Top Banner
Sulfophenylated Terphenylene Copolymer Membranes and Ionomers Thomas J. G. Skalski, [a] Michael Adamski, [a] Benjamin Britton, [a] Eric M. Schibli, [b] Timothy J. Peckham, [a] Thomas Weissbach, [a] Takashi Moshisuki, [c] Sandrine Lyonnard, [d] Barbara J. Frisken, [b] and Steven Holdcroft* [a] Introduction Acid-bearing polymers have been of high interest in recent de- cades owing to their potency as ion conducting media in vari- ous electrochemical applications, including the electrolysis of water, water purification, redox flow batteries, and hydrogen fuel cells (FCs). [1, 2] Since their conception in the 1960s, perfluor- osulfonic acid (PFSA) ionomers such as Nafion have been the technological and commercial benchmark for ion-conducting polymers. Despite their popularity, these materials suffer from drawbacks, namely: challenging syntheses and associated pro- duction costs; environmental concerns associated with (per)- fluorinated chemicals; and high reactant permeabilities, which may lead to system inefficiencies and failure. [3–5] Numerous sul- fonated polymers have been reported as alternatives to tradi- tional PFSAs, with emphasis on those containing aromatic groups within the polymer main chain. [5, 6] Materials such as poly(arylene ethers), [7, 8] poly(arylene ether ketones), [9–11] poly(ar- ylene sulfones), [11, 12] and poly(phenylenes), [9, 13–16] have been in- vestigated over recent decades. Historically, acid functionalization in these systems has been achieved through post-functionalization, typically by subject- ing a polymeric substrate to aggressive sulfonating reagents with variable concentrations or reaction times. [9] Although ef- fective, post-functionalization strategies are synonymous with a lack of reproducibility, with limited control over the location of the hydrophilic functional group on the resulting functional- ized polymer, and non-integer degrees of functionaliza- tion. [14, 17, 18] Synthesis of ion-conducting polymers from prefunc- tionalized monomers has been demonstrated to offer a distinct structural advantage over their post-functionalized analogues in that the degree of functionalization and polymeric structure are precisely controllable. [9, 13, 19–21] The presence of both hydrophilic and hydrophobic moieties on ion-conducting polymers drives phase segregation into re- spective domains, [6, 22–24] which facilitates the formation of ionic clusters and channels throughout the materials prepared therefrom. These nano- and microscopic morphological fea- tures are critical to their properties as ion-conducting media. [6, 24–27] A highly prospective approach to emphasize and exploit phase behavior in ion-containing polymers is copoly- merization, which has been frequently shown to be advanta- geous for the fabrication of membranes for electrochemical applications. [21, 28–30] A critical factor for such membranes is the total ion content, or ion exchange capacity (IEC). [31] There exists an intimate structure–morphology–property relationship The copolymerization of a prefunctionalized, tetrasulfonated oligophenylene monomer was investigated. The corresponding physical and electrochemical properties of the polymers were tuned by varying the ratio of hydrophobic to hydrophilic units within the polymers. Membranes prepared from these poly- mers possessed ion exchange capacities ranging from 1.86 to 3.50 meq g À1 and exhibited proton conductivities of up to 338 mS cm À1 (80 8C, 95 % relative humidity). Small-angle X-ray scat- tering and small-angle neutron scattering were used to eluci- date the effect of the monomer ratios on the polymer mor- phology. The utility of these materials as low gas crossover, highly conductive membranes was demonstrated in fuel cell devices. Gas crossover currents through the membranes of as low as 4 % (0.16 Æ 0.03 mA cm À2 ) for a perfluorosulfonic acid reference membrane were demonstrated. As ionomers in the catalyst layer, the copolymers yielded highly active porous electrodes and overcame kinetic losses typically observed for hydrocarbon-based catalyst layers. Fully hydrocarbon, nonfluo- rous, solid polymer electrolyte fuel cells are demonstrated with peak power densities of 770 mW cm À2 with oxygen and 456 mW cm À2 with air. [a] T. J.G. Skalski, M. Adamski, Dr. B. Britton, Dr. T. J. Peckham, Dr. T. Weissbach, Prof. S. Holdcroft Department of Chemistry, Simon Fraser University Burnaby, British Columbia V5A 1S6 (Canada) E-mail : [email protected] [b] E. M. Schibli, Prof. B. J. Frisken Department of Physics, Simon Fraser University Burnaby, British Columbia, V5A 1S6 (Canada) [c] Dr. T. Moshisuki Interdisciplinary Graduate School of Medicine and Engineering University of Yamanashi, 4 Takeda, Kofu, Yamanashi 400-8510 (Japan) [d] Prof. S. Lyonnard INAC, SPrAM, UMR 5819 CEA-CNRS-UJF CEA Grenoble, 17 rue des Martyrs, 38054 Grenoble cedex 9 (France) Supporting information and the ORCID identification number(s) for the author(s) of this article can be found under: https://doi.org/10.1002/cssc.201801965. ChemSusChem 2018, 11, 4033 – 4043 # 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 4033 Full Papers DOI: 10.1002/cssc.201801965
11

Sulfophenylated Terphenylene Copolymer Membranes and … · media.[6,24–27] Ahighly prospective approach to emphasize and exploit phase behavior in ion-containing polymers is...

Jun 09, 2020

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Sulfophenylated Terphenylene Copolymer Membranes and … · media.[6,24–27] Ahighly prospective approach to emphasize and exploit phase behavior in ion-containing polymers is copoly-merization,which

Sulfophenylated Terphenylene Copolymer Membranes andIonomersThomas J. G. Skalski,[a] Michael Adamski,[a] Benjamin Britton,[a] Eric M. Schibli,[b]

Timothy J. Peckham,[a] Thomas Weissbach,[a] Takashi Moshisuki,[c] Sandrine Lyonnard,[d]

Barbara J. Frisken,[b] and Steven Holdcroft*[a]

Introduction

Acid-bearing polymers have been of high interest in recent de-cades owing to their potency as ion conducting media in vari-

ous electrochemical applications, including the electrolysis of

water, water purification, redox flow batteries, and hydrogenfuel cells (FCs).[1, 2] Since their conception in the 1960s, perfluor-

osulfonic acid (PFSA) ionomers such as Nafion have been thetechnological and commercial benchmark for ion-conducting

polymers. Despite their popularity, these materials suffer fromdrawbacks, namely: challenging syntheses and associated pro-duction costs; environmental concerns associated with (per)-

fluorinated chemicals; and high reactant permeabilities, whichmay lead to system inefficiencies and failure.[3–5] Numerous sul-fonated polymers have been reported as alternatives to tradi-tional PFSAs, with emphasis on those containing aromatic

groups within the polymer main chain.[5, 6] Materials such aspoly(arylene ethers),[7, 8] poly(arylene ether ketones),[9–11] poly(ar-

ylene sulfones),[11, 12] and poly(phenylenes),[9, 13–16] have been in-

vestigated over recent decades.Historically, acid functionalization in these systems has been

achieved through post-functionalization, typically by subject-ing a polymeric substrate to aggressive sulfonating reagents

with variable concentrations or reaction times.[9] Although ef-fective, post-functionalization strategies are synonymous witha lack of reproducibility, with limited control over the location

of the hydrophilic functional group on the resulting functional-ized polymer, and non-integer degrees of functionaliza-tion.[14, 17, 18] Synthesis of ion-conducting polymers from prefunc-tionalized monomers has been demonstrated to offer a distinct

structural advantage over their post-functionalized analoguesin that the degree of functionalization and polymeric structure

are precisely controllable.[9, 13, 19–21]

The presence of both hydrophilic and hydrophobic moietieson ion-conducting polymers drives phase segregation into re-

spective domains,[6, 22–24] which facilitates the formation of ionicclusters and channels throughout the materials prepared

therefrom. These nano- and microscopic morphological fea-tures are critical to their properties as ion-conducting

media.[6, 24–27] A highly prospective approach to emphasize and

exploit phase behavior in ion-containing polymers is copoly-merization, which has been frequently shown to be advanta-

geous for the fabrication of membranes for electrochemicalapplications.[21, 28–30] A critical factor for such membranes is the

total ion content, or ion exchange capacity (IEC).[31] Thereexists an intimate structure–morphology–property relationship

