You are on page 1of 103

Journal of Power Sources

Ether-free sulfonated poly(fluorene biphenyl indole) membranes and ionomer binders


for proton exchange membrane fuel cells
--Manuscript Draft--

Manuscript Number: POWER-D-22-04425R2

Article Type: Research Paper

Keywords: fuel cell; proton exchange membrane; Ionomer binder; Chemical stability; Ether-free

Corresponding Author: Dukjoon Kim, PhD


Sungkyunkwan University
Suwon, KOREA, REPUBLIC OF

First Author: Thuc Vu Dong

Order of Authors: Thuc Vu Dong

Vo Dinh Cong Tinh

Dukjoon Kim

Abstract: Because of the vulnerability of ether bonds to oxidative radicals, the ether-free polymer
electrolytes were synthesized to prepare chemically more stable proton exchange
membranes for long time durable fuel cell application. Here, we presented a series of
ether-free sulfonated poly(fluorene biphenyl indole)s (SPFBI) synthesized by a feasible
and cost-effective process. Owing to the ether-free structure, SPFBI membranes
displayed great chemical stability comparable to the commercial Pemion membranes
from Fenton’s test. SPFBI membranes exhibited high proton conductivity up to 0.1956
S cm -1 at 80 o C under the limited swelling ratio lower than 16.5%. Consequently,
while the SPFBI-0.4 membrane showed the excellent cell performance reaching the
maximum power density of 365 mW cm -2 , it also played a decent role as an ionomer
binder for both SPFBI and Nafion membranes. In open circuit voltage stability test, the
membrane electrode assemblies using SPFBI-0.2 membrane sustained more than 620
h without any severe defects at 90 o C under low relative humidity of 30%,
outweighing the Nafion counterpart. This study provides a remarkable approach for
optimizing the ether-free hydrocarbon-based polymer structure for the application of
both polymer electrolyte membrane and ionomer binder with high proton conductivity
and chemical stability.

Suggested Reviewers: Simon Thiele


University of Freiburg Department of Microsystems Engineering
simon.thiele@imtek.de

Byungchan Bae
Korea Institute of Energy Research
bcbae@kier.re.kr

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Cover Letter

School of Chemical Engineering


Sungkyunkwan University
Suwon 16419, Republic of Korea
Tel: +82-31-290-7240 Fax: +82-31-290-7272
http://chemeng.skku.edu/

September, 2022

Dear Editor:

We are pleased to submit the paper entitled, “Ether-free sulfonated poly(fluorene biphenyl indole)

membranes and ionomer binders for proton exchange membrane fuel cells” by Vu Dong Thuc, Vo Dinh

Cong Tinh and Dukjoon Kim to Journal of Power Sources.

Fabrication of chemically stable membranes is crucial to extend the lifetime of polymer electrolyte

membrane fuel cells. Therefore, ether bonds that exist in many hydrocarbon-based proton exchange

membranes (PEMs) have raised concerns about oxidative stability, since those bonds are vulnerable to radical

attacks. In this work, we rationally designed a novel ether-free polymer, named sulfonated poly(fluorene

biphenyl indole) (SPFBI), consisting of a rigid polyaromatic main chain and long sulfonated alkyl side chains.

Accordingly, a series of SPFBI polymers were synthesized by a facile and cost-effective process using

superacid-catalyzed polymerization. Thanks to the decent solubility of SPFBI in alcohol and water mixture,

the polymers were employed to prepare both PEM and ionomer binder for fuel cell tests. In general, SPFBI

PEMs possessed high oxidative stability in comparison to Pemion and other reported PEMs. They also

exhibited excellent proton conductivities along with a limited swelling ratio, resulting in the enhanced fuel

cell performance of SPFBI membranes, compared to that of Nafion membrane. The open circuit voltage

stability test demonstrated superior long-term in-situ stability of SPFBI-0.2 PEM that remained intact after

over 620 h test. On the other hand, SPFBI-0.4 ionomer binder provides decent cell performances and did not

affect the stability of the cell. This study provides the possibility of developing ether-free PEMs with

excellent properties as well as merits further improvement of hydrocarbon ionomer binders for fuel cell

applications.
Ethics Statement

This work has not been published previously (except in the form of an abstract or as part of a published

lecture or academic thesis), and it is not under consideration for publication elsewhere. Also, its publication

is approved by all authors and tacitly or explicitly by the responsible authorities where the work was carried

out. If accepted, it will not be published elsewhere in the same form, in English or in any other language,

without the written consent of the Publisher.

Thank you for your consideration. We look forward to hearing a positive response from you.

Sincerely yours,

Dukjoon Kim

Ph.D., Professor, SKKU Fellow

Executive Director of Korea Polymer Society

National Research Laboratory, School of Chemical Engineering, Sungkyunkwan University

Suwon, 16419, Republic of Korea

[Email] djkim@skku.edu; [Phone] +82-10-3017-7250


Response to Reviewers

RESPONSE TO EDITOR’S COMMENT

Manuscript Number: POWER-D-22-04425R1

Title: Ether-free sulfonated poly(fluorene biphenyl indole) membranes and ionomer binders for
proton exchange membrane fuel cells.

The authors appreciate reviewers’ positive comments which recommend the publication of our
work in Journal of Power Sources. We also thank Editor for the comment on the format of Nyquist
plots:
Editor: the EIS Nyquist plots shall be traced on orthonormal sets of axes and parametered in
frequency as detailed in [https://doi.org/10.1016/j.jpowsour.2019.227635].
Response: Based on the comment and the practical guide, we modified the Nyquist plots in Fig.
6b, Fig 7d, e and f. In detail, all Nyquist plots were changed to be orthonormal (same length on
the x and y axes for 0.1 Ω cm2). To parameterize the graphs in frequency, we added the label and
a curved arrow to indicate the frequency range, as shown below:
Fig.6:

has changed to
Fig.7:

has changed to:


Previously rejected submission: Response to Editor/Reviewer
comments (Marked with highlighted changes)

Ether-free sulfonated poly(fluorene biphenyl indole) membranes and ionomer binders for

proton exchange membrane fuel cells

Vu Dong Thuca, Vo Dinh Cong Tinha, Dukjoon Kima*


a
School of Chemical Engineering, Sungkyunkwan University, 2066 Seobu-ro, Jangan-gu,

Suwon, Gyeonggi 16419, Republic of Korea


*
Email: djkim@skku.edu

Abstract

Because of the vulnerability of ether bonds to oxidative radicals, the ether-free polymer

electrolytes are synthesized to prepare chemically more stable proton exchange membranes for

long time durable fuel cell application. Here, we present a series of ether-free sulfonated

poly(fluorene biphenyl indole)s (SPFBI) synthesized by a feasible and cost-effective process.

Owing to the ether-free structure, SPFBI membranes display great chemical stability comparable

to the commercial Pemion membranes from Fenton’s test. SPFBI membranes exhibit high proton

conductivity up to 0.1956 S cm-1 at 80 oC under the limited swelling ratio lower than 16.5%.

Consequently, while the SPFBI-0.4 membrane shows the excellent cell performance reaching the

maximum power density of 365 mW cm-2, it also plays a decent role as an ionomer binder for both

SPFBI and Nafion membranes. In open circuit voltage stability test, the membrane electrode

assemblies using SPFBI-0.2 membrane can sustain more than 620 h without any severe defects at

90 oC under low relative humidity of 30%, outweighing the Nafion counterpart. This study

provides a remarkable approach for optimizing the ether-free hydrocarbon-based polymer

structure for the application of both polymer electrolyte membrane and ionomer binder with high

proton conductivity and chemical stability.

1
Keywords: Fuel cell; Proton exchange membrane; Ionomer binder; Chemical stability; Ether-free

1. Introduction.

To satisfy the growing demand for the green energy with no harm to the environment,

proton-exchange membrane fuel cell (PEMFC) has been recognized as one of the promising

electrochemical energy sources due to its eco-friendliness and high energy density [1–4]. As a key

component of PEMFC, proton-exchange membrane (PEM) needs to fulfill various requirements,

such as high mechanical and chemical stability and low gas crossover in addition to high proton

conductivity [5,6]. While many attempts have been launched to fabricate high-quality PEMs with

enhanced properties so far, the perfluorosulfonic acid-based membranes including Nafion are the

most successful and popular PEMs for commercial applications. Nafion possesses good

mechanical and oxidative stability as well as a decent proton conductivity in humidified conditions

due to the formation of ion clusters by well-separated hydrophobic and hydrophilic phases [4,7,8].

Despite those advantages, Nafion has major drawbacks, including high gas permeability, poor

water retention, and high cost [6,9]. Because of the perfluorinated backbone structure, Nafion also

causes environmental concerns in its disposal process [10,11].

In order to replace the perfluorinated polymers with cost and disposal ineffectiveness,

hydrocarbon-based polymers with acidic functional groups have been developed. Those generally

consist of the hydrophobic aromatic backbones functionalized with the proton-conducting moieties

such as sulfonic acid or phosphonic acid groups statistically distributed. Depending on the polymer

designs, these acid groups can be either directly attached to the polymer backbone [12,13], or

located on the long alkyl chains grafted onto the polymer backbone [14,15]. Therefore, similar to

the perfluorinated polymer membranes, the phase-separated hydrophilic domains in hydrocarbon-

2
based PEMs allow the construction of interconnected water channels to facilitate the proton

transport. With many attempts to improve membrane properties, the hydrocarbon-based PEMs

currently not only exhibit high proton conductivity and cell performance [16–18], but also offer

cost reduction compared to perfluorinated PEMs.

Despite the aforementioned advantages, several limitations of hydrocarbon-based PEMs

hinder their commercialization. A major concern is about their chemical stability, since the

polymer structure typically contains more reactive functional groups and thus possesses less

oxidative resistance than Nafion composed of an inert Teflon-like backbone [19,20]. In such

polymer structures, the ether bond is the most commonly encountered heteroatom bond in the main

chain, usually resulting from the polymerization reaction to link monomer molecules together

[14,21]. Unfortunately, ether groups are very prone to attacking of ·OH radicals continuously

generated from the formation of reactive oxygen species and fuel crossover during the cell

operation [22,23]. The low oxidative stability of ether group is because the oxygen in ether

increases the electron density of neighboring aryl rings via the electron-donating effect and thus

weakens the resistance against oxidative species [1,24,25]. Since the ether groups do not have

crucial role in PEM properties, their elimination from the polymer structure can be considered an

attractive way to enhance the chemical stability of PEMs. Accordingly, many studies have been

conducted to synthesize ether-free PEMs. For instance, Miyake et al. reported a flexible

polyphenylene PEM (SPP-QP) with excellent flexibility and stability [26]. Pemion is also a

commercially available ether-free hydrocarbon-based PEM [27,28]. The synthetic process of

polymer electrolytes without ether linkages is usually quite complicated with the requirement of

costly metallic catalysts [27–29]. The superacid-catalyzed polymerization reaction was employed

as a good method, characterized by a short reaction time without the requirement of heating and

3
expensive catalysts [30,31]. This method has been widely used to synthesize ether-free polymers

for anion exchange membrane (AEM) [32,33], but just a few publications have been reported for

PEMs [34,35]. As the ether-free PEMs synthesized from the previous studies showed only

mediocre properties without profound investigation of chemical stability enhancement, it is

difficult to draw the potential of its methodological effectiveness.

In this work, the synthesis and characterization of sulfonated poly(fluorene biphenyl indole)

(SPFBI) are reported as a novel ether-free polymer electrolyte with improved PEM properties.

Compared to the previous ether-free polymers synthesized by superacid-catalyzed reaction, this

polymer structure was rationally designed from common monomers to establish the mechanically

robust rigid aromatic backbone grafted with long sulfonated alkyl chains to improve proton

conductivity. The synthetic procedure of SPFBI involves two main steps, starting from the

synthesis of poly(fluorene biphenyl indole) (PFBI) from isatin, biphenyl, and fluorene under

superacid conditions [31]. In the second step, the sulfonated alkyl groups were grafted to PFBI

backbone using 1,3-propane sultone and tetrabutylammonium bromide catalyst under basic

conditions [36,37]. By varying the monomer ratios, a series of SPFBI PEMs with different sulfonic

acid contents were prepared for an intensive investigation on proton conductivity, oxidative

stability, and PEMFC performance. Additionally, the solubility of SPFBI in isopropanol and water

mixture allowed its utilization as the ionomer binder for the preparation of catalyst ink in

membrane electrode assembly (MEA) process.

2. Experimental Section.

2.1. Materials.

Fluorene, isatin, trifluoroacetic acid (TFA), trifluoromethanesulfonic acid (TFSA), 1,3

4
propane sultone (99.0%) and tetrabutylammonium bromide (TBAB) were purchased from Tokyo

Chemical Industry (TCI, Japan). Biphenyl, iron(II) sulfate heptahydrate (FeSO4.7H2O) and

hydrogen peroxide (H2O2) solution were obtained from Sigma-Aldrich (Milwaukee, WI, USA).

Nafion alcohol-based dispersion (5 wt%, EW1100) was supplied from Chemours (Wilmington,

DE, USA), as well as platinum (nominally 40% on carbon black) (Pt/C) was provided by Alfa

Aesar (Ward Hill, MA, USA) to prepare MEAs. Other common solvents (99.5% purity) were

obtained from Samchun Chemicals (Korea).

2.2. Synthesis of poly(fluorene biphenyl indole) (PFBI).

The synthetic process of PFBI is briefly represented in Scheme 1. In detail, isatin (150

mmol) was added to a three-neck round bottom flask, along with a total of 150 mmol of two other

monomers (biphenyl and fluorene). The ratio of biphenyl and fluorene was varied to control the

sulfonation degree of the final polymer product. Then, 15 mL of dichloromethane (CH2Cl2) was

poured into the flask to dissolve the monomers under magnetic stirring. Subsequently, TFA (10

mL) and TFSA (15 mL) were dropwise added, and the mixture was stirred and kept at 0-5 oC for

12 h. After the reaction was finished, the dark purple mixture was slowly poured into methanol to

obtain PFBI precipitation. The polymer was washed with methanol and water several times, and

then dried in a vacuum oven at 80 oC for 24 h. It is designated as PFBI-x, where x represents the

proportion of fluorene over the total amount of biphenyl + fluorene used for the synthesis. The

yield of the PFBI synthetic process was about 93-96%. 1H NMR (300 MHz, DMSO-d6): 10.78

ppm (Hi), 7.73 ppm (Hb), 7.55 ppm (Hc), 7.21 – 7.31 ppm (Hd,e,g), 6.93 – 7.19 ppm (Hf,h) and 3.74

ppm (Ha).

5
2.3. Synthesis of sulfonated poly(fluorene biphenyl indole) (SPFBI).

The second step in Scheme 1 illustrates the synthetic procedure of SPFBI. For each

synthesis, 3 g of PFBI and 0.116 g of TBAB were simultaneously dissolved in 70 mL of dimethyl

sulfoxide (DMSO) at 60 oC and then poured into a three-neck round bottom flask. Next, nitrogen

gas was continuously supplied into the flask to remove the oxygen, and 4 mL of 50 wt% KOH

was slowly added to the polymer solution by syringe, followed by adding an excess amount of 1,3-

propane sultone. The reaction mixture was stirred at 60 oC for 6 h, under a nitrogen atmosphere.

Afterward, the product was precipitated by dropwise adding into methanol, and then washed with

water and isopropyl alcohol (IPA) in sequence and dried. The obtained polymer is named SPFBI-

x, where x refers to the PFBI-x precursor. The yield of the sulfonation process is 85-88 %. 1H

NMR (300 MHz, DMSO-d6): 7.55 – 7.61 ppm (Hb,c), 7.21 – 7.40 ppm (Hd,e,g), 7.07 – 7.19 ppm

(Hf,h), 3.83 ppm (Hi), 2.34 – 2.46 ppm (Ha,l) and 1.89 ppm (Hk).

6
Scheme 1. Two-step synthetic process of SPFBI-x.

2.4. Preparation of membrane and catalyst ink.

To prepare the SPFBI membrane, 0.3 g of SPFBI was completely dissolved in DMSO at

60 oC, and then the polymer solution was poured into a petri dish and placed in a vacuum oven at

80 oC for 24 h. After that, the polymer membrane was peeled off to be immersed in 1 M H2SO4

solution for 24 h to convert the sulfonated groups to H+ form. Finally, the membrane was washed

with deionized water for 24 h. For membrane comparison, Nafion PEMs were prepared by casting

method. Nafion alcohol-based dispersion (5 wt%) was poured into a flat petri dish, then dried at

7
80 oC for 24 h. The formed film was peeled off from the dish to obtain a Nafion pristine membrane.

The thickness of all fabricated membranes in this work was controlled within a range of 80-90 µm.

For catalyst ink preparation, 5 wt% solution of SPFBI-0.4 was prepared in-house by

acidifying the polymer solid in 1 M H2SO4 solution. After rinsing it with water, it was dissolved

in a mixture of IPA and water (4:6 w/w). For comparison, the preparation of catalyst ink using

SPFBI-0.4 ionomer followed a similar procedure as that of Nafion ionomer. Pt/C catalyst was

mixed with deionized water, IPA, and 5 wt% polymer solution (Nafion or SPFBI-0.4) and

subsequently ultra-sonicated to obtain a colloidal catalyst slurry. In all catalyst ink solutions, the

ionomer binder:catalyst mass ratio was fixed at 1:4.

2.5. Characterizations.

Chemical structure. To confirm the synthesis of the polymers, the proton nuclear

magnetic resonance (1H NMR) spectra were obtained by a Varian Oxford 300 MHz NMR

spectrometer (CA, USA) using DMSO-d6 as a solvent to dissolve polymer samples. The standard

chemical shift of DMSO-d6 is determined at 2.50 ppm.

Intrinsic viscosity. The intrinsic viscosity ([η]) of SPFBI was determined by a method

using a Ubbelohde viscometer. The polymers were dissolved in DMSO to obtain solutions with

five different concentrations, and the efflux time of each solution was recorded at 20 oC. By

drawing a plot of reduced viscosity versus concentration, we obtained [η] at the y-intercept value.

Solubility testing. The solubility of polymers was measured in various common organic

solvents, water and a mixture of IPA and water (4:6 w/w). The tests were performed at room

temperature and 60 oC (only for solvents with a high boiling point). The polymer was considered

insoluble in a certain solvent if it cannot be dissolved in both room and elevated temperatures.

8
Membrane hydration properties and IEC. The water uptake (WU) and swelling ratio

(SR) of SPFBI membranes were measured using the weight and length change of membrane

samples in fully dry and wet states at different temperatures. The membrane was dried at 80 oC for

48 h, and then cut into a 1 x 3 cm rectangular piece to measure the mass (md) and length (ld) of the

dry sample. Next, the sample was immersed in deionized water for 48 h under stirring. After that,

the sample was removed, the excess water on the membrane surface was wiped off with tissue

papers to measure the mass and length of the membrane at wet state (mw and lw). The test was

repeated 3 times for each sample to obtain the average value of WU and SR according to the

following equations:

𝑊𝑈(%) = (𝑚𝑤 − 𝑚𝑑 )/𝑚𝑑 × 100 (1)

𝑆𝑅(%) = (𝑙𝑤 − 𝑙𝑑 )/𝑙𝑑 × 100 (2)

The experimental ion exchange capacity (IECex) was measured by acid-base titration

method. A dry membrane sample with the mass m (g) was submerged in standard NaOH solution

with the known concentration and volume (CNaOH, VNaOH) for 48 h. The base solution after

removing the membrane was titrated with standard HCl solution (CHCl, VHCl) to determine IECex

from equation (3):


𝐶𝑁𝑎𝑂𝐻 𝑉𝑁𝑎𝑂𝐻 − 𝐶𝐻𝐶𝑙 𝑉𝐻𝐶𝑙
𝐼𝐸𝐶𝑒𝑥 (meg g −1 ) = × 1000 (3)
𝑚

For comparison, the theoretical ion exchange capacity (IECtheo) was also calculated from

the chemical structure of SPFBI, based on the number of acid groups and the theoretical molecular

weight of the repeating unit of polymer.

