9
Contents lists available at ScienceDirect Journal of Membrane Science journal homepage: www.elsevier.com/locate/memsci Anionic multiblock copolymer membrane based on vinyl addition polymerization of norbornenes: Applications in anion-exchange membrane fuel cells Mrinmay Mandal, Garrett Huang, Paul A. Kohl School of Chemical and Biomolecular Engineering Georgia Institute of Technology, Atlanta, GA 30332-0100, United States ABSTRACT Stable, non-hydrolysable, hydroxide-conducting polymers are of interest for electrochemical devices such as fuel cells, electrolyzers and ow-batteries. In this study, a series of tetrablock copolymers containing an all-hydrocarbon backbone were synthesized. The synthesis is based on vinyl addition polymerization of norbornene using (η 3 -allyl)Pd( i Pr 3 P)Cl as the catalyst and lithium tetrakis(pentauorophenyl)-borate·(2.5Et 2 O) (Li[FABA]) as the activator. The tetrablock polymers were cast into membranes with an ion-exchange capacity (IEC) between 1.55 and 2.60 milliequivalents per gram (meq/g). The number of bound and unbound water molecules per ion pair was measured and correlated with conductivity. The membrane with the highest IEC (2.60 meq/g) had 10.6 unbound (i.e. free) and 17.9 bound water molecules per ion pair within the polymer. The presence of excess unbound water has led to ooding of the ion conductive channels and low hydroxide ion conductivity. The optimal anion conductivity was found with about 6.7 unbound and 11.9 bound water molecules per ion pair. Excellent hydroxide conductivity (> 120 mS/cm at 80 °C) was obtained at an ion-exchange capacity of 1.88 meq/g. The results show that hydroxide mobility is aided by phase segregation within the copolymer. The ion conducting polymer was stable in 1 M NaOH solution at 80 °C, showing essentially no loss in ion conductivity in 1200 h. Thermogravimetric analysis showed that the membrane backbone was stable up to 400 °C, consistent with previous polynorbornene-based materials. The peak power density of a H 2 /O 2 , hydroxide conducting fuel cell containing one of these membranes was 542.57 mW/cm 2 at 0.43 V and current density of 1.26 A/cm 2 at 60 °C. 1. Introduction Anion-exchange membranes (AEM) are of interest for a number of electrochemical devices (i.e. fuel cells, electrolyzers, redox ow battery) [1,2]. The development of AEMs with long-term alkaline stability and high hydroxide ion conductivity is of current interest [36]. The op- eration of fuel cells and electrolyzers at high pH allows the use of non- precious metal catalysts, especially for oxygen reduction and evolution, and reduced fuel crossover compared to proton-exchange membranes (PEM) devices [711]. Early AEMs have had low ionic conductivity, poor stability at high pH, and high water uptake [1214]. However, recent studies have shown that the tethered, long-chain trimethyl am- monium (TMA) cation is stabile in alkaline environments at elevated temperature [1518]. High hydroxide conductivity is critical in membranes for fuel cells, batteries and electrolyzers. The use of multiblock copolymers can im- prove the hydroxide mobility compared to random copolymers [19]. This is due to the high degree of phase segregation in block copolymers leading to ecient ion-channel formation, compared to random copo- lymers. In addition, water management within the membrane can play an important role in controlling the mechanical deformation (i.e. water swelling) and ionic conductivity. Water is needed in the membrane to form the ionic hydration shell for the mobile hydroxide ion and sta- tionary cation. However, excessive water uptake can lead to swelling of the ion conduction channels resulting in lower ionic conductivity (i.e. lower ion mobility) and softening of the membrane [19]. Thus, it is necessary to optimize the ion conducting channel size so that the amount of free, unbound (unproductive) water is minimized. The long-term stability of AEMs largely depends on the chemical nature of the polymer backbone, position of the cation within the polymer architecture, and chemical nature of the xed cation. In the past, polymers based on poly(arylene ether sulfone)s [2022] and poly (arylene ether ketone)s [23,24] were investigated as AEMs. Signicant degradation was observed at high pH. Polysulfone and polyketone groups in the polymer backbone were susceptible to nucleophilic attack by hydroxide. The poly(aryl ether) backbone undergoes cleavage of the CO bonds at high pH which limits the long-term use [2527]. In addition to the polymer backbones, nucleophilic attack of the xed cation headgroups results in degradation. The mechanisms for degradation of quaternary ammonium headgroups include β-hydrogen Hofmann elimination, direct nucleophilic substitution (S N 2) and elim- ination via ylide formation [2729]. Hibbs et al. found that a hexam- ethylene spacer between the trimethylammonium (TMA) cation and the polymer backbone results in better stability than the https://doi.org/10.1016/j.memsci.2018.10.041 Received 9 August 2018; Received in revised form 15 October 2018; Accepted 16 October 2018 Corresponding author. E-mail address: [email protected] (P.A. Kohl). Journal of Membrane Science 570–571 (2019) 394–402 Available online 18 October 2018 0376-7388/ © 2018 Elsevier B.V. All rights reserved. T

Journal of Membrane Sciencekohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal #1 JMS.pdf · In this study, a series of tetrablock copolymers containing an all-hydrocarbon backbone

  • Upload
    others

  • View
    8

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Journal of Membrane Sciencekohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal #1 JMS.pdf · In this study, a series of tetrablock copolymers containing an all-hydrocarbon backbone

Contents lists available at ScienceDirect

Journal of Membrane Science

journal homepage: www.elsevier.com/locate/memsci

Anionic multiblock copolymer membrane based on vinyl additionpolymerization of norbornenes: Applications in anion-exchange membranefuel cells

Mrinmay Mandal, Garrett Huang, Paul A. Kohl⁎

School of Chemical and Biomolecular Engineering Georgia Institute of Technology, Atlanta, GA 30332-0100, United States

A B S T R A C T

Stable, non-hydrolysable, hydroxide-conducting polymers are of interest for electrochemical devices such as fuel cells, electrolyzers and flow-batteries. In this study,a series of tetrablock copolymers containing an all-hydrocarbon backbone were synthesized. The synthesis is based on vinyl addition polymerization of norborneneusing (η3-allyl)Pd(iPr3P)Cl as the catalyst and lithium tetrakis(pentafluorophenyl)-borate·(2.5Et2O) (Li[FABA]) as the activator. The tetrablock polymers were castinto membranes with an ion-exchange capacity (IEC) between 1.55 and 2.60 milliequivalents per gram (meq/g). The number of bound and unbound water moleculesper ion pair was measured and correlated with conductivity. The membrane with the highest IEC (2.60meq/g) had 10.6 unbound (i.e. free) and 17.9 bound watermolecules per ion pair within the polymer. The presence of excess unbound water has led to flooding of the ion conductive channels and low hydroxide ionconductivity. The optimal anion conductivity was found with about 6.7 unbound and 11.9 bound water molecules per ion pair. Excellent hydroxide conductivity(> 120mS/cm at 80 °C) was obtained at an ion-exchange capacity of 1.88meq/g. The results show that hydroxide mobility is aided by phase segregation within thecopolymer. The ion conducting polymer was stable in 1M NaOH solution at 80 °C, showing essentially no loss in ion conductivity in 1200 h. Thermogravimetricanalysis showed that the membrane backbone was stable up to 400 °C, consistent with previous polynorbornene-based materials. The peak power density of a H2/O2,hydroxide conducting fuel cell containing one of these membranes was 542.57 mW/cm2 at 0.43 V and current density of 1.26 A/cm2 at 60 °C.

