1H NMR

Figure 2. 1H NMR spectra comparison of (a) PB-b-P4MS, (b) PE-bP4MS, and (c) PE-b-PVBBr copolymers in CDCl3.

polybutadiene and poly(4-methylstyrene). The peaks at 4.91− 5.02 and 5.53−5.63 ppm are signals from the 1,2-structured PB. The peaks at 5.32 and 5.49 ppm correspond to the two olefinic protons from the cis and trans 1,4-structured PB.48−51 The microstructure in the PB block can be calculated based on the integration of 1,2-structured PB and 1,4-structured PB. The amount of 1,4 configuration is predominant in the microstructure of PB as expected from the synthesis in cyclohexane, and the composition varies from 92% to 95% 1,4- for the four E

DOI: 10.1021/acs.macromol.5b01457 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

Downloaded by UNIV OF MASSACHUSETTS AMHERST on September 14, 2015 | http://pubs.acs.org Publication Date (Web): September 10, 2015 | doi: 10.1021/acs.macromol.5b01457

Scheme 2. Hydrogenation, Bromination, and Quaternization of Block Copolymers

Table 2. IEC, Water Uptake, and Conductivity of PE-Based Membranes copolymer AEMs

deg of brominationa (%)

deg of crystallinityb (%)

theor IECc (mmol/g)

titrated IECd (mmol/g)

water uptakee (%)

A B C D

82.5 84.6 83.1 81.8

24.1 25.7 25.1 24.3

1.38 1.78 2.07 2.30

1.17 1.66 1.87 1.92

15 26 35 37

σf (mS/cm) 24 47 70 73

(7.1) (11) (18) (19)

a

Determined by 1H NMR spectroscopy. bEstimated by DSC. c1H NMR spectroscopy and complete N(CH3)3 quaternization. dDetermined by Mohr’s titration. eMeasured in bromide form at room temperature. fHydroxide and bromide conductivity measured by EIS at 60 °C in water (bromide conductivity in parentheses). 1

different samples (Table 1). The chemical shift at 2.07−2.51 ppm is attributed to the benzylic methyl group from 4methylstyrene, and the peaks at 6.15−7.16 ppm represent aromatic protons of 4-methylstyrene. The molar composition of PB and P4MS can be readily estimated by comparing integration of characteristic peaks for PB and P4MS. For this sample (copolymer C), the mass composition of PB and P4MS is determined to be 60% and 40%, respectively. The diblock copolymer was subsequently reacted to hydrogenate the PB block, and then the quaternary ammonium functionality was introduced to the P4MS block according to the sequence in Scheme 2. The hydrogenation and bromination of para-methyl groups were each done in a homogeneous reaction, while the final conversion of bromomethyl groups to benzyltrimethylammonium groups was done as a heterogeneous reaction on a solution-cast film. Because the majority of the PB is 1,4-structured, hydrogenation of the PB block results in a predominantly polyethylene-like block with approximately 5−8% ethyl branches. The thermal decomposition of p-toluenesulfonhydrazide in xylenes is a common method for converting polydienes to hydrogenated polymers.52−54 A large excess (molar ratio of p-toluenesulfonhydrazide/butadiene repeat units = 8/1) of p-toluenesulfonhydrazide was added in this reaction to ensure complete hydrogenation. The resulting hydrogenated polymer remained soluble in xylenes at high temperature, so the polymer was precipitated into methanol while the solution was still hot. The resulting polymers were washed with boiling water for 30 min in order to remove ptoluenesulfonic acids formed during the reaction. The hydrogenated polymers were no longer soluble or only partially soluble in common solvents at room temperature, including THF, toluene, chloroform, tetrachloroethane, and xylenes. The polymer was soluble enough in chloroform at low concentration to perform 1H NMR spectroscopy in order to quantify the extent of the hydrogenation reaction. Figure 2b shows the