Thermal and mechanical properties. The thermal stability of the membrane in proton

form was investigated by thermogravimetric analysis (TGA) using a Seiko Exstar 6000 instrument

(Seiko Instruments, Japan). The temperature range was set from 100 to 700 oC at the heating rate

9
of 10 oC min-1 under nitrogen atmosphere. In terms of mechanical property, both dry and hydrated

membranes were cut into a 1 x 5 cm rectangular piece to measure their tensile stress vs. strain

behavior using a universal tensile machine (UTM, Model 5565, Lloyd, Fareham, UK) with a 250

N load cell.

Proton conductivity. Proton conductivity (σ, S cm-1) was measured in in-plane dimension

by an electrochemical workstation (ZIVE SP2, Wonatech, Korea) connected with a four-probe

single cell (Bekktech LLC, Loveland, CO, USA). Each membrane was cut into a 1 x 3 cm shape

to place inside the cell, and the test was repeated three times at 95% relative humidity (RH). The

testing temperature varied from 30 to 80 oC and the oxygen gas rate was 500 mL min-1. The proton

conductivity of the membranes was calculated from equation (4):

𝜎 (𝑆 𝑐𝑚−1 ) = 𝑙/(𝑅 × 𝑡 × 𝑤) (4)

where l is the distance between electrodes (cm), R is sample resistance (Ω), t and w are sample

thickness (cm) and width (cm), respectively.

Oxidative stability. Fenton’s reagent (an aqueous solution containing 3 wt% H2O2 and 3

ppm Fe2+) was used to study the oxidative stability of SPFBI membranes. The membrane sample

was dried, weighed, and then immersed in the Fenton solution at 80 oC for 1 to 5 h. After the test,

the sample was gently washed with water and dried in a vacuum oven at 80 oC. The residual weight

(RW) of the sample was calculated from the before-test weight (mb) and after-test weight (ma) of

the sample, using equation (5):

𝑅𝑊 (%) = (𝑚𝑎 /𝑚𝑏 ) × 100 (5)

After the test, the proton conductivity of the treated membrane was measured following

the aforementioned procedure. Moreover, the chemical structure of the membrane sample was

analyzed using Fourier-transform infrared (FT-IR) spectroscopy (Nicolet iS20, Thermo Scientific,

10
Waltham, MA, USA).

Catalyst particle size. The catalyst ink was prepared as described in Section 2.4, and then

IPA was added to dilute the ink two-fold. Subsequently, the resultant solution was ultra-sonicated

for 30 min and the particle size of the catalyst was identified by employing a dynamic light

scattering (DLS) analyzer (ELS-Z2000, Photal Otsuka Electronics, Japan).

PEMFC tests. Membrane electrode assembly (MEA) was prepared by spraying catalyst

ink onto both sides of the PEM surface. A vacuum plate controller system (CNL Energy, Korea)

and a metal mask were used to maintain the flatness of the membrane, and the temperature was

kept at 60 oC to completely remove moisture and solvents during the spraying process. The Pt

loading was fixed at 0.4 mg cm-2 on an active area of 5 cm2 for each side of the membrane. Next,

the MEA was assembled with two gas diffusion layers (GDL, SGL GDL 39BB, CNL Energy,

Korea, thickness ~325 µm) and then sealed with two Teflon gaskets. The formed stack was inserted

between two graphite plates of the cell test unit in a PEMFC station (CNL Energy, Korea) for in-

situ fuel cell tests.

The PEMFC performance tests were conduct at 80 oC and 100% RH. Fully humidified

hydrogen and oxygen gases were supplied at anode and cathode, respectively, at the same flow

rate of 300 mL min-1. This flow rate corresponds to the stoichiometry values of hydrogen (λH2) and

oxygen (λO2) of 8.6 and 17.2, respectively, at 1000 mA cm-2. Electrochemical impedance

spectroscopy (EIS) measurements were carried out at cell conditions as the cell performance test

by connecting the fuel cell stack with ZIVE SP2 electrochemical workstation (Wonatech, Korea).

The EIS tests used the potentiostatic mode in the frequency range of 0.1 Hz – 100 kHz at the

potential amplitude of 6 mV. Nyquist plots were constructed and fitted by ZIVE Smart Manager

software.

11
In-situ open-circuit voltage (OCV) test of MEAs was carried out at 90 oC and 30% RH to

accelerate the degradation of MEA. The hydrogen and oxygen gas inlet rates were the same at 300

mL min-1. The MEA stability was demonstrated, as the OCV decays as a function of operation

time.

3. Results and discussion.

3.1. Polymer properties.

In the present study, we synthesized a series of ether-free SPFBI by a simple two-step

procedure: the polymerization under superacid-catalyzed condition to form PFBI without ether

groups, followed by the sulfonation reaction between PFBI and 1,3-propane sultone in an alkaline

solution. As described in section 2.2. and 2.3, the synthetic process of SFPBI has its remarkable

advantages: facile and short synthesis time and heating requirement of no more than 60 oC. Three

SPFBI samples, SPFBI-0.2, 0.3 and 0.4 with different sulfonation degrees, were synthesized by

simply adjusting the fluorene/biphenyl feed ratio to fabricate PEMs.

The chemical structures of PFBI and SPFBI were confirmed by 1H NMR spectra in Fig. 1.

The SPFBI consists of a hydrophobic poly(phenylene indole) backbone grafted with long alkyl

sulfonated side chains. In terms of the polymer backbone, fluorene was rationally employed as a

monomer along with biphenyl to increase and regulate the sulfonation degree of the final product.

Fluorene has a higher rotation energy barrier than biphenyl because the carbon C9 restricts the

movement of phenyl rings, and thus fluorene molecule is more planar than biphenyl, resulting in

higher reactivity towards the electrophilic substitution reaction [38]. Therefore, the use of fluorene

is also expected to improve the molecular weight and rigidity of polymer backbone, which is

beneficial for physicochemical properties of SPFBI PEMs. As a result, SPFBI-0.2, 0.3 and 0.4

12
showed different intrinsic viscosities ([η]). The [η] value of SPFBI-0.2 was 5.35 dL g-1, lower than

6.18 and 6.36 dL g-1 of SPFBI-0.3 and SPFBI-0.4, respectively. The difference of [η] implies the

increase of polymer molecular weight (MW) from SPFBI-0.2 to 0.4, based on the relationship

between [η] and MW in the Mark–Houwink–Sakurada relationship [39]. This result might

correspond to the increase of fluorene segment in polymer structure from 0.2 to 0.4 which

improves the polymerization reactivity.

Fig. 1. 1H NMR spectra of (a) PFBI-0.4 and (b) SPFBI-0.4.

The solubility of polymers in common solvents is summarized in Table S1. When we

increased the fluorene content to 0.5, SPFBI-0.5 became so hydrophilic that it was dissolved in

water. Hence, the maximum fluorene ratio was fixed at 0.4, ensuring that the membrane can retain

and operate in fuel cell environments. Although the SPFBI with a lower sulfonation degree was

insoluble in individual water and IPA, it was surprising that it showed a decent solubility in a

mixture of IPA and water even at the ambient condition. The dissolution of SPFBI in IPA and water

mixture resulted in a light yellow and transparent solution (Fig. S1). When the solubility of this

polymer was further investigated in various IPA:H2O ratios (Table S2), the highest solubility was

13
exhibited at 4:6 w/w IPA:water mixture. The decent solubility in a mixture of those solvents

indicates that the polarity of the mixture is more suitable for SPFBI, and thus it allows the

utilization of the alcohol-based SPFBI solution as an ionomer binder. Because IPA and water are

easier to remove by heating after spraying catalyst onto a PEM, this alcohol-based polymer

solution is quite promising for the catalyst ink preparation compared to other attempts that

prepared the hydrocarbon ionomer solution in DMSO or dimethylacetamide [40,41]. In this study,

we used not only Nafion but also SPFBI as an ionomer binder to test the PEMFC performance for

comparison.

3.2. Physicochemical properties of membrane.

When the SPFBI membrane with H+ form was obtained after casting and ion-exchange in

acidic solution, it appeared as a yellow and transparent thin film. The thermal stability of the

membranes was studied by thermogravimetric analysis (TGA), as shown in Fig. 2a. All

membranes were thermally stable until the first significant degradation from 200 oC, and they

exhibited 28 – 30 % weight loss at 700 oC. The weight loss behavior of SPFBI membranes was

more clearly demonstrated by differential thermogravimetric (DTG) curves in Fig. 2b. In general,

the degradation behaviors of the membranes were similar to one another, illustrating two main

stages corresponding to two peaks in DTG - the thermal decomposition of sulfonic acid groups in

alkyl side chains (from 200 to around 425 oC) and the disintegration of the polymer main chain

(from 475 oC) [5,42]. The magnitude of DTG peaks was different among three membranes,

indicating more or less sulfonic acid group degradation compared to the backbone degradation.

This result was well consistent with the difference in sulfonation degree among the three polymers.

Overall, the high decomposition temperature of the membranes demonstrates that the thermal

14
stability of SPFBI PEMs satisfies the practical operation requirement of PEMFCs.

Fig. 2. (a) TGA curves and (b) DTG curves of SPFBI membranes in nitrogen atmosphere.

Ion exchange capacity (IEC) is a key parameter to evaluate the hydration properties and

proton-conducting ability of PEMs. For SPFBI membranes, we determined IEC values from both

polymer structure based theoretical and the acid-base titration-based experimental methods. As

listed in Table S3, the experimental IEC values followed a consistent order with theoretical ones,

and as the higher sulfonic acid content increased IEC, the SPFBI-0.4 membrane exhibited the

highest value of 3.25.

The water uptake of SPFBI membranes was gravimetrically measured in the temperature

range of 30 – 80 oC as presented in Fig. 3a. In general, water uptake behavior of SPFBI membranes

followed two major trends. Firstly, the water uptake of all membranes increased with temperature,

because higher temperature increases polymer free volume and polymer - water interaction [43].

Secondly, the sulfonic acid groups enhanced the hydrophilicity of PEMs, allowing the membrane

to absorb more water. As a result, the SPFBI-0.4 membrane showed the highest water uptake of

46.03 % compared to 41.69 % of SPFBI-0.2 at 80 oC. It is worth mentioning that the increase of

water content from 30 to 80 oC was not significant (less than 10 %), due to the rigidity of SPFBI

15
backbone and high intrinsic viscosity as described [38]. Notably, SPFBI membranes synthesized

in this study exhibited relatively low water uptake compared to other PEMs, taking into account

very high IEC values that often lead to an excessive water content of PEM[27,44].

Fig. 3b shows the swelling ratio of SPFBI membranes. Similar to the water uptake behavior,

the membranes were more swollen at higher temperatures and higher sulfonation degrees. For

instance, while the swelling ratio of SPFBI-0.4 was 13.98 % at 30 oC, it was 16.12 % at 80 oC.

Moreover, taking the benefits from limited water uptake, the membranes were dimensionally quite

stable, showing the swelling ratio lower than 14 % at 30 oC. These relatively low water uptake and

swelling ratios of SPFBI membranes compared with the previously reported hydrocarbon

membranes [12,14,21], indicate a remarkable advantage of their dimensional stability for PEMFC

applications.

Fig. 3. (a) Water uptake, (b) swelling ratio of SPFBI membranes at different temperatures, (c), (d)

16
stress-strain curves of the membranes at fully hydrated and dry forms, respectively.

Mechanical properties of SPFBI membranes were analyzed by UTM and the resulting

stress-strain behavior is depicted in Fig. 3c and 3d. With increasing sulfonation degree, the

membrane became more flexible in the hydrated state and thus the maximum tensile strength

decreased (from 35.4 MPa of SPFBI-0.2 to around 30 MPa of SPFBI-0.4 membrane), while the

elongation at break opposingly increased (from 24.2 % of SPFBI-0.2 to 29.6 % of SPFBI-0.4

membrane). Despite the highest intrinsic viscosity, the SPFBI-0.4 membrane was less rigid under

humidified conditions than other membranes because of the plasticizing effect of high water uptake.

To remove the effect of hydration, we also conducted the mechanical property test for dry

membranes. As expected, the tensile strength of membranes increased following the order 0.2 <

0.3 < 0.4, which is a different trend from the fully humidified test results. In the dry state, the

higher content of fluorene backbone contributes to the mechanical durability of the membranes,

even though it was outweighed by the effect of water uptake in the hydrated state. Due to the

highest tensile strength under humidified conditions, SPFBI-0.2 PEM was considered for use in

long-term stability test, as described later in this study.

3.3. Proton conductivity.

The proton conductivity of SPFBI membranes was measured in-plane at 95% RH and

temperatures between 30 and 80 oC, as depicted in Fig. 4a. We also measured the conductivity of

Nafion membrane for comparison. In general, as the hopping and/or vehicle movement of protons

is thermally activated in more hydrated state of the membranes, the proton conductivity of all

membranes gradually increased with temperature. Also, the membranes with higher IEC showed

17
higher proton conductivity; for instance, the SPFBI-0.4 membrane with the highest IEC exhibited

the highest proton conductivity of 0.121 S cm-1 at 30 oC and 0.1956 S cm-1 at 80 oC. This result

supports the positive effect of high sulfonic acid content on the proton conduction. Notably, the

SPFBI-0.3 and SPFBI-0.4 membranes outperformed the Nafion membrane in terms of proton

conductivity, and even the SPFBI-0.2 membrane reached a proton conductivity from 0.063 to

0.105 S cm-1, which is comparable to that of Nafion membrane.

Fig. 4. (a) Proton conductivity of Nafion and SPFBI membranes at different temperatures, (b) the

corresponding Arrhenius plots of SPFBI membranes, (c) comparison of proton conductivity

correlating with IEC value and in-plane swelling ratio for various reported PEMs (details of test

18
conditions are summarized in Table 1).

To obtain more intensive information on the proton conduction behavior of SPFBI PEMs,

the activation energy (Ea) was calculated from the Arrhenius plots relating ln σ and T-1 (Fig. 4b).

As the gradual decrease in Ea was observed from 9.87 kJ mol-1 of SPFBI-0.2 to 8.67 kJ mol-1 of

SPFBI-0.4, it demonstrates that protons are more feasibly transported when the sulfonic acid

density increases. This activation energy behavior implies that the proton conduction behavior in

SPFBI membranes complies with the well-known Grotthuss and vehicular mechanisms [45].

Because the water uptake increased with the sulfonation degree as described, the expanded ionic

cluster dimension provided lower energy barriers for proton transport. Besides, the large amount

of sulfonic acids are beneficial for proton conduction via Grotthuss mechanism by shortening the

distance between acid groups for proton hopping [46,47]. In summary, as the increase of sulfonic

acid density reduces the energy barrier of proton-transfer pathways, the SPFBI-0.4 membrane

showed the highest conductivity. For comparison, we summarized the proton conductivity of the

0.4 membrane and various commercial and lab-synthesized hydrocarbon-based PEMs as a

function of swelling ratio and IEC as three-dimensionally illustrated in Fig. 4c. Table 1 provides

a more detailed comparison, including test results, test conditions, and polymer types. In general,

the SPFBI-0.4 membrane illustrated an excellent proton conductivity under relatively low swelling

when it was placed alongside the non-composite PEMs such as Nafion and Pemion.

Table 1. Summary of proton conductivity, IECex and swelling ratio of SPFBI-04 membrane and

some typical commercial and lab-synthesized PEMs.

19
IECex In-plane SRb
PEM σa (S cm-1) Polymer type Ref.
(meq g-1) (%)

SPFBI-0.4 0.196 ± 0.005 3.25 ± 0.31 16.12 ± 0.16 Ether-free This work
Nafion 0.121 ±0.007 0.92 ± 0.24 6.74 ± 0.09 Perfluorinated This work
Pemion
~ 0.265 3.28 ± 0.06 44.87c Ether-free [27]
(sPPN-H+)
Pemion
~ 0.170 3.19 ± 0.05 33.3c Ether-free [27]
(sPPB-H+)
SPPBP 40 0.162 2.45 10.8 Ether-free [48]
SIPiP 0.0975 2.50 11.5 Ether-free [49]
SPAES160 ~ 0.150 2.10 ± 0.24 19.57 Ether-containing [14]
SPEEK 0.170 1.654 51.3d Ether-containing [50]
SPFAE 0.1507 1.62 5.2 Ether-containing [51]
a
Measured at 80 oC, RH varies from 90 – 100%
b
Measured at 80 oC in water
c
Mathematically calculated from reported area swelling ratios, assuming that the sample shape
was square.
d
Measured at 60 oC and 100% RH

3.4. Oxidative stability.

The oxidative stability of SPFBI membranes was investigated by subjecting the membrane

samples to Fentons’s test where the chemical degradation of membrane was stimulated by •OH

radical generation. The membranes were treated with Fenton’s solution (3 wt% H2O2, 3 ppm Fe2+)

for 1 h and 5 h. During the test, membrane samples were gradually dissolved in the solution due

to polymer decomposition under ·OH radical bombardment. The residual weights of the samples

were determined and listed in Table 2, indicating how much the membrane retained after the test.

While the SPFBI-0.2 sample exhibited the highest residual weight of 98.95 % after 1 h and 43.33%

after 5 h, the SPFBI sample lost 5.73% and 81.73% of its original weight after 1 h and 5 h treatment,

20
respectively. For all samples, no significant mass loss and visual changes were observable after

the exposure for 1 h. As the oxidative stability of the membrane decreased with increasing

sulfonation degree, the ether-free phenylene backbone might be less prone to radical attack than

highly reactive -SO3H groups as reported in several studies [14,27,52]. To more precisely evaluate

the oxidative stability of SPFBI membranes, their Fenton’s test results were compared with those

of other hydrocarbon PEMs obtained under similar test conditions in Table S4 and S5. For 1 h test,

the stability of all SPFBI PEMs surpasses that of the common ether-containing hydrocarbon PEMs

even under harsher conditions. Even though Pemion illustrated better oxidative stability than

SPFBI PEMs for 1 h test, their residual weight difference was not significant. For 5 h test, although

the stability of PEMs varied with the sulfonation degree, SPFBI PEMs showed quite good

oxidative stability that could compete with Pemion, a state-of-the-art ether-free PEM. It is

necessary to mention that these 5 h Fenton’s test results under the similar experimental conditions

have been barely reported for other hydrocarbon PEMs because it often led to the entire

decomposition of membrane samples [46]. The improved chemical stability of this ether-free

SPFBI membrane compared to the ether-containing PEMs can be explained by the different bond

dissociation energy between aryl-aryl (C6H5-C6H5) segment (476 ± 6 kJ mol-1) and aryl-ether-aryl

(C6H5-O-C6H5) segments (355 ± 6 kJ mol-1)[53]. More importantly, the ether group creates

vulnerable sites on the neighboring aromatic rings due to the electron-donating effect [1,25,54], as

illustrated in Fig. S2. The attack of the oxidative radicals on those sites leads to the cleavage and

scission of the polymer backbone, resulting in poor chemical stability of the polymer membrane

in Fenton’s test.

Table 2. Residual weights of SPFBI samples after 1 h and 5 h Fenton’s test.