1. Introduction

Anion-exchange membranes (AEM) are of interest for a number ofelectrochemical devices (i.e. fuel cells, electrolyzers, redox flow battery)[1,2]. The development of AEMs with long-term alkaline stability andhigh hydroxide ion conductivity is of current interest [3–6]. The op-eration of fuel cells and electrolyzers at high pH allows the use of non-precious metal catalysts, especially for oxygen reduction and evolution,and reduced fuel crossover compared to proton-exchange membranes(PEM) devices [7–11]. Early AEMs have had low ionic conductivity,poor stability at high pH, and high water uptake [12–14]. However,recent studies have shown that the tethered, long-chain trimethyl am-monium (TMA) cation is stabile in alkaline environments at elevatedtemperature [15–18].

High hydroxide conductivity is critical in membranes for fuel cells,batteries and electrolyzers. The use of multiblock copolymers can im-prove the hydroxide mobility compared to random copolymers [19].This is due to the high degree of phase segregation in block copolymersleading to efficient ion-channel formation, compared to random copo-lymers. In addition, water management within the membrane can playan important role in controlling the mechanical deformation (i.e. waterswelling) and ionic conductivity. Water is needed in the membrane to

form the ionic hydration shell for the mobile hydroxide ion and sta-tionary cation. However, excessive water uptake can lead to swelling ofthe ion conduction channels resulting in lower ionic conductivity (i.e.lower ion mobility) and softening of the membrane [19]. Thus, it isnecessary to optimize the ion conducting channel size so that theamount of free, unbound (unproductive) water is minimized.

The long-term stability of AEMs largely depends on the chemicalnature of the polymer backbone, position of the cation within thepolymer architecture, and chemical nature of the fixed cation. In thepast, polymers based on poly(arylene ether sulfone)s [20–22] and poly(arylene ether ketone)s [23,24] were investigated as AEMs. Significantdegradation was observed at high pH. Polysulfone and polyketonegroups in the polymer backbone were susceptible to nucleophilic attackby hydroxide. The poly(aryl ether) backbone undergoes cleavage of theC−O bonds at high pH which limits the long-term use [25–27].

In addition to the polymer backbones, nucleophilic attack of thefixed cation headgroups results in degradation. The mechanisms fordegradation of quaternary ammonium headgroups include β-hydrogenHofmann elimination, direct nucleophilic substitution (SN2) and elim-ination via ylide formation [27–29]. Hibbs et al. found that a hexam-ethylene spacer between the trimethylammonium (TMA) cation and thepolymer backbone results in better stability than the

https://doi.org/10.1016/j.memsci.2018.10.041Received 9 August 2018; Received in revised form 15 October 2018; Accepted 16 October 2018

⁎ Corresponding author.E-mail address: [email protected] (P.A. Kohl).

Journal of Membrane Science 570–571 (2019) 394–402

Available online 18 October 20180376-7388/ © 2018 Elsevier B.V. All rights reserved.

T

Page 2: Journal of Membrane Sciencekohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal #1 JMS.pdf · In this study, a series of tetrablock copolymers containing an all-hydrocarbon backbone

trimethylbenzylammonium (BTMA) cation in 4M KOH at 80 °C [30].Later, Mohanty et al. reported that a quaternary ammonium headgrouptethered to the backbone by a long alkyl chain had the best alkalinestability by comparing the stability of small molecules [15]. A cationattached to a long alkyl chain increases the barrier for the Hofmannelimination reaction and minimizes the risk of degradation [31]. It wasobserved that the polymers with the combination of an all-hydrocarbonbackbone and tethered quaternary ammonium group on a long alkylchain have the best long-term alkaline stability [32–34].

Price et al. synthesized hydrogenated poly(norbornene) as an anion-exchange membrane via ring-opening metathesis polymerization(ROMP) with high ionic conductivity, 177mS/cm at 80 °C. Althoughthe conductivity was high, the membranes were mechanically weak andnot stable under alkaline condition [35]. Register et al. described thesynthesis of block copolymers by vinyl addition polymerization ofsubstituted norbornenes in a living polymerization and their use as apervaporation membrane [36–38]. In a previous study, high Tg (385 °C)polynorbornene was shown to have excellent stability [39]. In thispaper, a facile synthetic strategy was used to prepare a series of tetra-block AEM copolymers based on vinyl addition polymerization ofnorbornenes, Scheme 1. The AEMs were cast from solution and theimpact of bound and unbound water on conductivity was evaluated.High thermal stability, excellent mechanical properties and negligiblelong-term degradation at high pH (1M NaOH solution at 80 °C) weredemonstrated. The AEMs were used as membranes in an all-alkalinefuel cell.

2. Experimental

2.1. Materials

1-hexene, 5-bromo-1-pentene and dicyclopentadiene were pur-chased from Alfa Aesar and used as-received. The monomers, butylnorbornene (BuNB) and bromopropyl norbornene (BPNB), were syn-thesized via a Diels-Alder reaction at high-temperature according to apublished procedure [40]. Prior to polymerization, the monomers werepurified by distillation over sodium and degassed by three freeze-pump-thaw cycles. All polymerization reactions were performed under a dryargon atmosphere in a glove box with rigorous care to avoid moistureand air. Toluene was dried by heating under reflux for 6 h over sodiumand benzophenone. Toluene was freshly distilled prior to use. Triiso-propylphosphine and [(η3-allyl)Pd(Cl)]2 were purchased from Sigma-Aldrich and used as-received. The catalyst, (allyl)palladium(triisopro-pylphosphine)chloride ((η3-allyl)Pd(iPr3P)Cl), was prepared accordingto a previously published report [41]. Lithium tetrakis(penta-fluorophenyl)-borate·(2.5Et2O) (Li[FABA]) was purchased fromBoulder Scientific Co. and used as-received. α,α,α-trifluorotoluene(TFT), anhydrous, ≥ 99% and tetrahydrofuran (THF) were purchased

from Sigma-Aldrich and used as-received.