The copolymerization of a prefunctionalized, tetrasulfonatedoligophenylene monomer was investigated. The corresponding

physical and electrochemical properties of the polymers weretuned by varying the ratio of hydrophobic to hydrophilic unitswithin the polymers. Membranes prepared from these poly-mers possessed ion exchange capacities ranging from 1.86 to3.50 meq g@1 and exhibited proton conductivities of up to 338

mS cm@1 (80 8C, 95 % relative humidity). Small-angle X-ray scat-tering and small-angle neutron scattering were used to eluci-

date the effect of the monomer ratios on the polymer mor-

phology. The utility of these materials as low gas crossover,

highly conductive membranes was demonstrated in fuel celldevices. Gas crossover currents through the membranes of as

low as 4 % (0.16:0.03 mA cm@2) for a perfluorosulfonic acidreference membrane were demonstrated. As ionomers in thecatalyst layer, the copolymers yielded highly active porous

electrodes and overcame kinetic losses typically observed forhydrocarbon-based catalyst layers. Fully hydrocarbon, nonfluo-

rous, solid polymer electrolyte fuel cells are demonstrated withpeak power densities of 770 mW cm@2 with oxygen and

456 mW cm@2 with air.

[a] T. J. G. Skalski, M. Adamski, Dr. B. Britton, Dr. T. J. Peckham,Dr. T. Weissbach, Prof. S. HoldcroftDepartment of Chemistry, Simon Fraser UniversityBurnaby, British Columbia V5A 1S6 (Canada)E-mail : [email protected]

[b] E. M. Schibli, Prof. B. J. FriskenDepartment of Physics, Simon Fraser UniversityBurnaby, British Columbia, V5A 1S6 (Canada)

[c] Dr. T. MoshisukiInterdisciplinary Graduate School of Medicine and EngineeringUniversity of Yamanashi, 4 Takeda, Kofu, Yamanashi 400-8510 (Japan)

[d] Prof. S. LyonnardINAC, SPrAM, UMR 5819 CEA-CNRS-UJFCEA Grenoble, 17 rue des Martyrs, 38054 Grenoble cedex 9 (France)

Supporting information and the ORCID identification number(s) for theauthor(s) of this article can be found under :https://doi.org/10.1002/cssc.201801965.

ChemSusChem 2018, 11, 4033 – 4043 T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim4033

Full PapersDOI: 10.1002/cssc.201801965

Breanna Hanson
Reprinted with permission from Structurally-Defined, Sulfo-Phenylated, Oligophenylenes and PolyphenylenesThomas J. G. Skalski, Benjamin Britton, Timothy J. Peckham, and Steven HoldcroftJournal of the American Chemical Society 2015 137 (38), 12223-12226DOI: 10.1021/jacs.5b07865. Copyright 2015 American Chemical Society.�
Page 2: Sulfophenylated Terphenylene Copolymer Membranes and … · media.[6,24–27] Ahighly prospective approach to emphasize and exploit phase behavior in ion-containing polymers is copoly-merization,which

between the IEC, water sorption, connectivity of hydrophilicchannels, degree of dissociation of acidic groups, and ultimate-

ly, ion conductivity of the membrane.[6, 32, 33] The developmentof new polymers and related materials with a focus on enhanc-

ing these aspects in unison are likely to enhance the propertiesof proton exchange membranes (PEM) to unprecedented

levels of performance over current standards.In the context of electrochemical technologies, equally im-

portant to the transport properties of materials is durability

and longevity. For example, the incorporation of ion-containingpolymers in FCs is considered a particularly harsh application

owing to the exposure to reactive free radicals at elevatedtemperatures. Therefore, the examination of ion-containing

polymers in the context of FCs is particularly useful for assess-ing their chemical stability in addition to being technologicallyrelevant, as these applications are the most demanding on the

ion-conducting polymeric medium. Here, the chemical stabilityis typically assessed ex situ through a Fenton’s reagent test,and in situ through a PEMFC open circuit voltage (OCV) accel-erated stress tests (ASTs).[34–36]

With PEMFC technology entering a growth in commercializa-tion, there is significant interest in the design of ion-conduct-

ing polymers with improved stabilities,[15, 37–40] the most promis-

ing of which is based on the use of polyarylene backbonesdevoid of labile linkages, such as poly(phenylene)s.[41, 42] Func-

tionalized polymers comprised entirely of para-phenylene link-ages have been reported by several groups, dating back to

Litt’s original polyphenylene sulfonic acid.[43] However, thedegree of polymerization achievable in such systems appears

to be limited, yielding materials with poor mechanical proper-

ties and solubility characteristics, preventing critical research inboth ex situ and in situ characterizations.[15, 44] An alternative

synthetic route to poly(phenylene)s involving two monomersand free of additional reagents or catalysts, which may circum-

vent the molecular weight limitations of polycondensations ofdihalogenated aromatics,[15] is the [4++2] Diels–Alder cycloaddi-tion.[42, 45, 46] Nonfunctionalized, phenylated poly(phenylene)s

(PPPs) prepared through this route possess excellent thermalstability[47] up to 550 8C in air and 575 8C under a nitrogen at-

mosphere.[41] The typical yields reported are near-quantitative,reinforcing the efficacy and potential industrial feasibility of

the synthetic approach.[41] The functionalized analogue, sulfo-nated phenylated poly(phenylene) (sPPP), was first prepared

through post-functionalization by Stille and co-workers[48] toenhance the solubility of the highly insoluble PPP, and more

recently, by the Sandia National Laboratory group for the pur-pose of preparing PEMs.[14] The latter group demonstrated that

post-functionalized sPPPs afford tough membranes possessingconsiderable electrochemical properties, and the group has

made significant contributions towards total polymer morpho-logical characterization.[16, 49–51] Owing to the high stability ofthe backbone, simple synthesis, and promising initial PEM

properties, sPPP became a topic of interest.[16, 49–51]

Unfortunately, despite promising initial reports, according tothe published literature the overall synthetic strategy had sev-eral limitations. First, polymerization of the PPP backbone

yielded a wide range of molecular weights (Mw = 12 300–172 000 Da) with dispersities (W) ranging from 1.9–4.0.[14, 41, 52]

Second, despite reporting a synthetic approach that allowed

for tuning the polymer IEC, the random and irreproduciblenature of postsulfonation gave irregular functionalizations on

the polymer backbone, in the range of 0.8–2.1 sulfonic acidgroups per repeat unit. These irregularities made it difficult to

provide definitive structural characterization. Structural regular-ity is important in obtaining well-defined, reproducible poly-

mers, as this has been linked to membranes with improved

properties.[15, 53, 54]

Recently, our group reported the synthesis of a prefunction-

alized, tetrasulfonated oligophenylene monomer based on theoriginal bis-tetracyclone (BTC),[55] which was used to prepare a

series of functionalized sPPP homopolymers and oligomerswith a precisely defined molecular structure.[13, 19] Additionally,

by modifying the dienophile monomer, we prepared homopol-

ymers with different hydrophobic segments to control the IEC,lower their water uptake and swelling, and to tune proton con-

ductivity.Herein we report a copolymer-based approach to control

and tune the properties of the ion-containing polymers forelectrochemical characteristics (Figure 1). By introducing a hy-

drophobic group (bis-tetracyclone, 6) into the polymer back-

bone through copolymerization using our previously describedpresulfonated monomer (8) in varied molar ratios, six statistical

copolymers sPPP(m)-y were prepared; here, m refers to thehydrophilic content in molar fraction, and y is the respectivecation, either (H+) or (HNEt3

+). Clear trends on the physico-chemical properties of interest, such as water sorption, dimen-

Figure 1. Preparation of sPPP(m)-H++ copolymers with tunable ion exchange capacity (IEC) values through the introduction of a hydrophobic monomer.