H NMR spectrum of PE-b-P4MS copolymer. Comparison with the PB-b-P4MS in Figure 2a shows that olefinic proton signals of PB at 4.91−5.63 ppm (including both 1,2-structured and 1,4-structured PB) completely disappear and new peaks form at 1.54 ppm corresponding to the methylene protons from polyethylene, demonstrating the complete hydrogenation. Bromination of PE-b-P4MS copolymer was conducted after the hydrogenation reaction to ensure the selective bromination of the 4-methyl groups without brominating the PB backbone. Bromination conditions and bromination results were examined with a P4MS homopolymer to determine that bromination could be controlled to the 4-methyl groups without reaction at the backbone benzylic position (see Supporting Information). The bromination reaction was carried out at 72 °C in carbon tetrachloride, and the copolymer remained in solution during the reaction. The reaction was carried out in the presence of AIBN with less than the stoichiometric equivalent amount (NBS/4-methyl group = 0.9/1 molar ratio) in order to selectively brominate the 4-methyl group without reacting the tertiary benzylic protons on the backbone. The solution color changed to a light yellow when NBS was added, and after 12 h of reaction, the light yellow polymer (PE-b-PVBBr) was obtained by precipitation in methanol and isolation by filtration. Figure 2c shows the 1H NMR spectrum of the PEb-PVBBr, and it is compared with the 1H NMR spectrum of PE-b-P4MS (Figure 2b). The chemical shift at 2.07−2.51 ppm corresponding to the 4-methyl group reduces significantly after bromination, while at the same time, a new peak at 4.20−4.76 ppm corresponding to benzyl bromide is observed in Figure 2c. It is expected to observe some remaining 4-methyl groups after bromination since less than 1 equiv of NBS relative to P4MS repeat units was used in this reaction. The degree of bromination of the P4MS block can be determined by the integration ratio between the benzyl bromide groups and the aromatic groups. The degree of bromination is calculated to be in a range of 81.8%−84.6% (Table 2) depending on copolymer F

DOI: 10.1021/acs.macromol.5b01457 Macromolecules XXXX, XXX, XXX−XXX

Article

Downloaded by UNIV OF MASSACHUSETTS AMHERST on September 14, 2015 | http://pubs.acs.org Publication Date (Web): September 10, 2015 | doi: 10.1021/acs.macromol.5b01457

Macromolecules

the hygroscopic quaternary ammonium cation in the membrane. The theoretical IECs of the quaternized membranes are compared with their titrated IECs in Table 2. The theoretical IECs were calculated based on the composition of 4-MS in the copolymer and the degree of bromination in each of the different copolymers. The titrated IEC was determined by the well-established Mohr’s titration method. The comparison shows that the quaternization reaction reached high conversion of 83−93%, in agreement with the IR spectroscopy results. Small-angle X-ray scattering (SAXS) was used to examine microphase separation of the hydrophobic and hydrophilic components. The quaternized block copolymer membranes were measured in the bromide counterion form, and SAXS experiments were performed in an environmental chamber at 60 °C and with relative humidity varied from 25% to 95%. Ordered morphologies were observed in all membranes tested, and all displayed a distinct first-order peak and a weak secondorder peak as shown in Figure 4. The d-spacing of the membranes increased from 48.3 to 69.8 nm on going from the lowest cationic functionalized sample to the most highly functionalized sample. All SAXS profiles displayed little response to the change in humidity, indicating that all membranes are structurally stable under the changing humidity conditions. This could be due to the semicrystalline structure of the membrane matrix causing restricted structural swelling. Although structural features were observed in all SAXS profiles, an unambiguous interpretation of the membrane morphologies is not possible based on the SAXS data alone. High-angle annular dark field scanning TEM (HAADF STEM) was performed on each membrane to further study the phaseseparated morphology in the PE block copolymer membranes. The corresponding micrographs, in Figure 5, show distinct phase separation of hydrophilic (bright) and hydrophobic (dark) regions. Membranes A through C show structures typical of bicontinuous polymer structures, with correlation lengths of 43, 42, and 57 nm, respectively. Membrane D also shows a bicontinuous morphology, but in this case the domains are very sharply defined, are cylindrical in nature, and have a correlation length of approximately 67 nm. This cylindrical structure is similar in character to previous studies involving PSb-PE block copolymers, particularly as reported by Uehara and co-workers.55,56 These domains appear to form cylinders running through the film, with many shown in cross section (round domains). (See the Supporting Information for a lower magnification micrograph.) Sample D is the only one of the four compositions to show morphological behavior close to the classical phase behavior of model diblock copolymers, despite the relatively high segregation strength of these materials. These findings are in contrast to published results on PE-bPS copolymers, where classical block copolymer morphological behavior has generally been observed. Lorenzo et al. prepared films of PE-b-PS by solution casting from toluene at 70 °C over the course of a week and then gently annealing at 150 °C. Wellordered PS cylinders, alternating lamellae, and PE cylinders were observed in the block copolymers having total molecular weights ranging from 41 000 up to 244 000 g/mol.57,58 Uehara and co-workers found a bicontinuous morphology for a PE-bPS copolymer with a 67 000 g/mol PE block and a 54 000 g/ mol PS where the samples were prepared by cooling from a 180 °C melt.55,56 The sample was either crystallized for 24 h at various temperatures or crystallized at 90 °C for various times, but both approaches produced the same morphology, with