2.2. Synthesis of tetrablock copolymer [Poly(BuNB-b-BPNB-b-BuNB-b-BPNB)]

The tetrablock copolymer, consisting of alternating butyl norbor-nene (BuNB) and bromopropyl norbornene (BPNB) blocks (two blockseach), was synthesized by the sequential addition of the monomers (oneafter the other) at room temperature in an inert atmosphere glove box.The monomers were divided into four round-bottomed flask (two foreach monomer) and toluene was added to make a 5 wt% solution ineach flask. The catalyst solution was prepared separately in a vial bydissolving (η3-allyl)Pd(iPr3P)Cl (12mg, 0.03mmol) and Li[FABA](28mg, 0.03mmol) in a solution composed of 0.5 g toluene and 0.5 gTFT. The catalyst solution was stirred for 20min. BuNB (0.45 g,3.00mmol) and toluene (10mL) were added to a 100mL round-bot-tomed flask fitted with a magnetic stir bar. The catalyst solution wasinjected into the flask under vigorous stirring. After 20min the BuNBpolymerization was complete. A small aliquot was removed andquenched with CH3CN for gel permeation chromatography (GPC)analysis. Then, a mixture of BPNB (0.64 g, 3.00mmol) and toluene(12mL) was added to the reaction flask, still containing the catalyst,and stirred for 3 h to incorporate the BPNB block onto the BuNBpolymer. After the 3 h reaction time (complete consumption of BPNB), asmall aliquot was taken out and quenched with CH3CN for GPC ana-lysis. Next, BuNB (0.45 g, 3.00mmol) and toluene (10mL) were addedto the reaction flask and allowed to react for 20min for incorporation ofthe third block. A small aliquot was again taken out and quenched withCH3CN for GPC analysis. Finally, a mixture of BPNB (0.64 g,3.00mmol) and toluene (12mL) was added to the flask and stirred for3 h to incorporate the fourth block onto the polymer. After completion,the reaction mixture was quenched and the polymer precipitated byaddition of methanol. The resulting polymer was dissolved in THF andstirred over activated charcoal. The solution was passed through analumina filter to remove any palladium residue. The resulting productwas precipitated from THF by addition of methanol. The polymerproduct was dried under vacuum at 60 °C. Tetrablock copolymers withdifferent hydrophobic and hydrophilic chain lengths were synthesizedby changing the monomer to catalyst feed ratio.

2.3. Nuclear magnetic resonance (NMR) spectra and GPC

The polymer samples were analyzed by 1H NMR using BrukerAvance 400MHz NMR instrument using. CDCl3 was the solvent. Thenumber average molecular weight (Mn) and polydispersity index (Mw/Mn) of the polymer samples were determined by GPC (Shimadzu)equipped with an LC-20 CE HPLC pump and a refractive index detector(RID-20 A, 120 V). Measurements were performed in THF with the

Scheme 1. Synthesis of tetrablock copolymer.

M. Mandal et al. Journal of Membrane Science 570–571 (2019) 394–402

395

Page 3: Journal of Membrane Sciencekohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal #1 JMS.pdf · In this study, a series of tetrablock copolymers containing an all-hydrocarbon backbone

eluent flow rate of 1.0mL/min at 30 °C. A polystyrene standard wasused.

2.4. Membrane casting and ion-exchange

The tetrablock copolymer (0.20 g) was dissolved in 5mL chloroformand the resulting solution was filtered through a 0.2 µm poly(tetra-fluoroethylene) (PTFE) membrane syringe filter into a 4 cm diameteraluminum dish. The solvent was evaporated at room temperature in anitrogen gas stream. The membrane was dried overnight under va-cuum. The membranes were colorless, flexible and free-standing with athickness of ca. 50 µm. Next, the bromobutyl headgroup was qua-ternized by immersion of the membrane in 45 wt% aqueous trimethy-lamine solution for 48 h at room temperature. The quaternized mem-brane with bromide counter-ion was removed from solution andwashed thoroughly with DI water. The membranes were then soaked in1M NaOH solution under nitrogen for 24 h to exchange the bromideions for hydroxide ions. The membranes were stored in DI water afterbeing washed with DI water three times.

2.5. Morphological characterization

Small angle X-ray scattering (SAXS) was used to analyze the mor-phology of AEMs. Hydrated membranes in bromide form were tested inair using either a Malvern Panalytical Empyrean XRD (Netherlands)instrument with a Pixel 3D detector or the NSLS-II beamline at theCenter for Functional Nanomaterials (Brookhaven National Laboratory,Upton, NY). The wave vector (q) was calculated using Eq. (1), where 2θis the scattering angle.

=q πlsin θ

42 (1)

The characteristic separation length, or inter-domain spacing (d)(i.e. the Bragg spacing) was calculated using Eq. (2).

=d πq

2(2)

Transmission electron microscopy (TEM) was also used to analyzethe morphology of membranes. TEM was performed with a JEOL JEM-1400 Transmission Electron Microscope. Dry membranes with a bro-mide counter ion were stained by fuming with osmium tetroxide atroom temperature prior to TEM examination. The stained membraneswere embedded within an epoxy resin, sectioned into ca. 50 nm thicksamples with a Leica UC6rt Ultramicrotome, and placed on a coppergrid for observation.

2.6. Hydroxide conductivity and alkaline stability

The ionic resistance of the membranes was measured using a four-point, in-plane probe and electrochemical impedance spectrometry(1 Hz to 1MHz) with a PAR 2273 potentiostat. All samples were testedin HPLC-grade water under a nitrogen purge to minimize the detri-mental effects of CO2. The samples were allowed to equilibrate for30min prior to each measurement. The in-plane ionic conductivity wascalculated using Eq. (3).

=σ LWTR (3)

Fig. 1. 1H NMR spectrum of tetrablock PNB-X54-Y46 in CDCl3.

Fig. 2. Representative GPC trace of PNB-X67-Y33, showing the sequentialgrowth of each block during the formation of tetrablock copolymer.

M. Mandal et al. Journal of Membrane Science 570–571 (2019) 394–402

396

Page 4: Journal of Membrane Sciencekohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal #1 JMS.pdf · In this study, a series of tetrablock copolymers containing an all-hydrocarbon backbone

In Eq. (3), σ is the ionic conductivity in S/cm, L is the length be-tween sensing electrodes in cm, W and T are the width and thickness ofthe membrane in cm, respectively, and R is the resistance measured inOhms. Long-term (up to 1200 h) alkaline stability testing was per-formed by storing the membrane in 1M NaOH solution at 80 °C in ateflon-lined Parr reactor. The ionic conductivity was measured peri-odically by taking the membranes out of solution and thoroughlywashing them with DI water before measuring the conductivity. Aftereach measurement, the membranes were placed back into the reactorswith a freshly prepared NaOH solution.

2.7. Ion exchange capacity (IEC), water uptake (WU), hydration number(λ), number of freezable water (Nfree) and bound, nonfreezable water(Nbound) molecules

The ion exchange capacity was calculated using NMR data which isdiscussed in detail in the next section. In addition, the membrane IECwas also measured by titration [42]. The membrane in Br- form was

first immersed in 0.1M NaCl solution for 24 h to exchange the bromideions for chloride ions. Next, the membrane in chloride form was thor-oughly washed with DI water and dried under vacuum for 24 h to ob-tain the dry weight. The dried membrane was immersed in a fixedvolume of 0.5M aqueous NaNO3 solution for 24 h. The Cl- ions releasedfrom the membrane were titrated with 0.05M AgNO3 using K2CrO4

(10 wt%) as the indicator. The IEC was calculated using Eq. (4).

=

×

IECc V

MAgNO AgNO

d

3 3

(4)

In Eq. (4), VAgNO3 (mL) is the volume of AgNO3 solution, CAgNO3

(0.05 mol L−1) is the concentration of AgNO3 solution, andMd (g) is theweight of the dried membrane sample.

The water uptake of the membranes was calculated using Eq. (5).

