georgia institute of technology - the importance of water...

11
Journal of The Electrochemical Society, 2020 167 054501 The Importance of Water Transport in High Conductivity and High-Power Alkaline Fuel Cells Mrinmay Mandal, 1 Garrett Huang, 1, Noor Ul Hassan, 2 Xiong Peng, 2, Taoli Gu, 3 Ahmon H. Brooks-Starks, 4 Bamdad Bahar, 3 William E. Mustain, 2, ∗∗ and Paul A. Kohl 1, ∗∗∗, z 1 School of Chemical and Biomolecular Engineering, Georgia Institute of Technology, Atlanta, Georgia, USA 2 Department of Chemical Engineering, University of South Carolina, Columbia, South Carolina, USA 3 Xergy Inc., Harrington, Delaware, USA 4 Chemical & Materials Engineering Department, California State Polytechnic University, Pomona, Pomona, California, USA High ionic conductivity membranes can be used to minimize ohmic losses in electrochemical devices such as fuel cells, flow batteries, and electrolyzers. Very high hydroxide conductivity was achieved through the synthesis of a norbornene-based tetrablock copolymer with an ion-exchange capacity of 3.88 meq/g. The membranes were cast with a thin polymer reinforcement layer and lightly cross- linked with N,N,N ,N -tetramethyl-1,6-hexanediamine. The norbornene polymer had a hydroxide conductivity of 212 mS/cm at 80°C. Light cross-linking helped to control the water uptake and provide mechanical stability while balancing the bound (i.e. waters of hydration) vs. free water in the films. The films showed excellent chemical stability with <1.5% conductivity loss after soaking in 1 M NaOH for 1000 h at 80°C. The aged films were analyzed by FT-IR before and after aging to confirm their chemical stability. AH 2 /O 2 alkaline polymer electrolyte fuel cell was fabricated and was able to achieve a peak power density of 3.5 W/cm 2 with a maximum current density of 9.7 A/cm 2 at 0.15 V at 80°C. The exceptionally high current and power densities were achieved by balancing and optimizing water removal and transport from the hydrogen negative electrode to the oxygen positive electrode. High water transport and thinness are critical aspects of the membrane in extending the power and current density of the cells to new record values. © The Author(s) 2019. Published by ECS. This is an open access article distributed under the terms of the Creative Commons Attribution 4.0 License (CC BY, http://creativecommons.org/licenses/by/4.0/ ), which permits unrestricted reuse of the work in any medium, provided the original work is properly cited. [DOI: 10.1149/2.0022005JES] Manuscript submitted August 16, 2019; revised manuscript received September 20, 2019. Published October 9, 2019. This paper is part of the JES Focus Issue on Heterogeneous Functional Materials for Energy Conversion and Storage. Anion-exchange membranes (AEMs) are a key component in al- kaline exchange membrane fuel cells (AEMFCs), flow batteries and electrolyzers. 1 Alkaline conditions are attractive because of the facile electrochemical reaction kinetics at high pH for oxygen reduction and water oxidation. 27 Device operation at high pH allows for the use of non-precious metal catalysts, simpler design for the balance of plant, and reduced fuel crossover. 912 However, it is imperative that the AEMs are thin, have long-term alkaline stability, and high hy- droxide ion conductivity. 8 There have been issues in the past with AEMs showing low ionic conductivity, poor stability at high pH, and high water uptake (leading to dimensional change); however, these issues are being systematically addressed. 1315 The formation of multi-block copolymers (BCP) are a means to achieve phase segregation within polymers in order to create high- mobility ion conduction channels within the hydrophilic phase of the polymer membrane. 1620 Previously, we have reported AEMs con- sisting of poly(norbornene) BCPs with record high hydroxide con- ductivity, 198 mS/cm, and very high peak power density in a hy- drogen/oxygen fuel cell, 3.4 W/cm 2 at 80°C. 2124 In addition, the poly(norbornene) polymer, as well as the membranes made from the polymer, were shown to have excellent thermal and mechanical prop- erties. The S N 2 substitution and Hoffmann elimination degradation routes were suppressed by tethering the quaternary ammonium head- group to the all-hydrocarbon poly(norbornene) backbone via a long alkyl hydrocarbon chain. 2528 Trimethyl quaternary ammonium cation head-groups have been shown to be stable cations in AEMs, and their low molecular weight enables high ion exchange capacity (IEC). 29 Water management is a key factor in achieving high AEMFC per- formance. The AEM plays a key role in balancing the water con- tent and distribution during device operation. 30,31 It has been shown that a significant majority of the reacting water at the AEMFC cath- ode is provided by back-diffusion of water produced at the AEMFC anode. 32 This suggests that high AEM water permeability is bene- Electrochemical Society Student Member. ∗∗ Electrochemical Society Member. ∗∗∗ Electrochemical Society Fellow. z E-mail: [email protected] ficial in AEMFCs. However, excessive AEM water uptake can flood the ion conducting channels within the polymer and lead to membrane softening and mechanical failure. 33,34 Thus, high water permeability without high water solubility appears to be a critical feature for AEM- FCs. Waters of hydration are necessary for hydroxide ion conduction; however, excessive unbound (i.e. free) water leads to low hydroxide mobility and membrane distortion. 23 Hence, it is necessary to balance the amount of free and bound water inside the membranes to yield the maximum hydroxide mobility and water transport. 2123 Cross-linking is an effective way to reduce water uptake and swelling. 24 However, AEMs with high cross-linking density can become too rigid, lead- ing to poor ion mobility, mechanical properties and water diffusivity (i.e. high water solubility without high diffusivity). 3537 In the case of polymers with high IEC, light cross-linking is an effective strat- egy to balance the high conductivity and water uptake (WU) without sacrificing IEC. 22,23 In addition, thinner membranes can enable rapid water transport without high water uptake, and enable high current density AEMFCs. Conveniently, light cross-linking also helps in the production of thin membranes with good mechanical properties. AEM carbonation upon exposure to CO 2 is another important fac- tor in AEMFC performance. When CO 2 -containing air is fed to an operating AEMFC, the hydroxide anions produced by the reduction of oxygen at the positive electrode react with carbon dioxide to produce carbonated anions with lower mobility than hydroxide (i.e. carbonate or bicarbonate), increasing ohmic-related losses. 38 Additionally, these carbonated anions can rapidly populate the AEM and AEMFC anode, leading to significant thermodynamic and kinetic-related losses. 32,39 It has been stated that the adverse effects of membrane carbonation could be minimized by using AEMs with very high ionic conductivity so that the decrease in mobility upon carbonation can be mitigated and cell decarbonation during operation through the so-called “self- purging” mechanism can occur more rapidly. Hence, AEMs with very high conductivity are most desirable. 23 In this study, the synthesis of chemically stable AEMs with record high conductivity, 212 mS/cm at 80°C, and their implementation into AEMFCs are described. This new AEM enables record performance in a hydrogen/oxygen AEMFC with a peak power density of 3.5 W/cm 2 and maximum current density of 9.7 A/cm 2 at 0.15 V at 80°C when ) unless CC License in place (see abstract). ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 130.207.74.224 Downloaded on 2019-10-10 to IP

Upload: others

Post on 10-Jul-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Georgia Institute of Technology - The Importance of Water …kohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal... · 2019-10-10 · Journal of The Electrochemical Society, 2020

Journal of The Electrochemical Society, 2020 167 054501

The Importance of Water Transport in High Conductivity andHigh-Power Alkaline Fuel CellsMrinmay Mandal, 1 Garrett Huang, 1,∗ Noor Ul Hassan,2 Xiong Peng,2,∗ Taoli Gu,3Ahmon H. Brooks-Starks,4 Bamdad Bahar,3 William E. Mustain, 2,∗∗and Paul A. Kohl 1,∗∗∗,z

1School of Chemical and Biomolecular Engineering, Georgia Institute of Technology, Atlanta, Georgia, USA2Department of Chemical Engineering, University of South Carolina, Columbia, South Carolina, USA3Xergy Inc., Harrington, Delaware, USA4Chemical & Materials Engineering Department, California State Polytechnic University, Pomona, Pomona,California, USA

High ionic conductivity membranes can be used to minimize ohmic losses in electrochemical devices such as fuel cells, flow batteries,and electrolyzers. Very high hydroxide conductivity was achieved through the synthesis of a norbornene-based tetrablock copolymerwith an ion-exchange capacity of 3.88 meq/g. The membranes were cast with a thin polymer reinforcement layer and lightly cross-linked with N,N,N′,N′-tetramethyl-1,6-hexanediamine. The norbornene polymer had a hydroxide conductivity of 212 mS/cm at 80°C.Light cross-linking helped to control the water uptake and provide mechanical stability while balancing the bound (i.e. waters ofhydration) vs. free water in the films. The films showed excellent chemical stability with <1.5% conductivity loss after soaking in1 M NaOH for 1000 h at 80°C. The aged films were analyzed by FT-IR before and after aging to confirm their chemical stability.A H2/O2 alkaline polymer electrolyte fuel cell was fabricated and was able to achieve a peak power density of 3.5 W/cm2 with amaximum current density of 9.7 A/cm2 at 0.15 V at 80°C. The exceptionally high current and power densities were achieved bybalancing and optimizing water removal and transport from the hydrogen negative electrode to the oxygen positive electrode. Highwater transport and thinness are critical aspects of the membrane in extending the power and current density of the cells to new recordvalues.© The Author(s) 2019. Published by ECS. This is an open access article distributed under the terms of the Creative CommonsAttribution 4.0 License (CC BY, http://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse of the work in anymedium, provided the original work is properly cited. [DOI: 10.1149/2.0022005JES]

Manuscript submitted August 16, 2019; revised manuscript received September 20, 2019. Published October 9, 2019. This paper ispart of the JES Focus Issue on Heterogeneous Functional Materials for Energy Conversion and Storage.