samples. Therefore, quantitative control over bromination of the P4MS block was successfully achieved. The PE-b-PVBBr was solution-cast into a film prior to the conversion into a quaternary ammonium functionalized polymer, since the amphiphilic material after quaternization is difficult to process into a membrane. PE-b-PVBBr powder was dissolved in 90 °C xylenes and cast on a hot plate held at 80 °C, keeping the copolymer in solution while the solvent slowly evaporated. The PE-b-PVBBr membrane remained transparent and homogeneous while holding at 80 °C for 1 h. The membrane was then dried in a vacuum oven at 80 °C to remove remaining xylenes. The PE-b-PVBBr copolymer membranes were examined by DSC to determine the degree of crystallinity of the PE component in the film. All membranes were found to have a similar amount of crystalline structure in the range of 24.1%−25.7% (Table 2). The amination reaction was accomplished by soaking the PEb-PVBBr membrane in trimethylamine aqueous solution at room temperature for 48 h. The resulting PE-b-P[VBTMA][Br] membrane was immersed in deionized water, then rinsed repeatedly with water, and dried in a vacuum oven to remove residual trimethylamine. Examination of the quaternization reaction was done by IR spectroscopy (Figure 3). The characteristic peak at 618 cm−1 corresponding to a C−Br stretch mostly disappears after amination, indicating high conversion in the quaternization reaction. Additionally, a new peak forms at 3360 cm−1 that can be attributed to water from

Figure 3. Infrared spectra of PE-b-PVBBr and PE-b-P[VBTMA][Br] (sample C). G

DOI: 10.1021/acs.macromol.5b01457 Macromolecules XXXX, XXX, XXX−XXX

Article

Downloaded by UNIV OF MASSACHUSETTS AMHERST on September 14, 2015 | http://pubs.acs.org Publication Date (Web): September 10, 2015 | doi: 10.1021/acs.macromol.5b01457

Macromolecules

Figure 4. SAXS profiles of quaternized polyethylene block copolymer AEMs (A−D).

groups,15,59−61 which could be attributed to the semicrystalline structure of the hydrophobic matrix. The bromide and hydroxide conductivities of all block copolymer AEMs were measured in ultrapure deionized water by electrochemical impedance spectroscopy and are reported for measurements at 60 °C in Table 2. The ionic conductivity is found to increase with increasing IEC and the concomitant water uptake as expected, due to the increase in the number of charge carriers. The membranes C and D show close values in ionic conductivity, in agreement with their similar ion exchange capacities. The hydroxide conductivities are approximately 4 times higher than the bromide conductivities. The hydroxide conductivities of block copolymer membranes were further examined in Figure 6 by sweeping temperature from 50 to 90 °C, and the conductivity increases with increasing temperature in all membranes. Furthermore, this method allows the calculation of activation energy for ion conduction in block copolymer membranes by an Arrhenius relationship. The activation energy was found to be dependent on the ion exchange capacity. The higher functionalized membrane shows lower activation energy, and the activation energy calculated for hydroxide membranes ranged from 27 to12 kJ/mol going from the least functionalized to the most highly functionalized membranes. The activation energies of these membranes are only slightly higher compared to the