21
Sample RW(%) (1 h) RW(%) (5 h)

SPFBI-0.2 98.85 ± 1.25 43.33 ± 2.39


SPFBI-0.3 95.00 ± 0.52 29.07 ± 3.05
SPFBI-0.4 94.27 ± 0.40 18.27 ± 3.56

The proton conductivity test and FT-IR were employed to analyze the after-treatment

membrane samples. Because all membrane samples were almost intact after 1 h treatment, it was

available to measure the proton conductivity of SPFBI membranes with and without 1 h exposure

with Fenton’s solution for comparison (Fig. 5a). Despite slight conductivity losses due to the

chemical degradation, the membranes still exhibited a decent proton conductivity at various

temperatures, implying that most of the polymer structure and membrane morphology were not

significantly damaged. The recovered pieces of samples were also assessed by Fourier-transform

infrared (FT-IR) spectroscopy (Fig. 5b-d). For all samples, there were no serious changes in the

chemical structure after Fenton’s treatment. The IR bands originating from -SO3H groups in the

pristine membranes still can be clearly observed in the spectra of the membranes after treatment

for both 1 h and 5 h.

22
Fig. 5. (a) Proton conductivity of SPFBI membranes before and after 1 h Fenton’s test; FT-IR

spectra of (b) SPFBI-0.2, (c) SPFBI-0.3 and (d) SPFBI-0.4 membrane before and after 1 h and 5

h Fenton’s test.

3.5. PEMFC performance.

Compared with many other hydrocarbon polymers used to prepare PEMs, it was discovered

that SPFBIs have a decent solubility in a mixture of IPA and water, as previously presented. As

this behavior has been rarely reported from the hydrocarbon-based polymers for PEMs, this unique

solubility property of SPFBI was more detail exploited for its application as an ionomer binder as

well as PEM. As presented, SPFBI-0.4 showed the highest proton conductivity, compared to

SPFBI-0.2 and SPFBI-0.3 PEMs. When the size of particles agglomerated in catalyst ink was

analyzed using the dynamic light scattering (DLS) method, SPFBI-0.4 was more compatible with

23
Pt/C catalyst than the other two SPFBI samples (Fig. S3) [55]. Accordingly, taking into account

both membrane and ionomer properties, SPFBI-0.4 was chosen to prepare the ionomer solution,

along with the conventional Nafion ionomer as a reference. Using these ionomer binders, we

fabricated MEAs consisting of SPFBI and Nafion membranes for PEMFC operation.

Fig. 6. (a) H2/O2 fuel cell performances of different MEAs using Nafion ionomer binder, (b)

corresponding Nyquist plots of MEAs using Nafion ionomer binder. (Test conditions: 80 oC, 100%

RH, and gas inlet rate of 300 mL min-1 at both anode and cathode).

First, we compared the cell performance of MEAs of SPFBI and Nafion PEMs fabricated

using the same ionomer binder. Fig. 6a shows the polarization curves of MEAs using the Nafion

binder at 80 oC under the fully humidified H2/O2 supply. As the thickness of membranes was

controlled by casting process, its difference among membranes was less than 3 µm as shown in

Table 3 to minimize its effect on cell performance. While the maximum power density of SPFBI-

0.2 MEA was quite similar to that of Nafion MEA, that of SPFBI-0.3 and SPFBI-0.4 MEAs

exhibited higher maximum power densities of 346 and 365 mW cm-2, respectively. This

performance improvement can be attributed to the superior proton conductivity of SPFBI

membranes. To further understand the effect of types of PEMs on PEMFC performance, the

24
electrochemical impedance spectroscopy (EIS) measurements were performed to obtain the

corresponding Nyquist plots for each MEA, as shown in Fig. 6b. All MEAs showed a semicircular

pattern of Nyquist plots, in which the high-frequency intercept (left intersection of the real axis)

provided the electrolyte resistance (Rel), and the semicircle diameter provided the charge-transfer

resistance (Rct) of MEA [56]. It can be seen that the differences in Rct among MEAs were small

because of the utilization of the same catalyst and ionomer binder for all MEAs. On the other hand,

the Rel values of MEAs varied quite a lot, following the order of SPFBI-0.4 < SPFBI-0.3 < Nafion

< SPFBI-0.2, which is in compliance with the cell performance results. These results demonstrate

the benefit of SPFBI PEMs that reduce the membrane resistance and thus lead to the enhancement

of cell performance.

25
Fig. 7. Comparison of H2/O2 cell performance of the MEAs fabricated with different PEMs and

ionomer: (a) Nafion membrane, (b) SPFBI-0.2 membrane, (c) SPFBI-0.4 membrane. The

corresponding Nyquist plots for (d) Nafion membrane, (e) SPFBI-0.2 membrane and (f) SPFBI-

0.4 membrane. (Test conditions: 80 oC, 100% RH, and gas inlet rates of 300 mL min-1 at both anode

and cathode).

26
Table 3. Summary of power density and EIS for SPFBI and Nafion MEAs.

Membrane Electrode Assembly (MEA)


Maximum power
Rel (Ω cm2) Rct (Ω cm2)
density (mW cm-2)
Thickness
Membrane Ionomer binder
(µm)

Nafion 318 0.17373 0.39523


Nafion 90
SPFBI-0.4 294 0.18020 0.45080
Nafion 321 0.18121 0.43527
SPFBI-0.2 88
SPFBI-0.4 296 0.18682 0.48073
Nafion 346 0.13311 0.44089
SPFBI-0.3 87
SPFBI-0.4 - - -
Nafion 365 0.12421 0.43427
SPFBI-0.4 87
SPFBI-0.4 346 0.11330 0.47170

Next, PEMFC performance tests for Nafion, SPFBI-0.2 and SPFBI-0.4 MEAs using the

SPFBI-0.4 ionomer were also conducted. Their polarization curves were depicted along with those

of MEAs using the Nafion ionomer (Fig. 7a-c) and the corresponding EIS spectra (Fig. 7d-f) for

comparison. Additionally, some important experimental results including the maximum power

density, Rel and Rct, are summarized in Table 3. For all membranes, SPFBI-0.4 ionomer still

achieved decent cell performances, even though the Nafion ionomer showed better performances

with a 5 – 8% higher maximum power density. A similar conclusion can be drawn from the

comparison of Nyquist plots for MEAs fabricated with different ionomers. In each EIS curve,

although Rel was not significantly varied because the same types of PEMs were employed, Rct

increased when SPFBI-0.4 ionomer was used instead of Nafion. For instance, the MEA fabricated

using Nafion ionomer and Nafion PEM exhibited Rct of 0.39523 Ω cm2, which was lower than that

27
of the MEA using SPFBI-0.4 ionomer and Nafion PEM, 0.45080 Ω cm2. This difference probably

stems from the difference in proton transport in the catalyst layer affected by ionomer [8,55], as

the SPFBI-0.4 ionomer was less compatible with the catalyst than the conventional ionomer

according to DLS analysis. Those results, however, still point out that the SPFBI-0.4 ionomer can

be utilized to prepare MEAs for not only SPFBI membranes but also Nafion membrane to obtain

acceptable cell performance. It must also state that hydrocarbon ionomers like SPFBI may behave

differently from Nafion ionomer, thus other factors (e.g., catalysts, gas flow rate or ionomer

loading) can be further optimized to be more suitable for SPFBI ionomer usage. Therefore, the cell

performance of SPFBI-0.4 ionomer MEAs can be improved with more suited MEAs designs,

which would be the target of our future works. To the extent of our knowledge, the application of

low-cost hydrocarbon polymers as ionomer binder in PEMFC has been rarely reported, as it mostly

resulted in inferior cell performances to Nafion [57] and the solvent for preparation of ionomer

solution was mostly DMSO which is aprotic and thus not sufficient for fuel cell manufacture

[58,59]. As this result demonstrate the application of SPFBI-0.4 ionomer binder for high PEMFC

performance, it encourages further developments of hydrocarbon polymers not only for PEM but

also for ionomer binder to completely replace the perfluorinated polymers.

28
Fig. 8. OCV decay over time for various MEAs fabricated with different membrane/ionomer

binders. (Test conditions: 90 oC, 30% RH, and gas inlet rate: 300 mL min-1 at both anode and

cathode).

Finally, the OCV in-situ stability tests were conducted to assess the effect of the SPFBI

membrane and ionomer on the long-term durability of the MEA. Under the harsh temperature and

humidity conditions, the OCV decay of MEAs was plotted over test time, as shown in Fig. 8. The

end-of-life point of MEAs was chosen as the time when OCV drops to 0 V. For MEAs using Nafion

membrane, the voltage could not sustain over 220 h, as it completely dropped after around 210 -

220 h test. On the contrary, the MEAs using SPFBI-0.2 membrane exhibited a much lower OCV

decay rate, resulting in the end-of-life point over 620 h, much higher than those of Nafion

membrane-used MEAs. This in-situ cell test illustrated surprisingly better durability of SPFBI-0.2

than Nafion, being different from the Fenton’s test results. Athough Nafion PEMs are well-known

for their excellent stability against chemical degradation [60,61], this mismatched correlation

between ex-situ Fenton’s and in-situ cell tests has been stated in many reports, especially when the

Nafion MEAs decay faster than the hydrocarbon MEAs in OCV test [14,62]. This difference is

due to the additional factors provided by in-situ cell test that cannot be obtained from ex-situ

Fenton’s test. For instance, the fuel (hydrogen and oxygen) crossover is one of the important

factors to be additionally considered in this in-situ test. Due to higher gas permeability of Nafion

than hydrocarbon membranes, the generation rate of OH radicals is much more higher in the

Nafion based MEA system than the hydrocarbon based one during cell operation. This provides

much more complex and preferential radical attacking situations for the chemical degradation of

membranes than the simple Fenton’s test conducted in the solution with the fixed OH radical

29
concentration [20,24]. Therefore, while the Fenton’s test provided good information about the

chemical degradation behavior of SPFBI membrane itself, the OCV test demonstrated much more

reliable durability of the SPFBI membrane in practical MEA operation.

The optical images of MEAs before and after the stability test were collected and shown in

Fig. S4. While the SPFBI-0.2 membrane-used MEAs have no observable defects even after over

620 h, some pinholes were found on the Nafion membrane-used MEAs after the test, indicating

severe damages caused by chemical degradation. This is in good accordance with the OCV decay

rate of the MEAs, demonstrating the excellent in-situ chemical stability of ether-free SPFBI-0.2

membrane. Besides, compared with Nafion ionomer, SPFBI-0.4 ionomer did not significantly

affect the OCV decay for both Nafion and SPFBI-0.2 membrane-based MEAs, since the MEAs

consisted of SPFBI-0.2 membrane and SPFBI-0.4 ionomer still illustrated an end-of-life point over

620 h. Overall, the in-situ OCV test results provide the application of SPFBI membrane and

ionomer for the long-term operation of PEMFC.

4. Conclusions.

To synthesize the ether-free polymer for the fabrication of high chemical stability proton

exchange membrane, we proposed a novel polymer synthesis via the superacid-catalyzed

polymerization of common and low-cost monomers. As a result, a series of ether-free sulfonated

poly(fluorene biphenyl indole) (SPFBI) were successfully synthesized with various sulfonation

degrees. Experimental results revealed that the excellent proton conductivity of SPFBI membranes

up to 0.1956 S cm-1 at 80 oC was observed under the low water uptake and swelling ratio compared

to other reported membranes. Owing to the free structure of SPFBI, its oxidative stability was

comparable to that of Pemion - the state-of-the-art ether-free polymer lectrolyte membrane and

30
superior to many common hydrocarbon membranes. Due to the excellent proton conductivity and

chemical stability, SPFBI-0.3 and SPFBI-0.4 membranes exhibited better cell performances than

the Nafion membrane, showing the sustainability of more than 620 h of the open-circuit voltage

test. More importantly, SPFBI-0.4 was possibly employed to prepare the ionomer binder due to

the solubility in the isopropanol and water mixture. Despite minor losses in the maximum power

density, the performance and stability of membrane electrode assemblies fabricated with SPFBI-

0.4 ionomer showed the potential promises for research on hydrocarbon ionomer binders. We

believe that the results in this study merit the further development of hydrocarbon-based ether-free

polymer lectrolyte membranes and ionomer binders for the fabrication of cost-effective and long-

term durable membrane electrode assembly for fuel cells.

Acknowledgments

This work was sponsored by the National Research Foundation of Korea Grant, funded by

the Korean Government (MEST) (NRF 2018M3D1A1058624).

31
References

[1] A. Kraytsberg, Y. Ein-Eli, Review of Advanced Materials for Proton Exchange Membrane

Fuel Cells, Energy Fuels. 28 (2014) 7303–7330. https://doi.org/10.1021/ef501977k.

[2] S.J. Peighambardoust, S. Rowshanzamir, M. Amjadi, Review of the proton exchange

membranes for fuel cell applications, International Journal of Hydrogen Energy. 35 (2010) 9349–

9384. https://doi.org/10.1016/j.ijhydene.2010.05.017.

[3] Y. Wang, D.F. Ruiz Diaz, K.S. Chen, Z. Wang, X.C. Adroher, Materials, technological status,

and fundamentals of PEM fuel cells – A review, Materials Today. 32 (2020) 178–203.

https://doi.org/10.1016/j.mattod.2019.06.005.

[4] Y. Prykhodko, K. Fatyeyeva, L. Hespel, S. Marais, Progress in hybrid composite Nafion®-

based membranes for proton exchange fuel cell application, Chemical Engineering Journal. 409

(2021) 127329. https://doi.org/10.1016/j.cej.2020.127329.

[5] M. Vinothkannan, S. Ramakrishnan, A.R. Kim, H.-K. Lee, D.J. Yoo, Ceria Stabilized by

Titanium Carbide as a Sustainable Filler in the Nafion Matrix Improves the Mechanical Integrity,

Electrochemical Durability, and Hydrogen Impermeability of Proton-Exchange Membrane Fuel

Cells: Effects of the Filler Content, ACS Appl. Mater. Interfaces. 12 (2020) 5704–5716.

https://doi.org/10.1021/acsami.9b18059.

[6] L. Zhang, S.-R. Chae, Z. Hendren, J.-S. Park, M.R. Wiesner, Recent advances in proton

exchange membranes for fuel cell applications, Chemical Engineering Journal. 204–206 (2012)

87–97. https://doi.org/10.1016/j.cej.2012.07.103.

[7] M.B. Karimi, F. Mohammadi, K. Hooshyari, Recent approaches to improve Nafion

performance for fuel cell applications: A review, International Journal of Hydrogen Energy. 44

(2019) 28919–28938. https://doi.org/10.1016/j.ijhydene.2019.09.096.

32
[8] J. Choi, J.H. Yeon, S.H. Yook, S. Shin, J.Y. Kim, M. Choi, S. Jang, Multifunctional

Nafion/CeO2 Dendritic Structures for Enhanced Durability and Performance of Polymer

Electrolyte Membrane Fuel Cells, ACS Appl. Mater. Interfaces. 13 (2021) 806–815.

https://doi.org/10.1021/acsami.0c21176.

[9] K.-D. Kreuer, Ion Conducting Membranes for Fuel Cells and other Electrochemical Devices,

Chem. Mater. 26 (2014) 361–380. https://doi.org/10.1021/cm402742u.

[10] M. Feng, R. Qu, Z. Wei, L. Wang, P. Sun, Z. Wang, Characterization of the thermolysis

products of Nafion membrane: A potential source of perfluorinated compounds in the environment,

Scientific Reports. 5 (2015) 9859. https://doi.org/10.1038/srep09859.

[11] M.J. Parnian, S. Rowshanzamir, J. Alipour Moghaddam, Investigation of physicochemical

and electrochemical properties of recast Nafion nanocomposite membranes using different loading

of zirconia nanoparticles for proton exchange membrane fuel cell applications, Materials Science

for Energy Technologies. 1 (2018) 146–154. https://doi.org/10.1016/j.mset.2018.06.008.

[12] T. Miyahara, T. Hayano, S. Matsuno, M. Watanabe, K. Miyatake, Sulfonated

Polybenzophenone/Poly(arylene ether) Block Copolymer Membranes for Fuel Cell Applications,

ACS Appl. Mater. Interfaces. 4 (2012) 2881–2884. https://doi.org/10.1021/am300821v.

[13] R.P. Pandey, A.K. Thakur, V.K. Shahi, Sulfonated Polyimide/Acid-Functionalized Graphene

Oxide Composite Polymer Electrolyte Membranes with Improved Proton Conductivity and Water-

Retention Properties, ACS Appl. Mater. Interfaces. 6 (2014) 16993–17002.

https://doi.org/10.1021/am504597a.

[14] V.D.C. Tinh, V.D. Thuc, D. Kim, Chemically sustainable fuel cells via layer-by-layer

fabrication of sulfonated poly(arylene ether sulfone) membranes containing cerium oxide

nanoparticles, Journal of Membrane Science. (2021) 119430.

33
https://doi.org/10.1016/j.memsci.2021.119430.

[15] C. Wang, Y. Zhou, B. Shen, X. Zhao, J. Li, Q. Ren, Proton-conducting poly(ether sulfone

ketone)s containing a high density of pendant sulfonic groups by a convenient and mild post-

sulfonation, Polymer Chemistry. 9 (2018) 4984–4993. https://doi.org/10.1039/c8py00996a.

[16] S. Chandra Sutradhar, M.M. Rahman, F. Ahmed, T. Ryu, Jin lei, S. Yoon, S. Lee, Y. Jin, W.

Kim, Improved proton conductive membranes from poly(phenylenebenzophenone)s with pendant

sulfonyl imide acid groups for fuel cells, Journal of Power Sources. 442 (2019) 227233.

https://doi.org/10.1016/j.jpowsour.2019.227233.

[17] H. Nguyen, F. Lombeck, C. Schwarz, P.A. Heizmann, M. Adamski, H.-F. Lee, B. Britton, S.

Holdcroft, S. Vierrath, M. Breitwieser, Hydrocarbon-based PemionTM proton exchange membrane

fuel cells with state-of-the-art performance, Sustainable Energy Fuels. 5 (2021) 3687–3699.

https://doi.org/10.1039/D1SE00556A.

[18] K. Oh, K. Ketpang, H. Kim, S. Shanmugam, Synthesis of sulfonated poly(arylene ether

ketone) block copolymers for proton exchange membrane fuel cells, Journal of Membrane Science.

507 (2016) 135–142. https://doi.org/10.1016/j.memsci.2016.02.027.

[19] J. Zhang, H. Zhang, J. Wu, J. Zhang, Chapter 11 - Fuel Cell Degradation and Failure Analysis,

in: J. Zhang, H. Zhang, J. Wu, J. Zhang (Eds.), Pem Fuel Cell Testing and Diagnosis, Elsevier,

Amsterdam, 2013: pp. 283–335. https://doi.org/10.1016/B978-0-444-53688-4.00011-5.

[20] C. Gittleman, Membrane Durability: Physical and Chemical Degradation, in: 2011: pp. 15–

88.

[21] S. Yang, Y. Ahn, D. Kim, Poly(arylene ether ketone) proton exchange membranes grafted

with long aliphatic pendant sulfonated groups for vanadium redox flow batteries, J. Mater. Chem.

A. 5 (2017) 2261–2270. https://doi.org/10.1039/C6TA07456A.

34
[22] Z. Rui, J. Liu, Understanding of free radical scavengers used in highly durable proton

exchange membranes, Progress in Natural Science: Materials International. 30 (2020) 732–742.

https://doi.org/10.1016/j.pnsc.2020.08.013.

[23] V.D. Cong Tinh, D. Kim, Enhancement of oxidative stability of PEM fuel cell by introduction

of HO• radical scavenger in Nafion ionomer, Journal of Membrane Science. 613 (2020) 118517.

https://doi.org/10.1016/j.memsci.2020.118517.