Anion-exchange membranes (AEMs) are a key component in al-kaline exchange membrane fuel cells (AEMFCs), flow batteries andelectrolyzers.1 Alkaline conditions are attractive because of the facileelectrochemical reaction kinetics at high pH for oxygen reduction andwater oxidation. 2–7 Device operation at high pH allows for the useof non-precious metal catalysts, simpler design for the balance ofplant, and reduced fuel crossover.9–12 However, it is imperative thatthe AEMs are thin, have long-term alkaline stability, and high hy-droxide ion conductivity.8 There have been issues in the past withAEMs showing low ionic conductivity, poor stability at high pH, andhigh water uptake (leading to dimensional change); however, theseissues are being systematically addressed.13–15

The formation of multi-block copolymers (BCP) are a means toachieve phase segregation within polymers in order to create high-mobility ion conduction channels within the hydrophilic phase of thepolymer membrane.16–20 Previously, we have reported AEMs con-sisting of poly(norbornene) BCPs with record high hydroxide con-ductivity, 198 mS/cm, and very high peak power density in a hy-drogen/oxygen fuel cell, 3.4 W/cm2 at 80°C.21–24 In addition, thepoly(norbornene) polymer, as well as the membranes made from thepolymer, were shown to have excellent thermal and mechanical prop-erties. The SN2 substitution and Hoffmann elimination degradationroutes were suppressed by tethering the quaternary ammonium head-group to the all-hydrocarbon poly(norbornene) backbone via a longalkyl hydrocarbon chain.25–28 Trimethyl quaternary ammonium cationhead-groups have been shown to be stable cations in AEMs, and theirlow molecular weight enables high ion exchange capacity (IEC).29

Water management is a key factor in achieving high AEMFC per-formance. The AEM plays a key role in balancing the water con-tent and distribution during device operation.30,31 It has been shownthat a significant majority of the reacting water at the AEMFC cath-ode is provided by back-diffusion of water produced at the AEMFCanode.32 This suggests that high AEM water permeability is bene-

∗Electrochemical Society Student Member.∗∗Electrochemical Society Member.

∗∗∗Electrochemical Society Fellow.zE-mail: [email protected]

ficial in AEMFCs. However, excessive AEM water uptake can floodthe ion conducting channels within the polymer and lead to membranesoftening and mechanical failure.33,34 Thus, high water permeabilitywithout high water solubility appears to be a critical feature for AEM-FCs. Waters of hydration are necessary for hydroxide ion conduction;however, excessive unbound (i.e. free) water leads to low hydroxidemobility and membrane distortion.23 Hence, it is necessary to balancethe amount of free and bound water inside the membranes to yield themaximum hydroxide mobility and water transport.21–23 Cross-linkingis an effective way to reduce water uptake and swelling.24 However,AEMs with high cross-linking density can become too rigid, lead-ing to poor ion mobility, mechanical properties and water diffusivity(i.e. high water solubility without high diffusivity).35–37 In the caseof polymers with high IEC, light cross-linking is an effective strat-egy to balance the high conductivity and water uptake (WU) withoutsacrificing IEC.22,23 In addition, thinner membranes can enable rapidwater transport without high water uptake, and enable high currentdensity AEMFCs. Conveniently, light cross-linking also helps in theproduction of thin membranes with good mechanical properties.

AEM carbonation upon exposure to CO2 is another important fac-tor in AEMFC performance. When CO2-containing air is fed to anoperating AEMFC, the hydroxide anions produced by the reductionof oxygen at the positive electrode react with carbon dioxide to producecarbonated anions with lower mobility than hydroxide (i.e. carbonateor bicarbonate), increasing ohmic-related losses.38 Additionally, thesecarbonated anions can rapidly populate the AEM and AEMFC anode,leading to significant thermodynamic and kinetic-related losses.32,39

It has been stated that the adverse effects of membrane carbonationcould be minimized by using AEMs with very high ionic conductivityso that the decrease in mobility upon carbonation can be mitigatedand cell decarbonation during operation through the so-called “self-purging” mechanism can occur more rapidly. Hence, AEMs with veryhigh conductivity are most desirable.23

In this study, the synthesis of chemically stable AEMs with recordhigh conductivity, 212 mS/cm at 80°C, and their implementation intoAEMFCs are described. This new AEM enables record performance ina hydrogen/oxygen AEMFC with a peak power density of 3.5 W/cm2

and maximum current density of 9.7 A/cm2 at 0.15 V at 80°C when

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 130.207.74.224Downloaded on 2019-10-10 to IP

Page 2: Georgia Institute of Technology - The Importance of Water …kohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal... · 2019-10-10 · Journal of The Electrochemical Society, 2020

Journal of The Electrochemical Society, 2020 167 054501

water transport from the anode to cathode is controlled. Light cross-linking was used to control the WU at high IEC in this polymer,3.88 meq/g. It is shown that thin, reinforced, high-conductivity mem-branes with excellent water transport are critical components in pro-ducing high-power AEMFCs. A comparison of AEMFC membranethickness supports the hypothesis that water transport from anode tocathode is a critical factor in the achievable current and power density.

Experimental

Materials.—The monomers, butyl norbornene (BuNB), bro-mobutyl norbornene (BBNB) and bromopropyl norbornene (BPNB),were supplied by Promerus, LLC (Brecksville, OH). Prior topolymerization, the monomers were purified by distillation oversodium and degassed in three freeze-pump-thaw cycles. The pro-cedures were carried out under a dry argon atmosphere in aglove box with the rigorous exclusion of moisture and air.The catalyst [(allyl)palladium(triisopropylphosphine)chloride, (η3-allyl)Pd(iPr3P)Cl)] was prepared following a previously publishedprocedure.40 Lithium tetrakis(pentafluorophenyl)-borate·(2.5Et2O)(Li[FABA]) was purchased from Boulder Scientific Co. and used as-received. N,N,N′,N′-Tetramethyl-1,6-hexanediamine (TMHDA), an-hydrous toluene (99.9%), anhydrous α,α,α-trifluorotoluene (TFT,≥99%) and tetrahydrofuran (THF) were purchased from Sigma-Aldrich and used as-received.

Tetrablock copolymer synthesis and characterization.—Thetetrablock copolymers, GT82, GT78, and GT64 (polymers contain-ing 82 mol%, 78mol%, and 64 mol% of the halogenated monomer,respectively) were synthesized following a previously reported pro-cedure using bromobutyl norbornene (BBNB) and bromopropyl nor-bornene (BPNB).21,23

The polymer samples were analyzed by 1H NMR using a BrukerAvance 400 MHz NMR instrument using CDCl3 as the solvent. Thenumber average molecular weight (Mn) and dispersity (Ð) of GT82,GT78, and GT64 were determined by gel permeation chromatography(GPC) (Shimadzu) equipped with an LC-20 AD HPLC pump and arefractive index detector (RID-20 A, 120 V). GPC measurements wereperformed in THF with the eluent flow rate of 1.0 mL/min at 30°Cand calibrated against a polystyrene standard.

Film and membrane casting.—In this paper, the non-reinforcedmaterials formed into sheets for testing will be referred to as ‘films’and the reinforced polymer materials will be referred to as ‘mem-branes’. The precursor tetrablock copolymers were dissolved in chlo-roform (2 to 3 wt%) and cast into freestanding films for conductivitytests. The cross-linking agent, TMHDA, was added to the solutionin different molar ratios (5 to 20 mol%) with respect to the molesof halogenated monomer in the polymer (determined by 1H NMR).After stirring for 10 min at room temperature, the solution was fil-tered through a 0.45 μm poly(tetrafluoroethylene) (PTFE) membranesyringe filter and poured into a 4 cm diameter aluminum dish. Thealuminum dish was placed in a preheated oven at 60°C and dried for24 h. The colorless, transparent, flexible and freestanding films were10 μm to 50 μm thick. The films were quaternized with 50 wt% aque-ous trimethylamine solution for 48 h at room temperature to convertthe bromobutyl moieties to quaternary ammonium head-groups. Thequaternized films were ion-exchanged from bromide ions to hydrox-ide ions after immersing in 1 M NaOH solution under nitrogen for24 h.