slight variations in PE-domain continuity as a function of annealing protocol. Uehara and co-workers also found that postprocessing sulfonation of their material did not produce a substantial change in morphology.49,56 In light of the prior work on PE-based block copolymers, the findings here suggest that the morphological behavior of the membranes in this work is largely determined when the PE-bPVBBr block copolymers are drop-cast from xylenes solution onto an 80 °C preheated glass slide. Such a sample preparation procedure can be expected to generate a kinetically trapped, nonequilibrium morphology, even for a block copolymer in the strong segregation limit. After drop-casting, the samples are annealed overnight at 80 °C. Amination of the films does not redissolve the membranes, and they are subsequently dried overnight at only 60 °C, well below the Tm of PE. The four different polymer membranes were prepared with a range of ion exchange capacities that is dependent on the block composition of the copolymer precursor and the extent of bromine functionalization. The water uptake of the block copolymer membranes increases with increasing ion exchange capacity from 15% to 37% going from the low functionalized membrane to the most highly functionalized membrane. The water uptake values of the PE-based block copolymer membranes are relatively lower than other AEMs with similar ion exchange capacity and the same cationic functional H

DOI: 10.1021/acs.macromol.5b01457 Macromolecules XXXX, XXX, XXX−XXX

Article

Downloaded by UNIV OF MASSACHUSETTS AMHERST on September 14, 2015 | http://pubs.acs.org Publication Date (Web): September 10, 2015 | doi: 10.1021/acs.macromol.5b01457

Macromolecules

Figure 5. HAADF STEM micrographs of PE block copolymer membranes. The bright regions are the P[VBTMA][Br] domains, while the dark regions are the PE domains.

activation energies reported for Nafion (5−20 kJ/mol).62−64 However, the activation energies determined for the block copolymer membranes are very comparable with the values for other AEMs.34,64−66 The tensile properties of the PE-based block copolymer AEMs (A, C, and D) were measured at room temperature under dry conditions (Figure 7). The tensile stress and strain at break were found to decrease with the increase in ion exchange capacity, while the Young’s modulus follows the reverse order. The Young’s modulus increases from 200 to 324 MPa going from the least functionalized membrane (A) to the most highly functionalized membrane (D). The PE-based block copolymer AEMs exhibit excellent mechanical properties with tensile stress at break ranging from 13.5 to 22.5 MPa and strain at break in the range of 115% to 205%, suggesting the membranes are durable for use as fuel cell membranes and validate the concept of using semicrystalline PE as matrix materials in anion exchange membrane applications. Further study of the

mechanical properties under wet and changing temperature conditions will be reported in the future.



CONCLUSIONS Polyethylene block copolymer AEMs were prepared by sequential anionic polymerization of butadiene and 4methylstyrene and subsequent postfunctionalization reactions. High 1,4-PB is obtained through anionic polymerization under the appropriately designed polymerization conditions, and hydrogenation to a semicrystalline, hydrophobic polyethylene block is accomplished. The polyethylene block copolymer AEMs form tough films with high strain at break indicative of their ductile behavior. Block copolymer compositions can be readily controlled to vary the IECs, and the hydrophobic− hydrophilic nature of the materials results in strong segregation to form ordered phase separated materials. The activation energy for conductivity is dependent on IEC and the sample with the highest IEC displayed values comparable to Nafion. I

DOI: 10.1021/acs.macromol.5b01457 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

Downloaded by UNIV OF MASSACHUSETTS AMHERST on September 14, 2015 | http://pubs.acs.org Publication Date (Web): September 10, 2015 | doi: 10.1021/acs.macromol.5b01457

made possible through a grant from the NSF MRI program (grant # CHW-0923537). This research used resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract No. DE-AC02-06CH11357. A.M.S. was supported by the Postgraduate Research Participation Program at the US Army Research Laboratory, administered by the Oak Ridge Institute of Science and Education through an interagency agreement between the US Department of Energy and Army Research Laboratory (Contract ORISE1120-1120-99). The authors also want to thank Blake Whitley and Prof. Kip Findley for their assistance with membrane tensile testing and Drs. Scott Walck and Aaron Jackson for assistance with transmission electron microscopy.