[24] R. Borup, J. Meyers, B. Pivovar, Y.S. Kim, R. Mukundan, N. Garland, D. Myers, M. Wilson,

F. Garzon, D. Wood, P. Zelenay, K. More, K. Stroh, T. Zawodzinski, J. Boncella, J.E. McGrath, M.

Inaba, K. Miyatake, M. Hori, K. Ota, Z. Ogumi, S. Miyata, A. Nishikata, Z. Siroma, Y. Uchimoto,

K. Yasuda, K. Kimijima, N. Iwashita, Scientific Aspects of Polymer Electrolyte Fuel Cell

Durability and Degradation, Chem. Rev. 107 (2007) 3904–3951.

https://doi.org/10.1021/cr050182l.

[25] M. Adamski, N. Peressin, S. Holdcroft, On the evolution of sulfonated polyphenylenes as

proton exchange membranes for fuel cells, Mater. Adv. 2 (2021) 4966–5005.

https://doi.org/10.1039/D1MA00511A.

[26] Miyake Junpei, Taki Ryunosuke, Mochizuki Takashi, Shimizu Ryo, Akiyama Ryo, Uchida

Makoto, Miyatake Kenji, Design of flexible polyphenylene proton-conducting membrane for next-

generation fuel cells, Science Advances. 3 (n.d.) eaao0476. https://doi.org/10.1126/sciadv.aao0476.

[27] M. Adamski, T.J.G. Skalski, B. Britton, T.J. Peckham, L. Metzler, S. Holdcroft, Highly Stable,

Low Gas Crossover, Proton-Conducting Phenylated Polyphenylenes, Angewandte Chemie

International Edition. 56 (2017) 9058–9061. https://doi.org/10.1002/anie.201703916.

[28] T.J.G. Skalski, M. Adamski, B. Britton, E.M. Schibli, T.J. Peckham, T. Weissbach, T.

Moshisuki, S. Lyonnard, B.J. Frisken, S. Holdcroft, Sulfophenylated Terphenylene Copolymer

35
Membranes and Ionomers, ChemSusChem. 11 (2018) 4033–4043.

https://doi.org/10.1002/cssc.201801965.

[29] T.J.G. Skalski, B. Britton, T.J. Peckham, S. Holdcroft, Structurally-Defined, Sulfo-

Phenylated, Oligophenylenes and Polyphenylenes, J. Am. Chem. Soc. 137 (2015) 12223–12226.

https://doi.org/10.1021/jacs.5b07865.

[30] D.R. Nieto, S. Fomine, M.G. Zolotukhin, L. Fomina, M. del C.G. Hernandez,

Superelectrophilic Activation of N-Substituted Isatins: Implications for Polymer Synthesis, a

Theoretical Study, Macromolecular Theory and Simulations. 18 (2009) 138–144.

https://doi.org/10.1002/mats.200800075.

[31] M.C.G. Hernandez, M.G. Zolotukhin, S. Fomine, G. Cedillo, S.L. Morales, N. Fröhlich, E.

Preis, U. Scherf, M. Salmón, M.I. Chávez, J. Cárdenas, A. Ruiz-Trevino, Novel, Metal-Free,

Superacid-Catalyzed “Click” Reactions of Isatins with Linear, Nonactivated, Multiring Aromatic

Hydrocarbons, Macromolecules. 43 (2010) 6968–6979. https://doi.org/10.1021/ma101048z.

[32] R. Ren, S. Zhang, H.A. Miller, F. Vizza, J.R. Varcoe, Q. He, Facile Preparation of an Ether-

Free Anion Exchange Membrane with Pendant Cyclic Quaternary Ammonium Groups, ACS Appl.

Energy Mater. 2 (2019) 4576–4581. https://doi.org/10.1021/acsaem.9b00674.

[33] J. Wang, Y. Zhao, B.P. Setzler, S. Rojas-Carbonell, C. Ben Yehuda, A. Amel, M. Page, L.

Wang, K. Hu, L. Shi, S. Gottesfeld, B. Xu, Y. Yan, Poly(aryl piperidinium) membranes and

ionomers for hydroxide exchange membrane fuel cells, Nature Energy. 4 (2019) 392–398.

https://doi.org/10.1038/s41560-019-0372-8.

[34] H. Nederstedt, P. Jannasch, Poly(p-terphenyl alkylene)s grafted with highly acidic sulfonated

polypentafluorostyrene side chains for proton exchange membranes, Journal of Membrane Science.

647 (2022) 120270. https://doi.org/10.1016/j.memsci.2022.120270.

36
[35] T. Ryu, H. Jang, F. Ahmed, N.S. Lopa, H. Yang, S. Yoon, I. Choi, W. Kim, Synthesis and

characterization of polymer electrolyte membrane containing methylisatin moiety by

polyhydroalkylation for fuel cell, International Journal of Hydrogen Energy. 43 (2018) 5398–5404.

https://doi.org/10.1016/j.ijhydene.2017.12.164.

[36] Y. Cui, B. Xu, B. Yang, H. Yao, S. Li, J. Hou, A Novel pH Neutral Self-Doped Polymer for

Anode Interfacial Layer in Efficient Polymer Solar Cells, Macromolecules. 49 (2016) 8126–8133.

https://doi.org/10.1021/acs.macromol.6b01595.

[37] T. Shiraki, Y. Tsuchiya, T. Noguchi, S. Tamaru, N. Suzuki, M. Taguchi, M. Fujiki, S. Shinkai,

Creation of Circularly Polarized Luminescence from an Achiral Polyfluorene Derivative through

Complexation with Helix-Forming Polysaccharides: Importance of the meta-Linkage Chain for

Helix Formation, Chemistry – An Asian Journal. 9 (2014) 218–222.

https://doi.org/10.1002/asia.201301216.

[38] N. Chen, H.H. Wang, S.P. Kim, H.M. Kim, W.H. Lee, C. Hu, J.Y. Bae, E.S. Sim, Y.-C. Chung,

J.-H. Jang, S.J. Yoo, Y. Zhuang, Y.M. Lee, Poly(fluorenyl aryl piperidinium) membranes and

ionomers for anion exchange membrane fuel cells, Nature Communications. 12 (2021) 2367.

https://doi.org/10.1038/s41467-021-22612-3.

[39] Dilute Solution Thermodynamics, Molecular Weights, and Sizes, in: Introduction to Physical

Polymer Science, John Wiley & Sons, Ltd, 2005: pp. 71–143.

https://doi.org/10.1002/0471757128.ch3.

[40] T. Omata, M. Tanaka, K. Miyatake, M. Uchida, H. Uchida, M. Watanabe, Preparation and

Fuel Cell Performance of Catalyst Layers Using Sulfonated Polyimide Ionomers, ACS Appl. Mater.

Interfaces. 4 (2012) 730–737. https://doi.org/10.1021/am201360j.

[41] J.-S. Park, P. Krishnan, S.-H. Park, G.-G. Park, T.-H. Yang, W.-Y. Lee, C.-S. Kim, A study on

37
fabrication of sulfonated poly(ether ether ketone)-based membrane-electrode assemblies for

polymer electrolyte membrane fuel cells, Journal of Power Sources. 178 (2008) 642–650.

https://doi.org/10.1016/j.jpowsour.2007.08.008.

[42] S. Lee, Y. Lim, M.A. Hossain, H. Jang, Y. Jeon, S. Lee, L. Jin, W. Kim, Synthesis and

properties of grafting sulfonated polymer containing isatin by super acid-catalyzed

polyhydroxyalkylation reaction for PEMFC, Renewable Energy. 79 (2015) 72–77.

https://doi.org/10.1016/j.renene.2014.08.023.

[43] Glass-Rubber Transition Behavior, in: Introduction to Physical Polymer Science, John Wiley

& Sons, Ltd, 2005: pp. 349–425. https://doi.org/10.1002/0471757128.ch8.

[44] N. Peressin, M. Adamski, E.M. Schibli, E. Ye, B.J. Frisken, S. Holdcroft, Structure–Property

Relationships in Sterically Congested Proton-Conducting Poly(phenylene)s: the Impact of

Biphenyl Linearity, Macromolecules. 53 (2020) 3119–3138.

https://doi.org/10.1021/acs.macromol.0c00310.

[45] D.J. Kim, M.J. Jo, S.Y. Nam, A review of polymer–nanocomposite electrolyte membranes

for fuel cell application, Journal of Industrial and Engineering Chemistry. 21 (2015) 36–52.

https://doi.org/10.1016/j.jiec.2014.04.030.

[46] L. Liu, Y. Lu, Y. Pu, N. Li, Z. Hu, S. Chen, Highly sulfonated carbon nano-onions as an

excellent nanofiller for the fabrication of composite proton exchange membranes with enhanced

water retention and durability, Journal of Membrane Science. 640 (2021) 119823.

https://doi.org/10.1016/j.memsci.2021.119823.

[47] V.D. Thuc, V.D. Cong Tinh, D. Kim, Simultaneous improvement of proton conductivity and

chemical stability of Nafion membranes via embedment of surface-modified ceria nanoparticles

in membrane surface, Journal of Membrane Science. 642 (2022) 119990.

38
https://doi.org/10.1016/j.memsci.2021.119990.

[48] S.C. Sutradhar, H. Jang, N. Banik, J. Yoo, T. Ryu, H. Yang, S. Yoon, W. Kim, Synthesis and

characterization of proton exchange poly (phenylenebenzophenone)s membranes grafted with

propane sulfonic acid on pendant phenyl groups, International Journal of Hydrogen Energy. 42

(2017) 12749–12758. https://doi.org/10.1016/j.ijhydene.2016.11.054.

[49] L. Jin, Y. Lee, D. Kim, F. Ahmed, T. Ryu, W. Kim, H. Jang, Comparative study of chemically

different structured sulfonic acid and sulfonimide acid of Poly(isatine-phenylene) electrolyte for

PEMFC, International Journal of Hydrogen Energy. 46 (2021) 6762–6774.

https://doi.org/10.1016/j.ijhydene.2020.11.160.

[50] P. Wei, Y. Sui, X. Li, Q. Liu, B. Zhu, C. Cong, X. Meng, Q. Zhou, Sandwich-structure

PI/SPEEK/PI proton exchange membrane developed for achieving the high durability on excellent

proton conductivity and stability, Journal of Membrane Science. 644 (2022) 120116.

https://doi.org/10.1016/j.memsci.2021.120116.

[51] Y. Lu, Y. Liu, N. Li, Z. Hu, S. Chen, Sulfonated graphitic carbon nitride nanosheets as proton

conductor for constructing long-range ionic channels proton exchange membrane, Journal of

Membrane Science. 601 (2020) 117908. https://doi.org/10.1016/j.memsci.2020.117908.

[52] S.-W. Lee, Z.G. Abdi, J.-C. Chen, K.-H. Chen, Optimal method for preparing sulfonated

polyaryletherketones with high ion exchange capacity by acid-catalyzed crosslinking for proton

exchange membrane fuel cells, Journal of Polymer Science. 59 (2021) 706–720.

https://doi.org/10.1002/pol.20200872.

[53] Handbook of Bond Dissociation Energies in Organic Compounds By Yu-Ran Luo (University

of South Florida, St. Petersburg). CRC Press LLC: Boca Raton. 2003. xii + 380 pp. $159.95. ISBN

0-8493-1589-1., J. Am. Chem. Soc. 126 (2004) 982–982. https://doi.org/10.1021/ja0336224.

39
[54] J. Walkowiak-Kulikowska, J. Wolska, H. Koroniak, Polymers application in proton exchange

membranes for fuel cells (PEMFCs), 2 (2017). https://doi.org/10.1515/psr-2017-0018.

[55] J.E. Chae, S.J. Yoo, J.Y. Kim, J.H. Jang, S.Y. Lee, K.H. Song, H.-J. Kim, Hydrocarbon-based

electrode ionomer for proton exchange membrane fuel cells, International Journal of Hydrogen

Energy. 45 (2020) 32856–32864. https://doi.org/10.1016/j.ijhydene.2020.03.003.

[56] X.-Z. Yuan, C. Song, H. Wang, J. Zhang, eds., Impedance and its Corresponding

Electrochemical Processes, in: Electrochemical Impedance Spectroscopy in PEM Fuel Cells:

Fundamentals and Applications, Springer London, London, 2010: pp. 95–138.

https://doi.org/10.1007/978-1-84882-846-9_3.

[57] A. Strong, B. Britton, D. Edwards, T.J. Peckham, H.-F. Lee, W.Y. Huang, S. Holdcroft,

Alcohol-Soluble, Sulfonated Poly(arylene ether)s: Investigation of Hydrocarbon Ionomers for

Proton Exchange Membrane Fuel Cell Catalyst Layers, Journal of The Electrochemical Society.

162 (2015) F513–F518. https://doi.org/10.1149/2.0251506jes.

[58] S. Holdcroft, Fuel Cell Catalyst Layers: A Polymer Science Perspective, Chem. Mater. 26

(2014) 381–393. https://doi.org/10.1021/cm401445h.

[59] R. Sood, C. Iojoiu, E. Espuche, F. Gouanvé, G. Gebel, H. Mendil-Jakani, S. Lyonnard, J.

Jestin, Comparative Study of Proton Conducting Ionic Liquid Doped Nafion Membranes

Elaborated by Swelling and Casting Methods: Processing Conditions, Morphology, and Functional

Properties, J. Phys. Chem. C. 118 (2014) 14157–14168. https://doi.org/10.1021/jp502454m.

[60] H. Jang, T. Ryu, S.C. Sutradhar, F. Ahmed, K. Choi, H. Yang, S. Yoon, W. Kim, Studies of

sulfonated poly(phenylene)-block-poly(ethersulfone) for proton exchange membrane fuel cell,

International Journal of Hydrogen Energy. 42 (2017) 12768–12776.

https://doi.org/10.1016/j.ijhydene.2017.01.112.

40
[61] A.Z. Al Munsur, B.-H. Goo, Y. Kim, O.J. Kwon, S.Y. Paek, S.Y. Lee, H.-J. Kim, T.-H. Kim,

Nafion-Based Proton-Exchange Membranes Built on Cross-Linked Semi-Interpenetrating

Polymer Networks between Poly(acrylic acid) and Poly(vinyl alcohol), ACS Appl. Mater.

Interfaces. 13 (2021) 28188–28200. https://doi.org/10.1021/acsami.1c05662.

[62] R. Shimizu, J. Tsuji, N. Sato, J. Takano, S. Itami, M. Kusakabe, K. Miyatake, A. Iiyama, M.

Uchida, Durability and degradation analysis of hydrocarbon ionomer membranes in polymer

electrolyte fuel cells accelerated stress evaluation, Journal of Power Sources. 367 (2017) 63–71.

https://doi.org/10.1016/j.jpowsour.2017.09.025.

41
Table 1

Table 1. Summary of proton conductivity, IECex and swelling ratio of SPFBI-04 membrane and

some typical commercial and lab-synthesized PEMs.

a -1 IECex In-plane SRb


PEM σ (S cm ) Polymer type Ref.
(meq g-1) (%)

SPFBI-0.4 0.196 ± 0.005 3.25 ± 0.31 16.12 ± 0.16 Ether-free This work
Nafion 0.121 ±0.007 0.92 ± 0.24 6.74 ± 0.09 Perfluorinated This work
Pemion
~ 0.265 3.28 ± 0.06 44.87c Ether-free [27]
(sPPN-H+)
Pemion
~ 0.170 3.19 ± 0.05 33.3c Ether-free [27]
(sPPB-H+)
SPPBP 40 0.162 2.45 10.8 Ether-free [48]
SIPiP 0.0975 2.50 11.5 Ether-free [49]
SPAES160 ~ 0.150 2.10 ± 0.24 19.57 Ether-containing [14]
SPEEK 0.170 1.654 51.3d Ether-containing [50]
SPFAE 0.1507 1.62 5.2 Ether-containing [51]
a
Measured at 80 oC, RH varies from 90 – 100%
b
Measured at 80 oC in water
c
Mathematically calculated from reported area swelling ratios, assuming that the sample shape
was square.
d
Measured at 60 oC and 100% RH
Table 2

Table 2. Residual weights of SPFBI samples after 1 h and 5 h Fenton’s test.

Sample RW(%) (1 h) RW(%) (5 h)

SPFBI-0.2 98.85 ± 1.25 43.33 ± 2.39


SPFBI-0.3 95.00 ± 0.52 29.07 ± 3.05
SPFBI-0.4 94.27 ± 0.40 18.27 ± 3.56
Table 3

Table 3. Summary of power density and EIS for SPFBI and Nafion MEAs.

Membrane Electrode Assembly (MEA)


Maximum power
Rel (Ω cm2) Rct (Ω cm2)
density (mW cm-2)
Thickness
Membrane Ionomer binder
(µm)

Nafion 318 0.17373 0.39523


Nafion 90
SPFBI-0.4 294 0.18020 0.45080
Nafion 321 0.18121 0.43527
SPFBI-0.2 88
SPFBI-0.4 296 0.18682 0.48073
Nafion 346 0.13311 0.44089
SPFBI-0.3 87
SPFBI-0.4 - - -
Nafion 365 0.12421 0.43427
SPFBI-0.4 87
SPFBI-0.4 346 0.11330 0.47170
Fig. 1 Click here to access/download;Figure(s) - provided separately;Figure 1.tif
Fig. 2 Click here to access/download;Figure(s) - provided separately;Figure 2.tif
Fig. 3 Click here to access/download;Figure(s) - provided separately;Figure 3 revised.tif
Fig. 4 Click here to access/download;Figure(s) - provided separately;Figure 4 revised.tif
Fig. 5 Click here to access/download;Figure(s) - provided separately;Figure 5 revised.tif
Fig.6 Click here to access/download;Figure(s) - provided separately;Figure 6-revised2.tif
Fig. 7 Click here to access/download;Figure(s) - provided
separately;Figure 7-revised2.tif
Fig. 8 Click here to access/download;Figure(s) - provided separately;Figure 8.tif
Supplementary Materials

Click here to access/download


Supplementary Materials
Supplementary information-revised(no marked
changes).docx
Credit Author Statement

FORM FOR AUTHOR STATEMENT


Manuscript ID number: POWER-D-22-04425
Manuscript Title: Ether-free sulfonated poly(fluorene biphenyl indole) membranes and ionomer
binders for proton exchange membrane fuel cells.
Details of authors to be added to the manuscript:

Name Email address Organization name City/Coun


try
First author Vu Dong Thuc dongthucv@gmail.com Chemical Engineering/ Republic
Sungkyunkwan University of Korea
Second author Vo Dinh Cong vodinhcongtinh@gmail. Chemical Engineering/ Republic
Tinh com Sungkyunkwan University of Korea
Corresponding Dukjoon Kim djkim@skku.edu Chemical Engineering/ Republic
author Sungkyunkwan University of Korea

CrediT author statement:


Vu Dong Thuc: Conceptualization, Investigation, Writing-Original draft preparation, Writing-
Review & Editing.
Vo Dinh Cong Tinh: Conceptualization, Investigation.
Dukjoon Kim: Supervision, Resources, Funding acquisition, Writing-Review & Editing.
Highlights

Highlights
 A series of ether-free SPFBI was synthesized by superacid-catalyzed polymerization.
 Both SPFBI proton exchange membrane and ionomer binder were studied.
 Chemically stable SPFBI membrane reached the proton conductivity of 0.1956 S cm-1.
 In-situ stability test demonstrated high durability of SPFBI membrane (> 620 h).
Declaration of Interest Statement

Declaration of interests

☒The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:
Revised Cleaned Manuscript (with no marked changes) Click here to view linked References