Polymer films were made for testing properties of the pure poly-mer. Reinforced structures (i.e. “membranes”) were made for fuel celltesting. Reinforced membranes were formed by dissolving the syn-thesized precursor polymer in toluene, forming a 10 wt% solution.Toluene was used rather than chloroform to better control the evapo-ration rate of the composite membranes, however the casting solventdid not affect the quality of the film or membrane. The cross-linkingagent, TMHDA, was added to the solution in different molar ratios(5 to 20 mol%) with respect to the moles of halogenated monomer

in the polymer (determined by 1H NMR). After stirring for 30 min atroom temperature until homogenous and translucent, the solution wasfiltered through a 0.45 μm PTFE membrane syringe filter. The precur-sor polymer solution was applied to a thin, microporous PTFE supportmaterial for reinforcement. The membranes, nominally 10 μm thick,were formed by drying at room temperature. The membranes werequaternized with 50 wt% aqueous trimethylamine solution for 48 hat room temperature to convert the bromobutyl/bromopropyl moietiesto quaternary ammonium head-groups. The quaternized membraneswere ion-exchanged from bromide ions to hydroxide ions by immer-sion in 1 M NaOH solution under nitrogen for 24 h.

Hydroxide conductivity.—The ionic conductivity of the films wasmeasured using four-point probe electrochemical impedance spec-troscopy (1 Hz to 1 MHz) with a PAR 2273 potentiostat. The filmswere cut into 1 × 4 cm strips and tested in HPLC-grade water undera nitrogen purge to minimize carbonation. The films were allowed toequilibrate for 30 min prior to each measurement. The in-plane ionicconductivity was calculated using Equation 1.

σ = L

WtR[1]

In Eq. 1, σ is the ionic conductivity, L is the length between sensingelectrodes, W and t are the width and thickness of the films, respec-tively, and R is the measured resistance. The CO3

2− conductivity wasmeasured after storing the films (initially in hydroxide form) in sat-urated aqueous K2CO3 for 24 h. The long-term (>1000 h) alkalinestability was determined by periodically measuring the ionic conduc-tivity after soaking the films in hydroxide-form in aqueous 1 M NaOHat 80°C in a Teflon-lined Parr reactor. Prior to each measurement, thefilms were taken out of solution and thoroughly washed with DI wa-ter. After each measurement, the films were stored in the reactors withfresh NaOH solution. After equilibration, each data point was mea-sured in triplicate and the average value is reported here. The deviationin the measurement of each data point was <0.5%. Long-term chem-ical stability was probed by FT-IR analysis at the same state as theconductivity was measured using a Nicolet 6700 FT-IR spectrometer.

IEC, WU, hydration number (λ), number of freezable (Nfree) andnon-freezable water (Nbound) molecules.—The ion-exchange capac-ity (IEC) of the precursor polymers were determined by 1H NMR.Mohr’s titration method was used to measure the IEC of the filmsafter cross-linking and quaternization to confirm the accuracy of theNMR results.41 In a typical procedure, the material in Br− form wasconverted to Cl− form by soaking in 0.1 M NaCl solution for 24 h.The film was removed from the NaCl solution, thoroughly washedwith DI water and dried in vacuum for >24 h. The dry film weightwas recorded. Next, the films were immersed in a fixed volume of0.05 M NaNO3 for 24 h. Finally, the released Cl− was titrated with0.05M AgNO3 using K2CrO4 (10 wt%) as the indicator. Measure-ments were performed in triplicate to ensure repeatability. The IECwas calculated using Equation 2.

IEC = CAgNO3 × VAgNO3

Md[2]

In Eq. 2, VAgNO3 (mL) is the volume of AgNO3 solution, CAgNO3

(0.05 mol·L−1) is the concentration of AgNO3 solution, and Md (g) isthe weight of the dried sample. The water uptake was calculated usingEquation 3.

WU (%) = Mw − Md

Md× 100 [3]

In Eq. 3, Md is the dry mass of the film and Mw is the wet mass ofthe film after dabbing off excess surface water with a dry towel. Themass of the films was measured at room temperature in OH− form.The hydration number (λ) or the number of water molecules per ionicgroup, was calculated using Equation 4.

λ = 1000 × WU%

IEC × 18[4]

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 130.207.74.224Downloaded on 2019-10-10 to IP

Page 3: Georgia Institute of Technology - The Importance of Water …kohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal... · 2019-10-10 · Journal of The Electrochemical Society, 2020

Journal of The Electrochemical Society, 2020 167 054501

The number of freezable water (Nfree) and bound water (or non-freezable water) (Nbound) was determined by differential scanningcalorimetry (DSC). DSC measurements were carried out on a Discov-ery DSC with an autosampler (TA Instruments). The samples werefully hydrated by soaking in deionized water for one week. After ex-cess water on the surface was removed, a 5 to 10 mg sample wasquickly cut and sealed in an aluminum DSC pan. The sample was firstcooled to −70°C at 5°C/min and then heated to 30°C at 5°C/min underN2 (20 mL/min). The quantity of freezable and non-freezable waterwas determined by Equations 5 to 7.42–44

Nfree = Mfree

Mtot× λ [5]

Mfree is the mass of freezable water and Mtot is the total mass ofwater absorbed in the film. The weight fraction of freezable water wascalculated using Equation 6.

Mfree

Mtot= Hf/Hice

(MW − Md ) /Mw[6]

Mw is the wet mass of the film obtained after gently wiping excesswater from the surface. Md is the dry mass of the film. Hf is theenthalpy obtained by the integration of the DSC freezing peak andHice is the enthalpy of water fusion, corrected for the subzero freezingpoint according to Equation 7.

Hice = Hoice − �Cp�Tf [7]

�Cp is the difference between the specific heat capacity of liquidwater and ice. �Tf is the freezing point depression.

Thermal and mechanical stability.—The thermal stability of thedried film in bromide ion form was studied using thermogravimetricanalysis (TGA) on a TA Instruments Q50 analyzer. The temperaturewas ramped at 10°C/min up to 800°C in a nitrogen atmosphere. Thestorage modulus of the reinforced composite membranes was mea-sured by dynamic mechanical analysis (DMA) using a TA InstrumentsQ800 under a 1 Hz single-frequency strain mode in air at 25°C. A fullyhydrated, rectangular sample was loaded into the DMA with tensionclamps after removing surface water. Experiential parameters for theDMA were set to 0.1% strain and a preload force of 0.01 N with aforce track of 125%.

Fuel cell testing.—Catalyst inks were made from an ETFE-g-poly(VBTMAC) powder ionomer45 and Pt-based electrocatalysts.40% Pt/C (Alfa Aesar HiSPEC 4000, Pt nominally 40%wt, supportedon Vulcan XC-72R carbon) was used at the cathode and 60% Pt-Ru/C(Alfa Aesar HiSPEC 10000, Pt nominally 40 wt%., and Ru, nominally20 wt%., supported on Vulcan XC-72R carbon) was used as the an-ode. The details for ink formulation and fabrication of gas diffusionelectrodes (GDEs) have been previously reported.30,31 AEMFCs with5 cm2 active area were assembled by placing the anode and cathodeGDEs on opposite sides of the AEM in Scribner fuel cell hardwarewith a single channel serpentine flow field. Fuel cell experiments werecontrolled using a Scribner 850e fuel cell test station.

At cell startup, N2 at 100% relative humidity was flowed throughthe anode and cathode until a cell temperature of 60°C was reached.Then, the N2 feeds were switched to UHP H2 and O2 and a constantvoltage of 0.5V was applied to allow the cell to break-in. After a stablecurrent density was established, the dew points of the anode and cath-ode reacting gases were optimized. The cell temperature was gradu-ally increased to 80°C in 5°C increments, with the anode/cathode dewpoints being simultaneously optimized with the cell temperature toavoid membrane dry out. After the cell temperature reached 80°C, thedegree of back-pressurization was also set. After the cell equilibratedat the desired conditions, a polarization curve was collected by slowlysweeping the voltage from open circuit to 0.1 V at 10 mV/s scan rate.Cell performance was also investigated with CO2-free air fed to thecathode. In this case, after the H2/O2 cell testing, the cathode gas wasswitched to air and the system was allowed to equilibrate at the samereacting gas dew points and back pressurization for 10 min. Then, the

Figure 1. GPC traces of GT82 and GT64 tetrablock copolymers.

reacting gas dew points and back pressurization were again optimizedusing the same procedure as above before the polarization curve wascollected.