Figure 6. Hydroxide ion conductivity temperature dependence of PEbased block copolymer anion exchange membranes A−D.

(1) Abdel Rahim, M. A.; Abdel Hameed, R. M.; Khalil, M. W. J. Power Sources 2004, 134, 160−169. (2) Varcoe, J. R.; Slade, R. C. T. Fuel Cells 2005, 5, 187−200. (3) Couture, G.; Alaaeddine, A.; Boschet, F.; Ameduri, B. Prog. Polym. Sci. 2011, 36, 1521−1557. (4) Merle, G.; Wessling, M.; Nijmeijer, K. J. Membr. Sci. 2011, 377, 1−35. (5) McLean, G. F.; Niet, T.; Prince-Richard, S.; Djilali, N. Int. J. Hydrogen Energy 2002, 27, 507−526. (6) Jaeger, W.; Bohrisch, J.; Laschewsky, A. Prog. Polym. Sci. 2010, 35, 511−577. (7) Einsla, B. R.; Chempath, S.; Pratt, L.; Boncella, J.; Rau, J.; Macomber, C.; Pivovar, B. ECS Trans. 2007, 11, 1173−1180. (8) Pan, J.; Lu, S.; Li, Y.; Huang, A.; Zhuang, L.; Lu, J. Adv. Funct. Mater. 2010, 20, 312−319. (9) Li, N.; Leng, Y.; Hickner, M. A.; Wang, C.-Y. J. Am. Chem. Soc. 2013, 135, 10124−10133. (10) Hibbs, M. R. J. Polym. Sci., Part B: Polym. Phys. 2013, 51, 1736− 1742. (11) Gu, S.; Cai, R.; Luo, T.; Chen, Z.; Sun, M.; Liu, Y.; He, G.; Yan, Y. Angew. Chem., Int. Ed. 2009, 48, 6499−6502. (12) Noonan, K. J. T.; Hugar, K. M.; Kostalik, H. A.; Lobkovsky, E. B.; Abruña, H. D.; Coates, G. W. J. Am. Chem. Soc. 2012, 134, 18161− 18164. (13) Yildirim, M. H.; Stamatialis, D.; Wessling, M. J. Membr. Sci. 2008, 321, 364−372. (14) Mokrini, A.; Huneault, M. A.; Shi, Z.; Xie, Z.; Holdcroft, S. J. Membr. Sci. 2008, 325, 749−757. (15) Kostalik, H. A.; Clark, T. J.; Robertson, N. J.; Mutolo, P. F.; Longo, J. M.; Abruña, H. c. D.; Coates, G. W. Macromolecules 2010, 43, 7147−7150. (16) Faraj, M.; Boccia, M.; Miller, H.; Martini, F.; Borsacchi, S.; Geppi, M.; Pucci, A. Int. J. Hydrogen Energy 2012, 37, 14992−15002. (17) Sherazi, T. A.; Yong Sohn, J.; Moo Lee, Y.; Guiver, M. D. J. Membr. Sci. 2013, 441, 148−157. (18) Schauer, J.; Hnát, J.; Brožová, L.; Ž itka, J.; Bouzek, K. J. Membr. Sci. 2012, 401−402, 83−88. (19) Elabd, Y. A.; Hickner, M. A. Macromolecules 2011, 44, 1−11. (20) Weber, R. L.; Ye, Y.; Schmitt, A. L.; Banik, S. M.; Elabd, Y. A.; Mahanthappa, M. K. Macromolecules 2011, 44, 5727−5735. (21) Ye, Y.; Choi, J.-H.; Winey, K. I.; Elabd, Y. A. Macromolecules 2012, 45, 7027−7035. (22) Ye, Y.; Sharick, S.; Davis, E. M.; Winey, K. I.; Elabd, Y. A. ACS Macro Lett. 2013, 2, 575−580. (23) Tsai, T.-H.; Maes, A. M.; Vandiver, M. A.; Versek, C.; Seifert, S.; Tuominen, M.; Liberatore, M. W.; Herring, A. M.; Coughlin, E. B. J. Polym. Sci., Part B: Polym. Phys. 2013, 51, 1751−1760. (24) Rebeck, N. T.; Li, Y.; Knauss, D. M. J. Polym. Sci., Part B: Polym. Phys. 2013, 51, 1770−1778.