ChemSusChem 2018, 11, 4033 – 4043 www.chemsuschem.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim4034

Full Papers

Page 3: Sulfophenylated Terphenylene Copolymer Membranes and … · media.[6,24–27] Ahighly prospective approach to emphasize and exploit phase behavior in ion-containing polymers is copoly-merization,which

sional stability, hydrolytic, and oxidative degradation stabilitywere observed. By varying the degree of hydrophobicity

within the polymer backbone, we show precise tunability overelectrochemically important characteristics such as IEC and

proton conductivity, which we extend to demonstrations inFCs. Finally, we report X-ray and neutron diffraction scattering

structural characterization to further elucidate the effects thatIEC tuning has on the morphology of these ionic copolymers.

Experimental Section

The overall synthetic routes for the reported monomers and result-ing polymers are outlined in Scheme 1 a and Scheme 1 b, respec-tively. The syntheses of compounds 3 and 4 are detailed in theSupporting Information, whereas the syntheses of compounds 6, 7,8, 10, and 11 have been reported in our previous work.[13] Sonoga-shira cross-coupling reactions (syntheses of compounds 3 and 10)were conducted by using standard Schlenk manifold techniques.

Protodesilylation (compound 11) and sulfonation (compound 7)were performed under an argon atmosphere.

General procedure for the syntheses of sPPP(m)-H++

The general synthetic route for the synthesis of sPPP(m)-HNEt3++ is

shown in Scheme 1 b. We began with a polymerization optimiza-tion study using sPPP(0.8)-HNEt3

++ . The effects of temperature,time, solvent concentration, and dienophile/diene ratio on polymeryield and molecular weight were investigated. The obtained poly-mers were analyzed with size exclusion chromatography (SEC) indimethylformamide (DMF) and all subsequent polymerizationswere performed based on these optimizations.

General polymerization method for the synthesis ofsPPP(0.9)-H++

To a 250 mL Schlenk flask containing a stir bar, monomers 6(0.217 g, 0.314 mmol, 0.1 equiv) and 8 (4.00 g, 2.83 mmol,

Scheme 1. The synthetic pathways for the syntheses of (a) the reported monomers, and (b) the sPPP(m)-y copolymers. (i) Pd(PPh3)2Cl2, CuI, HNEt2, 56 8C, 6 h;(ii) I2, DMSO, 150 8C, 8 h; (iii) KOH, EtOH, 80 8C, 3 h; (iv) trimethylsilyl chlorosulfonate (TMS-O-SOCl2), dichloroethane (DCE), room temperature, 8 h; (v) NEt3,nBuOH, 4 h; (vi) K2CO3, Et2O/MeOH (3/1), room temperature, 6 h; (vii) nitrobenzene, 195–220 8C, 120 h; (viii) NaOH, MeOH, 4 h, room temperature; (ix) H2SO4(aq),H2O, 4 h, RT. Herein, Me = methyl, Et = ethyl, Bu = butyl.

ChemSusChem 2018, 11, 4033 – 4043 www.chemsuschem.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim4035

Full Papers

Page 4: Sulfophenylated Terphenylene Copolymer Membranes and … · media.[6,24–27] Ahighly prospective approach to emphasize and exploit phase behavior in ion-containing polymers is copoly-merization,which

0.9 equiv), were added, followed by nitrobenzene (44 mL). The so-lution was gently stirred at room temperature while it was de-gassed with argon for 30 min. Freshly sublimated 1,4-diethynylben-zene (11; 0.402 g, 3.19 mmol, 1.02 equiv) was then added, and thereaction mixture was further stirred at room temperature for30 min. The reaction vessel was then inserted into a sand bath pre-set to 220 8C, with medium stirring, for 120 h. After cooling, the so-lution was poured into boiling ethyl acetate (400 mL) and refluxedfor 4 h. The solution was filtered without cooling and the collectedpolymer precipitate was washed with boiling ethyl acetate (3 V100 mL). The obtained brown powder was dissolved in methanol(15 mL) and poured into ethyl acetate (500 mL). The resulting pre-cipitate was collected by filtration and dried under vacuum at120 8C overnight, to give sPPP(0.9)-HNEt3

++ as an off-white, fibroussolid product. Yield: 4.17 g (84.7 %). The polymers were character-ized by 1H NMR and SEC in DMF prior to further use.

Cation exchange from the triethylammonium salt to the acid formhas been outlined in our previous work.[13] The triethylammoniumsalt (@SO3

@HNEt3+) was removed first by soaking the polymer in a

basic methanolic solution (NaOH in MeOH), resulting in conversionof the polymer to an alkaline form (@SO3

@Na+). The polymer wasthen subject to an aqueous acid wash, converting it to its acidicform (@SO3

@H+). Following this conversion, the polymer was againcharacterized using 1H NMR and SEC in DMF.

Membrane preparation and characterization

The polymer acidic forms sPPP(m)-H++ were cast from 7 % w/w sol-utions in dimethyl sulfoxide (DMSO). After complete dissolution ofthe polymer, each solution was thoroughly filtered through a boro-silicate glass filtration Buchner funnel with a sintered disc (coarsefrit) by vacuum filtration, and the resulting polymer solution wascoated onto a glass plate using an adjustable doctor blade. Afterheating in an oven at 86 8C for 8 h to evaporate the DMSO, theglass plate was removed and cooled to ambient conditions. The re-sulting membrane was released from the glass plate by immersionin dilute acid, rinsed thoroughly with deionized water, and driedovernight at 80 8C under vacuum. Membranes were obtained withan average thickness of 45 mm, from which samples were then cutfor further characterization.

Mechanical properties

Mechanical properties, such as tensile strength, Young’s modulus,and elongation at break, are typically assessed by a standardizedtensile test.[56] These parameters are essential for evaluating amembrane material’s overall physical robustness and durability, es-pecially to elastic deformation.[56] Barbell-shaped membrane sam-ples were cut using a standard ASTM D638 type IV specimen cut-ting die from polymer membranes equilibrated at ambient condi-tions for a minimum of 24 h. Mechanical strength measurementswere obtained on an Instron 3344 Series single column system,with a set operational crosshead speed of 5 mm min@1. Each valuerepresents an average of at minimum three sample measurements.Error is reported as the standard deviation.

Dimensional stability, ion exchange capacity, and watersorption characteristics

Previously published[19] methodologies were used to determine thepolymer water sorption characteristics, including water uptake,water content, dimensional stability (volumetric expansion or

volume uptake), ion exchange capacity (IEC), number of water mol-ecules per sulfonic group (l, mol H2O per mol of @SO3H), mem-brane analytical acid concentration [SO3H], and effective protonmobility (meff). A detailed outline can be found in the SupportingInformation.

Stability measurements

The oxidative stability of the polymers was examined by subjectingmembrane samples to Fenton’s reagent test. The metal-catalyzeddisproportionation of H2O2 is an effective method of generatingoxygen-containing free radical species in solution, serving ascommon ex situ accelerated degradation testing for PEMs. Mem-brane samples were exposed to Fenton’s reagent (80 8C, 3 % H2O2,3 ppm Fe2 +) and removed periodically for analysis until no mem-brane visibly remained.

Proton conductivity

In-plane proton conductivity (s) was measured by AC impedancespectroscopy using conductivity cells assembled from membranesamples with a two-electrode configuration, according to a proce-dure described elsewhere.[57] Measurements were performed insidea humidity chamber held at 30 or 80 8C at various relative humidity(RH) values: 95 %, 90 %, 70 %, 50 %, and 30 % RH. Resistance datawere plotted as a Nyquist plot, which was fit to a Randles equiva-lent electrical circuit to calculate a value for membrane ionic resist-ance, from which s was calculated. Additional information can befound in the Supporting Information.