Constant current density stability tests were performed for ca.100 hours at 600 mA/cm2. These cells were broken-in using the sameprocedure described above. UHP H2 was fed to the anode and CO2-free air was fed to the cathode in lieu of UHP O2. The dew points ofthe reacting gases were adjusted periodically during durability testingto ensure adequate membrane hydration. Also, the cell high frequencyresistance (HFR) was monitored (at a frequency of 7 kHz) by the fuelcell test station throughout the test.

Results and Discussion

Synthesis and characterization of poly(norbornene) tetrablockcopolymer.—The tetrablock copolymers were synthesized using ahydrophilic monomer, either BPNB or BBNB, and a hydrophobicmonomer, BuNB. The polymers were synthesized following a previ-ous reported procedure.21,23 The number average molecular weights(Mn) and dispersity (Ɖ) of the tetrablock copolymers were determinedby GPC. The Mn and Ɖ of the tetrablock copolymers were 57.7 kDaand 1.41, respectively, for GT82, 50.96 kDa and 2.02, respectively,for GT64, and 103.6 kDa and 1.3, respectively, for GT78. Previously,the successful formation of tetrablock copolymers by the sequentialgrowth of each block after the addition of each individual monomerover time was demonstrated.21,23 GPC traces of the two polymers areshown in Figure 1. The precursor block copolymers with a range ofIECs (3.37–3.88 meq/g) were crossed-linked with TMHDA at dif-ferent molar concentration (5 to 20 mol%) with respect to the molesof brominated sites in the tetrablock polymer, Scheme 1. A previ-ous study showed that the conductivity was lower if the cross-linkerconcentration was <5 mol%.23 Also, membranes and films withoutcross-linking are unusable for fuel cell testing due to their gel-likecharacter and excess WU after soaking in TMA. A summary of filmproperties as the IEC and degree of crosslinking were changed is pro-vided in Table I. The mole percent of the cross-linker with respectto the available cross-linking sites is indicated by the sample name.For example, GT82-5 has 82% of the monomers in the polymer func-tionalized with the quaternary ammonium cation and 5 mol% addedcross-linker with respect to the original moles of bromobutyl sites.

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 130.207.74.224Downloaded on 2019-10-10 to IP

Page 4: Georgia Institute of Technology - The Importance of Water …kohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal... · 2019-10-10 · Journal of The Electrochemical Society, 2020

Journal of The Electrochemical Society, 2020 167 054501

Scheme 1. Synthesis of cross-linked AEMs using BBNB and BPNB as the halogenated block.

The IEC of the precursor polymers was evaluated by 1H NMR, andan example for GT82 is shown in Figure 2.21,23 The IEC values ob-tained by 1H NMR were spot-checked by titration, as described in theExperimental Section. From Figure 2, the combined mol% and wt%of hydrophobic and halogenated block were calculated. The mass of

TMHDA was included in the calculation of the final IEC values, whichranged from 3.25 to 3.84 meq/g. In addition, the IEC values were alsomeasured by titration. The IEC values measured by NMR and titrationmatched well as shown in Table I. In order to confirm that the bromi-nated head-groups were fully quaternized, the films were analyzed by

Table I. Properties of cross-linked poly(butyl norbornene-b-bromobutyl norbornene-b-butyl norbornene-b-bromobutyl norbornene) andpoly(butyl norbornene-b-bromopropyl norbornene-b-butyl norbornene-b-bromopropyl norbornene) films in hydroxide form.

OH−Conductivity

Block Cross-linker (mS/cm)b σ/IECe Ionic ASRf Water Hydration

copolymer Concentration(mol%) 25°C 80°C IEC (meq/g)c IEC (meq/g)d (S g/cm eq) (Ohm-cm2) Uptakeg (%) number λ Nfree Nbound

GT82-5a 5 109 212 3.84 3.76 55.2 0.04 122 17.6 7.59 10.1GT82-10a 10 90 193 3.80 3.71 50.8 0.05 87 12.7 1.04 11.7GT82-15a 15 67 147 3.76 3.70 39.1 0.06 67 9.90 1.41 8.49GT82-20a 20 55 128 3.72 3.69 34.4 0.08 61 9.11 0.00 9.11GT78-5a 5 96 162 3.70 3.61 43.8 0.06 163 24.5 12.7 11.0

GT78-15a 15 65 138 3.62 3.52 38.1 0.08 65 9.98 1.71 8.26GT64-5 5 84 181 3.34 3.30 54.2 0.05 90 15.0 3.18 11.8

GT64-10 10 74 162 3.31 3.24 48.9 0.05 61 10.2 0.13 10.1GT64-15 15 62 142 3.28 3.26 43.3 0.07 52 8.81 0.05 8.76GT64-20 20 52 119 3.25 3.16 36.7 0.07 43 7.35 0.09 7.26

aBlock copolymers using bromobutyl norbornene (BBNB) as the halogenated block.bOH− conductivity was measured by four-probe conductivity cell.cIEC was determined by 1H NMR.dIEC was determined by titration.eIonic conductivity at 80°C/IEC.fIonic ASR was calculated using the following equation: ASR = L/σ where L = film thickness in cm; σ = ion conductivity in S/cm (at 80°C).gWater uptake was measured at room temperature. GT82-5 can be explained as the polymer with 82 mol% of the halogenated monomer and the polymeris crossed-linked with 5 mol % of TMHDA.

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 130.207.74.224Downloaded on 2019-10-10 to IP

Page 5: Georgia Institute of Technology - The Importance of Water …kohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal... · 2019-10-10 · Journal of The Electrochemical Society, 2020

Journal of The Electrochemical Society, 2020 167 054501

Figure 2. 1H NMR spectrum of tetrablock GT82 in CDCl3.

FT-IR in precursor polymer form after cross-linking, and after cross-linking and quaternization. The disappearance of the alkyl bromidepeak at 646 cm−1 and the appearance of quaternary ammonium peaksat 908 cm−1, 968 cm−1, and 1481 cm−1 show complete quaternization,within the sensitivity of the FTIR, Figure 3. In addition, the successfulincorporation of TMHDA after cross-linking was also indicated by the

Figure 3. FT-IR spectra of the GT82 film in precursor polymer form, aftercross-linking, and after cross-linking and quaternization.

appearance of a new quaternary ammonium peak at 1481 cm−1 in theIR spectra, Figure 3.

Hydroxide conductivity (σ) and ionic area specific resistance(ASR).—As discussed in the Introduction, AEMs with high hydroxideconductivity are desired for electrochemical devices. The hydroxideconductivity was measured at four temperatures between 25°C and80°C. As shown in Figure 4 (left), the conductivity increased withtemperature. The slope of ln(σ) vs. 1/T was used to calculate an ef-fective activation energy (Ea) using the Arrhenius Equation, Figure 4(right). The calculated activation energy, 10.8 to 13.6 kJ/mol, is com-parable to previously reported AEMs and Nafion-117 materials.46–48

It was previously reported that a maximum conductivity was obtainedat a particular TMHDA concentration.23 At 80°C, the GT82 conduc-tivity decreased from 212 to 128 mS/cm with an increase in TMHDAconcentration from 5 mol% to 20 mol% due to an insufficient amountof bound and free water for hydration and ion conduction inside thefilm. For GT64, the conductivity was 181 and 119 mS/cm at 80°Cat 5 mol% and 20 mol% TMHDA, respectively. The conductivity ofGT78 at 5 mol% and 15 mol% TMHDA was 162 and 138 mS/cm,respectively, at 80°C. The lower conductivity at 5 mol% TMHDA forGT78 in comparison to GT64 (162 vs. 181 mS/cm at 80°C) shows theimportance of proper water management inside the films. GT78-5 had12.7 free waters per ion pair (a relatively high value), whereas GT64-5 had 3.2 free waters per ion pair (more optimized). The excess freewater in GT78-5 led to lower mobility (43.8 S g/cm eq) compared toGT64-5 (54.2 S g/cm eq).

The ionic area specific resistance (ASR) is an important membranemetric for electrochemical devices. The ASR can be calculated bymultiplying the film resistance by the measurement area. Ex-situ ASRvalues for each of the films are shown in Table I. It was observed thatthe ASR values were able to meet the US Department of Energy AEMtarget of ≤0.04 ohm cm2.

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 130.207.74.224Downloaded on 2019-10-10 to IP

Page 6: Georgia Institute of Technology - The Importance of Water …kohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal... · 2019-10-10 · Journal of The Electrochemical Society, 2020

Journal of The Electrochemical Society, 2020 167 054501

Figure 4. Plot of ionic conductivity of cross-linked AEMs at different temperature (left), Arrhenius plot of ln(σ) vs. inverse temperature (right).

The effect of carbonation on anion conductivity, Equations 8 and9, was also investigated.