Figure 7. Stress vs strain curves of membranes A, C, and D at room temperature and dry conditions.

The results suggest promise for further investigation of semicrystalline polyethylene block copolymers for AEM applications.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.macromol.5b01457. Bromination of P4MS; transmission electron micrograph of sample D at lower magnification (PDF)



REFERENCES

AUTHOR INFORMATION

Corresponding Author

*(D.M.K.) E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was funded by the Army Research Office through a MURI (grant # W911NF-10-1-0520). NMR spectroscopy was J

DOI: 10.1021/acs.macromol.5b01457 Macromolecules XXXX, XXX, XXX−XXX

Article

Downloaded by UNIV OF MASSACHUSETTS AMHERST on September 14, 2015 | http://pubs.acs.org Publication Date (Web): September 10, 2015 | doi: 10.1021/acs.macromol.5b01457

Macromolecules (25) Vinodh, R.; Ilakkiya, A.; Elamathi, S.; Sangeetha, D. Mater. Sci. Eng., B 2010, 167, 43−50. (26) Sun, L.; Guo, J.; Zhou, J.; Xu, Q.; Chu, D.; Chen, R. J. Power Sources 2012, 202, 70−77. (27) Zeng, Q. H.; Liu, Q. L.; Broadwell, I.; Zhu, A. M.; Xiong, Y.; Tu, X. P. J. Membr. Sci. 2010, 349, 237−243. (28) Ma, Y.; Wu, G.; Yang, W. J. Polym. Sci., Part A: Polym. Chem. 2003, 41, 2755−2764. (29) Štěpánek, M.; Košovan, P.; Procházka, K.; Janata, M.; Netopilík, M.; Pleštil, J.; Šlouf, M. Langmuir 2010, 26, 9289−9296. (30) Kim, H. K.; Zhang, M.; Yuan, X.; Lvov, S. N.; Chung, T. C. M. Macromolecules 2012, 45, 2460−2470. (31) Brandrup, J.; Immergut, E. H. Polymer Handbook; Interscience Publishers: New York, 1966. (32) Wunderlich, B. Thermal Analysis; Academic Press: New York, 1990. (33) Skoog, D. A.; West, D. M.; Holler, F. J. Fundamentals of Analytical Chemistry, 7th ed.; Thomson Learning, Inc.: Stamford, CT, 1996. (34) Yan, J.; Hickner, M. A. Macromolecules 2010, 43, 2349−2356. (35) Pandey, T. P.; Maes, A. M.; Sarode, H. N.; Peters, B. D.; Lavina, S.; Vezzu, K.; Yang, Y.; Poynton, S. D.; Varcoe, J. R.; Seifert, S.; Liberatore, M. W.; Di Noto, V.; Herring, A. M. Phys. Chem. Chem. Phys. 2015, 17, 4367−4378. (36) Liu, Y.; Horan, J. L.; Schlichting, G. J.; Caire, B. R.; Liberatore, M. W.; Hamrock, S. J.; Haugen, G. M.; Yandrasits, M. A.; Seifert, S.; Herring, A. M. Macromolecules 2012, 45, 7495−7503. (37) Schlichting, G. J.; Horan, J. L.; Jessop, J. D.; Nelson, S. E.; Seifert, S.; Yang, Y.; Herring, A. M. Macromolecules 2012, 45, 3874− 3882. (38) Mitchell, D. R. G. Microsc. Res. Technol. 2008, 71, 588−593. (39) Han, X.; Zhou, L.; Liu, H.; Hu, Y. Polym. Degrad. Stab. 2007, 92, 75−85. (40) Hou, H. C.; Tsiang, R. C. C.; Hsieh, H. C. C. J. Polym. Sci., Part A: Polym. Chem. 1997, 35, 2969−2980. (41) Quirk, R. P. Abstr. Pap. Am. Chem. Soc. 1985, 190, 74-POY. (42) Mays, J. W.; Hadjichristidis, N. Polym. Bull. 1989, 22, 471−474. (43) Quirk, R. P.; Sarkis, M. T. Phase-Selective Curing of Poly(pMethylstyrene-b-p-Methylstyrene). In Advances in Elastomers and Rubber Elasticity; Lal, J., Mark, J. E., Eds.; Plenum Press: New York, 1986; pp 143−156. (44) Chen, J.; Fetters, L. J. Polym. Bull. 1981, 4, 275−280. (45) Gatzke, A. L. J. Polym. Sci., Part A-1: Polym. Chem. 1969, 7, 2281−2292. (46) Bender, J. T.; Knauss, D. M. J. Polym. Sci., Part A: Polym. Chem. 2006, 44, 828−836. (47) Hsieh, H. L.; Quirk, R. P. Anionic Polymerization: Principles and Practical Applications; Marcel Dekker, Inc.: New York, 1996. (48) Sato, H.; Tanaka, Y. J. Polym. Sci., Polym. Chem. Ed. 1979, 17, 3551−3558. (49) Worsfold, D. J.; Bywater, S. Macromolecules 1978, 11, 582−586. (50) Sato, H.; Ono, A.; Tanaka, Y. Polymer 1977, 18, 580−585. (51) Tanaka, Y.; Sato, H. Polymer 1976, 17, 113−116. (52) Dounis, P.; James Feast, W. Polymer 1996, 37, 2547−2554. (53) Grassi, A.; Caprio, M.; Zambelli, A.; Bowen, D. E. Macromolecules 2000, 33, 8130−8135. (54) Phinyocheep, P.; Pasiri, S.; Tavichai, O. J. Appl. Polym. Sci. 2003, 87, 76−82. (55) Uehara, H.; Yoshida, T.; Kakiage, M.; Yamanobe, T.; Komoto, T.; Nomura, K.; Nakajima, K.; Matsuda, M. Macromolecules 2006, 39, 3971−3975. (56) Uehara, H.; Kakiage, M.; Sekiya, M.; Yamagishi, T.; Yamanobe, T.; Nakajima, K.; Watanabe, T.; Nomura, K.; Hase, K.; Matsuda, M. Macromolecules 2009, 42, 7627−7630. (57) Lorenzo, A. T.; Arnal, M. L.; Müller, A. J.; Boschetti de Fierro, A.; Abetz, V. Eur. Polym. J. 2006, 42, 516−533. (58) Lorenzo, A. T.; Arnal, M. L.; Müller, A. J.; Boschetti-de-Fierro, A.; Abetz, V. Macromolecules 2007, 40, 5023−5037.