X-ray scattering

A SAXSLab Ganesha 300XL instrument was used for X-ray scatter-ing measurements. The instrument utilizes a copper anode sourceoperating at 50 kV and 0.6 mA, which produces radiation with awavelength of 1.54 a. Measurements were performed utilizing amobile PILATUS3 R 300 K photon counting detector. Small-angle(SAXS), medium-angle (MAXS), and wide-angle (WAXS) X-ray scat-tering configurations were employed to provide a q-range of rang-ing from 0.06 to 27 nm@1. Measurements were performed on sam-ples in vacuum and on hydrated samples. Samples measured invacuum were allowed to equilibrate within the evacuated instru-ment chamber for one hour. Hydrated samples were immersed indeionized water for one hour, patted dry with tissue, and mea-sured in sealed cells. Low-q features in the SAXS data were fit tothe correlation length model (CLM) as implemented in the NISTsmall-angle scattering package provided for Igor Pro [Eq. (1)] ,[58]

I qð Þ ¼ AQn þ

C1þ Qx1ð Þm þ B ð1Þ

in which x1 is the correlation length, m and n are Porod exponents,A and C are scale parameters and B is the q-independent back-ground. The first term is motivated by the power-law behavior atlow-q, the second term by scattering from regions in which thereare correlations in the electron density fluctuations, and the thirdterm, B, accounts for incoherent scattering. Mid-q features werefitted to the broad peak model [BPM, Eq. (2)] ,

I qð Þ ¼ AQn þ

C1þ Q@ Q2j jx2ð Þm þ B ð2Þ

ChemSusChem 2018, 11, 4033 – 4043 www.chemsuschem.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim4036

Full Papers

Page 5: Sulfophenylated Terphenylene Copolymer Membranes and … · media.[6,24–27] Ahighly prospective approach to emphasize and exploit phase behavior in ion-containing polymers is copoly-merization,which

also as implemented in the NIST package. Although most termsare the same as in the CLM, the model includes a peak centered atQ2, which indicates a characteristic spacing d2 is present in the ma-terial. This length scale is often approximated by the Bragg length;however, this can underestimate the characteristic length by up to22 % in polymer systems [Eq. (3)] .[59]

d2 ¼2p

Q2

ð3Þ

Neutron scattering

Neutron scattering was performed at the D22 beamline at the In-stitut Laue-Langevin in Grenoble, France. Measurements were per-formed using two detector positions, which provided a q-range of0.06 to 6 nm@1. Samples were measured after equilibration with at-mospheres of controlled humidity and temperature, or with D2O atcontrolled temperature. Samples were humidified in sealed con-tainers containing saturated salt solutions at 20 8C for 72 h.[60] Sam-ples were hydrated by immersion in D2O for at least 3 h at 20, 40,or 60 8C. These samples were loaded into sealable sample cellswith two drops of excess D2O. The 2D scattering patterns were ra-dially averaged to obtain the 1D spectra, which were then correct-ed for detector efficiency, transmission, and empty-cell backgroundscattering. Mid-q features in ambient humidity and fully hydratedsamples were fit to the BPM [Eq. (2)] .

Fuel cell characterization

Membrane-electrode assemblies (MEAs) incorporating sPPP(m)-H++

polymers were formed, integrated into FC systems, and character-ized in situ as described in the Supporting Information. PolymerssPPP(0.6)-H++ , sPPP(0.8)-H++ , or sPPP(1.0)-H++ were incorporatedinto these systems as both ionomer in the catalyst layers (20 wt %ionomer of solids to a total loading of 0.4 mgPt cm@2) and mem-brane, yielding an assessment of each sPPP(0.6)-H++ , sPPP(0.8)-H++ ,and sPPP(1.0)-H++ as fully hydrocarbon solid polymer electrolyteFCs. These were compared with a wholly PFSA reference MEA com-posed of Nafion D520 in the catalyst layers (30 wt % ionomer ofsolids to a total loading of 0.4 mgPt cm@2) and Nafion 212 mem-brane.

Electrochemical characterization was performed by using a com-bined potentiostat/frequency-response analyzer (Princeton AppliedResearch VersaSTAT 4) after equilibration of the fuel cell to a stablelow potential <0.15 V under fully humidified 0.25/0.5 slpm H2/N2

anode/cathode gas feeds. Chronoamperometry was performed at100 mV step@1 from 0 to 600 mV at 30 s step@1, and linear sweepvoltammetry at 2 mV s@1 from OCV to 600 mV for the measurementof fuel crossover and detection of shorting, respectively. Cathodeionomer conductivity was measured by electrochemical impedancespectroscopy. Using an inert (0 slpm) N2 feed, cyclic voltammetrywas performed from 0.05 to 0.8 V up to 20 cycles or measurementdestabilization for the measurement of the electrocatalyticallyactive surface area. Electrochemical data was processed as de-scribed previously.[61] After disassembly, MEAs were freeze-fracturedwith liquid N2 and the membrane and electrode thicknesses weremeasured by scanning electron microscopy (FEI/Aspex Explorer).

Results and Discussion

Synthesis, casting, and characterizations

Five random copolymers sPPP(m)-HNEt3++ and one homopoly-

mer sPPP(1.0)-HNEt3++ were prepared from the presulfonated

diene monomer BTC(SO3HNEt3+) (8), its unsulfonated diene an-

alogue BTC (6), and dienophile 1,4-diethynylbenzne (11). Inthis study, we investigated the properties of these polymers,

from sPPP(1.0) to the copolymer sPPP(0.5), which contains50 % of each presulfonated and unsulfonated monomers. Theoptimum polymerization conditions were 120 h reaction timein nitrobenzene solvent at a monomer concentration of96 g L@1 and temperature of 220 8C. Numerous other solvents,namely DMSO, DMF, N-methyl-2-pyrrolidone (NMP), dimethyl-

acetamide (DMAc), and sulfolane, were ineffective at yielding

polymers with sufficient molecular weights. The same syntheticstrategy was employed for each copolymer. The yield de-

creased with increasing hydrophobic content, which was at-tributed to poorer solubility of the hydrophobic backbone in

nitrobenzene. Polymer purity was assessed by 1H NMR spec-troscopy using the triethylammonium cations on sPPP(m)-HNEt3

++ as internal probes. Terminal ethynyl functional groups

were observed (singlet at 4.12 ppm), which were used to esti-mate the molecular weight (detailed in the Supporting Infor-

mation). Triethylammonium cations were removed through atwo-step process involving exchange to the sodium form

sPPP(m)-Na++ by immersion in methanolic sodium hydroxidesolution, followed by immersion in an aqueous sulfuric acid so-

lution, yielding the respective acidic form sPPP(m)-H++ . After

drying under vacuum at 120 8C to remove the residual water,the polymer molecular weights were investigated. Both the

triethylammonium and acid forms were characterized usingsize-exclusion chromatography (SEC) in DMF containing LiBr

(0.01 mol L@1). The average molecular weights are reported inTable 1.

These data show that the molecular weights of sPPP(m)-HNEt3

++ decreased with increasing hydrophobic character. Thisobservation suggested that polymer solubility in the nitroben-zene solvent system may be a limiting factor with increasinghydrophobic character. To further probe this hypothesis, the

fully hydrophobic PPP was synthesized from BTC (6) and 1,4-

Table 1. Yield and molecular weights of sPPP(m)-HNEt3++ and sPPP(m)-

H++ , as determined by SEC in DMF.