OH− + CO2 ↼⇁ HCO3− [8]

OH− + HCO3− ↼⇁ CO3

2− + H2O [9]

GT64-5 and GT64-10 were selected for study of the carbonate con-ductivity because they have a good balance of conductivity and wateruptake. The comparison of OH− and CO3

2− conductivity for GT64-5and GT64-10 with temperature is shown in Figure 5 (left). The OH−

conductivity for GT64-5 was 84 mS/cm and 181 mS/cm at 25°C and80°C, respectively. In the CO3

2− form, the conductivity of GT64-5 was18 mS/cm and 54 mS/cm at 25°C and 80°C, respectively. For GT64-10, the conductivity before and after carbonation was 74 mS/cm vs.15 mS/cm at 25°C, and 162 mS/cm vs. 47 mS/cm at 80°C. The ratioof the OH− to CO3

2− mobility was 4.7 (GT64-5) and 4.9 (GT64-10)at 25°C. This is in agreement with a previous reported mobility ratioof 5.39 That is, carbonate has only about 20% to 22% of the mobilityof hydroxide at 25°C. At 80°C, the mobility ratio decreased to 3.4 and

3.7 for GT64-5 and GT64-10, respectively. Hence, the adverse effectof carbonation on anion mobility becomes comparatively less signifi-cant at fuel cell-relevant temperatures. It is noted AEMFC carbonationhas other negative effects on fuel cell performance, such as loweringthe pH at the negative electrode, inducing kinetic losses at the anode,and buildup of CO2 in the hydrogen fuel if it is recirculated.32,39 Theactivation energy (Ea) for OH− and CO3

2− transport was calculatedfrom the lnσ vs. 1/T for GT64-5 and GT64-10, Figure 5 (right). The Ea

values were 12.3 kJ/mol (OH−) and 17.6 kJ/mol (CO32−) for GT64-5,

and 12.6 kJ/mol (OH−) vs. 18.4 kJ/mol (CO32−) for GT64-10.

Water Uptake (WU).—The conductivity of a film is closely relatedto its water uptake per ion because the water is needed for ion hydrationand transport.49 The WU of the films was measured at room tempera-ture and the results are shown in Table I. The WU of the film decreasedwith TMHDA concentration. For GT82-5, the WU was 122% and theAEM showed a conductivity of 212 mS/cm at 80°C. The films withthe highest TMHDA concentration, GT82-20, had only 61% WU anda much lower conductivity, 128 mS/cm at 80°C. Clearly, both WU and

Figure 5. The measurement of OH− and CO32− conductivity as a function of temperature (left), Arrhenius plot of ln(σ) vs. inverse temperature for cross-linked

AEMs in OH− and CO32− form (right).

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 130.207.74.224Downloaded on 2019-10-10 to IP

Page 7: Georgia Institute of Technology - The Importance of Water …kohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal... · 2019-10-10 · Journal of The Electrochemical Society, 2020

Journal of The Electrochemical Society, 2020 167 054501

Figure 6. Variation of water uptake and IEC with TMHDA concentration of cross-linked GT82 AEMs (left). Variation of conductivity and hydration number withIEC of cross-linked GT82 AEMs (right).

conductivity decrease with cross-link density, as expected. A similartrend was observed for GT64. However, interestingly, GT78-5 had thehighest WU, 163%, which is likely due to its high molecular weight,Mn = 103.6 kDa, resulting in larger channels that can accumulate ex-cess free water; however, this led GT78-5 to have the lowest measuredconductivity for a film with 5% TMHDA. Figure 6 (left) shows thevariation of WU and IEC with TMHDA concentration for GT82. Asshown in Figure 6 (left), WU decreased significantly with TMHDAconcentration, although the IEC dropped marginally. Hence, it is clearthat light cross-linking is an effective strategy to mitigate WU withoutsacrificing IEC. The WU can also be viewed in terms of hydrationnumber (λ), which is a measure of the number of water moleculesper ionic pair. As shown in Figure 6 (right), for GT82 films, λ andconductivity were 9.11 to 17.6, and 128 to 212 mS/cm at 80°C, re-spectively. It is noted that although the IEC changed marginally (3.72to 3.84 meq/g), λ and conductivity changed dramatically.

The number of bound or non-freezable (Nbound) water and unboundor freezable (Nfree) water molecules per ion pair (i.e. anion-cation pair)was calculated from λ using DSC measurements. It is shown in Table Ithat increasing the TMHDA concentration decreased both Nbound andNfree. For example, GT82-20 had the lowest Nbound and Nfree waters,0 and 9, respectively. For GT82-5, the Nbound and Nfree were 7.6 and10, respectively. The ionic conductivity (σ) and mobility, as measuredby σ/IEC, were related to the amount of Nbound and Nfree waters in thefilm. For GT82-5, the conductivity and mobility were 212 mS/cm and55 (S g)/(eq cm), respectively, at 80°C. Both σ and σ/IEC decrease to128 mS/cm and 34 (S g)/(eq cm), respectively, for GT82-20 due toinsufficient Nbound and Nfree waters inside the film. As shown in Table Ifor GT64-5, σ and σ/IEC decrease with TMHDA concentration. ForGT78-5, the Nbound and Nfree were 11 and 12.7, respectively. The highervalue of Nfree, 12.7, caused flooding of the ion conductive channels andthe conductivity was lower than GT64-5 (181 vs. 162 mS/cm at 80°C).

At high TMHDA concentration, the film is more rigid and it isdifficult for water molecules to populate the film. The lack of suffi-cient free and bound water inside the film causes the conductivity todecrease.23 The membrane with an optimum amount of bound andfree water is expected to produce the most effective membrane in afuel cell.21–23

Thermal, alkaline stability and mechanical properties.—Ther-mogravimetric analysis (TGA) was used to measure the thermal sta-bility of the films in Br− form. Four significant weight loss regionswere observed, Figure 7. The weight loss in the first temperature stepbelow 100°C corresponds to evaporation of water or organic solventsfrom the film.14,50 The weight loss resulting from the decomposition of

the quaternary ammonium group was observed at ca. 275°C. The thirdstep at 350°C was due to the decomposition of the alkyl side chains inthe polymer. In the fourth step, polymer backbone degradation tookplace at >400°C. The variation in IEC had no significant effect onthe thermal stability. The results were in-line with the previously pub-lished norbornene polymers.21,22

The durability and lifespan of alkaline electrochemical devices de-pends on the chemical stability of the AEMs. The films were aged in1 M NaOH for more than 1000 h and the conductivity was measuredat regular intervals. Each data point was measured in triplicate andthere was <0.5% deviation in the individual measurements. The lossin hydroxide ion conductivity was plotted as a function of aging time.As shown in Figure 8 (left), the AEMs had <1.5% loss in ionic con-ductivity over 1000 h. This is in agreement with previously publishedresults.21–23 The alkaline stability was further confirmed by monitoringthe structural change before and after the alkaline treatment using FT-IR spectroscopy. Figure 8 (right) shows the appearance/disappearanceof new/old peaks in the IR spectrum. The unchanged C-N stretchingfrequencies at 911, 971, and 1483 cm−1 indicates that the head-groupmoiety was intact after alkaline aging.51–53

Figure 7. TGA traces of the cross-linked AEMs under nitrogen atmosphere.

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 130.207.74.224Downloaded on 2019-10-10 to IP

Page 8: Georgia Institute of Technology - The Importance of Water …kohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal... · 2019-10-10 · Journal of The Electrochemical Society, 2020

Journal of The Electrochemical Society, 2020 167 054501

Figure 8. Alkaline stability analysis of cross-linked AEMs in 1 M NaOH solution at 80°C. OH− conductivity loss over time (left). Structural characterizationusing FT-IR spectroscopy (right).

The elastic modulus of the GT-64 and GT-82 polymer films (with-out reinforcement) was compared to the modulus of the reinforcedGT-64 membranes, which was previously reported.24 The modulus ofthe membranes increases with cross-linking. PTFE reinforcement ofthe polymer significantly increases the elastic modulus. For example,the GT82 films with 10 mol% or 15 mol% crosslinking had an elasticmodulus of 45 to 50 MPa. GT64 with similar crosslinker had about halfthe value of GT82; however, the results are not directly comparablebecause of the different molecular weight of the polymers.