=−

×M M

MWU(%) 100w d

d (5)

In Eq. (5), Md is the dry mass of the membrane and Mw is the wetmass of the membrane after removing excess surface water. Themembranes were in OH- form and measured at room temperature. Thehydration number (λ), number of water molecules per ionic group, wascalculated using Eq. (6).

×

WUIEC

λ 1000 %18 (6)

The number of freezable water (Nfree) and bound water (or non-freezable water) (Nbound) were determined by differential scanningcalorimetry (DSC). DSC measurements were carried out on a DiscoveryDSC with autosampler (TA Instruments). The membrane samples werefully hydrated by soaking in deionized water for one week. After thewater on the membrane surface was dabbed off, a 5–10mg sample wasquickly sealed in an aluminum pan. The sample was cooled to −50 °Cand then heated to 30 °C at a rate of 5 °C/min under N2 (20mL/min).The quantity of freezable and non-freezable water was determined byEqs. (7) – (9) [43–45].

= ×NMM

λfreefree

tot (7)

Mfree is the mass of freezable water and Mtot is the total mass ofwater absorbed in the membrane. The weight fraction of freezablewater was calculated using Eq. (8).

=

MM

H HM M M

/( )/

free

tot

f ice

W d w (8)

Hf is the enthalpy obtained by the integration of the DSC freezing

Table 1Properties of poly(butyl norbornene-b-quaternary ammonium propyl norbornene-b-butyl norbornene-b-quaternary ammonium propyl norbornene) membranes inhydroxide form.

Block copolymer Molecular Weighta

(kg/mol)Mw/Mn IEC (Ion Exchange

Capacity) (meq./g)bOH- Conductivity (mS/cm)c

σ/IECd WaterUptakee (%)

Hydrationnumber λ

Nfree Nbound Inter-domainspacing, d (nm)f

25 °C 80 °C

PNB-X74-Y26 39.58 1.42 1.55 23.4 61.3 39.5 26.2 9.41 0.91 8.50 NDPNB-X70-Y30 37.97 1.29 1.77 27.0 67.4 38.1 59.6 18.71 5.42 13.29 37.2PNB-X67-Y33 38.86 1.28 1.92 32.2 71.8 37.4 68.8 19.91 7.81 12.10 44.2PNB-X62-Y38 50.77 1.54 2.21 50.9 101.9 46.1 71.0 17.85 7.61 10.24 49.9PNB-X54-Y46 45.33 1.55 2.60 44.9 80.0 30.8 133.6 28.55 10.65 17.90 86.4PNB-X68-Y32 114.9 1.42 1.88 62.0 122.7 65.2 63.0 18.62 6.74 11.88 ND

a Measured in bromopropyl form by gel permeation chromatography at RT in THF relative to polystyrene standards.b IEC (Ion Exchange Capacity) was calculated via 1H NMR results in bromopropyl form.c OH- conductivity was measured by four-probe conductivity cell.d Ionic conductivity at 80 °C/IEC.e Water uptake was measured at room temperature.f Inter-domain spacing measured using small angle X-ray scattering (SAXS) in bromide form; ND =not determined. PNB =polynorbornene; X=hydrophobic

block; Y= hydrophilic block; numbers in the subscript indicate the molar ratio of each block.

Fig. 3. SAXS spectra of hydrated tetrablock copolymer poly(BuNB-b-BPNB-b-BuNB-b-BPNB) membranes in bromide form.

M. Mandal et al. Journal of Membrane Science 570–571 (2019) 394–402

397

Page 5: Journal of Membrane Sciencekohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal #1 JMS.pdf · In this study, a series of tetrablock copolymers containing an all-hydrocarbon backbone

peak and Hice is enthalpy of fusion for water, corrected for the subzerofreezing point according to Eq. (9).

= −H H C TΔ Δice iceo

p f (9)

ΔCp is the difference between the specific heat capacity of liquidwater and ice. ΔTf is the freezing point depression.

The thermal stability of the dried membranes in bromide ion formwas studied using thermogravimetric analysis (TGA) on a TAInstruments Q50 analyzer. The temperature was ramped at 10 °C/min

up to 800 °C in a nitrogen atmosphere.

2.8. Membrane electrode assembly (MEA) fabrication and single-cell testing

One of the best performing membrane in this study (PNB-X62-Y38)was selected for testing in an alkaline exchange membrane fuel cell(AEMFC). The AEM anode and cathode were fabricated via the slurrymethod and were identical. A lower molecular weight (20.5 kg/mol)version of the poly(BuNB-b-BPNB-b-BuNB-b-BPNB) tetrablock

Fig. 4. TEM micrographs of a) PNB-X74-Y26, b) PNB-X70-Y30, c) PNB-X67-Y33, d) PNB-X62-Y38, e) PNB-X54-Y46 in bromide form.

Fig. 5. Ionic conductivity of polynorbornene AEMs at different temperature. Fig. 6. Arrhenius plot of lnσ vs. inverse temperature for polynorbornene AEMs.

M. Mandal et al. Journal of Membrane Science 570–571 (2019) 394–402

398

Page 6: Journal of Membrane Sciencekohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal #1 JMS.pdf · In this study, a series of tetrablock copolymers containing an all-hydrocarbon backbone

copolymer anion-exchange ionomer powder was first synthesized usingthe same method as the membranes discussed in this paper. Low mo-lecular weight ionomer material was previously found to be

advantageous for use as a polymeric binder in fuel cell and electrolyzerelectrode fabrication [46]. The dry ionomer powder and 50% platinumon Vulcan XC-72 (carbon) catalyst was ground together with a mortarand pestle in 1.5mL of isopropyl alcohol (IPA) for 10min to producefiner particles. An additional 2mL of IPA was added and the mixtureand ground for another 5min to achieve the proper slurry viscosity forspraying. The catalyst and ionomer slurry was further sonicated in awater bath at room temperature for 30min to ensure even dispersion.The homogenized catalyst and ionomer slurry was sprayed onto 1%water-proofed Toray TGPH-060 carbon paper and dried for 24 h atroom temperature. The platinum loading was approximately 2.1mg/cm2 and an ionomer/carbon ratio of 40% was used. The high metalloading was chosen intentionally to minimize any kinetic losses causedby the non-optimized catalyst.

Prior to MEA testing, the electrodes and membranes were soaked in1M NaOH for 1 h (replacing the solution every 20min) in a nitrogenatmosphere to convert the membrane and ionomer to hydroxide form.The MEA was placed into Fuel Cell Technologies hardware betweensingle-pass serpentine graphite plates with 6mil PTFE gaskets. TheMEAs were tested in a Scribner 850e Fuel Cell Test Station at a celltemperature of 60 °C. Humidified H2 and O2 gas feeds were supplied atthe anode and cathode, respectively, at 0.5 L/min. The dew points ofthe anode and cathode streams were adjusted throughout the course oftesting in order to optimize the water balance within the AEMFC.