Polymer sPPP(m)-HNEt3++ sPPP(m)-H++

Yield [%][a] Mw[a]

[Da]W[a] Yield [%][b] Mw

[b]

[Da]W[b]

sPPP(1.0) 89.1 363 800 1.86 92.1 290 900 1.90sPPP(0.9) 81.1 347 000 2.41 89.2 221100 1.83sPPP(0.8) 89.0 240 000 1.96 >99 225 000 2.01sPPP(0.7) 79.1 267 800 2.63 >99 176 000 2.43sPPP(0.6) 85.6 217 000 2.75 93.2 103 550 3.50sPPP(0.5) 83.2 198 000 3.43 92.8 161 400 2.87PPP 67.2 77 600 2.57 N/A N/A N/A

ChemSusChem 2018, 11, 4033 – 4043 www.chemsuschem.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim4037

Full Papers

Page 6: Sulfophenylated Terphenylene Copolymer Membranes and … · media.[6,24–27] Ahighly prospective approach to emphasize and exploit phase behavior in ion-containing polymers is copoly-merization,which

diethynylbenzene (11), which resulted in material with aweight average molecular weight (Mw) of only 77 600 Da. The

molecular weights of functionalized polymers decreased afterconversion from triethylammonium to acid form for each poly-

mer, supporting a successful exchange to a repeat unit oflower molecular weight (removal of 4 triethylamine molecules

per repeat unit).Polymer acidic forms sPPP(m)-H++ were cast into membranes

using the aforementioned techniques, and their membraneproperties were investigated. Mechanical strength measure-ment data showed that all sPPP(m)-H++ membranes had similar

tensile strength (43.8:5.1 MPa) and Young’s modulus (1228:191 MPa), larger than that of NR-211 (17.3:0.4 MPa and 270:17 MPa, respectively), but displayed lower elongation at breakvalues (Table S9 and Figure S24). The data obtained from these

measurements suggested that the rigid-rod polyphenylene

backbone employed provides significant mechanical strengthand resistance to elastic deformation,[42] properties that are in-

tegral to robustness and longevity in electrochemical devicemanufacturing and operation.[35] Ionic content does not appear

to significantly affect the material mechanical properties. Thefindings presented herein are consistent with previously pub-

lished mechanical properties of post-sulfonated phenylated

polyphenylenes,[14] as well as presulfonated derivatives.[19]

Theoretical (IECth) and experimental (IECxp) ion exchange ca-

pacity values, calculated and determined by titration experi-ments, respectively, are shown in Figure 2. They follow a linear

trend from sPPP(0.5)-H++ (IECth = 2.17 meq g@1) to the homopo-

lymer sPPP(1.0)-H++ (IECth = 3.70 meq g@1). The experimentalIECs closely matched their theoretical counterparts, as previ-

ously observed.[13, 19] These data provide insight to polymertunability according to the ratios of hydrophilic and hydropho-

bic monomers used. The water uptake (WU%), water content

(WC%), volume uptake (VU%), water sorption (l, number ofprotons per sulfonic acid: mol H2OCmol@1 SO3H), acid concen-

tration ([@SO3H]), and material densities (g cm@3) of each poly-mer membrane are summarized in Table 2. High hydrophilic

content polymers sPPP(1.0)-H++–sPPP(0.9)-H++ showed exces-sive swelling and eventual dissolution in water at 80 8C but re-

mained intact at room temperature. sPPP(0.8)-H++–sPPP(0.5)-H++ were fully insoluble.

sPPP(1.0)-H++ showed the largest water and volume uptakes,

which declined significantly with increasing hydrophobic char-acter. Similarly, polymer water content showed a gradual de-

cline with increasing hydrophobic character from sPPP(1.0)-H++

to sPPP(0.5)-H++ . The number of water molecule per sulfonicacid was highest in sPPP(1.0)-H++ , which consequently pos-

sessed one of the lowest acid concentrations.In-plane ionic conductivity measurements were conducted

at 30 and 80 8C between 30 % and 95 % RH. Data are summar-ized in Figures 3 and 4 and Tables S7 and S8. There was a clear

trend with increasing hydrophobic character and decreasing

Figure 2. Theoretical and experimental IEC values for sPPP(m)-H++ polymers.

Table 2. Properties of sPPP(m)-H++ membranes.

Polymer WU [%] WC [%] VU [%] l [@SO3H] Density [g cm@3]

sPPP(1.0)(H++) 319.3 76.1 364.3 50.7 0.85 1.17sPPP(0.9)(H ++) 278.3 73.6 348.8 47.2 0.84 1.16sPPP(0.8)(H ++) 191.7 65.7 200.0 37.0 1.10 1.15sPPP(0.7)(H ++) 125.4 55.6 141.9 25.0 1.34 1.11sPPP(0.6)(H ++) 90.5 47.5 93.9 20.2 1.35 1.09sPPP(0.5)(H ++) 65.1 40.3 67.7 19.4 1.51 1.06

Figure 3. (a) Proton conductivities of sPPP(m)-H++ and Nafion NRE 211 at30 8C and variable RH and (b) the respective log plot.

ChemSusChem 2018, 11, 4033 – 4043 www.chemsuschem.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim4038

Full Papers

Page 7: Sulfophenylated Terphenylene Copolymer Membranes and … · media.[6,24–27] Ahighly prospective approach to emphasize and exploit phase behavior in ion-containing polymers is copoly-merization,which

proton conductivities in the polymers examined. At 30 8C and

95 % RH, sPPP(1.0)-H++ displayed a noteworthy conductivity(120 mS cm@1) compared with Nafion 211 (79 mS cm@1). Theconductivity of the highly hydrophobic sPPP(0.5)-H++

(0.5 mS cm@1) was significantly lower under identical condi-tions. At 80 8C and 95 % RH, sPPP(1.0)-H++ showed excellent

conductivity (338 mS cm@1), and polymers with increasing hy-drophobic character up to sPPP(0.6)-H++ (133 mS cm@1) hadhigher conductivity than Nafion 211 (113 mS cm@1). These data,along with the trends observed in polymer water sorption,

suggested that the introduction of hydrophobic co-monomersinto the prefunctionalized proton conducting polymers is aneffective means of tuning both their physicochemical and elec-trochemical properties.

Membrane stability

To demonstrate the resilience of a wholly aromatic polymer

backbone, a polymer oxidative stability test was conducted bysubjecting membrane samples to Fenton’s reagent. In the case

of sPPP(1.0)-H++ and sPPP(0.9)-H++ , the membrane samplesdissolved prior to completion of the analysis, likely owing to

their high hydrophilic ratios, as previously reported.[13] Howev-er, in these instances, it was possible to quench the Fenton’s

reagent solutions using excess sodium sulfite, which resultedin precipitation of the dissolved polymers from solution and al-lowed for collection of material for subsequent analysis. After1 h exposure, all sPPP(m)-H++ membrane samples were visually

intact, displayed no observable mass loss (Figure S17), andwere chemically unchanged when assessed by 1H NMR (Figur-es S18–S23). After 3 h total exposure time, sample dissolutionhindered our ability to accurately determine the residual massfor sPPP(1.0)-H++ and sPPP(0.9)-H++ samples, but recoveredmaterial further showed no changes to chemical structurewhen assessed by 1H NMR. The residual masses of membrane

samples (excluding precipitated materials) are summarized inFigure S17. These data reflect the high oxidative stability of a

polyphenylene backbone and expand on previously publishedresults for these types of systems.[13, 19, 62] In stark contrast, a re-

cently reported class of highly phenylated, postsulfonated pol-

y(arylene ether)s with similar IECs retained as little as 80 % ofinitial sample masses, or dissolved entirely, after just 1 h under

these conditions.[63]

Morphological analysis

Small angle X-ray scattering data are presented in Figure 5, inwhich vacuum measurements in vacuum are shown in Fig-

ure 5 a and measurements in ambient humidity are shown inFigure 5 b, for samples containing monomers that are 50 to

100 % sulfonated. All samples show a mid-q feature (Feature 2,at approximately 3 nm@1) and a pair of high-q peaks (Feature 3,

q>10 nm@1). Feature 2 and the second peak in Feature 3 are

Figure 4. (a) Proton conductivities of sPPP(m)-H++ and Nafion NRE 211 at80 8C and variable RH and (b) the respective log plot.

Figure 5. SAXS measurements for sPPP(m)-H++ samples with different chargecontent (a) in vacuum, and (b) in paste cells at ambient humidity. Fits of theCLM [Eq. (1)] to Feature 1 for sPPP(0.5), sPPP(0.6), and sPPP(0.7) are shownas dashed curves, whereas fits of the BPM [Eq. (2)] to Feature 2 for all sam-ples are shown as solid curves.

ChemSusChem 2018, 11, 4033 – 4043 www.chemsuschem.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim4039

Full Papers

Page 8: Sulfophenylated Terphenylene Copolymer Membranes and … · media.[6,24–27] Ahighly prospective approach to emphasize and exploit phase behavior in ion-containing polymers is copoly-merization,which

more prominent in the ambient humidity measurements, butotherwise differences between measurements in the two envi-

ronments are not remarkable. Feature 2 reveals the presenceof nanoscale organization in the range of a few nanometers,

whereas Feature 3 probably arises from interchain correlations.In addition, samples with fewer sulfonated monomers,sPPP(0.5)-H++ , sPPP(0.6)-H++ , and sPPP(0.7)-H++ , show a low-qfeature (Feature 1) located at approximately 0.5 nm@1. The

small angle behavior dominated in all samples by a steep in-tensity upturn, which was related to the presence of long-range inhomogeneity, a typical behavior observed in randomlydistributed polymer aggregates or bundles.