Fuel cell performance.—GT82-15 was selected for detailed anal-ysis in an alkaline fuel cell because of its high conductivity and bal-anced WU. The GT82-15 membrane was loaded into a 5 cm2 singlecell AEMFC with identical electrodes to those published previously.24

In these electrodes, a small amount of PTFE is added to both the an-ode and cathode electrodes in order to help with water management.This allows the results obtained in this study to be directly comparedto the literature state-of-the-art. Polarization curves were collectedwith both H2/O2 and H2/air (CO2-free) feed gases entering the cellat a volumetric flowrate of 1 L/min, and the results are shown inFigure 9. Positively, the GT82-15 based AEMFC was able to achievea power density of 3.5 W/cm2 and a maximum current density of 9.7A/cm2 at 80°C, making this cell the highest-performing AEMFC re-ported to-date.24,54,55 The high performance was due to a combinationof factors: (i) very high ionic conductivity (i.e. low internal resistanceloss) and (ii) very high water transport through the AEM from thehydrogen anode to the oxygen cathode during operation. A criticalcurrent-limiting factor in AEMFC’s is the rate of water transport fromthe water-producing negative electrode to the water-consuming posi-tive electrode.32

In order to show the influence of anode-cathode water transporton the achievable current density and peak power more definitively,two variables were changed in the next batch of fuel cell tests. First,the electrode composition was changed to be identical to a previousstudy30 where no PTFE was added to the catalyst layers. This is im-portant because in these tests, the electrode itself does little to removewater from the anode. This places an even greater burden on the mem-brane to transport water from the anode to cathode. This change inthe electrodes is expected to reduce the overall achievable current andpower, but the purpose of this experiment is to show the importance ofwater transport through the membrane, not just achieving the highestpower density. Second, the membrane and membrane thickness wasvaried; GT78-15 membranes with thicknesses of 10 μm, 20 μm and30 μm were used. Testing different thicknesses is important because

the diffusional flux of water from the hydrogen anode to the oxygencathode is inversely proportional to thickness. The cells were oper-ated under various conditions and their performance and operatingHFR were measured in order to understand the water transport behav-ior during operation. Figure 10 shows the results of AEMFC tests withGT78-15 AEMs with different thicknesses. The cell with the 10 μmmembrane has a lower peak power density and achievable current thanthe cell in Figure 9, due to the experimental conditions noted above.However, the performance was still very high, and higher than othercells with identical electrodes.30 The optimized dew points for the cellwith the 10 μm AEM were 75°C and 77°C at the anode and cathode,respectively. The peak power density was 1.95 W/cm2 and the cellwas able to support a current density of 6.3 A/cm2 at 0.1 V. Also, theaverage HFR during operation was 5.8 m�, which corresponds to anASR of 0.029 � cm2, which is somewhat lower than the ex-situ valuefor this AEM reported in Table I. The in-situ ASR of this membrane isalso the lowest of any AEM reported to date in an operating AEMFC.

0 2000 4000 6000 8000 100000.0

0.2

0.4

0.6

0.8

1.0

1.2

GT82-15 H2/O2

GT82-15 H2/Air

Current Density (mA/cm2)

Cel

l Vol

tage

(V)

0

500

1000

1500

2000

2500

3000

3500

Pow

er D

ensi

ty (m

W/c

m2 )

Figure 9. Current-voltage (filled) and current-power density (empty) curvesfor AEMFCs with a 10 μm GT82-15 AEM, operating with H2/O2 (blue)and H2/air (red) feeds at 1 L/min. The cell temperature was 80°C. The an-ode/cathode dew points for the cells operating with O2 and air (CO2-free)at the cathode were 66°C/75°C and 70°C/78°C, respectively. The Anode:0.70 mgPtRu/cm2; 0.05 MPa backpressure. Cathode: 0.60 mgPt/cm2; 0.1 MPabackpressure.

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 130.207.74.224Downloaded on 2019-10-10 to IP

Page 9: Georgia Institute of Technology - The Importance of Water …kohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal... · 2019-10-10 · Journal of The Electrochemical Society, 2020

Journal of The Electrochemical Society, 2020 167 054501

0 1000 2000 3000 4000 5000 6000 70000.0

0.2

0.4

0.6

0.8

1.0

1.2 GT78-15 10 µm GT78-15 20 µm GT78-15 30 µm GT78-15 30 µm (RDP)

Current Density (mA/cm2)

Cel

l Vol

tage

(V)

0

500

1000

1500

2000

Pow

er D

ensi

ty (m

W/c

m2 )

Figure 10. Current-voltage (filled) and current-power density (empty) curvesfor AEMFCs assembled with GT78-15 AEMs with three thicknesses: 10 μm,20 μm and 30 μm. All cells were operated with H2/O2 reacting gases at 80°Cwith no back pressurization. The anode and cathode catalyst loadings were0.70 mgPtRu/cm2 and 0.60 mgPt/cm2, respectively. The AEMFCs operated with10 μm (blue) and 20 μm (red) were operated at anode/cathode dew pointsof 75°C/77°C. AEMFCs assembled with the 30 μm AEM were operated atanode/cathode dew points of 75°C/77°C (cyan) and 70°C/72°C (purple).

As shown in Figure 10, as the thickness of the AEM increased, theachievable peak power density significantly decreased. However, thedecrease in current and power density were not caused by an increasein the voltage drop across the thicker membrane because the ASRof the 20 μm GT78-15 AEM was slightly lower (5.6 m�.) than theHFR of the 10 μm GT78-15 AEM (5.8 m�). Although the ASR isproportional to the membrane thickness, the membrane conductivitydepends on the WU. During operation in a fuel cell, the hydration stateof the membrane varies across the membrane as water is consumed atthe cathode and produced at the anode. This means that λ in an operat-ing cell will have a spatial dependence with a lower value at the oxygencathode and a higher value at the hydrogen anode. Thus, in an operat-ing cell, very thin membranes can have higher HFR because the localhydration state is less favorable. This local hydration gradient is likelymore exaggerated in very thin membranes than thicker membranes.This is supported by the observation that during operation, the HFRfor the 30 μm was 80% higher than the HFR for the 20μm AEM (i.e.50% higher thickness led to 80% higher HFR). The gradient in hydra-tion number does call into question the validity of ex-situ hydroxideaging of AEMs since the hydration number (i.e. local hydroxide con-centration within the membrane) in operating cells is non-uniform andis likely to be low at the oxygen cathode. No degradation in perfor-mance with time was observed, however, and the long-term durabilityof these AEMs will be the subject of a future study.

The values for the optimized dew points of the feed gases weredifferent for the 10 μm and 20 μm thick membranes. Figure 10 showedthe performance for all cells at the optimized dew points for the 10 μmAEM, 75°C and 77°C, respectively, for the anode and cathode. Theoptimized dew points for the 20 μm membrane were 70°C for theanode and 73°C for the cathode. The difference in the optimized dewpoints is important because the dew point helps to set the rate ofwater evaporation from the anode where water is produced. The waterformed at the anode either evaporates into the anode reacting gas ordiffuses across the membrane to the cathode. The lower the dew pointof the incoming anode, the higher the rate of evaporation. If the rateof water removal from the anode is not fast enough, anode floodingoccurs. Combined with the lower achievable power density and lowercurrent density for the thicker membrane, the lower values for theoptimum dew points shows that the rate of water transport throughthe AEM is lower, resulting in lower water flux and current at theoxygen cathode. This is the first time that the AEM thickness and water

transport rates have been linked, which is important for understandingthe design and operation of AEMs and AEMFCs.

The distortion of the polarization and power curves (Fig. 10) athigher thicknesses and higher dew points also suggests that water dy-namics are playing a significant role in device performance. In the30 μm thick AEM, there is an initial polarization that reaches an ap-parent mass transport limit at ca. 1.5 A/cm2. Then, there is an inflectionpoint as the cell passes 2 A/cm2 where this limitation appears to besomewhat alleviated. This can be explained by observing the rate ofwater consumption at the cathode at high current. Water consumptionat the cathode dries out the electrode and lowers the local hydrationnumber of the AEM, but a higher water flux from the anode through themembrane allows the cell to support a higher current density withoutanode-flooding. It should be noted that the distorted current-voltageshape can be avoided by increasing the amount of water removed fromthe cell by lowering the oxygen dew point. Also shown in Figure 10are polarization and power curves for a cell with the 30 μm GT78-15AEM operated at anode and cathode dew points of 70°C and 72°C,respectively. The cell performance at the reduced reacting gas dewpoints curves are denoted as “GT78-15 30 μm (RDP)”. As shown, thepolarization curve has a more typical shape and a higher peak powerdensity was achieved. Lowering the feed gas dew points in this celldoes come at the expense of reduced performance in the kinetic re-gion and higher HFR (ca. 14 m�) and ASR (0.07 � cm2) due to therelatively lower rate of water transport through the thicker AEM.

The dynamics between the cell performance at one static set ofdew points and optimal dew points for each AEM thickness shows thedew point sensitivity of AEMFCs. In general, the thicker the AEM, themore sensitive the AEMFC performance was to the reacting gas dewpoints. With the thickest AEM tested here (30 μm), even the shape ofthe polarization curve changed when the reacting gas dew points weretoo high. In addition, the distorted shape was still prevalent when theanode and cathode dew points were each only 2°C higher that theiroptimized values. With thinner AEMs, the shapes of the polarizationcurves were less sensitive to the dew points, as was the achievablecurrent and peak power density. For example, when the dew pointsfor the 20 μm AEMs were increased by 2°C, the peak power densitydecreased by ca. 5% from its maximum value and the current densityat 0.1 V was nearly identical. It was only when the dew points wereincreased by 6°C to 8°C that the peak power and achievable currentdensity were significantly affected. For the cells operating with the10 μm AEM, their performance was slightly more sensitive to thedew point values than the 20 μm membranes, but the shape of thepolarization curve did not change as a result of altering the dew points.For these cells, if the dewpoints were either raised or lowered by 2°C,there was around a 10% drop in the peak power density, although theachievable current density at 0.1 V was similar.