3. Results and discussion

3.1. Synthesis and characterization of tetrablock copolymer

The monomers (BuNB and BPNB) were synthesized by a previouslydescribed procedure [40]. The catalyst, (η3-allyl)Pd(iPr3P)Cl was pre-pared in high yield and purity following a previous report [41]. Thereactivity of the BuNB monomer ([M]0/[Pd] = 100: 1) was higher thanthat of the BPNB momomer. The reaction time for each block varied(20min for BuNB and 3 h for BPNB) in order to achieve completeconversion for each block. A one-to-one mole ratio of Li[FABA] to thecatalyst was sufficient to generate the cationic Pd complex for thepolymer initiation. The absence of olefinic protons in the 1H NMRspectra of the polymer produced shows that the polymerization reactionproceeded through the vinyl addition pathway, eliminating the occur-rence of ring-opening metathesis polymerization (ROMP), Fig. 1. The1H NMR spectrum of the other materials is provided in the SupportingInformation (Figs. S1-S5). The polymerization time for the individualmonomers for the different monomer-to-initiator feed ratios ([M]0/[Pd]) was optimized in order to avoid branching side-reactions [36].Incomplete conversion of the polymer into monomer would also createproblems in synthesizing the block copolymer via the sequential addi-tion of different monomers [37].

A series of tetrablock copolymers (PNB-Xa-Yb; PNB is poly-norbornene, a is the mole percent of combined hydrophobic blocks, X,and b is the mole percent of combined hydrophilic blocks, Y) weresynthesized by varying the length of the hydrophobic block (BuNB) andthe hydrophilic block (BPNB). The block sizes were changed by ad-justing the monomer-to-initiator feed ratio ([M]0/[Pd]) in tolueneduring polymerization, Scheme 1. A representative GPC trace of PNB-X67-Y33 is shown in Fig. 2 to demonstrate the sequential growth of eachblock in the formation of the tetrablock copolymer. The numberaverage molecular weight (Mn) of the first block (BuNB) was de-termined by GPC analysis, as described in the Experimental Section. Mn

was found to be 12.32 kDa, Fig. 2. The second monomer was thenadded to the reaction mixture and allowed to react for 3 h. The Mn ofthe combined first and second block was 19.80 kDa, making the Mn ofthe second block 7.48 kDa. Similarly, the Mn of the third and fourthblocks were found to be 9.25 kDa and 9.81 kDa, respectively. Theproperties of the various membranes are given in Table 1. The Mn of thesynthesized polymers ranged from 38 to 114.9 kDa with polydispersity

Fig. 7. The alkaline stability of polynorbornene AEMs at 1M NaOH solution at80 °C.

Fig. 8. TGA traces of the polynorbornene AEMs under nitrogen atmosphere.

Fig. 9. Polarization data for PNB-X62-Y38 AEMFC (2.1mg/cm2 Pt, 40% io-nomer to carbon ratio) with and without iR correction. Cell temperature was60 °C with anode and cathode dewpoints both set at 46 °C. Flow rates of hu-midified H2 and O2 were both 0.5 L/min.

M. Mandal et al. Journal of Membrane Science 570–571 (2019) 394–402

399

Page 7: Journal of Membrane Sciencekohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal #1 JMS.pdf · In this study, a series of tetrablock copolymers containing an all-hydrocarbon backbone

of 1.28–1.55. The IEC value was determined by 1H NMR spectroscopyand found to be 1.55–2.60meq/g, as discussed later. The methyl pro-tons of the hydrophilic block resonate at 0.89 ppm. The methyleneprotons adjacent to the bromine atom in the hydrophilic block appearat 3.41 ppm. The X and Y values within poly(BuNB-b-BPNB-b-BuNB-b-BPNB) was analyzed via NMR by comparing the integration ratio of Ha

and Hb, Fig. 1. The mol% and wt% of the combined hydrophilic andhydrophobic blocks in the tetrablock copolymer were calculated via theNMR spectra.

3.2. Morphology and characterization

SAXS and/or TEM were used to investigate the microstructure of thepolynorbornene membranes synthesized here. The inter-domain spa-cing (d), or the average separation length between inhomogeneities inthe membranes, was determined from the Bragg spacing of the primaryscattering peak in the SAXS spectra, as shown in Fig. 3. The inter-do-main spacing values are listed in Table 1. The domain size range was37.2–86.4 nm, which directly correlated with water uptake. More spe-cifically, it can be seen that the number of unbound waters tracks withthe domain size. Larger domains provide regions where free (unbound)water can populate. For example, PNB-X54-Y46 showed a very highwater uptake, 133.6%, and a d-spacing, 86.4 nm. By comparison, thesevalues are about twice as large as the ones for PNB-X67-Y33, which hada water uptake of 68.8% and a d-spacing of 44.2 nm.

PNB-X74-Y26, was examined using transmission electron micro-scopy (TEM) because the X-ray scattering was not as definitive as withother samples. TEM analysis was performed in bromide form ratherthan in hydroxide form to avoid inadvertent degradation of the mem-branes due to concentrated hydroxide in dry membranes. It was ob-served that the size of the phases increased with the hydrophilicity ofthe membranes. This is consistent with the increasing inter-domainspacing observed by SAXS. The ion channels also appear to lose theirdefinition and well-defined structure as the channel size became larger.Fig. 4 shows TEM micrographs for five of the membranes. The darkregions correspond to the hydrophilic domains with bromide counterions and the bright regions correspond to hydrophobic domains.Membranes with higher hydrophilicity also contain more dark regionsin the TEM micrographs. It should be noted that trends observed usingTEM in the dry state would likely be more pronounced if the mem-branes could be observed swollen with water as in the SAXS measure-ments.

3.3. Ion Exchange Capacity (IEC), hydroxide conductivity

The IEC is an important parameter in determining the membraneionic conductivity and water uptake. The IEC was evaluated from 1HNMR spectroscopy and found to be between 1.55 and 2.60meq/g.Taking PNB-X54-Y46 as the representative sample, the IEC was calcu-lated by comparing the integration ratio of the methylene protons ad-jacent to the bromine atom of the hydrophilic block at 3.41 ppm andmethyl protons of the hydrophobic block at 0.89 ppm, Fig. 1. FromFig. 1, the integration ratio of Ha to Hb was 3.46:2 which is equal to1.153:1 (for 1 proton). To further confirm the extent of the quaterni-zation reaction, the IEC was measured by titration. The titration pro-cedure involved converting the counter anion in the membrane tochloride, followed by titration of the amount of chloride present in themembrane, as described in the Experimental Section. An excellentcorrelation for IEC was found between the two methods, titration andNMR. For PNB-X67-Y33, the measured IECs by titration and NMR werefound to be 1.90meq/g and 1.92meq/g, respectively.

High hydroxide conductivity (σ) is desired in membranes used inelectrochemical devices [14]. Fig. 5 shows that the hydroxide con-ductivity increased with temperature from 25 °C to 80 °C and followedan Arrhenius relationship. In the case of PNB-X68-Y32, the conductivitywas 122.7 mS/cm at 80 °C. Membrane (PNB-X54-Y46) had the highest

IEC, 2.60meq/g, but a modest conductivity, 80mS/cm at 80 °C, incomparison to PNB-X68-Y32 (IEC = 1.88meq/g). Fig. 6 shows the plotof lnσ vs. 1000/T for all the membranes. The hydroxide transport ac-tivation energy (Ea) was calculated from the slope in Fig. 6 and wasfound to be 9.33–15.28 kJmol-1. These Ea values are close to that ofNafion-117, 12.75 kJmol-1 [47].