The SAXS data were complemented by small angle neutronscattering (SANS) to further evaluate the origin of the variouspeaks, with particular focus on the nanoscale organization.

Neutron scattering results are presented in Figure 6, in which

measurements on samples of different charge content hydrat-ed in D2O are shown in Figure 6 a, whereas Figures 6 b and 6c

show data for sPPP(0.7)-H++ and sPPP(0.9)-H++ at various levelsof hydration. Features 1 and 2 are also visible in the SANS

data. Feature 1 is only clearly visible in the data for fully hy-drated sPPP(0.5)-H++ and sPPP(0.7)-H++ ; for the latter it is clear-

er in Figure 6 b than in Figure 6 a. In Figure 6 a, Feature 2 ap-

pears in samples with higher charge content, generally shiftingto lower-q as the charge content is increased except for

sPPP(0.9)-H++ , in which the feature is less pronounced. Fea-ture 2 clearly strengthens and shifts to lower q as the level of

hydration is increased for both sPPP(0.7)-H++ and sPPP(0.9)-H++

, as shown in Figures 6 b and 6 c. The SANS measurements

were not made to high enough q for the high-q feature seen

in WAXS (Feature 3) to be visible. Samples were also hydratedin mixtures of H20 and D20 (data not shown). All features com-

pletely disappeared in mixtures that were 60:40 H2O/D2O. Cal-culations indicate that the scattering length density of the

polymer should match that of a 60:40 H2O/D2O mixture.The correlation length model (CLM) was fit to Feature 1 in

the X-ray data for sPPP(0.5)-H++ , sPPP(0.6)-H++ , and sPPP(0.7)-H++ , yielding correlation lengths ranging from 1.53:0.08 to1.36:0.01 nm, as shown in Table 3. This length scale is not sig-nificantly different in vacuum or ambient measurements butappears to decrease with increasing sulfonation. Because this

feature is not present in samples that contain few unsulfonat-ed monomers, and because it is independent of water content,

we believe that this length scale corresponds to individual

monomers. Scattering from sulfonated and unsulfonated mon-omers will be different; we estimated a X-ray scattering length

density of 10.35 V 10@6 a@2 for the fully sulfonated monomers[64]

and 9.49 V 10@6 a@2 for the unsulfonated monomers.[65] The size

of the correlation length was consistent with an estimated mo-nomer length of 1.5 nm; this feature was not present in an

analysis of randomly postsulfonated polyphenylenes,[16] in

which the monomers are not distinct. The difference of theneutron scattering length densities is smaller, consistent with

the fact that this feature is less clear in the SANS data. We alsoattempted to fit the BPM to Feature 1 in the data. However,

the feature was not distinct enough to resolve both formfactor and structure factor.

The broad peak model (BPM) was fit to Feature 2 in the X-ray and neutron data, except for sPPP(0.5)-H++ , for which Fea-

ture 2 was not visible. The correlation length associated with

the size of the clusters, x2, was approximately 1 nm for all fits,but the characteristic spacing associated with this feature, d2,

varied with hydration and charge content. The results for sam-ples at different levels of hydration are shown in Figure 7 a.

The characteristic spacing, d2, was just under 2 nm until thehumidity reached approximately 50 %. At this point, it in-

Figure 6. SANS measurements for (a) samples of different charge content hy-drated in D2O at 20 oC, (b) sPPP(0.7)-H++ at various levels of hydration, and(c) sPPP(0.9)-H++ at various levels of hydration; both parts (b) and (c) alsocompare the data for the fully hydrated sample at 20 and 40 8C. In (a), the q-independent background terms, as determined from fits to the data, havebeen subtracted. Measurements of the fully hydrated samples have beenvertically offset for clarity. The fits of the BPM to Feature 2 are included in (a)as solid curves.

Table 3. Results of fitting the correlation length model [Eq. (1)] to Fea-ture 1 in the SAXS data.

Material x1 , vacuum [nm] x1 , ambient [nm]

sPPP(0.5)(H ++) 1.46:0.03 1.53:0.07sPPP(0.6)(H ++) 1.42:0.08 1.40:0.05sPPP(0.7)(H ++) 1.36:0.01 1.38:0.01

ChemSusChem 2018, 11, 4033 – 4043 www.chemsuschem.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim4040

Full Papers

Page 9: Sulfophenylated Terphenylene Copolymer Membranes and … · media.[6,24–27] Ahighly prospective approach to emphasize and exploit phase behavior in ion-containing polymers is copoly-merization,which

creased with increasing humidity by an amount that depended

on the charge content. In the fully hydrated samples (Fig-ure 7 b), we observed length scales ranging from 2.67:0.07 to4.3:0.5 nm. Given that Feature 2 strengthens and shifts to

larger length scales with increasing hydration as the humiditywas varied and with increasing charge content as measured infully hydrated samples, we believe that this was associatedwith the spacing between small nanometer clusters of ions

and water in the polymer. A similar feature was observed inprevious measurements of randomly postsulfonated polyphe-

nylenes; however, the length scale associated with the feature

did not change systematically with degree of sulfonation inthe postsulfonated materials.[60] Our observations are summar-

ized visually in the schematic in Figure 8.

Analyses in fuel cells

Fully hydrocarbon sPPP-based MEAs were formed using

sPPP(0.6)-H++ , sPPP(0.8)-H++ , and sPPP(1.0)-H++ as both mem-brane and ionomer in the catalyst layer and operated as FCs.

The composition, construction, and conditioning of MEAs areas described in the Supporting Information. sPPP-based mem-

branes exhibited very high in situ ionic conductivity (Fig-ure 9 a): for sPPP(0.6)-H++ , sPPP(0.8)-H++ , and sPPP(1.0)-H++

in situ conductivities at 80 8C, 100 % RH were measured to be74:2, 244:28, and 256:34 mS cm@1, respectively, compared

with 81:1 mS cm@1 for Nafion 212. These values correspondedwell with the conductivities measured ex situ at 95:5 % RH

(see Figure 4), previous reports,[13] and similar materials.[19]

H2 fuel crossover currents for sPPP(0.6)-H++ , sPPP(0.8)-H++ ,and sPPP(1.0)-H++ were very low: 0.41:0.02, 0.57:0.01, and

0.16:0.03 mA cm@2, respectively, compared with the 3.73:0.06 mA cm@2 measured for the PFSA reference. Thus, fully hy-

drocarbon sPPP-type FCs exhibited only 4–15 % of the fuel (H2)crossover of the PFSA reference membrane of similar thick-

nesses, as per Figure 9 b.Polarization data was collected with cathode gas feeds of

oxygen, Figure 10 a, and air, Figure 10 b. For sPPP(0.6)-H++ ,sPPP(0.8)-H++ , and sPPP(1.0)-H++ , peak power densities inPFSA-optimized conditions in oxygen were 500, 770, and

418 mW cm@2, respectively, compared with 792 mW cm@2 forthe PFSA reference (Figure S32 b); in air, 244, 456, and

256 mW cm@2, respectively, compared with 455 mW cm@2 forthe PFSA reference (Figure S32 c). Time scales relevant to the

determination of equilibrated water transport (5 min pt@1), and

all sPPP-based systems show equivalent water transport toPFSAs, thus showing stability, repeatability, and equal or lower

mass-transport losses for equivalent non-MEA components.Both sPPP(0.8)-H++ and sPPP(1.0)-H++ formed low-resistance

membranes (see Figure S32 a). However, sPPP(1.0)-H++ exhibit-ed significant kinetic-region losses which we attributed to sig-

Figure 7. The broad peak model was fit to Feature 2 in the X-ray data[Eq. (2)] , in which d2 was calculated by using Equation (3). Part (a) shows fitresults for the SAXS (vacuum and ambient) and SANS measurements per-formed at different levels of humidity. Part (b) shows the SANS measure-ments of fully hydrated samples at 20 8C with varying degrees of hydrophilicmonomer content (denoted sulfonation %). Open markers are used to repre-sent SAXS results, whereas closed markers are used to represent SANSresults.