Finally, in-situ stability is an important factor in all fuel cells.Constant-current, short term durability tests were performed with theGT82-15 membrane-based AEMFC, Figure 11 (identical configura-tion to the cell in Figure 9). Remarkably, at a constant current densityof 600 mA/cm2 under H2/air (CO2-free), the cell voltage loss wasnegligible over 100 h. The cell voltage fluctuation at the 80 h pointwas due to a change in the feed dew points caused by automatic fillingof cold water into the test station humidifier. The high frequency resis-tance (HFR) of the cell was nearly constant over the entire experiment.The extremely stable HFR suggests very little, if any, degradation ofthe GT82-15 membrane during testing, which is in good agreementwith the ex-situ chemical stability tests for this AEM, discussed above.The most likely reason drop in the maximum current with air (about3.5 A/cm2) compared to the maximum current density with pure oxy-gen (about 9.5 A/cm2) is permeation of oxygen through the ionomerat the positive electrode.

Discussion

High ionic conductivity, rapid water transport, long-term alkalinestability and the ability to make thin, robust membranes are the keyproperties that are needed from AEMs in AEMFCs. The effect of car-

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 130.207.74.224Downloaded on 2019-10-10 to IP

Page 10: Georgia Institute of Technology - The Importance of Water …kohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal... · 2019-10-10 · Journal of The Electrochemical Society, 2020

Journal of The Electrochemical Society, 2020 167 054501

0 20 40 60 80 1000.3

0.4

0.5

0.6

0.7

0.8

Cell Voltage (V)HFR (mOhm)

Time (h)

Cel

l Vol

tage

(V)

0

5

10

15

20

25

HFR

(mO

hm)

Figure 11. Cell voltage at constant current (left axis) and HFR (right axis) forGT82-15 AEMFC.

bonation has an impact in fuel cell operation because the mobility ofcarbonate is significantly lower than hydroxide. Thus, membranes withvery high hydroxide conductivities (e.g. >200 mS/cm at 80°C) are fa-vored. The properties of several AEMs reported in the literature, andtheir comparison to this work are shown in Table II. Previously, Wanget al. synthesized radiation-grafted HDPE and LDPE AEMs with highhydroxide conductivity, 202 to 214 mS/cm at 80°C; however, therewas 6.2 to 8% decrease in conductivity after 500 h ex-situ aging.55,56

Later, Zhu et al. developed crossed-linked, comb-shaped PPO filmswith ionic conductivity of 200 mS/cm (80°C) that showed a 27% de-crease in conductivity after 500 h aging.57 Recently, Zhu et al. reportedan AEM able to achieve 201 mS/cm (80°C) and 15.8% loss in con-ductivity over 1000 h with poly(olefin)-based AEMs.58 Jannasch re-ported QPip-tethered PPO based AEMs with high ionic conductivity of221 mS/cm at 80°C, but 15% degradation occurred after only 240 h.59

Recently, we have reported the synthesis of poly(norbornene) films byvinyl addition and ROMP.22,23 The films have high ionic conductivity,

ca. 200 mS/cm (80°C), and no detectable loss in ionic conductivityin ex-situ aging at 80°C. The hydroxide conducting films synthesizedin this study had little or no degradation (<1.5% degradation over1000 h at 80°C in 1 M NaOH) and record high ionic conductivity,212 mS/cm at 80°C for a stable AEM. The simple backbone structurecontaining all sp3 hybridized carbons provides a hydrophobic back-bone with minimal bond polarity (inhibiting hydroxide attack) and isa lower-cost alternative to AEMs with fluorinated backbones.14,25,29,60

This study also shows that the critical AEM features responsible forthe leap in fuel cell performance shown here are the mechanical prop-erties (i.e. ability to fabricate ultra-thin membranes) and water trans-port (i.e. transporting water from the anode to the water-consumingcathode). Water removal from the anode is critical for preventing an-ode flooding and supplying water to the oxygen cathode. Humidifi-cation/dehumidification of the feed gases alone is not sufficient forachieving high current density because the inner portion of the elec-trodes (closest to the AEM) are the most active regions. Thus, highwater flux from the anode to the cathode is the most effective meansof mitigating anode flooding and supplying the water to the cathode.Ultra-thin membranes facilitate high flux rates for water and mini-mize the parasitic voltage drop across the membrane. Very light cross-linking in the AEMs used here provided enough control over WU sothat very high IEC values could be used without mechanical distortionof the membranes.

Conclusions

A series of cross-linked poly(norbornene) membranes were syn-thesized for electrochemical devices, and were demonstrated in AEM-FCs. The precursor polymers were cross-linked with a hexyl spacerdiamine (TMHDA) in different mol% with respect to the halogenatedmonomer in the polymer chain. The IECs after cross-linking were3.25 to 3.84 meq/g. Light cross-linking was found to be effective inmitigating the effect of high IECs toward high water uptake and lowmechanical stability. The 10.1 bound water molecules and 7.6 free wa-ter molecules were optimum for hydration and effective conductionof ions through the film (GT82-5) to produce maximum ionic con-ductivity. The record high ionic conductivity of 212 mS/cm at 80°Cwas achieved with only <1.5% loss in ionic conductivity over 1000 hof aging in 1 M NaOH at 80°C. The films were stable up to 400°C.

Table II. Properties (hydroxide ion conductivity, alkaline stability, water uptake etc.) of highly conducting (>190 mS/cm at 80°C) AEMs.

Membrane

OH−Conductivity

(mS/cm)b IEC (meq/g)d Water Uptakee(%) Alkaline stabilityf Reference

Cross-linked membranes using VApoly(norbornene)

198 3.43 69 <1.5% degradation after 1000 h 23

Cross-linked membranes using ROMPpoly(norbornene)

195 4.51 115 Retained its initial conductivity after 792 h 22

Cross-linked membranes using VApoly(norbornene)a

212 3.84 (3.76) 122 <1.5% degradation after 1000 h This work

HDPE-based radiation-grafted AEM 214c (2.44) 155 8% drop in ionic conductivity after 500 h 55LDPE-based radiation-grafted AEM 202c (2.54) 149 6.2% drop in ionic conductivity after 500 h 56Crossed-linked comb-shaped PPO

membranes200 2.63 (2.35) 144 27% decrease in conductivity for similar

membrane after 500 h57

Poly(olefin)-based AEMs 201 2.76 (2.41) 193 15.8% degradation after 1000 h 58AEM Based on QPip-Tethered PPO 221 2.8 (2.6) 115 15% degradation after 240 hg 59

aBlock copolymers using bromobutyl norbornene (BBNB) as the halogenated block.bOH− conductivity was measured at 80°C.cMeasured at 100% relative humidity.dTheoretical IECs was determined by 1H NMR or feed ratio, numbers in the brackets indicate the IECs by titration.eWater uptake was measured at room temperature.fStability assessment was performed at 80°C in 1 M NaOH or 1 M KOH.gMeasurement was performed at 90°C. VA = vinyl addition; ROMP = ring-opening metathesis polymerization.

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 130.207.74.224Downloaded on 2019-10-10 to IP

Page 11: Georgia Institute of Technology - The Importance of Water …kohl.chbe.gatech.edu/sites/default/files/Mrinmay Mandal... · 2019-10-10 · Journal of The Electrochemical Society, 2020

Journal of The Electrochemical Society, 2020 167 054501

The H2/O2 fuel cell performance at 80°C showed a peak power den-sity of 3.5 W/cm2 and maximum current density of 9.7 A/cm2. It wasfound that the mechanical properties and ability to form ultra-thinmembranes is critical to obtaining high fuel cell performance.

ORCID

Mrinmay Mandal https://orcid.org/0000-0002-3404-9588Garrett Huang https://orcid.org/0000-0003-0100-4161William E. Mustain https://orcid.org/0000-0001-7804-6410Paul A. Kohl https://orcid.org/0000-0001-6267-3647

References

1. G. Merle, M. Wessling, and K. Nijmeijer, J. Membr. Sci., 377, 35 (2011).2. N. Li and M. D. Guiver, Macromolecules, 47, 2175 (2014).3. N. Li, T. Yan, Z. Li, T. Thurn-Albrecht, and W. H. Binder, Energy Environ. Sci., 5,

7888 (2012).4. 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, and L. Zhuang,Energy Environ. Sci., 7, 3135 (2014).