The σ/IEC ratio is a measure of the hydroxide mobility. By thismetric, PNB-X68-Y32 had the highest hydroxide mobility, despitehaving a more modest IEC (Table 1). Conversely, the hydroxide mobi-lity in PNB-X54-Y46 was the lowest among the membranes even thoughthe IEC was the highest (Table 1). A higher IEC may be expected to havehigher mobility because there would be a greater tendency for phasesegregation. A possible explanation for the high anion mobility withPNB-X68-Y32 is its molecular weight. The higher molecular weight ofPNB-X68-Y32, compared to the other samples, contributed to its lowerwater uptake and greater chain entanglement, compared to the othersamples. The effect of molecular weight is particularly clear whencomparing the PNB-X68-Y32 (last line of Table 1) to the third and fourthlines (-Y33 and -Y38, respectively). The number of blocks is the same andthe IEC values are close (ca. 1.88–2.21). However, the σ/IEC is almostdouble for PNB-X68-Y32 due to its longer block length which enablesbetter ion channel formation. Further investigation of the effect ofmolecular weight is underway.

3.4. Water uptake (WU), hydration number (λ), number of freezable watermolecules (Nfree) and bound, non-freezable water molecules (Nbound)

Water uptake is a key parameter in determining the conductivityand mechanical stability of the AEMs. An adequate amount of water isnecessary for ion hydration and conduction. However, excess water inthe form of free water can lead to swelling and poor performance of themembrane electrode assembly (MEA) due to membrane softening andchannel flooding. Hence, an optimum amount of bound water in themembrane is required to from the ion solvent shell [16,17,48]. Asshown in Table 1, the water uptake of the membranes increased withincreasing IEC, reaching up to 133.6%. The best performing membrane,PNB-X68-Y32 had 63% WU and conductivity of 122.7 mS/cm at 80 °C.PNB-X54-Y46 had the highest IEC and consequently the highest WU of133.6% but lower ionic conductivity (80mS/cm at 80 °C) compared toPNB-X68-Y32. The hydration number (λ) is the number of water mole-cules per ionic head-groups. The high hydration number of PNB-X54-Y46 was the result of the presence of unproductive water and largerchannel size. The inter-domain spacing was 86.4 nm, as measured bySAXS. The free water content increased with domain size, as shown inTable 1. More specifically, it can be seen that the number of unboundwaters tracks with the domain size. This shows that the unproductive,free-water can populate ion channels when they are larger than theoptimum size [19]. Previously, it was shown that ion channels that aretoo small had low water uptake resulting in poor ion mobility andconductivity.

Differential scanning calorimetry was used to measure the numberof freezable water molecules (Nfree) and bound, non-freezable watermolecules (Nbound) in the membrane. In the DSC thermogram, freewater freezes just below 0 °C. Using the hydration number, the numberof free water Nfree can be obtained by subtracting the number of boundwaters. The results of all the membranes are shown in Table 1. In thecase of PNB-X68-Y32, the amount of Nfree and Nbound in the membranewas 6.7 and 11.9, respectively. This was close to the optimum numberof bound waters (9–10 per ionic pair), as previously reported [16].Consequently, this membrane also showed the highest conductivity,122.7 mS/cm at 80 °C. The presence of higher free (10.6) and bound(17.9) water for PNB-X54-Y46 resulted in a decrease in the conductivity,80mS/cm at 80 °C, although the IEC (2.60 vs. 1.88) was higher thanPNB-X68-Y32. This can be attributed to the formation of overly large ionconducting channels flooded with unproductive water. The lower freewater content (0.9) for PNB-X74-Y26 was not sufficient to support

M. Mandal et al. Journal of Membrane Science 570–571 (2019) 394–402

400

Page 8: Journal of Membrane Sciencekohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal #1 JMS.pdf · In this study, a series of tetrablock copolymers containing an all-hydrocarbon backbone

effective ion transport. Hence, the conductivity was lower (61.3mS/cmat 80 °C). In case of PNB-X67-Y33, even though the free water (7.8) wasacceptable, the higher bound water (12.1) resulted in lower con-ductivity due to the larger number of waters of hydration. The lowerfree water (5.4) and higher bound water (13.3) for PNB-X70-Y30 wasthe reason for lower ionic conductivity in the membrane.

3.5. Alkaline and thermal stability

Membrane durability is essential for long operational life in elec-trochemical devices. The alkaline stability assessment of the currentAEMs was performed by soaking the membranes in 1M NaOH solutionat 80 °C. The loss of ionic conductivity was measured vs. time for1200 h. No detectable (< 1%) loss was observed in the ionic con-ductivity over 1200 h, Fig. 7. Hence, it can be concluded that the non-hydrolysable polymer backbone with cations tethered via long sidechains displays adequate alkaline stability compared to hydrolysablepolymer backbones (e.g. polysulfone, polyketone, polyethers) and thebenzyl attachment of cations to the polymer chains.

The thermal stability of the membranes was investigated by ther-mogravimetric analysis (TGA), Fig. 8. Four degradation stages wereobserved. The first stage of decomposition, below 100 °C, was due towater loss from the membrane. The second stage around 250 °C was dueto the decomposition of the quaternary ammonium group. The thirdstage from 300 °C to 400 °C resulted from the degradation of the alkylside chains in the polymer. The fourth stage, above 400 °C, was due tothe decomposition of polymer backbone [17,37]. Our results suggestthat the membranes were sufficiently stable at the operating conditionsfor low temperature AEM fuel cells or electrolyzers, which typicallyoperate below 80 °C.

3.6. Fuel cell testing

PNB-X62-Y38 was selected for single-cell alkaline exchange mem-brane fuel cell testing because of its high ionic conductivity and ex-cellent alkaline stability. The free-standing membrane was also me-chanically robust and withstood compression in the fuel cell hardwarewithout damage.

The fuel cell was operated at 60 °C to be comparable to many otherAEMFCs found in literature. The MEA underwent a break-in startupprocedure, where the cell was discharged at 0.5 V for one hour followedby an additional hour at 0.2 V. The anode and cathode dew point wereset to 50 °C (69.51% RH) to avoid flooding of the catalyst layer. Theopen circuit voltage (OCV) was found to be 1.028 V after break-in.

After the initial conditioning period, water balance within the cellwas optimized by adjusting the dew point of the humidified H2 and O2

streams at a constant cell voltage of 0.2 V. After each dew point ad-justment, a discharge curve from open circuit to 0.2 V was recorded.Omasta et al. previously found that proper water management wasimportant for both stability and performance in AEMFCs [49]. Becausethe hydrogen oxidation reaction at the anode produces water and someof it diffuses through the AEM to hydrate the cathode, excess watercontent (< 100% RH) in the H2 input stream should be avoided. Afteradjusting the anode and cathode dew points, it was found that bothanode and cathode dew points of 46 °C (59.83% RH) provided the op-timal stability and power output. As seen in Fig. 9, a peak power densityof 542.57mW/cm2 was obtained at 0.43 V and 1.26 A/cm2. Whencorrected for iR loss across the membrane (HFR= 123mΩ cm2), the iR-corrected peak power density of the AEMFC was 713.04mW/cm2. Thisohmic resistance is higher than that reported by Omasta et al. (ca. HFR= 50mΩ cm2) and can also be attributed to other factors such mem-brane thickness (t= 115 µm) and interfacial contact resistance due to anon-optimized ionomer and catalyst layer. Nonetheless, these resultsare promising and refinements to the MEA could lead to even higherperformance.