1
2
3
4
5
6
Ether-free sulfonated poly(fluorene biphenyl indole) membranes and ionomer binders for
7
8 proton exchange membrane fuel cells
9
10 Vu Dong Thuca, Vo Dinh Cong Tinha, Dukjoon Kima*
11
12 a
13 School of Chemical Engineering, Sungkyunkwan University, 2066 Seobu-ro, Jangan-gu,
14
15 Suwon, Gyeonggi 16419, Republic of Korea
16
17 *
18 Email: djkim@skku.edu
19
20
21
22
23 Abstract
24
25 Because of the vulnerability of ether bonds to oxidative radicals, the ether-free polymer
26
27
electrolytes are synthesized to prepare chemically more stable proton exchange membranes for
28
29
30 long time durable fuel cell application. Here, we present a series of ether-free sulfonated
31
32 poly(fluorene biphenyl indole)s (SPFBI) synthesized by a feasible and cost-effective process.
33
34
35 Owing to the ether-free structure, SPFBI membranes display great chemical stability comparable
36
37 to the commercial Pemion membranes from Fenton’s test. SPFBI membranes exhibit high proton
38
39
40 conductivity up to 0.1956 S cm-1 at 80 oC under the limited swelling ratio lower than 16.5%.
41
42 Consequently, while the SPFBI-0.4 membrane shows the excellent cell performance reaching the
43
44
45
maximum power density of 365 mW cm-2, it also plays a decent role as an ionomer binder for both
46
47 SPFBI and Nafion membranes. In open circuit voltage stability test, the membrane electrode
48
49 assemblies using SPFBI-0.2 membrane can sustain more than 620 h without any severe defects at
50
51
52 90 oC under low relative humidity of 30%, outweighing the Nafion counterpart. This study
53
54 provides a remarkable approach for optimizing the ether-free hydrocarbon-based polymer
55
56
57 structure for the application of both polymer electrolyte membrane and ionomer binder with high
58
59 proton conductivity and chemical stability.
60
61
62 1
63
64
65
1
2
3
4
5
6
Keywords: Fuel cell; Proton exchange membrane; Ionomer binder; Chemical stability; Ether-free
7
8
9
10 1. Introduction.
11
12
13
14
To satisfy the growing demand for the green energy with no harm to the environment,
15
16 proton-exchange membrane fuel cell (PEMFC) has been recognized as one of the promising
17
18 electrochemical energy sources due to its eco-friendliness and high energy density [1–4]. As a key
19
20
21 component of PEMFC, proton-exchange membrane (PEM) needs to fulfill various requirements,
22
23 such as high mechanical and chemical stability and low gas crossover in addition to high proton
24
25
26 conductivity [5,6]. While many attempts have been launched to fabricate high-quality PEMs with
27
28 enhanced properties so far, the perfluorosulfonic acid-based membranes including Nafion are the
29
30
31 most successful and popular PEMs for commercial applications. Nafion possesses good
32
33 mechanical and oxidative stability as well as a decent proton conductivity in humidified conditions
34
35
36
due to the formation of ion clusters by well-separated hydrophobic and hydrophilic phases [4,7,8].
37
38 Despite those advantages, Nafion has major drawbacks, including high gas permeability, poor
39
40 water retention, and high cost [6,9]. Because of the perfluorinated backbone structure, Nafion also
41
42
43 causes environmental concerns in its disposal process [10,11].
44
45 In order to replace the perfluorinated polymers with cost and disposal ineffectiveness,
46
47
48 hydrocarbon-based polymers with acidic functional groups have been developed. Those generally
49
50 consist of the hydrophobic aromatic backbones functionalized with the proton-conducting moieties
51
52
53
such as sulfonic acid or phosphonic acid groups statistically distributed. Depending on the polymer
54
55 designs, these acid groups can be either directly attached to the polymer backbone [12,13], or
56
57 located on the long alkyl chains grafted onto the polymer backbone [14,15]. Therefore, similar to
58
59
60 the perfluorinated polymer membranes, the phase-separated hydrophilic domains in hydrocarbon-
61
62 2
63
64
65
1
2
3
4
5
6
based PEMs allow the construction of interconnected water channels to facilitate the proton
7
8 transport. With many attempts to improve membrane properties, the hydrocarbon-based PEMs
9
10 currently not only exhibit high proton conductivity and cell performance [16–18], but also offer
11
12
13 cost reduction compared to perfluorinated PEMs.
14
15 Despite the aforementioned advantages, several limitations of hydrocarbon-based PEMs
16
17
18 hinder their commercialization. A major concern is about their chemical stability, since the
19
20 polymer structure typically contains more reactive functional groups and thus possesses less
21
22
23 oxidative resistance than Nafion composed of an inert Teflon-like backbone [19,20]. In such
24
25 polymer structures, the ether bond is the most commonly encountered heteroatom bond in the main
26
27
chain, usually resulting from the polymerization reaction to link monomer molecules together
28
29
30 [14,21]. Unfortunately, ether groups are very prone to attacking of ·OH radicals continuously
31
32 generated from the formation of reactive oxygen species and fuel crossover during the cell
33
34
35 operation [22,23]. The low oxidative stability of ether group is because the oxygen in ether
36
37 increases the electron density of neighboring aryl rings via the electron-donating effect and thus
38
39
40 weakens the resistance against oxidative species [1,24,25]. Since the ether groups do not have
41
42 crucial role in PEM properties, their elimination from the polymer structure can be considered an
43
44
45
attractive way to enhance the chemical stability of PEMs. Accordingly, many studies have been
46
47 conducted to synthesize ether-free PEMs. For instance, Miyake et al. reported a flexible
48
49 polyphenylene PEM (SPP-QP) with excellent flexibility and stability [26]. Pemion is also a
50
51
52 commercially available ether-free hydrocarbon-based PEM [27,28]. The synthetic process of
53
54 polymer electrolytes without ether linkages is usually quite complicated with the requirement of
55
56
57 costly metallic catalysts [27–29]. The superacid-catalyzed polymerization reaction was employed
58
59 as a good method, characterized by a short reaction time without the requirement of heating and
60
61
62 3
63
64
65
1
2
3
4
5
6
expensive catalysts [30,31]. This method has been widely used to synthesize ether-free polymers
7
8 for anion exchange membrane (AEM) [32,33], but just a few publications have been reported for
9
10 PEMs [34,35]. As the ether-free PEMs synthesized from the previous studies showed only
11
12
13 mediocre properties without profound investigation of chemical stability enhancement, it is
14
15 difficult to draw the potential of its methodological effectiveness.
16
17
18 In this work, the synthesis and characterization of sulfonated poly(fluorene biphenyl indole)
19
20 (SPFBI) are reported as a novel ether-free polymer electrolyte with improved PEM properties.
21
22
23 Compared to the previous ether-free polymers synthesized by superacid-catalyzed reaction, this
24
25 polymer structure was rationally designed from common monomers to establish the mechanically
26
27
robust rigid aromatic backbone grafted with long sulfonated alkyl chains to improve proton
28
29
30 conductivity. The synthetic procedure of SPFBI involves two main steps, starting from the
31
32 synthesis of poly(fluorene biphenyl indole) (PFBI) from isatin, biphenyl, and fluorene under
33
34
35 superacid conditions [31]. In the second step, the sulfonated alkyl groups were grafted to PFBI
36
37 backbone using 1,3-propane sultone and tetrabutylammonium bromide catalyst under basic
38
39
40 conditions [36,37]. By varying the monomer ratios, a series of SPFBI PEMs with different sulfonic
41
42 acid contents were prepared for an intensive investigation on proton conductivity, oxidative
43
44
45
stability, and PEMFC performance. Additionally, the solubility of SPFBI in isopropanol and water
46
47 mixture allowed its utilization as the ionomer binder for the preparation of catalyst ink in
48
49 membrane electrode assembly (MEA) process.
50
51
52
53
54 2. Experimental Section.
55
56
57 2.1. Materials.
58
59 Fluorene, isatin, trifluoroacetic acid (TFA), trifluoromethanesulfonic acid (TFSA), 1,3
60
61
62 4
63
64
65
1
2
3
4
5
6
propane sultone (99.0%) and tetrabutylammonium bromide (TBAB) were purchased from Tokyo
7
8 Chemical Industry (TCI, Japan). Biphenyl, iron(II) sulfate heptahydrate (FeSO4.7H2O) and
9
10 hydrogen peroxide (H2O2) solution were obtained from Sigma-Aldrich (Milwaukee, WI, USA).
11
12
13 Nafion alcohol-based dispersion (5 wt%, EW1100) was supplied from Chemours (Wilmington,
14
15 DE, USA), as well as platinum (nominally 40% on carbon black) (Pt/C) was provided by Alfa
16
17
18 Aesar (Ward Hill, MA, USA) to prepare MEAs. Other common solvents (99.5% purity) were
19
20 obtained from Samchun Chemicals (Korea).
21
22
23
24
25 2.2. Synthesis of poly(fluorene biphenyl indole) (PFBI).
26
27
The synthetic process of PFBI is briefly represented in Scheme 1. In detail, isatin (150
28
29
30 mmol) was added to a three-neck round bottom flask, along with a total of 150 mmol of two other
31
32 monomers (biphenyl and fluorene). The ratio of biphenyl and fluorene was varied to control the
33
34
35 sulfonation degree of the final polymer product. Then, 15 mL of dichloromethane (CH2Cl2) was
36
37 poured into the flask to dissolve the monomers under magnetic stirring. Subsequently, TFA (10
38
39
40 mL) and TFSA (15 mL) were dropwise added, and the mixture was stirred and kept at 0-5 oC for
41
42 12 h. After the reaction was finished, the dark purple mixture was slowly poured into methanol to
43
44
45
obtain PFBI precipitation. The polymer was washed with methanol and water several times, and
46
47 then dried in a vacuum oven at 80 oC for 24 h. It is designated as PFBI-x, where x represents the
48
49 proportion of fluorene over the total amount of biphenyl + fluorene used for the synthesis. The
50
51
52 yield of the PFBI synthetic process was about 93-96%. 1H NMR (300 MHz, DMSO-d6): 10.78
53
54 ppm (Hi), 7.73 ppm (Hb), 7.55 ppm (Hc), 7.21 – 7.31 ppm (Hd,e,g), 6.93 – 7.19 ppm (Hf,h) and 3.74
55
56
57 ppm (Ha).
58
59
60
61
62 5
63
64
65
1
2
3
4
5
6
2.3. Synthesis of sulfonated poly(fluorene biphenyl indole) (SPFBI).
7
8 The second step in Scheme 1 illustrates the synthetic procedure of SPFBI. For each
9
10 synthesis, 3 g of PFBI and 0.116 g of TBAB were simultaneously dissolved in 70 mL of dimethyl
11
12
13 sulfoxide (DMSO) at 60 oC and then poured into a three-neck round bottom flask. Next, nitrogen
14
15 gas was continuously supplied into the flask to remove the oxygen, and 4 mL of 50 wt% KOH
16
17
18 was slowly added to the polymer solution by syringe, followed by adding an excess amount of 1,3-
19
20 propane sultone. The reaction mixture was stirred at 60 oC for 6 h, under a nitrogen atmosphere.
21
22
23 Afterward, the product was precipitated by dropwise adding into methanol, and then washed with
24
25 water and isopropyl alcohol (IPA) in sequence and dried. The obtained polymer is named SPFBI-
26
27
x, where x refers to the PFBI-x precursor. The yield of the sulfonation process is 85-88 %. 1H
28
29
30 NMR (300 MHz, DMSO-d6): 7.55 – 7.61 ppm (Hb,c), 7.21 – 7.40 ppm (Hd,e,g), 7.07 – 7.19 ppm
31
32 (Hf,h), 3.83 ppm (Hi), 2.34 – 2.46 ppm (Ha,l) and 1.89 ppm (Hk).
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 6
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39 Scheme 1. Two-step synthetic process of SPFBI-x.
40
41
42
43
44 2.4. Preparation of membrane and catalyst ink.
45
46 To prepare the SPFBI membrane, 0.3 g of SPFBI was completely dissolved in DMSO at
47
48
49 60 oC, and then the polymer solution was poured into a petri dish and placed in a vacuum oven at
50
51 80 oC for 24 h. After that, the polymer membrane was peeled off to be immersed in 1 M H2SO4
52
53
54
solution for 24 h to convert the sulfonated groups to H+ form. Finally, the membrane was washed
55
56 with deionized water for 24 h. For membrane comparison, Nafion PEMs were prepared by casting
57
58 method. Nafion alcohol-based dispersion (5 wt%) was poured into a flat petri dish, then dried at
59
60
61
62 7
63
64
65
1
2
3
4
5
6
80 oC for 24 h. The formed film was peeled off from the dish to obtain a Nafion pristine membrane.
7
8 The thickness of all fabricated membranes in this work was controlled within a range of 80-90 µm.
9
10 For catalyst ink preparation, 5 wt% solution of SPFBI-0.4 was prepared in-house by
11
12
13 acidifying the polymer solid in 1 M H2SO4 solution. After rinsing it with water, it was dissolved
14
15 in a mixture of IPA and water (4:6 w/w). For comparison, the preparation of catalyst ink using
16
17
18 SPFBI-0.4 ionomer followed a similar procedure as that of Nafion ionomer. Pt/C catalyst was
19
20 mixed with deionized water, IPA, and 5 wt% polymer solution (Nafion or SPFBI-0.4) and
21
22
23 subsequently ultra-sonicated to obtain a colloidal catalyst slurry. In all catalyst ink solutions, the
24
25 ionomer binder:catalyst mass ratio was fixed at 1:4.
26
27
28
29
30 2.5. Characterizations.
31
32 Chemical structure. To confirm the synthesis of the polymers, the proton nuclear
33
34
35 magnetic resonance (1H NMR) spectra were obtained by a Varian Oxford 300 MHz NMR
36
37 spectrometer (CA, USA) using DMSO-d6 as a solvent to dissolve polymer samples. The standard
38
39
40 chemical shift of DMSO-d6 is determined at 2.50 ppm.
41
42 Intrinsic viscosity. The intrinsic viscosity ([η]) of SPFBI was determined by a method
43
44
45
using a Ubbelohde viscometer. The polymers were dissolved in DMSO to obtain solutions with
46
47 five different concentrations, and the efflux time of each solution was recorded at 20 oC. By
48
49 drawing a plot of reduced viscosity versus concentration, we obtained [η] at the y-intercept value.
50
51
52 Solubility testing. The solubility of polymers was measured in various common organic
53
54 solvents, water and a mixture of IPA and water (4:6 w/w). The tests were performed at room
55
56
57 temperature and 60 oC (only for solvents with a high boiling point). The polymer was considered
58
59 insoluble in a certain solvent if it cannot be dissolved in both room and elevated temperatures.
60
61
62 8
63
64
65
1
2
3
4
5
6
Membrane hydration properties and IEC. The water uptake (WU) and swelling ratio
7
8 (SR) of SPFBI membranes were measured using the weight and length change of membrane
9
10 samples in fully dry and wet states at different temperatures. The membrane was dried at 80 oC for
11
12
13 48 h, and then cut into a 1 x 3 cm rectangular piece to measure the mass (md) and length (ld) of the
14
15 dry sample. Next, the sample was immersed in deionized water for 48 h under stirring. After that,
16
17
18 the sample was removed, the excess water on the membrane surface was wiped off with tissue
19
20 papers to measure the mass and length of the membrane at wet state (mw and lw). The test was
21
22
23 repeated 3 times for each sample to obtain the average value of WU and SR according to the
24
25 following equations:
26
27
28
𝑊𝑈(%) = (𝑚𝑤 − 𝑚𝑑 )/𝑚𝑑 × 100 (1)
29
30 𝑆𝑅(%) = (𝑙𝑤 − 𝑙𝑑 )/𝑙𝑑 × 100 (2)
31
32 The experimental ion exchange capacity (IECex) was measured by acid-base titration
33
34
35 method. A dry membrane sample with the mass m (g) was submerged in standard NaOH solution
36
37 with the known concentration and volume (CNaOH, VNaOH) for 48 h. The base solution after
38
39
40 removing the membrane was titrated with standard HCl solution (CHCl, VHCl) to determine IECex
41
42 from equation (3):
43
44 𝐶𝑁𝑎𝑂𝐻 𝑉𝑁𝑎𝑂𝐻 − 𝐶𝐻𝐶𝑙 𝑉𝐻𝐶𝑙
45 𝐼𝐸𝐶𝑒𝑥 (meg g −1 ) = × 1000 (3)
𝑚
46
47
48 For comparison, the theoretical ion exchange capacity (IECtheo) was also calculated from
49
50 the chemical structure of SPFBI, based on the number of acid groups and the theoretical molecular
51
52
53
weight of the repeating unit of polymer.
54
55 Thermal and mechanical properties. The thermal stability of the membrane in proton
56
57 form was investigated by thermogravimetric analysis (TGA) using a Seiko Exstar 6000 instrument
58
59
60 (Seiko Instruments, Japan). The temperature range was set from 100 to 700 oC at the heating rate
61
62 9
63
64
65
1
2
3
4
5
6
of 10 oC min-1 under nitrogen atmosphere. In terms of mechanical property, both dry and hydrated
7
8 membranes were cut into a 1 x 5 cm rectangular piece to measure their tensile stress vs. strain
9
10 behavior using a universal tensile machine (UTM, Model 5565, Lloyd, Fareham, UK) with a 250
11
12
13 N load cell.
14
15 Proton conductivity. Proton conductivity (σ, S cm-1) was measured in in-plane dimension
16
17
18 by an electrochemical workstation (ZIVE SP2, Wonatech, Korea) connected with a four-probe
19
20 single cell (Bekktech LLC, Loveland, CO, USA). Each membrane was cut into a 1 x 3 cm shape
21
22
23 to place inside the cell, and the test was repeated three times at 95% relative humidity (RH). The
24
25 testing temperature varied from 30 to 80 oC and the oxygen gas rate was 500 mL min-1. The proton
26
27
conductivity of the membranes was calculated from equation (4):
28
29
30 𝜎 (𝑆 𝑐𝑚−1 ) = 𝑙/(𝑅 × 𝑡 × 𝑤) (4)
31
32 where l is the distance between electrodes (cm), R is sample resistance (Ω), t and w are sample
33
34
35 thickness (cm) and width (cm), respectively.
36
37 Oxidative stability. Fenton’s reagent (an aqueous solution containing 3 wt% H2O2 and 3
38
39
40 ppm Fe2+) was used to study the oxidative stability of SPFBI membranes. The membrane sample
41
42 was dried, weighed, and then immersed in the Fenton solution at 80 oC for 1 to 5 h. After the test,
43
44