Figure 8. Schematic representations of our interpretation of the origins ofFeatures 1 and 2. (a) Feature 1 is associated with the size of the monomer, inwhich x1 is the size of a monomer. If the majority of monomers are thesame, this length scale cannot be observed, but when both monomers arepresent, this feature is observable because the two monomers scatter differ-ently. (b) Feature 2 is associated with the spacing and size of the water-richregions in the samples. This region facilitates ion conductivity. d2 is the dis-tance between water pockets. The minimum size for d2 is the distance be-tween two sulfonic acid groups (15 a), but even a “dry” membrane contains3–4 water molecules per sulfonic acid. As water content increases, d2 grows.

ChemSusChem 2018, 11, 4033 – 4043 www.chemsuschem.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim4041

Full Papers

Page 10: Sulfophenylated Terphenylene Copolymer Membranes and … · media.[6,24–27] Ahighly prospective approach to emphasize and exploit phase behavior in ion-containing polymers is copoly-merization,which

nificant in situ swelling of ionomer in the catalyst layer. Full po-

larization of fully hydrocarbon systems is rare in the literature,and the equivalence of the fully sPPP(0.8)-H++ system to the

PFSA reference in all of the kinetic, Ohmic, and mass transportregimes is a very promising result for future development.

Conclusions

The control and tunability of the physical and electrochemicalproperties of sulfonated and phenylated polyphenylenes was

demonstrated by copolymerization of ion-containing prefunc-tionalized and nonfunctionalized monomers. Variation of theratios of these monomers resulted in changes in the polymericstructure, which were elucidated through small-angle X-ray

and neutron scattering experiments. sPPP(m)-H++ polymerswith weight average molecular weights of up to 291 kDa wereachieved (DPn = 269), from which tough yet flexible mem-

branes were prepared and characterized for their physico-chemical and electrochemical properties. Experimental ion ex-

change capacities from 1.86 to 3.50 meq g@1 were observed,which corresponded to 50–100 % functionalized monomers

and displayed ex situ proton conductivities of 62–338 mS cm@1

at 80 8C and 95 % relative humidity. Ex situ oxidative stabilitymeasurements highlighted the chemical integrity of the purely

aromatic backbone free of labile linkages; after 3 h total expo-sure time, no changes to the chemical structure were ob-

served. Finally, in situ FC characterization yielded a promisingassessment of the sPPP(m)-H++ efficacy as fully hydrocarbon

solid polymer electrolyte FCs, with peak power densities meas-

uring 770 mW cm@2 in oxygen and 456 mW cm@2 in air. Cumula-

tively, the work reported herein demonstrates the feasibility ofpurely hydrocarbon approaches to electrochemical systems,such as FCs, as a true competitor to traditional PFSA-basedmaterials.

Acknowledgements

Financial support for this study was provided by Natural Sciences

and Engineering Research Council of Canada (NSERC) and theCatalysis Research for Polymer Electrolyte Fuel Cells (CaRPE-FC)

Network, administered by SFU and supported by AutomotivePartnership Canada (APC) Grant No. APCPJ417858-11. Research

described in this work made use of the 4D LABS shared facilities

at SFU supported by the Canada Foundation for Innovation (CFI),British Columbia Knowledge Development Fund (BCKDF), Western

Economic Diversification Canada (WD), and Simon Fraser Univer-sity (SFU). Neutron scattering experiments were performed at In-

stitut Laue-Langevin in Grenoble, France with help from Drs.Christina Iojiou and Lionel Porcar. This research was enabled, in

Figure 9. Membrane data for sPPP-based systems compared with a PFSA ref-erence for a) membrane conductivity from in situ data (iR drop method) andb) hydrogen fuel crossover data determined by chronoamperometry.

Figure 10. sPPP(1.0)-H++ , sPPP-(0.8)-H++ , and sPPP(0.6)-H++ as both mem-brane and ionomer in the catalyst layer, i.e. , hydrocarbon-based FCs, com-pared with a standard Nafion reference, operating at zero backpressure(100 kPaabs) and conditions optimized for the Nafion reference: 95 % RH,80 8C, 0.5/1.0 slpm gas flow of H2/O2 (a) or air (b) at the anode/cathode; alldata 5 min pt@1 at 100–200 mA cm@2 intervals with 1 min pt@1 at mA cm@2 in-tervals to resolve the kinetic region; 0.4 mgPt cm@2 cathode/anode catalystloadings with composition and conditioning as described above and in thesupporting information. with composition and conditioning as describedabove and in the supporting information.

ChemSusChem 2018, 11, 4033 – 4043 www.chemsuschem.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim4042

Full Papers

Page 11: Sulfophenylated Terphenylene Copolymer Membranes and … · media.[6,24–27] Ahighly prospective approach to emphasize and exploit phase behavior in ion-containing polymers is copoly-merization,which

part, by support provided by WestGrid (www.westgrid.ca) andCompute Canada Calcul Canada (www.computecanada.ca). The

authors acknowledge Dr. Alexandra Furtos-Matei and Marie-Christine Tang from the Centre R8gionale de Spectrom8trie de

Masse de l’Universit8 de Montr8al for high-resolution mass spec-trum analyses.

Conflict of interest

The authors declare no conflict of interest.

Keywords: copolymers · fuel cells · membranes · ionomers ·electrochemistry

[1] M. Carmo, D. L. Fritz, J. Mergel, D. Stolten, Int. J. Hydrogen Energy 2013,38, 4901 – 4934.

[2] H. Zhang, P. K. Shen, Chem. Rev. 2012, 112, 2780 – 2832.[3] J. Ere, D. J. Jones, Annu. Rev. Mater. Res. 2003, 33, 503 – 555.[4] K.-D. Kreuer, Chem. Mater. 2014, 26, 361 – 380.[5] J. Miyake, K. Miyatake, Polym. J. 2017, 49, 487 – 495.[6] T. J. Peckham, S. Holdcroft, Adv. Mater. 2010, 22, 4667 – 4690.[7] T. Miyahara, T. Hayano, S. Matsuno, M. Watanabe, K. Miyatake, ACS Appl.

Mater. Interfaces 2012, 4, 2881 – 2884.[8] H. F. Lee, P. H. Wang, Y. C. Huang, W. H. Su, R. Gopal, C. C. Lee, S. Hold-

croft, W. Y. Huang, J. Polym. Sci. Part A 2014, 52, 2579 – 2587.[9] G. Maier, J. Meier-Haack, Adv. Polym. Sci. Fuel Cells II 2008, 216, 1 – 62.

[10] J. Miyake, M. Watanabe, K. Miyatake, ACS Appl. Mater. Interfaces 2013, 5,5903 – 5907.

[11] H. F. Lee, B. Britton, Y. C. Huang, T. J. Peckham, Y. Y. Hsu, Y. C. Tseng, P. C.Huang, C. C. Lee, M. Y. Chang, S. Holdcroft, W.-Y. Huang, J. Mater. Sci.2016, 51, 9805 – 9821.

[12] W. L. Harrison, M. Hickner, Y. S. Kim, J. E. McGrath, Fuel Cells 2005, 5,201 – 212.

[13] T. J. G. Skalski, B. Britton, T. J. Peckham, S. Holdcroft, J. Am. Chem. Soc.2015, 137, 12223 – 12226.

[14] C. Fujimoto, M. Hickner, C. Cornelius, D. Loy, Macromolecules 2005, 38,5010 – 5016.

[15] K. Si, D. Dong, R. Wycisk, M. Litt, J. Mater. Chem. 2012, 22, 20907 –20917.

[16] L. He, C. Fujimoto, C. Cornelius, D. Perahia, Macromolecules 2009, 42,7084 – 7090.