4. Conclusion

A series of tetrablock copolymers containing all-hydrocarbonbackbone based on vinyl addition polymerization of norbornene weresynthesized for anion-exchange membranes. To the authors’ knowl-edge, this is first anion exchange membrane based on vinyl addition-type polynorbornene. These membranes displayed high thermal stabi-lity up to 400 °C. For PNB-X68-Y32, the ionic conductivity was 122.7mS/cm at 80 °C with an IEC (1.88 meq/g), which was less than PNB-X54-Y46 (2.6 meq/g). This shows the importance of optimizing thebound and unbound water content in the membrane. Water content inPNB-X68-Y32 was measured by DSC analysis and it was found that 6.7unbound water molecules and 11.9 bound water molecules in themembrane leading to the best ionic conductivity among the synthesizedsamples. The long-term alkaline stability test in 1M NaOH solution at80 °C showed exceptional chemical stability with no detectable de-gradation (< 1%) over a 1200 h period. PNB-X62-Y38 was used tofabricate an MEA for an alkaline fuel cell test which achieved excellentperformance with a peak power density of 542.57mW/cm2 at 0.43 Vand 1.26 A/cm2.

Acknowledgements

The authors gratefully acknowledge the financial support of theARPA-E IONICS (United States Department of Energy) program and thehelpful discussions of Dr. Edmund Elce.

Appendix A. Supplementary material

Supplementary data associated with this article can be found in theonline version at doi:10.1016/j.memsci.2018.10.041.

References

[1] J.R. Varcoe, P. Atanassov, D.R. Dekel, A.M. Herring, M.A. Hickner, P.A. Kohl,A.R. Kucernak, W.E. Mustain, K. Nijmeijer, K. Scott, T.W. Xu, L. Zhuang, Anion-exchange membranes in electrochemical energy systems, Energy Environ. Sci. 7(2014) 3135–3191.

[2] N.W. Li, M.D. Guiver, Ion transport by nanochannels in ion-containing aromaticcopolymers, Macromolecules 47 (2014) 2175–2198.

[3] B.C.H. Steele, A. Heinzel, Materials for fuel-cell technologies, Nature 414 (2001)345–352.

[4] Y. Leng, G. Chen, A.J. Mendoza, T.B. Tighe, M.A. Hickner, C.Y. Wang, Solid-statewater electrolysis with an alkaline membrane, J. Am. Chem. Soc. 134 (2012)9054–9057.

[5] K. Zeng, D. Zhang, Recent progress in alkaline water electrolysis for hydrogenproduction and applications, Prog. Energy Combust. Sci. 36 (2010) 307–326.

[6] P. Alotto, M. Guarnieri, F. Moro, Redox flow batteries for the storage of renewableenergy: a review, Renew. Sustain. Energy Rev. 29 (2014) 325–335.

[7] S. Lu, J. Pan, A. Huang, L. Zhuang, J. Lu, Alkaline polymer electrolyte fuel cellscompletely free from noble metal catalysts, Proc. Natl. Acad. Sci. USA 105 (2008)20611–20614.

[8] E.H. Yu, X. Wang, U. Krewer, L. Li, K. Scott, Direct oxidation alkaline fuel cells: frommaterials to systems, Energy Environ. Sci. 5 (2012) 5668–5680.

[9] J.R. Varcoe, R.C.T. Slade, Prospects for alkaline anion-exchange membranes in lowtemperature, Fuel Cells 5 (2005) 187–200.

[10] M.A. Hickner, A.M. Herring, E.B. Coughlin, Anion exchange membranes: currentstatus and moving forward, J. Polym. Sci. Part B: Polym. Phys. 51 (2013)1727–1735.

[11] J. Zhou, K. Joseph, J.M. Ahlfield, D.Y. Park, P.A. Kohl, Poly(arylene ether) ionomerswith pendant quinuclidium groups and varying molecular weight for alkalineelectrodes, J. Electrochem. Soc. 160 (2013) F573–F578.

[12] G. Merle, M. Wessling, K. Nijmeijer, Anion exchange membranes for alkaline fuelcells: a review, J. Membr. Sci. 377 (2011) 1–35.

[13] G. Couture, A. Alaaeddine, F. Boschet, B. Ameduri, Polymeric materials as anion-exchange membranes for alkaline fuel cells, Prog. Polym. Sci. 36 (2011)1521–1557.

[14] Y.J. Wang, J. Qiao, R. Baker, J. Zhang, Alkaline polymer electrolyte membranes forfuel cell applications, Chem. Soc. Rev. 42 (2013) 5768–5787.

[15] A.D. Mohanty, C. Bae, Mechanistic analysis of ammonium cation stability for al-kaline exchange membrane fuel cells, J. Mater. Chem. A 2 (2014) 17314–17320.

[16] L. Liu, P.A. Kohl, Anion conducting multiblock copolymers with different tetheredcations, J. Polym. Sci. Part A: Polym. Chem. 56 (2018) 1395–1403.

[17] L. Liu, J. Ahlfield, A. Tricker, D. Chu, P.A. Kohl, Anion conducting multiblock co-polymer membranes with partial fluorination and long head-group tethers, J.

M. Mandal et al. Journal of Membrane Science 570–571 (2019) 394–402

401

Page 9: Journal of Membrane Sciencekohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal #1 JMS.pdf · In this study, a series of tetrablock copolymers containing an all-hydrocarbon backbone

Mater. Chem. A 4 (2016) 16233–16244.[18] A.D. Mohanty, S.E. Tignor, M.R. Sturgeon, H. Long, B.S. Pivovar, C. Bae,

Thermochemical stability study of alkyl-tethered quaternary ammonium cations foranion exchange membrane fuel cells, J. Electrochem. Soc. 164 (2017)F1279–F1285.

[19] D.Y. Park, P.A. Kohl, H.W. Beckham, Anion-conductive multiblock aromatic copo-lymer membranes: structure-property relationships, J. Phys. Chem. C 117 (2013)15468–15477.

[20] X. Zhang, Q. Shi, P. Chen, J. Zhou, S. Li, H. Xu, X. Chen, Z. An, Block poly(aryleneether sulfone) copolymers tethering aromatic side-chain quaternary ammonium asanion exchange membranes, Polym. Chem. 9 (2018) 699–711.

[21] Q. Shi, P. Chen, X. Zhang, Q. Weng, X. Chen, Z. An, Synthesis and properties of poly(arylene ether sulfone) anion exchange membranes with pendant benzyl-quaternaryammonium groups, Polymer 121 (2017) 137–148.