45 the sample was gently washed with water and dried in a vacuum oven at 80 oC. The residual weight
46
47 (RW) of the sample was calculated from the before-test weight (mb) and after-test weight (ma) of
48
49
the sample, using equation (5):
50
51
52 𝑅𝑊 (%) = (𝑚𝑎 /𝑚𝑏 ) × 100 (5)
53
54 After the test, the proton conductivity of the treated membrane was measured following
55
56
57 the aforementioned procedure. Moreover, the chemical structure of the membrane sample was
58
59 analyzed using Fourier-transform infrared (FT-IR) spectroscopy (Nicolet iS20, Thermo Scientific,
60
61
62 10
63
64
65
1
2
3
4
5
6
Waltham, MA, USA).
7
8 Catalyst particle size. The catalyst ink was prepared as described in Section 2.4, and then
9
10 IPA was added to dilute the ink two-fold. Subsequently, the resultant solution was ultra-sonicated
11
12
13 for 30 min and the particle size of the catalyst was identified by employing a dynamic light
14
15 scattering (DLS) analyzer (ELS-Z2000, Photal Otsuka Electronics, Japan).
16
17
18 PEMFC tests. Membrane electrode assembly (MEA) was prepared by spraying catalyst
19
20 ink onto both sides of the PEM surface. A vacuum plate controller system (CNL Energy, Korea)
21
22
23 and a metal mask were used to maintain the flatness of the membrane, and the temperature was
24
25 kept at 60 oC to completely remove moisture and solvents during the spraying process. The Pt
26
27
loading was fixed at 0.4 mg cm-2 on an active area of 5 cm2 for each side of the membrane. Next,
28
29
30 the MEA was assembled with two gas diffusion layers (GDL, SGL GDL 39BB, CNL Energy,
31
32 Korea, thickness ~325 µm) and then sealed with two Teflon gaskets. The formed stack was inserted
33
34
35 between two graphite plates of the cell test unit in a PEMFC station (CNL Energy, Korea) for in-
36
37 situ fuel cell tests.
38
39
40 The PEMFC performance tests were conduct at 80 oC and 100% RH. Fully humidified
41
42 hydrogen and oxygen gases were supplied at anode and cathode, respectively, at the same flow
43
44
45
rate of 300 mL min-1. This flow rate corresponds to the stoichiometry values of hydrogen (λH2) and
46
47 oxygen (λO2) of 8.6 and 17.2, respectively, at 1000 mA cm-2. Electrochemical impedance
48
49 spectroscopy (EIS) measurements were carried out at cell conditions as the cell performance test
50
51
52 by connecting the fuel cell stack with ZIVE SP2 electrochemical workstation (Wonatech, Korea).
53
54 The EIS tests used the potentiostatic mode in the frequency range of 0.1 Hz – 100 kHz at the
55
56
57 potential amplitude of 6 mV. Nyquist plots were constructed and fitted by ZIVE Smart Manager
58
59 software.
60
61
62 11
63
64
65
1
2
3
4
5
6
In-situ open-circuit voltage (OCV) test of MEAs was carried out at 90 oC and 30% RH to
7
8 accelerate the degradation of MEA. The hydrogen and oxygen gas inlet rates were the same at 300
9
10 mL min-1. The MEA stability was demonstrated, as the OCV decays as a function of operation
11
12
13 time.
14
15
16
17
18 3. Results and discussion.
19
20 3.1. Polymer properties.
21
22
23 In the present study, we synthesized a series of ether-free SPFBI by a simple two-step
24
25 procedure: the polymerization under superacid-catalyzed condition to form PFBI without ether
26
27
groups, followed by the sulfonation reaction between PFBI and 1,3-propane sultone in an alkaline
28
29
30 solution. As described in section 2.2. and 2.3, the synthetic process of SFPBI has its remarkable
31
32 advantages: facile and short synthesis time and heating requirement of no more than 60 oC. Three
33
34
35 SPFBI samples, SPFBI-0.2, 0.3 and 0.4 with different sulfonation degrees, were synthesized by
36
37 simply adjusting the fluorene/biphenyl feed ratio to fabricate PEMs.
38
39
40 The chemical structures of PFBI and SPFBI were confirmed by 1H NMR spectra in Fig. 1.
41
42 The SPFBI consists of a hydrophobic poly(phenylene indole) backbone grafted with long alkyl
43
44
45
sulfonated side chains. In terms of the polymer backbone, fluorene was rationally employed as a
46
47 monomer along with biphenyl to increase and regulate the sulfonation degree of the final product.
48
49 Fluorene has a higher rotation energy barrier than biphenyl because the carbon C9 restricts the
50
51
52 movement of phenyl rings, and thus fluorene molecule is more planar than biphenyl, resulting in
53
54 higher reactivity towards the electrophilic substitution reaction [38]. Therefore, the use of fluorene
55
56
57 is also expected to improve the molecular weight and rigidity of polymer backbone, which is
58
59 beneficial for physicochemical properties of SPFBI PEMs. As a result, SPFBI-0.2, 0.3 and 0.4
60
61
62 12
63
64
65
1
2
3
4
5
6
showed different intrinsic viscosities ([η]). The [η] value of SPFBI-0.2 was 5.35 dL g-1, lower than
7
8 6.18 and 6.36 dL g-1 of SPFBI-0.3 and SPFBI-0.4, respectively. The difference of [η] implies the
9
10 increase of polymer molecular weight (MW) from SPFBI-0.2 to 0.4, based on the relationship
11
12
13 between [η] and MW in the Mark–Houwink–Sakurada relationship [39]. This result might
14
15 correspond to the increase of fluorene segment in polymer structure from 0.2 to 0.4 which
16
17
18 improves the polymerization reactivity.
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36 Fig. 1. 1H NMR spectra of (a) PFBI-0.4 and (b) SPFBI-0.4.
37
38
39
40
41 The solubility of polymers in common solvents is summarized in Table S1. When we
42
43
44
increased the fluorene content to 0.5, SPFBI-0.5 became so hydrophilic that it was dissolved in
45
46 water. Hence, the maximum fluorene ratio was fixed at 0.4, ensuring that the membrane can retain
47
48 and operate in fuel cell environments. Although the SPFBI with a lower sulfonation degree was
49
50
51 insoluble in individual water and IPA, it was surprising that it showed a decent solubility in a
52
53 mixture of IPA and water even at the ambient condition. The dissolution of SPFBI in IPA and water
54
55
56 mixture resulted in a light yellow and transparent solution (Fig. S1). When the solubility of this
57
58 polymer was further investigated in various IPA:H2O ratios (Table S2), the highest solubility was
59
60
61
62 13
63
64
65
1
2
3
4
5
6
exhibited at 4:6 w/w IPA:water mixture. The decent solubility in a mixture of those solvents
7
8 indicates that the polarity of the mixture is more suitable for SPFBI, and thus it allows the
9
10 utilization of the alcohol-based SPFBI solution as an ionomer binder. Because IPA and water are
11
12
13 easier to remove by heating after spraying catalyst onto a PEM, this alcohol-based polymer
14
15 solution is quite promising for the catalyst ink preparation compared to other attempts that
16
17
18 prepared the hydrocarbon ionomer solution in DMSO or dimethylacetamide [40,41]. In this study,
19
20 we used not only Nafion but also SPFBI as an ionomer binder to test the PEMFC performance for
21
22
23 comparison.
24
25
26
27
3.2. Physicochemical properties of membrane.
28
29
30 When the SPFBI membrane with H+ form was obtained after casting and ion-exchange in
31
32 acidic solution, it appeared as a yellow and transparent thin film. The thermal stability of the
33
34
35 membranes was studied by thermogravimetric analysis (TGA), as shown in Fig. 2a. All
36
37 membranes were thermally stable until the first significant degradation from 200 oC, and they
38
39
40 exhibited 28 – 30 % weight loss at 700 oC. The weight loss behavior of SPFBI membranes was
41
42 more clearly demonstrated by differential thermogravimetric (DTG) curves in Fig. 2b. In general,
43
44
45
the degradation behaviors of the membranes were similar to one another, illustrating two main
46
47 stages corresponding to two peaks in DTG - the thermal decomposition of sulfonic acid groups in
48
49 alkyl side chains (from 200 to around 425 oC) and the disintegration of the polymer main chain
50
51
52 (from 475 oC) [5,42]. The magnitude of DTG peaks was different among three membranes,
53
54 indicating more or less sulfonic acid group degradation compared to the backbone degradation.
55
56
57 This result was well consistent with the difference in sulfonation degree among the three polymers.
58
59 Overall, the high decomposition temperature of the membranes demonstrates that the thermal
60
61
62 14
63
64
65
1
2
3
4
5
6
stability of SPFBI PEMs satisfies the practical operation requirement of PEMFCs.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22 Fig. 2. (a) TGA curves and (b) DTG curves of SPFBI membranes in nitrogen atmosphere.
23
24
25
26
27 Ion exchange capacity (IEC) is a key parameter to evaluate the hydration properties and
28
29 proton-conducting ability of PEMs. For SPFBI membranes, we determined IEC values from both
30
31
32
polymer structure based theoretical and the acid-base titration-based experimental methods. As
33
34 listed in Table S3, the experimental IEC values followed a consistent order with theoretical ones,
35
36 and as the higher sulfonic acid content increased IEC, the SPFBI-0.4 membrane exhibited the
37
38
39 highest value of 3.25.
40
41 The water uptake of SPFBI membranes was gravimetrically measured in the temperature
42
43
44 range of 30 – 80 oC as presented in Fig. 3a. In general, water uptake behavior of SPFBI membranes
45
46 followed two major trends. Firstly, the water uptake of all membranes increased with temperature,
47
48
49 because higher temperature increases polymer free volume and polymer - water interaction [43].
50
51 Secondly, the sulfonic acid groups enhanced the hydrophilicity of PEMs, allowing the membrane
52
53
to absorb more water. As a result, the SPFBI-0.4 membrane showed the highest water uptake of
54
55
56 46.03 % compared to 41.69 % of SPFBI-0.2 at 80 oC. It is worth mentioning that the increase of
57
58 water content from 30 to 80 oC was not significant (less than 10 %), due to the rigidity of SPFBI
59
60
61
62 15
63
64
65
1
2
3
4
5
6
backbone and high intrinsic viscosity as described [38]. Notably, SPFBI membranes synthesized
7
8 in this study exhibited relatively low water uptake compared to other PEMs, taking into account
9
10 very high IEC values that often lead to an excessive water content of PEM[27,44].
11
12
13 Fig. 3b shows the swelling ratio of SPFBI membranes. Similar to the water uptake behavior,
14
15 the membranes were more swollen at higher temperatures and higher sulfonation degrees. For
16
17
18 instance, while the swelling ratio of SPFBI-0.4 was 13.98 % at 30 oC, it was 16.12 % at 80 oC.
19
20 Moreover, taking the benefits from limited water uptake, the membranes were dimensionally quite
21
22
23 stable, showing the swelling ratio lower than 14 % at 30 oC. These relatively low water uptake and
24
25 swelling ratios of SPFBI membranes compared with the previously reported hydrocarbon
26
27
membranes [12,14,21], indicate a remarkable advantage of their dimensional stability for PEMFC
28
29
30 applications.
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
Fig. 3. (a) Water uptake, (b) swelling ratio of SPFBI membranes at different temperatures, (c), (d)
61
62 16
63
64
65
1
2
3
4
5
6
stress-strain curves of the membranes at fully hydrated and dry forms, respectively.
7
8
9
10 Mechanical properties of SPFBI membranes were analyzed by UTM and the resulting
11
12
13 stress-strain behavior is depicted in Fig. 3c and 3d. With increasing sulfonation degree, the
14
15 membrane became more flexible in the hydrated state and thus the maximum tensile strength
16
17
18 decreased (from 35.4 MPa of SPFBI-0.2 to around 30 MPa of SPFBI-0.4 membrane), while the
19
20 elongation at break opposingly increased (from 24.2 % of SPFBI-0.2 to 29.6 % of SPFBI-0.4
21
22
23 membrane). Despite the highest intrinsic viscosity, the SPFBI-0.4 membrane was less rigid under
24
25 humidified conditions than other membranes because of the plasticizing effect of high water uptake.
26
27
To remove the effect of hydration, we also conducted the mechanical property test for dry
28
29
30 membranes. As expected, the tensile strength of membranes increased following the order 0.2 <
31
32 0.3 < 0.4, which is a different trend from the fully humidified test results. In the dry state, the
33
34
35 higher content of fluorene backbone contributes to the mechanical durability of the membranes,
36
37 even though it was outweighed by the effect of water uptake in the hydrated state. Due to the
38
39
40 highest tensile strength under humidified conditions, SPFBI-0.2 PEM was considered for use in
41
42 long-term stability test, as described later in this study.
43
44
45
46
47 3.3. Proton conductivity.
48
49 The proton conductivity of SPFBI membranes was measured in-plane at 95% RH and
50
51
52 temperatures between 30 and 80 oC, as depicted in Fig. 4a. We also measured the conductivity of
53
54 Nafion membrane for comparison. In general, as the hopping and/or vehicle movement of protons
55
56
57 is thermally activated in more hydrated state of the membranes, the proton conductivity of all
58
59 membranes gradually increased with temperature. Also, the membranes with higher IEC showed
60
61
62 17
63
64
65
1
2
3
4
5
6
higher proton conductivity; for instance, the SPFBI-0.4 membrane with the highest IEC exhibited
7
8 the highest proton conductivity of 0.121 S cm-1 at 30 oC and 0.1956 S cm-1 at 80 oC. This result
9
10 supports the positive effect of high sulfonic acid content on the proton conduction. Notably, the
11
12
13 SPFBI-0.3 and SPFBI-0.4 membranes outperformed the Nafion membrane in terms of proton
14
15 conductivity, and even the SPFBI-0.2 membrane reached a proton conductivity from 0.063 to
16
17
18 0.105 S cm-1, which is comparable to that of Nafion membrane.
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54 Fig. 4. (a) Proton conductivity of Nafion and SPFBI membranes at different temperatures, (b) the
55
56
57 corresponding Arrhenius plots of SPFBI membranes, (c) comparison of proton conductivity
58
59 correlating with IEC value and in-plane swelling ratio for various reported PEMs (details of test
60
61
62 18
63
64
65
1
2
3
4
5
6
conditions are summarized in Table 1).
7
8
9
10 To obtain more intensive information on the proton conduction behavior of SPFBI PEMs,
11
12
13 the activation energy (Ea) was calculated from the Arrhenius plots relating ln σ and T-1 (Fig. 4b).
14
15 As the gradual decrease in Ea was observed from 9.87 kJ mol-1 of SPFBI-0.2 to 8.67 kJ mol-1 of
16
17
18 SPFBI-0.4, it demonstrates that protons are more feasibly transported when the sulfonic acid
19
20 density increases. This activation energy behavior implies that the proton conduction behavior in
21
22
23 SPFBI membranes complies with the well-known Grotthuss and vehicular mechanisms [45].
24
25 Because the water uptake increased with the sulfonation degree as described, the expanded ionic
26
27
cluster dimension provided lower energy barriers for proton transport. Besides, the large amount
28
29
30 of sulfonic acids are beneficial for proton conduction via Grotthuss mechanism by shortening the
31
32 distance between acid groups for proton hopping [46,47]. In summary, as the increase of sulfonic
33
34
35 acid density reduces the energy barrier of proton-transfer pathways, the SPFBI-0.4 membrane
36
37 showed the highest conductivity. For comparison, we summarized the proton conductivity of the
38
39
40 0.4 membrane and various commercial and lab-synthesized hydrocarbon-based PEMs as a
41
42 function of swelling ratio and IEC as three-dimensionally illustrated in Fig. 4c. Table 1 provides
43
44
45
a more detailed comparison, including test results, test conditions, and polymer types. In general,
46
47 the SPFBI-0.4 membrane illustrated an excellent proton conductivity under relatively low swelling
48
49 when it was placed alongside the non-composite PEMs such as Nafion and Pemion.
50
51
52
53
54 Table 1. Summary of proton conductivity, IECex and swelling ratio of SPFBI-04 membrane and
55
56
57 some typical commercial and lab-synthesized PEMs.
58
59
60
61
62 19
63
64
65
1
2
3
4
5
6 IECex In-plane SRb
7 PEM σa (S cm-1) Polymer type Ref.
8 (meq g-1) (%)
9
10 SPFBI-0.4 0.196 ± 0.005 3.25 ± 0.31 16.12 ± 0.16 Ether-free This work
11
12
Nafion 0.121 ±0.007 0.92 ± 0.24 6.74 ± 0.09 Perfluorinated This work
13
14 Pemion
~ 0.265 3.28 ± 0.06 44.87c Ether-free [27]
15 (sPPN-H+)
16
Pemion
17 ~ 0.170 3.19 ± 0.05 33.3c Ether-free [27]
18 (sPPB-H+)
19
20
SPPBP 40 0.162 2.45 10.8 Ether-free [48]
21
22 SIPiP 0.0975 2.50 11.5 Ether-free [49]
23
24 SPAES160 ~ 0.150 2.10 ± 0.24 19.57 Ether-containing [14]
25
26 SPEEK 0.170 1.654 51.3d Ether-containing [50]
27
28 SPFAE 0.1507 1.62 5.2 Ether-containing [51]
29 a
30 Measured at 80 oC, RH varies from 90 – 100%
b
31 Measured at 80 oC in water
c
32 Mathematically calculated from reported area swelling ratios, assuming that the sample shape
33
34
was square.
d
35 Measured at 60 oC and 100% RH
36
37
38
39 3.4. Oxidative stability.
40
41 The oxidative stability of SPFBI membranes was investigated by subjecting the membrane
42
43
44
samples to Fentons’s test where the chemical degradation of membrane was stimulated by •OH
45
46 radical generation. The membranes were treated with Fenton’s solution (3 wt% H2O2, 3 ppm Fe2+)
47
48 for 1 h and 5 h. During the test, membrane samples were gradually dissolved in the solution due
49
50
51 to polymer decomposition under ·OH radical bombardment. The residual weights of the samples
52
53 were determined and listed in Table 2, indicating how much the membrane retained after the test.
54
55
56 While the SPFBI-0.2 sample exhibited the highest residual weight of 98.95 % after 1 h and 43.33%
57
58 after 5 h, the SPFBI sample lost 5.73% and 81.73% of its original weight after 1 h and 5 h treatment,