(59) Hibbs, M. R.; Hickner, M. A.; Alam, T. M.; McIntyre, S. K.; Fujimoto, C. H.; Cornelius, C. J. Chem. Mater. 2008, 20, 2566−2573. (60) Tanaka, M.; Fukasawa, K.; Nishino, E.; Yamaguchi, S.; Yamada, K.; Tanaka, H.; Bae, B.; Miyatake, K.; Watanabe, M. J. Am. Chem. Soc. 2011, 133, 10646−10654. (61) Price, S. C.; Ren, X.; Jackson, A. C.; Ye, Y.; Elabd, Y. A.; Beyer, F. L. Macromolecules 2013, 46, 7332−7340. (62) Halseid, R.; Vie, P. J. S.; Tunold, R. J. Electrochem. Soc. 2004, 151, A381−A388. (63) Saito, M.; Hayamizu, K.; Okada, T. J. Phys. Chem. B 2005, 109, 3112−3119. (64) Slade, R. C. T.; Varcoe, J. R. Solid State Ionics 2005, 176, 585− 597. (65) Varcoe, J. R. Phys. Chem. Chem. Phys. 2007, 9, 1479−1486. (66) Xiong, Y.; Fang, J.; Zeng, Q. H.; Liu, Q. L. J. Membr. Sci. 2008, 311, 319−325.

K

DOI: 10.1021/acs.macromol.5b01457 Macromolecules XXXX, XXX, XXX−XXX