[22] E.A. Weiber, D. Meis, P. Jannasch, Anion conducting multiblock poly(arylene ethersulfone)s containing hydrophilic segments densely functionalized with quaternaryammonium groups, Polym. Chem. 6 (2015) 1986–1996.

[23] R. Akiyama, N. Yokota, E. Nishino, K. Asazawa, K. Miyatake, Anion conductivearomatic copolymers from dimethylaminomethylated monomers: synthesis, prop-erties, and applications in alkaline fuel cells, Macromolecules 49 (2016)4480–4489.

[24] M. Shimada, S. Shimada, J. Miyake, M. Uchida, K. Miyatake, Anion conductivearomatic polymers containing fluorenyl groups: effect of the position and number ofammonium groups, J. Polym. Sci. Part A: Polym. Chem. 54 (2016) 935–944.

[25] C. Fujimoto, D.S. Kim, M. Hibbs, D. Wrobleski, Y.S. Kim, Backbone stability ofquaternized polyaromatics for alkaline membrane fuel cells, J. Membr. Sci.423–424 (2012) 438–449.

[26] S.A. Nuñez, M.A. Hickner, Quantitative 1H NMR analysis of chemical stabilities inanion-exchange membranes, ACS Macro Lett. 2 (2013) 49–52.

[27] C.G. Arges, V. Ramani, Two-dimensional NMR spectroscopy reveals cation-trig-gered backbone degradation in polysulfone-based anion exchange membranes,Proc. Natl. Acad. Sci. USA 110 (2013) 2490–2495.

[28] S. Chempath, J.M. Boncella, L.R. Pratt, N. Henson, B.S. Pivovar, Density functionaltheory study of degradation of tetraalkylammonium hydroxides, J. Phys. Chem. C114 (2010) 11977–11983.

[29] S. Chempath, B.R. Einsla, L.R. Pratt, C.S. Macomber, J.M. Boncella, J.A. Rau,B.S. Pivovar, Mechanism of tetraalkylammonium headgroup degradation in alka-line fuel cell membranes, J. Phys. Chem. C 112 (2008) 3179–3182.

[30] M.R. Hibbs, Alkaline stability of poly(phenylene)-based anion exchange membraneswith various cations, J. Polym. Sci. Part B: Polym. Phys. 51 (2013) 1736–1742.

[31] H. Long, K. Kim, B.S. Pivovar, Hydroxide degradation pathways for substitutedtrimethylammonium cations: a DFT study, J. Phys. Chem. C 116 (2012) 9419–9426.

[32] W.H. Lee, A.D. Mohanty, C. Bae, Fluorene-based hydroxide ion conducting poly-mers for chemically stable anion exchange membrane fuel cells, ACS Macro Lett. 4(2015) 453–457.

[33] H. Ono, T. Kimura, A. Takano, K. Asazawa, J. Miyake, J. Inukai, K. Miyatake,Robust anion conductive polymers containing perfluoroalkylene and pendant am-monium groups for high performance fuel cells, J. Mater. Chem. A 5 (2017)

24804–24812.[34] W.H. Lee, Y.S. Kim, C. Bae, Robust hydroxide ion conducting poly(biphenyl alky-

lene)s for alkaline fuel cell membranes, ACS Macro Lett. 4 (2015) 814–818.[35] S.C. Price, X. Ren, A.M. Savage, F.L. Beyer, Synthesis and characterization of anion-

exchange membranes based on hydrogenated poly(norbornene), Polym. Chem. 8(2017) 5708–5717.

[36] D.G. Kim, T. Takigawa, T. Kashino, O. Burtovyy, A. Bell, R.A. Register,Hydroxyhexafluoroisopropylnorbornene block and random copolymers via vinyladdition polymerization and their application as iobutanol pervaporation mem-branes, Chem. Mater. 27 (2015) 6791–6801.

[37] D.G. Kim, A. Bell, R.A. Register, Living vinyl addition polymerization of substitutednorbornenes by a t-Bu3P-ligated methylpalladium complex, ACS Macro Lett. 4(2015) 327–330.

[38] B.G. Kang, D.G. Kim, R.A. Register, Vinyl addition copolymers of norbornylnor-bornene and hydroxyhexafluoroisopropylnorbornene for efficient recovery of n-butanol from dilute aqueous solution via pervaporation, Macromolecules 51 (2018)3702–3710.

[39] N.R. Grove, P.A. Kohl, S.A.B. Allen, S. Jayaraman, R. Shick, Functionalized poly-norbornene dielectric polymers: adhesion and mechanical properties, J. Polym. Sci.Part B: Polym. Phys. 37 (1999) 3003–3010.

[40] S. Martínez-Arranz, A.C. Albéniz, P. Espinet, Versatile route to functionalizedVinylic Addition Polynorbornenes, Macromolecules 43 (2010) 7482–7487.

[41] J. Lipian, R.A. Mimna, J.C. Fondran, D. Yandulov, R.A. Shick, B.L. Goodall,L.F. Rhodes, J.C. Huffman, Addition polymerization of norbornene-type monomers.High activity cationic allyl palladium catalysts, Macromolecules 35 (2002)8969–8977.

[42] C. Wang, B. Mo, Z. He, Q. Shao, D. Pan, E. Wujick, J. Guo, X. Xie, X. Xie, Z. Guo,Crosslinked norbornene copolymer anion exchange membrane for fuel cells, J.Membr. Sci. 556 (2018) 118–125.

[43] S.J. Lue, S.J. Shieh, Water states in perfluorosulfonic acid membranes using dif-ferential scanning calorimetry, J. Macromol. Sci. Part B: Phys. 48 (2009) 114–127.

[44] B. Mecheri, V. Felice, Z. Zhang, A. D'Epifanio, S. Licoccia, A.C. Tavares, DSC andDVS investigation of water mobility in nafion/Zeolite composite membranes forfuel cell applications, J. Phys. Chem. C 116 (2012) 20820–20829.

[45] A.L. Moster, B.S. Mitchell, Hydration and proton conduction in Nafion/ceramicnanocomposite membranes produced by solid-state processing of powders frommechanical attrition, J. Appl. Polym. Sci. 113 (2009) 243–250.

[46] J. Ahlfield, G. Huang, L. Liu, Y. Kaburagi, Y. Kim, P.A. Kohl, Anion conductingionomers for fuel cells and electrolyzers, J. Electrochem. Soc. 164 (2017)F1648–F1653.

[47] B. Lin, L. Qiu, J. Lu, F. Yan, Cross-linked alkaline ionic liquid-based polymerelectrolytes for alkaline fuel cell applications, Chem. Mater. 22 (2010) 6718–6725.

[48] L. Liu, G. Huang, P.A. Kohl, Anion conducting multiblock copolymers with multiplehead-groups, J. Mater. Chem. A 6 (2018) 9000–9008.

[49] T.J. Omasta, L. Wang, X. Peng, C.A. Lewis, J.R. Varcoe, W.E. Mustain, Importanceof balancing membrane and electrode water in anion exchange membrane fuel cells,J. Power Sources 375 (2018) 205–213.

M. Mandal et al. Journal of Membrane Science 570–571 (2019) 394–402

402