[17] E. M. W. Tsang, Z. Zhang, A. C. C. Yang, Z. Shi, T. J. Peckham, R. Narimani,B. J. Frisken, S. Holdcroft, Macromolecules 2009, 42, 9467 – 9480.

[18] A. Kraytsberg, Y. Ein-Eli, Energy Fuels 2014, 28, 7303 – 7330.[19] M. Adamski, T. J. G. Skalski, B. Britton, T. J. Peckham, L. Metzler, S. Hold-

croft, Angew. Chem. Int. Ed. 2017, 56, 9058 – 9061; Angew. Chem. 2017,129, 9186 – 9189.

[20] K. B. Wiles, C. M. de Diego, J. de Abajo, J. E. McGrath, J. Membr. Sci.2007, 294, 22 – 29.

[21] J. R. Rowlett, Y. Chen, A. T. Shaver, O. Lane, C. Mittelsteadt, H. Xu, M.Zhang, R. B. Moore, S. Mecham, J. E. Mcgrath, Polymer 2013, 54, 6305 –6313.

[22] K. D. Kreuer, J. Membr. Sci. 2001, 185, 29 – 39.[23] T. J. Peckham, J. Schmeisser, M. Rodgers, S. Holdcroft, J. Mater. Chem.

2007, 17, 3255.[24] T. Weissbach, E. M. W. Tsang, A. C. C. Yang, R. Narimani, B. J. Frisken, S.

Holdcroft, J. Mater. Chem. 2012, 22, 24348 – 24355.[25] N. Li, M. D. Guiver, Macromolecules 2014, 47, 2175 – 2198.[26] G. He, Z. Li, J. Zhao, S. Wang, H. Wu, M. D. Guiver, Z. Jiang, Adv. Mater.

2015, 27, 5280 – 5295.[27] A. C. C. Yang, R. Narimani, B. J. Frisken, S. Holdcroft, J. Membr. Sci. 2014,

469, 251 – 261.[28] Y. A. Elabd, M. Hickner, Macromolecules 2011, 44, 1 – 11.[29] J. Ding, C. Chuy, S. Holdcroft, Chem. Mater. 2001, 13, 2231 – 2233.

[30] J. R. Rowlett, V. Lilavivat, A. T. Shaver, Y. Chen, A. Daryaei, H. Xu, C. Mit-telsteadt, S. Shimpalee, J. S. Riffle, J. E. McGrath, Polymer 2017, 122,296 – 302.

[31] Y. A. Elabd, E. Napadensky, C. W. Walker, K. I. Winey, Macromolecules2006, 39, 399 – 407.

[32] J. Peron, A. Mani, X. Zhao, D. Edwards, M. Adachi, T. Soboleva, Z. Shi, Z.Xie, T. Navessin, S. Holdcroft, J. Membr. Sci. 2010, 356, 44 – 51.

[33] D. W. Shin, M. D. Guiver, Y. M. Lee, Chem. Rev. 2017, 117, 4759 – 4805.[34] J. Wu, X. Z. Yuan, J. J. Martin, H. Wang, J. Zhang, J. Shen, S. Wu, W.

Merida, J. Power Sources 2008, 184, 104 – 119.[35] F. N. Bechi, M. Inaba, T. J. Schmidt, Polymer Electrolyte Fuel Cell Durability,

Springer, Berlin, 2009.[36] M. Mench, E. C. Kumbur, T. N. Veziroglu, Polymer Electrolyte Fuel Cell

Degradation, Elsevier Inc. , Amsterdam, NLD., 2011.[37] T. Mochizuki, M. Uchida, K. Miyatake, ACS Energy Lett. 2016, 1, 348 – 352.[38] K. Si, R. Wycisk, D. Dong, K. Cooper, M. Rodgers, P. Brooker, D. Slattery,

M. Litt, Macromolecules 2013, 46, 422 – 433.[39] S. Takamuku, P. Jannasch, Macromolecules 2012, 45, 6538 – 6546.[40] K. S. Yoon, J. Y. Lee, T. H. Kim, D. M. Yu, D. W. Seo, S. K. Hong, Y. T. Hong,

J. Ind. Eng. Chem. 2014, 20, 2310 – 2316.[41] H. Mukamal, F. W. Harris, J. K. Stille, J. Polym. Sci. A 1967, 5, 2721 – 2729.[42] A. J. Berresheim, M. Meller, K. Mellen, Chem. Rev. 1999, 99, 1747 – 1785.[43] S. Granados-focil, M. Litt, Prepr. Pap. Am. Chem. Soc. Div. Fuel Chem.

2004, 49, 528 – 529.[44] M. Litt, G.-F. Sergio, K. Junwon, S. Kun, R. Wycisk, RCS Trans. 2010, 33,

695 – 710.[45] K. Heinz, Chem. Ber. 1960, 93, 1769 – 1773.[46] W. Ried, D. Freitag, Naturwissenschaften 1966, 53, 306.[47] W. Ried, D. Freitag, Angew. Chem. Int. Ed. 1968, 7, 835 – 844; Angew.

Chem. 1968, 80, 932 – 942.[48] H. F. VanKerckhoven, Y. K. Gilliams, J. K. Stille, Macromolecules 1972, 5,

541 – 546.[49] B. Cherry, C. Fujimoto, C. Cornelius, T. Alam, Macromolecules 2005, 38,

1201 – 1206.[50] M. Hickner, C. Fujimoto, C. Cornelius, Polymer 2006, 47, 4238 – 4244.[51] R. J. Stanis, M. A. Yaklin, C. J. Cornelius, T. Takatera, A. Umemoto, A. Am-

brosini, C. Fujimoto, J. Power Sources 2010, 195, 104 – 110.[52] Z. B. Shifrina, M. S. Averina, A. L. Rusanov, M. Wagner, K. Mellen, Macro-

molecules 2000, 33, 3525 – 3529.[53] H. Pu, Polymers for PEM Fuel Cells, Wiley, Hoboken, NJ, 2014.[54] Z. Zhang, L. Wu, T. Xu, J. Mater. Chem. 2012, 22, 13996.[55] M. Ogliaruso, M. G. Romanelli, E. I. Becker, Chem. Rev. 1965, 65, 261 –

367.[56] D. Gay, J. Gambelin, 2008, 0 – 1.[57] Z. Xie, C. Song, B. Andreaus, T. Navessin, Z. Shi, J. Zhang, S. Holdcroft, J.

Electrochem. Soc. 2006, 153, E173.[58] S. R. Kline, J. Appl. Crystallogr. 2006, 39, 895 – 900.[59] R. J. Roe, Methods of X-Ray and Neutron Scattering in Polymer Science

Oxford University Press, Oxford, UK. , 2000.[60] J.-C. Perrin, S. Lyonnard, F. Volino, J. Phys. Chem. C 2007, 111, 3393 –

3404.[61] A. Strong, B. Britton, D. Edwards, T. Peckham, H. Lee, W. Huang, S. Hold-

croft, J. Electrochem. Soc. 2015, 162, F513 – F518.[62] J. Miyake, R. Taki, T. Mochizuki, R. Shimizu, R. Akiyama, M. Uchida, K.

Miyatake, Sci. Adv. 2017, 3, eaao0476.[63] H. F. Lee, Y. C. Huang, P. H. Wang, C. C. Lee, Y. S. Hung, R. Gopal, S. Hold-

croft, W. Y. Huang, Mater. Today Commun. 2015, 3, 114 – 121.[64] Calculations were based on a chemical formula of C60S4O12H41, a density

of 1.17 g cm@3 and using the NIST SLD Calculator; https://www.ncnr.nist.gov/resources/activation/.

[65] Calculations were based on a chemical formula of C60H41, a density of1.06 g cm@3, and using the NIST SLD Calculator, https://www.ncnr.nist.-gov/resources/activation/.

Manuscript received: August 28, 2018Revised manuscript received: September 21, 2018

Accepted manuscript online: September 24, 2018Version of record online: November 7, 2018

ChemSusChem 2018, 11, 4033 – 4043 www.chemsuschem.org T 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim4043

Full Papers