5. L. Carrette, K. A. Friedrich, and U. Stimming, Fuel Cells, 1, 5 (2001).6. M. Winter and R. J. Brodd, Chem. Rev., 104, 4245 (2004).7. B. C. H. Steele and A. Heinzel, Nature, 414, 345 (2001).8. C. G. Arges and L. Zhang, ACS Appl. Energy Mater., 1, 2991 (2018).9. S. Lu, J. Pan, A. Huang, L. Zhuang, and J. Lu, Proc. Natl. Acad. Sci. U. S. A., 105,

20611 (2008).10. E. H. Yu, X. Wang, U. Krewer, L. Li, and K. Scott, Energy Environ. Sci., 5, 5668

(2012).11. M. A. Hickner, A. M. Herring, and E. B. Coughlin, J. Polym. Sci., Part B: Polym.

Phys., 51, 1727 (2013).12. J. Zhou, K. Joseph, J. M. Ahlfield, D.-Y. Park, and P. A. Kohl, J. Electrochem. Soc.,

160, F573 (2013).13. A. D. Mohanty and C. Bae, J. Mater. Chem. A, 2, 17314 (2014).14. L. Liu, J. Ahlfield, A. Tricker, D. Chu, and P. A. Kohl, J. Mater. Chem. A, 4, 16233

(2016).15. A. D. Mohanty, S. E. Tignor, M. R. Sturgeon, H. Long, B. S. Pivovar, and C. Bae, J.

Electrochem. Soc., 164, F1279 (2017).16. D.-Y. Park, P. A. Kohl, and H. W. Beckham, J. Phys. Chem. C, 117, 15468 (2013).17. X. C. Chen, D. T. Wong, S. Yakovlev, K. M. Beers, K. H. Downing, and N. P. Balsara,

Nano Lett., 14, 4058 (2014).18. J. Sun, X. Liao, A. M. Minor, N. P. Balsara, and R. N. Zuckermann, J. Am. Chem.

Soc., 136, 14990 (2014).19. S. C. Price, X. Ren, A. C. Jackson, Y. Ye, Y. A. Elabd, and F. L. Beyer, Macro-

molecules, 46, 7332 (2013).20. K. M. Meek, S. Sharick, Y. Ye, K. I. Winey, and Y. A. Elabd, Macromolecules, 48,

4850 (2015).21. M. Mandal, G. Huang, and P. A. Kohl, J. Membr. Sci., 570−571, 394 (2019).22. W. Chen, M. Mandal, G. Huang, X. Wu, G. He, and P. A. Kohl, ACS Appl. Energy

Mater., 2, 2458 (2019).23. M. Mandal, G. Huang, and P. A. Kohl, ACS Appl. Energy Mater., 2, 2447 (2019).24. G. Huang, M. Mandal, X. Peng, A. C. Yang-Neyerlin, B. S. Pivovar, W. E. Mustain,

and P. A. Kohl, J. Electrochem. Soc., 166, F637 (2019).25. L. Liu, G. Huang, and P. A. Kohl, J. Mater. Chem. A, 6, 9000 (2018).

26. W.-H. Lee, E. J. Park, J. Han, D. W. Shin, Y. S. Kim, and C. Bae, ACS Macro Lett.,6, 566 (2017).

27. H.-S. Dang and P. Jannasch, J. Mater. Chem. A, 4, 11924 (2016).28. A. M. A. Mahmoud and K. Miyatake, J. Mater. Chem. A, 6, 14400 (2018).29. L. Liu and P. A. Kohl, J. Polym. Sci. Part A: Polym. Chem., 56, 1395 (2018).30. T. J. Omasta, A. M. Park, J. M. LaManna, Y. Zhang, X. Peng, L. Wang,

D. L. Jacobson, J. R. Varcoe, D. S. Hussey, B. S. Pivovar, and W. E. Mustain, EnergyEnviron. Sci., 11, 551 (2018).

31. T. J. Omasta, L. Wang, X. Peng, C. A. Lewis, J. R. Varcoe, and W. E. Mustain, J.Power Sources, 375, 205 (2018).

32. M. R. Gerhardt, L. M. Pant, and A. Z. Weber, J. Electrochem. Soc., 166, F3180 (2019).33. J. Pan, C. Chen, Y. Li, L. Wang, L. Tan, G. Li, X. Tang, L. Xiao, J. Lu, and L. Zhuang,

Energy Environ. Sci., 7, 354 (2014).34. M. M. Hossain, L. Wu, X. Liang, Z. Yang, J. Hou, and T. Xu, J. Power Sources, 390,

234 (2018).35. C. X. Lin, Y. Z. Zhuo, E. N. Hu, Q. G. Zhang, A. M. Zhu, and Q. L. Liu, J. Membr.

Sci., 539, 24 (2017).36. K. H. Lee, D. H. Cho, Y. M. Kim, S. J. Moon, J. G. Seong, D. W. Shin, J.-Y. Sohn,

J. F. Kim, and Y. M. Lee, Energy Environ. Sci., 10, 275 (2017).37. J. Wang, G. He, X. Wu, X. Yan, Y. Zhang, Y. Wang, and L. Du, J. Membr. Sci., 459,

86 (2014).38. N. Ziv, W. E. Mustain, and D. R. Dekel, ChemSusChem, 11, 1136 (2018).39. Y. Zheng, T. J. Omasta, X. Peng, L. Wang, J. R. Varcoe, B. S. Pivovar, and

W. E. Mustain, Energy Environ. Sci.40. J. Lipian, R. A. Mimna, J. C. Fondran, D. Yandulov, R. A. Shick, B. L. Goodall,

L. F. Rhodes, and J. C. Huffman, Macromolecules, 35, 8969 (2002).41. C. Wang, B. Mo, Z. He, Q. Shao, D. Pan, E. Wujick, J. Guo, X. Xie, X. Xie, and

Z. Guo, J. Membr. Sci., 556, 118 (2018).42. S. J. Lue and S. J. Shieh, J. Macromol. Sci. Part B: Phys., 48, 114 (2009).43. B. Mecheri, V. Felice, Z. Zhang, A. D’Epifanio, S. Licoccia, and A. C. Tavares, J.

Phys. Chem. C, 116, 20820 (2012).44. A. L. Moster and B. S. Mitchell, J. Appl. Polym. Sci., 113, 243 (2009).45. S. D. Poynton, R. C. T. Slade, T. J. Omasta, W. E. Mustain, R. Escudero-Cid, P. Ocón,

and J. R. Varcoe, J. Mater. Chem. A, 2, 5124 (2014).46. H.-S. Dang and P. Jannasch, Macromolecules, 48, 5742 (2015).47. J. Pan, S. Lu, Y. Li, A. Huang, L. Zhuang, and J. Lu, Adv. Funct. Mater., 20, 312

(2010).48. B. Lin, L. Qiu, J. Lu, and F. Yan, Chem. Mater., 22, 6718 (2010).49. C. Chen, Y.-L. S. Tse, G. E. Lindberg, C. Knight, and G. A. Voth, J. Am. Chem. Soc.,

138, 991 (2016).50. M. Zhu, X. Zhang, Y. Wang, Y. Wu, H. Wang, M. Zhang, Q. Chen, Z. Shen, and N. Li,

J. Membr. Sci., 554, 264 (2018).51. A. M. A. Mahmoud, A. M. M. Elsaghier, K. Otsuji, and K. Miyatake, Macro-

molecules, 50, 4256 (2017).52. A. Amel, N. Yitzhack, A. Beylin, J. Pan, M. A. Hickner, and Y. Ein-Eli, J. Electrochem.

Soc., 165, F1133 (2018).53. D. Chen and M. A. Hickner, ACS Appl. Mater. Interfaces, 4, 5775 (2012).54. Q. Li, H. Peng, Y. Wang, L. Xiao, J. Lu, and L. Zhuang, Angew. Chem. Int. Ed., 131,

1456 (2019).55. L. Wang, X. Peng, W. E. Mustain, and J. R. Varcoe, Energy Environ. Sci., 12, 1575

(2019).56. L. Wang, M. Bellini, H. A. Miller, and J. R. Varcoe, J. Mater. Chem. A, 6, 15404

(2018).57. L. Zhu, T. J. Zimudzi, N. Li, J. Pan, B. Lin, and M. A. Hickner, Polym. Chem., 7,

2464 (2016).58. L. Zhu, X. Yu, X. Peng, T. J. Zimudzi, N. Saikia, M. T. Kwasny, S. Song,

D. I. Kushner, Z. Fu, G. N. Tew, W. E. Mustain, M. A. Yandrasits, and M. A. Hickner,Macromolecules, 52, 4030 (2019).

59. H.-S. Dang and P. Jannasch, ACS Appl. Energy Mater., 1, 2222 (2018).60. J. Ahlfield, G. Huang, L. Liu, Y. Kaburagi, Y. Kim, and P. A. Kohl, J. Electrochem.

Soc., 164, F1648 (2017).

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 130.207.74.224Downloaded on 2019-10-10 to IP