59
60
61
62 20
63
64
65
1
2
3
4
5
6
respectively. For all samples, no significant mass loss and visual changes were observable after
7
8 the exposure for 1 h. As the oxidative stability of the membrane decreased with increasing
9
10 sulfonation degree, the ether-free phenylene backbone might be less prone to radical attack than
11
12
13 highly reactive -SO3H groups as reported in several studies [14,27,52]. To more precisely evaluate
14
15 the oxidative stability of SPFBI membranes, their Fenton’s test results were compared with those
16
17
18 of other hydrocarbon PEMs obtained under similar test conditions in Table S4 and S5. For 1 h test,
19
20 the stability of all SPFBI PEMs surpasses that of the common ether-containing hydrocarbon PEMs
21
22
23 even under harsher conditions. Even though Pemion illustrated better oxidative stability than
24
25 SPFBI PEMs for 1 h test, their residual weight difference was not significant. For 5 h test, although
26
27
the stability of PEMs varied with the sulfonation degree, SPFBI PEMs showed quite good
28
29
30 oxidative stability that could compete with Pemion, a state-of-the-art ether-free PEM. It is
31
32 necessary to mention that these 5 h Fenton’s test results under the similar experimental conditions
33
34
35 have been barely reported for other hydrocarbon PEMs because it often led to the entire
36
37 decomposition of membrane samples [46]. The improved chemical stability of this ether-free
38
39
40 SPFBI membrane compared to the ether-containing PEMs can be explained by the different bond
41
42 dissociation energy between aryl-aryl (C6H5-C6H5) segment (476 ± 6 kJ mol-1) and aryl-ether-aryl
43
44
45
(C6H5-O-C6H5) segments (355 ± 6 kJ mol-1)[53]. More importantly, the ether group creates
46
47 vulnerable sites on the neighboring aromatic rings due to the electron-donating effect [1,25,54], as
48
49 illustrated in Fig. S2. The attack of the oxidative radicals on those sites leads to the cleavage and
50
51
52 scission of the polymer backbone, resulting in poor chemical stability of the polymer membrane
53
54 in Fenton’s test.
55
56
57
58
59 Table 2. Residual weights of SPFBI samples after 1 h and 5 h Fenton’s test.
60
61
62 21
63
64
65
1
2
3
4
5
6
7
Sample RW(%) (1 h) RW(%) (5 h)
8
9 SPFBI-0.2 98.85 ± 1.25 43.33 ± 2.39
10
11 SPFBI-0.3 95.00 ± 0.52 29.07 ± 3.05
12
13 SPFBI-0.4 94.27 ± 0.40 18.27 ± 3.56
14
15
16
17 The proton conductivity test and FT-IR were employed to analyze the after-treatment
18
19
20 membrane samples. Because all membrane samples were almost intact after 1 h treatment, it was
21
22 available to measure the proton conductivity of SPFBI membranes with and without 1 h exposure
23
24
25 with Fenton’s solution for comparison (Fig. 5a). Despite slight conductivity losses due to the
26
27 chemical degradation, the membranes still exhibited a decent proton conductivity at various
28
29
30
temperatures, implying that most of the polymer structure and membrane morphology were not
31
32 significantly damaged. The recovered pieces of samples were also assessed by Fourier-transform
33
34 infrared (FT-IR) spectroscopy (Fig. 5b-d). For all samples, there were no serious changes in the
35
36
37 chemical structure after Fenton’s treatment. The IR bands originating from -SO3H groups in the
38
39 pristine membranes still can be clearly observed in the spectra of the membranes after treatment
40
41
42 for both 1 h and 5 h.
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 22
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 Fig. 5. (a) Proton conductivity of SPFBI membranes before and after 1 h Fenton’s test; FT-IR
34
35 spectra of (b) SPFBI-0.2, (c) SPFBI-0.3 and (d) SPFBI-0.4 membrane before and after 1 h and 5
36
37
38
h Fenton’s test.
39
40
41
42 3.5. PEMFC performance.
43
44
45 Compared with many other hydrocarbon polymers used to prepare PEMs, it was discovered
46
47 that SPFBIs have a decent solubility in a mixture of IPA and water, as previously presented. As
48
49
50 this behavior has been rarely reported from the hydrocarbon-based polymers for PEMs, this unique
51
52 solubility property of SPFBI was more detail exploited for its application as an ionomer binder as
53
54
55 well as PEM. As presented, SPFBI-0.4 showed the highest proton conductivity, compared to
56
57 SPFBI-0.2 and SPFBI-0.3 PEMs. When the size of particles agglomerated in catalyst ink was
58
59
60
analyzed using the dynamic light scattering (DLS) method, SPFBI-0.4 was more compatible with
61
62 23
63
64
65
1
2
3
4
5
6
Pt/C catalyst than the other two SPFBI samples (Fig. S3) [55]. Accordingly, taking into account
7
8 both membrane and ionomer properties, SPFBI-0.4 was chosen to prepare the ionomer solution,
9
10 along with the conventional Nafion ionomer as a reference. Using these ionomer binders, we
11
12
13 fabricated MEAs consisting of SPFBI and Nafion membranes for PEMFC operation.
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29 Fig. 6. (a) H2/O2 fuel cell performances of different MEAs using Nafion ionomer binder, (b)
30
31 corresponding Nyquist plots of MEAs using Nafion ionomer binder. (Test conditions: 80 oC, 100%
32
33
34 RH, and gas inlet rate of 300 mL min-1 at both anode and cathode).
35
36
37
38
39 First, we compared the cell performance of MEAs of SPFBI and Nafion PEMs fabricated
40
41 using the same ionomer binder. Fig. 6a shows the polarization curves of MEAs using the Nafion
42
43
44
binder at 80 oC under the fully humidified H2/O2 supply. As the thickness of membranes was
45
46 controlled by casting process, its difference among membranes was less than 3 µm as shown in
47
48 Table 3 to minimize its effect on cell performance. While the maximum power density of SPFBI-
49
50
51 0.2 MEA was quite similar to that of Nafion MEA, that of SPFBI-0.3 and SPFBI-0.4 MEAs
52
53 exhibited higher maximum power densities of 346 and 365 mW cm-2, respectively. This
54
55
56 performance improvement can be attributed to the superior proton conductivity of SPFBI
57
58 membranes. To further understand the effect of types of PEMs on PEMFC performance, the
59
60
61
62 24
63
64
65
1
2
3
4
5
6
electrochemical impedance spectroscopy (EIS) measurements were performed to obtain the
7
8 corresponding Nyquist plots for each MEA, as shown in Fig. 6b. All MEAs showed a semicircular
9
10 pattern of Nyquist plots, in which the high-frequency intercept (left intersection of the real axis)
11
12
13 provided the electrolyte resistance (Rel), and the semicircle diameter provided the charge-transfer
14
15 resistance (Rct) of MEA [56]. It can be seen that the differences in Rct among MEAs were small
16
17
18 because of the utilization of the same catalyst and ionomer binder for all MEAs. On the other hand,
19
20 the Rel values of MEAs varied quite a lot, following the order of SPFBI-0.4 < SPFBI-0.3 < Nafion
21
22
23 < SPFBI-0.2, which is in compliance with the cell performance results. These results demonstrate
24
25 the benefit of SPFBI PEMs that reduce the membrane resistance and thus lead to the enhancement
26
27
of cell performance.
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 25
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46 Fig. 7. Comparison of H2/O2 cell performance of the MEAs fabricated with different PEMs and
47
48 ionomer: (a) Nafion membrane, (b) SPFBI-0.2 membrane, (c) SPFBI-0.4 membrane. The
49
50
51 corresponding Nyquist plots for (d) Nafion membrane, (e) SPFBI-0.2 membrane and (f) SPFBI-
52
53 0.4 membrane. (Test conditions: 80 oC, 100% RH, and gas inlet rates of 300 mL min-1 at both anode
54
55
56
and cathode).
57
58
59
60
61
62 26
63
64
65
1
2
3
4
5
6
Table 3. Summary of power density and EIS for SPFBI and Nafion MEAs.
7
8
9
10 Membrane Electrode Assembly (MEA)
11 Maximum power
Rel (Ω cm2) Rct (Ω cm2)
12 density (mW cm-2)
13 Thickness
14 Membrane Ionomer binder
15 (µm)
16
17 Nafion 318 0.17373 0.39523
18 Nafion 90
19 SPFBI-0.4 294 0.18020 0.45080
20
21 Nafion 321 0.18121 0.43527
22 SPFBI-0.2 88
SPFBI-0.4 296 0.18682 0.48073
23
24 Nafion 346 0.13311 0.44089
25 SPFBI-0.3 87
26 SPFBI-0.4 - - -
27
28 Nafion 365 0.12421 0.43427
29 SPFBI-0.4 87
30 SPFBI-0.4 346 0.11330 0.47170
31
32
33
34 Next, PEMFC performance tests for Nafion, SPFBI-0.2 and SPFBI-0.4 MEAs using the
35
36 SPFBI-0.4 ionomer were also conducted. Their polarization curves were depicted along with those
37
38
39
of MEAs using the Nafion ionomer (Fig. 7a-c) and the corresponding EIS spectra (Fig. 7d-f) for
40
41 comparison. Additionally, some important experimental results including the maximum power
42
43 density, Rel and Rct, are summarized in Table 3. For all membranes, SPFBI-0.4 ionomer still
44
45
46 achieved decent cell performances, even though the Nafion ionomer showed better performances
47
48 with a 5 – 8% higher maximum power density. A similar conclusion can be drawn from the
49
50
51 comparison of Nyquist plots for MEAs fabricated with different ionomers. In each EIS curve,
52
53 although Rel was not significantly varied because the same types of PEMs were employed, Rct
54
55
56 increased when SPFBI-0.4 ionomer was used instead of Nafion. For instance, the MEA fabricated
57
58 using Nafion ionomer and Nafion PEM exhibited Rct of 0.39523 Ω cm2, which was lower than that
59
60
61
62 27
63
64
65
1
2
3
4
5
6
of the MEA using SPFBI-0.4 ionomer and Nafion PEM, 0.45080 Ω cm2. This difference probably
7
8 stems from the difference in proton transport in the catalyst layer affected by ionomer [8,55], as
9
10 the SPFBI-0.4 ionomer was less compatible with the catalyst than the conventional ionomer
11
12
13 according to DLS analysis. Those results, however, still point out that the SPFBI-0.4 ionomer can
14
15 be utilized to prepare MEAs for not only SPFBI membranes but also Nafion membrane to obtain
16
17
18 acceptable cell performance. It must also state that hydrocarbon ionomers like SPFBI may behave
19
20 differently from Nafion ionomer, thus other factors (e.g., catalysts, gas flow rate or ionomer
21
22
23 loading) can be further optimized to be more suitable for SPFBI ionomer usage. Therefore, the cell
24
25 performance of SPFBI-0.4 ionomer MEAs can be improved with more suited MEAs designs,
26
27
which would be the target of our future works. To the extent of our knowledge, the application of
28
29
30 low-cost hydrocarbon polymers as ionomer binder in PEMFC has been rarely reported, as it mostly
31
32 resulted in inferior cell performances to Nafion [57] and the solvent for preparation of ionomer
33
34
35 solution was mostly DMSO which is aprotic and thus not sufficient for fuel cell manufacture
36
37 [58,59]. As this result demonstrate the application of SPFBI-0.4 ionomer binder for high PEMFC
38
39
40 performance, it encourages further developments of hydrocarbon polymers not only for PEM but
41
42 also for ionomer binder to completely replace the perfluorinated polymers.
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 28
63
64
65
1
2
3
4
5
6
Fig. 8. OCV decay over time for various MEAs fabricated with different membrane/ionomer
7
8 binders. (Test conditions: 90 oC, 30% RH, and gas inlet rate: 300 mL min-1 at both anode and
9
10 cathode).
11
12
13
14
15 Finally, the OCV in-situ stability tests were conducted to assess the effect of the SPFBI
16
17
18 membrane and ionomer on the long-term durability of the MEA. Under the harsh temperature and
19
20 humidity conditions, the OCV decay of MEAs was plotted over test time, as shown in Fig. 8. The
21
22
23 end-of-life point of MEAs was chosen as the time when OCV drops to 0 V. For MEAs using Nafion
24
25 membrane, the voltage could not sustain over 220 h, as it completely dropped after around 210 -
26
27
220 h test. On the contrary, the MEAs using SPFBI-0.2 membrane exhibited a much lower OCV
28
29
30 decay rate, resulting in the end-of-life point over 620 h, much higher than those of Nafion
31
32 membrane-used MEAs. This in-situ cell test illustrated surprisingly better durability of SPFBI-0.2
33
34
35 than Nafion, being different from the Fenton’s test results. Athough Nafion PEMs are well-known
36
37 for their excellent stability against chemical degradation [60,61], this mismatched correlation
38
39
40 between ex-situ Fenton’s and in-situ cell tests has been stated in many reports, especially when the
41
42 Nafion MEAs decay faster than the hydrocarbon MEAs in OCV test [14,62]. This difference is
43
44
45
due to the additional factors provided by in-situ cell test that cannot be obtained from ex-situ
46
47 Fenton’s test. For instance, the fuel (hydrogen and oxygen) crossover is one of the important
48
49 factors to be additionally considered in this in-situ test. Due to higher gas permeability of Nafion
50
51
52 than hydrocarbon membranes, the generation rate of OH radicals is much more higher in the
53
54 Nafion based MEA system than the hydrocarbon based one during cell operation. This provides
55
56
57 much more complex and preferential radical attacking situations for the chemical degradation of
58
59 membranes than the simple Fenton’s test conducted in the solution with the fixed OH radical
60
61
62 29
63
64
65
1
2
3
4
5
6
concentration [20,24]. Therefore, while the Fenton’s test provided good information about the
7
8 chemical degradation behavior of SPFBI membrane itself, the OCV test demonstrated much more
9
10 reliable durability of the SPFBI membrane in practical MEA operation.
11
12
13 The optical images of MEAs before and after the stability test were collected and shown in
14
15 Fig. S4. While the SPFBI-0.2 membrane-used MEAs have no observable defects even after over
16
17
18 620 h, some pinholes were found on the Nafion membrane-used MEAs after the test, indicating
19
20 severe damages caused by chemical degradation. This is in good accordance with the OCV decay
21
22
23 rate of the MEAs, demonstrating the excellent in-situ chemical stability of ether-free SPFBI-0.2
24
25 membrane. Besides, compared with Nafion ionomer, SPFBI-0.4 ionomer did not significantly
26
27
affect the OCV decay for both Nafion and SPFBI-0.2 membrane-based MEAs, since the MEAs
28
29
30 consisted of SPFBI-0.2 membrane and SPFBI-0.4 ionomer still illustrated an end-of-life point over
31
32 620 h. Overall, the in-situ OCV test results provide the application of SPFBI membrane and
33
34
35 ionomer for the long-term operation of PEMFC.
36
37
38
39
40 4. Conclusions.
41
42 To synthesize the ether-free polymer for the fabrication of high chemical stability proton
43
44
45
exchange membrane, we proposed a novel polymer synthesis via the superacid-catalyzed
46
47 polymerization of common and low-cost monomers. As a result, a series of ether-free sulfonated
48
49 poly(fluorene biphenyl indole) (SPFBI) were successfully synthesized with various sulfonation
50
51
52 degrees. Experimental results revealed that the excellent proton conductivity of SPFBI membranes
53
54 up to 0.1956 S cm-1 at 80 oC was observed under the low water uptake and swelling ratio compared
55
56
57 to other reported membranes. Owing to the free structure of SPFBI, its oxidative stability was
58
59 comparable to that of Pemion - the state-of-the-art ether-free polymer lectrolyte membrane and
60
61
62 30
63
64
65
1
2
3
4
5
6
superior to many common hydrocarbon membranes. Due to the excellent proton conductivity and
7
8 chemical stability, SPFBI-0.3 and SPFBI-0.4 membranes exhibited better cell performances than
9
10 the Nafion membrane, showing the sustainability of more than 620 h of the open-circuit voltage
11
12
13 test. More importantly, SPFBI-0.4 was possibly employed to prepare the ionomer binder due to
14
15 the solubility in the isopropanol and water mixture. Despite minor losses in the maximum power
16
17
18 density, the performance and stability of membrane electrode assemblies fabricated with SPFBI-
19
20 0.4 ionomer showed the potential promises for research on hydrocarbon ionomer binders. We
21
22
23 believe that the results in this study merit the further development of hydrocarbon-based ether-free
24
25 polymer lectrolyte membranes and ionomer binders for the fabrication of cost-effective and long-
26
27
term durable membrane electrode assembly for fuel cells.
28
29
30
31
32 Acknowledgments
33
34
35 This work was sponsored by the National Research Foundation of Korea Grant, funded by
36
37 the Korean Government (MEST) (NRF 2018M3D1A1058624).
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 31
63
64
65
1
2
3
4
5
6
References
7
8
9 [1] A. Kraytsberg, Y. Ein-Eli, Review of Advanced Materials for Proton Exchange Membrane
10
11 Fuel Cells, Energy Fuels. 28 (2014) 7303–7330. https://doi.org/10.1021/ef501977k.
12
13
14
[2] S.J. Peighambardoust, S. Rowshanzamir, M. Amjadi, Review of the proton exchange
15
16 membranes for fuel cell applications, International Journal of Hydrogen Energy. 35 (2010) 9349–
17
18 9384. https://doi.org/10.1016/j.ijhydene.2010.05.017.
19
20
21 [3] Y. Wang, D.F. Ruiz Diaz, K.S. Chen, Z. Wang, X.C. Adroher, Materials, technological status,
22
23 and fundamentals of PEM fuel cells – A review, Materials Today. 32 (2020) 178–203.
24
25
26 https://doi.org/10.1016/j.mattod.2019.06.005.
27
28 [4] Y. Prykhodko, K. Fatyeyeva, L. Hespel, S. Marais, Progress in hybrid composite Nafion®-
29
30
31 based membranes for proton exchange fuel cell application, Chemical Engineering Journal. 409
32
33 (2021) 127329. https://doi.org/10.1016/j.cej.2020.127329.
34
35
36
[5] M. Vinothkannan, S. Ramakrishnan, A.R. Kim, H.-K. Lee, D.J. Yoo, Ceria Stabilized by
37
38 Titanium Carbide as a Sustainable Filler in the Nafion Matrix Improves the Mechanical Integrity,
39
40 Electrochemical Durability, and Hydrogen Impermeability of Proton-Exchange Membrane Fuel
41
42
43 Cells: Effects of the Filler Content, ACS Appl. Mater. Interfaces. 12 (2020) 5704–5716.
44
45 https://doi.org/10.1021/acsami.9b18059.
46
47
48 [6] L. Zhang, S.-R. Chae, Z. Hendren, J.-S. Park, M.R. Wiesner, Recent advances in proton
49
50 exchange membranes for fuel cell applications, Chemical Engineering Journal. 204–206 (2012)
51
52
53
87–97. https://doi.org/10.1016/j.cej.2012.07.103.
54
55 [7] M.B. Karimi, F. Mohammadi, K. Hooshyari, Recent approaches to improve Nafion
56
57 performance for fuel cell applications: A review, International Journal of Hydrogen Energy. 44
58
59
60 (2019) 28919–28938. https://doi.org/10.1016/j.ijhydene.2019.09.096.
61
62 32
63
64
65
1
2
3
4
5
6
[8] J. Choi, J.H. Yeon, S.H. Yook, S. Shin, J.Y. Kim, M. Choi, S. Jang, Multifunctional
7
8 Nafion/CeO2 Dendritic Structures for Enhanced Durability and Performance of Polymer
9
10 Electrolyte Membrane Fuel Cells, ACS Appl. Mater. Interfaces. 13 (2021) 806–815.
11
12
13 https://doi.org/10.1021/acsami.0c21176.
14
15 [9] K.-D. Kreuer, Ion Conducting Membranes for Fuel Cells and other Electrochemical Devices,
16
17
18 Chem. Mater. 26 (2014) 361–380. https://doi.org/10.1021/cm402742u.
19
20 [10] M. Feng, R. Qu, Z. Wei, L. Wang, P. Sun, Z. Wang, Characterization of the thermolysis
21
22
23 products of Nafion membrane: A potential source of perfluorinated compounds in the environment,
24
25 Scientific Reports. 5 (2015) 9859. https://doi.org/10.1038/srep09859.
26
27
[11] M.J. Parnian, S. Rowshanzamir, J. Alipour Moghaddam, Investigation of physicochemical
28
29
30 and electrochemical properties of recast Nafion nanocomposite membranes using different loading
31
32 of zirconia nanoparticles for proton exchange membrane fuel cell applications, Materials Science
33
34
35 for Energy Technologies. 1 (2018) 146–154. https://doi.org/10.1016/j.mset.2018.06.008.
36
37 [12] T. Miyahara, T. Hayano, S. Matsuno, M. Watanabe, K. Miyatake, Sulfonated
38
39
40 Polybenzophenone/Poly(arylene ether) Block Copolymer Membranes for Fuel Cell Applications,
41
42 ACS Appl. Mater. Interfaces. 4 (2012) 2881–2884. https://doi.org/10.1021/am300821v.
43
44
45
[13] R.P. Pandey, A.K. Thakur, V.K. Shahi, Sulfonated Polyimide/Acid-Functionalized Graphene
46
47 Oxide Composite Polymer Electrolyte Membranes with Improved Proton Conductivity and Water-
48
49 Retention Properties, ACS Appl. Mater. Interfaces. 6 (2014) 16993–17002.
50
51
52 https://doi.org/10.1021/am504597a.
53
54 [14] V.D.C. Tinh, V.D. Thuc, D. Kim, Chemically sustainable fuel cells via layer-by-layer
55
56
57 fabrication of sulfonated poly(arylene ether sulfone) membranes containing cerium oxide
58
59 nanoparticles, Journal of Membrane Science. (2021) 119430.
60
61
62 33
63
64
65
1
2
3
4
5
6
https://doi.org/10.1016/j.memsci.2021.119430.
7
8 [15] C. Wang, Y. Zhou, B. Shen, X. Zhao, J. Li, Q. Ren, Proton-conducting poly(ether sulfone
9
10 ketone)s containing a high density of pendant sulfonic groups by a convenient and mild post-
11
12
13 sulfonation, Polymer Chemistry. 9 (2018) 4984–4993. https://doi.org/10.1039/c8py00996a.
14
15 [16] S. Chandra Sutradhar, M.M. Rahman, F. Ahmed, T. Ryu, Jin lei, S. Yoon, S. Lee, Y. Jin, W.
16
17
18 Kim, Improved proton conductive membranes from poly(phenylenebenzophenone)s with pendant
19
20 sulfonyl imide acid groups for fuel cells, Journal of Power Sources. 442 (2019) 227233.
21
22
23 https://doi.org/10.1016/j.jpowsour.2019.227233.
24
25 [17] H. Nguyen, F. Lombeck, C. Schwarz, P.A. Heizmann, M. Adamski, H.-F. Lee, B. Britton, S.
26
27
Holdcroft, S. Vierrath, M. Breitwieser, Hydrocarbon-based PemionTM proton exchange membrane
28
29
30 fuel cells with state-of-the-art performance, Sustainable Energy Fuels. 5 (2021) 3687–3699.
31
32 https://doi.org/10.1039/D1SE00556A.
33
34
35 [18] K. Oh, K. Ketpang, H. Kim, S. Shanmugam, Synthesis of sulfonated poly(arylene ether
36
37 ketone) block copolymers for proton exchange membrane fuel cells, Journal of Membrane Science.
38
39
40 507 (2016) 135–142. https://doi.org/10.1016/j.memsci.2016.02.027.
41
42 [19] J. Zhang, H. Zhang, J. Wu, J. Zhang, Chapter 11 - Fuel Cell Degradation and Failure Analysis,
43
44
45
in: J. Zhang, H. Zhang, J. Wu, J. Zhang (Eds.), Pem Fuel Cell Testing and Diagnosis, Elsevier,
46
47 Amsterdam, 2013: pp. 283–335. https://doi.org/10.1016/B978-0-444-53688-4.00011-5.
48
49 [20] C. Gittleman, Membrane Durability: Physical and Chemical Degradation, in: 2011: pp. 15–
50
51
52 88.
53
54 [21] S. Yang, Y. Ahn, D. Kim, Poly(arylene ether ketone) proton exchange membranes grafted
55
56
57 with long aliphatic pendant sulfonated groups for vanadium redox flow batteries, J. Mater. Chem.
58
59 A. 5 (2017) 2261–2270. https://doi.org/10.1039/C6TA07456A.
60
61
62 34
63
64
65
1
2
3
4
5
6
[22] Z. Rui, J. Liu, Understanding of free radical scavengers used in highly durable proton
7
8 exchange membranes, Progress in Natural Science: Materials International. 30 (2020) 732–742.
9
10 https://doi.org/10.1016/j.pnsc.2020.08.013.
11
12
13 [23] V.D. Cong Tinh, D. Kim, Enhancement of oxidative stability of PEM fuel cell by introduction
14
15 of HO• radical scavenger in Nafion ionomer, Journal of Membrane Science. 613 (2020) 118517.
16
17
18 https://doi.org/10.1016/j.memsci.2020.118517.
19
20 [24] R. Borup, J. Meyers, B. Pivovar, Y.S. Kim, R. Mukundan, N. Garland, D. Myers, M. Wilson,
21
22
23 F. Garzon, D. Wood, P. Zelenay, K. More, K. Stroh, T. Zawodzinski, J. Boncella, J.E. McGrath, M.
24
25 Inaba, K. Miyatake, M. Hori, K. Ota, Z. Ogumi, S. Miyata, A. Nishikata, Z. Siroma, Y. Uchimoto,
26
27
K. Yasuda, K. Kimijima, N. Iwashita, Scientific Aspects of Polymer Electrolyte Fuel Cell
28
29
30 Durability and Degradation, Chem. Rev. 107 (2007) 3904–3951.
31
32 https://doi.org/10.1021/cr050182l.
33
34
35 [25] M. Adamski, N. Peressin, S. Holdcroft, On the evolution of sulfonated polyphenylenes as
36
37 proton exchange membranes for fuel cells, Mater. Adv. 2 (2021) 4966–5005.
38
39
40 https://doi.org/10.1039/D1MA00511A.
41
42 [26] Miyake Junpei, Taki Ryunosuke, Mochizuki Takashi, Shimizu Ryo, Akiyama Ryo, Uchida
43
44
45
Makoto, Miyatake Kenji, Design of flexible polyphenylene proton-conducting membrane for next-
46
47 generation fuel cells, Science Advances. 3 (n.d.) eaao0476. https://doi.org/10.1126/sciadv.aao0476.
48
49 [27] M. Adamski, T.J.G. Skalski, B. Britton, T.J. Peckham, L. Metzler, S. Holdcroft, Highly Stable,
50
51
52 Low Gas Crossover, Proton-Conducting Phenylated Polyphenylenes, Angewandte Chemie
53
54 International Edition. 56 (2017) 9058–9061. https://doi.org/10.1002/anie.201703916.
55
56
57 [28] T.J.G. Skalski, M. Adamski, B. Britton, E.M. Schibli, T.J. Peckham, T. Weissbach, T.
58
59 Moshisuki, S. Lyonnard, B.J. Frisken, S. Holdcroft, Sulfophenylated Terphenylene Copolymer
60
61
62 35
63
64
65
1
2
3
4
5
6
Membranes and Ionomers, ChemSusChem. 11 (2018) 4033–4043.
7
8 https://doi.org/10.1002/cssc.201801965.
9
10 [29] T.J.G. Skalski, B. Britton, T.J. Peckham, S. Holdcroft, Structurally-Defined, Sulfo-
11
12
13 Phenylated, Oligophenylenes and Polyphenylenes, J. Am. Chem. Soc. 137 (2015) 12223–12226.
14
15 https://doi.org/10.1021/jacs.5b07865.
16
17
18 [30] D.R. Nieto, S. Fomine, M.G. Zolotukhin, L. Fomina, M. del C.G. Hernandez,
19
20 Superelectrophilic Activation of N-Substituted Isatins: Implications for Polymer Synthesis, a
21
22
23 Theoretical Study, Macromolecular Theory and Simulations. 18 (2009) 138–144.
24
25 https://doi.org/10.1002/mats.200800075.
26
27
[31] M.C.G. Hernandez, M.G. Zolotukhin, S. Fomine, G. Cedillo, S.L. Morales, N. Fröhlich, E.
28
29
30 Preis, U. Scherf, M. Salmón, M.I. Chávez, J. Cárdenas, A. Ruiz-Trevino, Novel, Metal-Free,
31
32 Superacid-Catalyzed “Click” Reactions of Isatins with Linear, Nonactivated, Multiring Aromatic
33
34
35 Hydrocarbons, Macromolecules. 43 (2010) 6968–6979. https://doi.org/10.1021/ma101048z.
36
37 [32] R. Ren, S. Zhang, H.A. Miller, F. Vizza, J.R. Varcoe, Q. He, Facile Preparation of an Ether-
38
39
40 Free Anion Exchange Membrane with Pendant Cyclic Quaternary Ammonium Groups, ACS Appl.
41
42 Energy Mater. 2 (2019) 4576–4581. https://doi.org/10.1021/acsaem.9b00674.
43
44
45
[33] J. Wang, Y. Zhao, B.P. Setzler, S. Rojas-Carbonell, C. Ben Yehuda, A. Amel, M. Page, L.
46
47 Wang, K. Hu, L. Shi, S. Gottesfeld, B. Xu, Y. Yan, Poly(aryl piperidinium) membranes and
48
49 ionomers for hydroxide exchange membrane fuel cells, Nature Energy. 4 (2019) 392–398.
50
51
52 https://doi.org/10.1038/s41560-019-0372-8.
53
54 [34] H. Nederstedt, P. Jannasch, Poly(p-terphenyl alkylene)s grafted with highly acidic sulfonated
55
56
57 polypentafluorostyrene side chains for proton exchange membranes, Journal of Membrane Science.
58
59 647 (2022) 120270. https://doi.org/10.1016/j.memsci.2022.120270.
60
61
62 36
63
64
65
1
2
3
4
5
6
[35] T. Ryu, H. Jang, F. Ahmed, N.S. Lopa, H. Yang, S. Yoon, I. Choi, W. Kim, Synthesis and
7
8 characterization of polymer electrolyte membrane containing methylisatin moiety by
9
10 polyhydroalkylation for fuel cell, International Journal of Hydrogen Energy. 43 (2018) 5398–5404.
11
12
13 https://doi.org/10.1016/j.ijhydene.2017.12.164.
14
15 [36] Y. Cui, B. Xu, B. Yang, H. Yao, S. Li, J. Hou, A Novel pH Neutral Self-Doped Polymer for
16
17
18 Anode Interfacial Layer in Efficient Polymer Solar Cells, Macromolecules. 49 (2016) 8126–8133.
19
20 https://doi.org/10.1021/acs.macromol.6b01595.
21
22
23 [37] T. Shiraki, Y. Tsuchiya, T. Noguchi, S. Tamaru, N. Suzuki, M. Taguchi, M. Fujiki, S. Shinkai,
24
25 Creation of Circularly Polarized Luminescence from an Achiral Polyfluorene Derivative through
26
27
Complexation with Helix-Forming Polysaccharides: Importance of the meta-Linkage Chain for
28
29
30 Helix Formation, Chemistry – An Asian Journal. 9 (2014) 218–222.
31
32 https://doi.org/10.1002/asia.201301216.
33
34
35 [38] N. Chen, H.H. Wang, S.P. Kim, H.M. Kim, W.H. Lee, C. Hu, J.Y. Bae, E.S. Sim, Y.-C. Chung,
36
37 J.-H. Jang, S.J. Yoo, Y. Zhuang, Y.M. Lee, Poly(fluorenyl aryl piperidinium) membranes and
38
39
40 ionomers for anion exchange membrane fuel cells, Nature Communications. 12 (2021) 2367.
41
42 https://doi.org/10.1038/s41467-021-22612-3.
43
44
45
[39] Dilute Solution Thermodynamics, Molecular Weights, and Sizes, in: Introduction to Physical
46
47 Polymer Science, John Wiley & Sons, Ltd, 2005: pp. 71–143.
48
49 https://doi.org/10.1002/0471757128.ch3.
50
51
52 [40] T. Omata, M. Tanaka, K. Miyatake, M. Uchida, H. Uchida, M. Watanabe, Preparation and
53
54 Fuel Cell Performance of Catalyst Layers Using Sulfonated Polyimide Ionomers, ACS Appl. Mater.
55
56
57 Interfaces. 4 (2012) 730–737. https://doi.org/10.1021/am201360j.
58
59 [41] J.-S. Park, P. Krishnan, S.-H. Park, G.-G. Park, T.-H. Yang, W.-Y. Lee, C.-S. Kim, A study on
60
61
62 37
63
64
65
1
2
3
4
5
6
fabrication of sulfonated poly(ether ether ketone)-based membrane-electrode assemblies for
7
8 polymer electrolyte membrane fuel cells, Journal of Power Sources. 178 (2008) 642–650.
9
10 https://doi.org/10.1016/j.jpowsour.2007.08.008.
11
12
13 [42] S. Lee, Y. Lim, M.A. Hossain, H. Jang, Y. Jeon, S. Lee, L. Jin, W. Kim, Synthesis and
14
15 properties of grafting sulfonated polymer containing isatin by super acid-catalyzed
16
17
18 polyhydroxyalkylation reaction for PEMFC, Renewable Energy. 79 (2015) 72–77.
19
20 https://doi.org/10.1016/j.renene.2014.08.023.
21
22
23 [43] Glass-Rubber Transition Behavior, in: Introduction to Physical Polymer Science, John Wiley
24
25 & Sons, Ltd, 2005: pp. 349–425. https://doi.org/10.1002/0471757128.ch8.
26
27
[44] N. Peressin, M. Adamski, E.M. Schibli, E. Ye, B.J. Frisken, S. Holdcroft, Structure–Property
28
29
30 Relationships in Sterically Congested Proton-Conducting Poly(phenylene)s: the Impact of
31
32 Biphenyl Linearity, Macromolecules. 53 (2020) 3119–3138.
33
34
35 https://doi.org/10.1021/acs.macromol.0c00310.
36
37 [45] D.J. Kim, M.J. Jo, S.Y. Nam, A review of polymer–nanocomposite electrolyte membranes
38
39
40 for fuel cell application, Journal of Industrial and Engineering Chemistry. 21 (2015) 36–52.
41
42 https://doi.org/10.1016/j.jiec.2014.04.030.
43
44
45
[46] L. Liu, Y. Lu, Y. Pu, N. Li, Z. Hu, S. Chen, Highly sulfonated carbon nano-onions as an
46
47 excellent nanofiller for the fabrication of composite proton exchange membranes with enhanced
48
49 water retention and durability, Journal of Membrane Science. 640 (2021) 119823.
50
51
52 https://doi.org/10.1016/j.memsci.2021.119823.
53
54 [47] V.D. Thuc, V.D. Cong Tinh, D. Kim, Simultaneous improvement of proton conductivity and
55
56
57 chemical stability of Nafion membranes via embedment of surface-modified ceria nanoparticles
58
59 in membrane surface, Journal of Membrane Science. 642 (2022) 119990.
60
61
62 38
63
64
65
1
2
3
4
5
6
https://doi.org/10.1016/j.memsci.2021.119990.
7
8 [48] S.C. Sutradhar, H. Jang, N. Banik, J. Yoo, T. Ryu, H. Yang, S. Yoon, W. Kim, Synthesis and
9
10 characterization of proton exchange poly (phenylenebenzophenone)s membranes grafted with
11
12
13 propane sulfonic acid on pendant phenyl groups, International Journal of Hydrogen Energy. 42
14
15 (2017) 12749–12758. https://doi.org/10.1016/j.ijhydene.2016.11.054.
16
17
18 [49] L. Jin, Y. Lee, D. Kim, F. Ahmed, T. Ryu, W. Kim, H. Jang, Comparative study of chemically
19
20 different structured sulfonic acid and sulfonimide acid of Poly(isatine-phenylene) electrolyte for
21
22
23 PEMFC, International Journal of Hydrogen Energy. 46 (2021) 6762–6774.
24
25 https://doi.org/10.1016/j.ijhydene.2020.11.160.
26
27
[50] P. Wei, Y. Sui, X. Li, Q. Liu, B. Zhu, C. Cong, X. Meng, Q. Zhou, Sandwich-structure
28
29
30 PI/SPEEK/PI proton exchange membrane developed for achieving the high durability on excellent
31
32 proton conductivity and stability, Journal of Membrane Science. 644 (2022) 120116.
33
34
35 https://doi.org/10.1016/j.memsci.2021.120116.
36
37 [51] Y. Lu, Y. Liu, N. Li, Z. Hu, S. Chen, Sulfonated graphitic carbon nitride nanosheets as proton
38
39
40 conductor for constructing long-range ionic channels proton exchange membrane, Journal of
41
42 Membrane Science. 601 (2020) 117908. https://doi.org/10.1016/j.memsci.2020.117908.
43
44
45
[52] S.-W. Lee, Z.G. Abdi, J.-C. Chen, K.-H. Chen, Optimal method for preparing sulfonated
46
47 polyaryletherketones with high ion exchange capacity by acid-catalyzed crosslinking for proton
48
49 exchange membrane fuel cells, Journal of Polymer Science. 59 (2021) 706–720.
50
51
52 https://doi.org/10.1002/pol.20200872.
53
54 [53] Handbook of Bond Dissociation Energies in Organic Compounds By Yu-Ran Luo (University
55
56
57 of South Florida, St. Petersburg). CRC Press LLC: Boca Raton. 2003. xii + 380 pp. $159.95. ISBN
58
59 0-8493-1589-1., J. Am. Chem. Soc. 126 (2004) 982–982. https://doi.org/10.1021/ja0336224.
60
61
62 39
63
64
65
1
2
3
4
5
6
[54] J. Walkowiak-Kulikowska, J. Wolska, H. Koroniak, Polymers application in proton exchange
7
8 membranes for fuel cells (PEMFCs), 2 (2017). https://doi.org/10.1515/psr-2017-0018.
9
10 [55] J.E. Chae, S.J. Yoo, J.Y. Kim, J.H. Jang, S.Y. Lee, K.H. Song, H.-J. Kim, Hydrocarbon-based
11
12
13 electrode ionomer for proton exchange membrane fuel cells, International Journal of Hydrogen
14
15 Energy. 45 (2020) 32856–32864. https://doi.org/10.1016/j.ijhydene.2020.03.003.
16
17
18 [56] X.-Z. Yuan, C. Song, H. Wang, J. Zhang, eds., Impedance and its Corresponding
19
20 Electrochemical Processes, in: Electrochemical Impedance Spectroscopy in PEM Fuel Cells:
21
22
23 Fundamentals and Applications, Springer London, London, 2010: pp. 95–138.
24
25 https://doi.org/10.1007/978-1-84882-846-9_3.
26
27
[57] A. Strong, B. Britton, D. Edwards, T.J. Peckham, H.-F. Lee, W.Y. Huang, S. Holdcroft,
28
29
30 Alcohol-Soluble, Sulfonated Poly(arylene ether)s: Investigation of Hydrocarbon Ionomers for
31
32 Proton Exchange Membrane Fuel Cell Catalyst Layers, Journal of The Electrochemical Society.
33
34
35 162 (2015) F513–F518. https://doi.org/10.1149/2.0251506jes.
36
37 [58] S. Holdcroft, Fuel Cell Catalyst Layers: A Polymer Science Perspective, Chem. Mater. 26
38
39
40 (2014) 381–393. https://doi.org/10.1021/cm401445h.
41
42 [59] R. Sood, C. Iojoiu, E. Espuche, F. Gouanvé, G. Gebel, H. Mendil-Jakani, S. Lyonnard, J.
43
44
45
Jestin, Comparative Study of Proton Conducting Ionic Liquid Doped Nafion Membranes
46
47 Elaborated by Swelling and Casting Methods: Processing Conditions, Morphology, and Functional
48
49 Properties, J. Phys. Chem. C. 118 (2014) 14157–14168. https://doi.org/10.1021/jp502454m.
50
51
52 [60] H. Jang, T. Ryu, S.C. Sutradhar, F. Ahmed, K. Choi, H. Yang, S. Yoon, W. Kim, Studies of
53
54 sulfonated poly(phenylene)-block-poly(ethersulfone) for proton exchange membrane fuel cell,
55
56
57 International Journal of Hydrogen Energy. 42 (2017) 12768–12776.
58
59 https://doi.org/10.1016/j.ijhydene.2017.01.112.
60
61
62 40
63
64
65
1
2
3
4
5
6
[61] A.Z. Al Munsur, B.-H. Goo, Y. Kim, O.J. Kwon, S.Y. Paek, S.Y. Lee, H.-J. Kim, T.-H. Kim,
7
8 Nafion-Based Proton-Exchange Membranes Built on Cross-Linked Semi-Interpenetrating
9
10 Polymer Networks between Poly(acrylic acid) and Poly(vinyl alcohol), ACS Appl. Mater.
11
12
13 Interfaces. 13 (2021) 28188–28200. https://doi.org/10.1021/acsami.1c05662.
14
15 [62] R. Shimizu, J. Tsuji, N. Sato, J. Takano, S. Itami, M. Kusakabe, K. Miyatake, A. Iiyama, M.
16
17
18 Uchida, Durability and degradation analysis of hydrocarbon ionomer membranes in polymer
19
20 electrolyte fuel cells accelerated stress evaluation, Journal of Power Sources. 367 (2017) 63–71.
21
22
23 https://doi.org/10.1016/j.jpowsour.2017.09.025.
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 41
63
64
65

You might also like