Sie sind auf Seite 1von 7

international journal of hydrogen energy 36 (2011) 18131819

Available at www.sciencedirect.com

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

Synthesis and characterization of crosslinked sulfonated


poly(arylene ether sulfone) membranes for high temperature
PEMFC applications
Ki Tae Park, Jeong Hwan Chun, Sang Gon Kim, Byung-Hee Chun, Sung Hyun Kim*
Department of Chemical & Biological Engineering, Korea University, 1 Anam-Dong, Seongbuk-Gu, Seoul 136-701, Republic of Korea

article info

abstract

Article history:

Sulfonated poly(arylene ether sulfone) copolymers containing carboxyl groups are

Received 15 September 2009

prepared by an aromatic substitution polymerization reaction using phenolphthalin, 3,30 -

Received in revised form

disulfonated-4,40 -dichlorodiphenyl sulfone, 4,40 -dichlorodiphenyl sulfone and 4,40 -bisphenol

1 February 2010

A as polymer electrolyte membranes for the development of high temperature polymer

Accepted 2 February 2010

electrolyte membrane fuel cells. Thin, ductile films are fabricated by the solution casting

Available online 3 March 2010

method, which resulted in membranes with a thickness of approximately 50 mm. Hydroquinone is used to crosslink the prepared copolymer in the presence of the catalyst, sodium

Keywords:

hypophosphite. The synthesized copolymers and membranes are characterized by 1H NMR,

Sulfonated

FT-IR, TGA, ion exchange capacity, water uptake and proton conductivity measurements.

Poly(arylene ether sulfone)

The water uptake and proton conductivity of the membranes are decreased with increasing

Carboxyl group

the degree of crosslinking which is determined by phenolphthalin content in the copol-

High temperature

ymer (015 mol%). The prepared membranes are tested in a 9 cm2 commercial single cell at

Low humidity

80  C and 120  C in humidified H2/air under different relative humidity conditions. The

Polymer electrolyte membrane fuel

uncrosslinked membrane is found to perform better than the crosslinked membranes at

cell

80  C; however, the crosslinked membranes perform better at 120  C. The crosslinked


membrane containing 10 mol% of phenolphthalin (CPS-PP10) shows the best performance
of 600 mA cm2 at 0.6 V and better performance than the commercial Nafion 112
(540 mA cm2 at 0.6 V) at 120  C and 30 % RH.
2010 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved.

1.

Introduction

Polymer electrolyte membranes (PEMs) are key components


that determine the cost and performance of polymer electrolyte membrane fuel cells (PEMFCs). The most recent
studies on PEMFCs have primarily focused on developing new
proton conducting membranes for operation at higher
temperatures under lower humidification conditions [112].
Higher operation temperatures (e.g., >100  C) have many
advantages such as faster electrode kinetics, higher tolerance

to CO poisoning, smaller heat exchanger, and easier water


thermal management [13]. Most commercially available
proton exchange membranes use a perfluorosulfonic acid
(PFSA) ionomer, such as Nafion, which is produced by
DuPont. However, Nafion-type membranes have several
disadvantages including high cost, fuel permeability and
decreased performance accompanying dehydration above
80  C [14]. As alternatives to Nafion, many sulfonated polymers such as sulfonated poly(ether ether ketone), sulfonated
poly(arylene ether sulfone), sulfonated poly(phenylene

* Corresponding author. Tel.: 82 2 3290 3297; fax: 82 2 926 6102.


E-mail address: kimsh@korea.ac.kr (S.H. Kim).
0360-3199/$ see front matter 2010 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2010.02.019

1814

international journal of hydrogen energy 36 (2011) 18131819

sulfide) and sulfonated polystyrene have been developed for


fuel cell applications [1524]. Recently, sulfonated poly(arylene ether sulfone) copolymers have been proposed as one
of the most attractive new polymer materials due to their
excellent thermal and oxidative stability, good mechanical
properties, superior processability and exceptional hydrolytic
stability [25,26]. One major advantage of using direct
condensation polymerization rather than conventional
preparation, which uses post sulfonation reactions on
homopolymers, for preparing these membranes, is that two
sulfonic acid groups are attached at the meta position of the
deactivating sulfonyl group [27]. It has been recognized that
sulfonated aromatic copolymers exhibit less microscopic
phase separation between hydrophobic domains and hydrophilic ionic clusters than PFSA polymers, because of the
lower hydrophobicity of the aromatic backbone, increased
stiffness of the aromatic backbone, shorter ionic side chains
and lower acidity of the sulfonic acid group [28,29]. The
hydrophilic/hydrophobic domain structure of these copolymers strongly depends on the composition of the sulfonated
monomer. Generally, proton conductivity and water retention capacity of membranes are enhanced when the degree
of sulfonation is increased. However, when the degree of
sulfonation exceeds a certain level, the membranes are
highly swollen and/or soluble in water. One way to overcome
this problem is to crosslink the polymer. To date, there have
been a few studies that examined crosslinked poly(arylene
ether sulfone) copolymers [19,30,31]. Kim et al. reported on
the preparation of sulfonated poly(arylene ether sulfone)
copolymers containing carboxylic acid groups using phenolphthalin, 3,30 -disulfonated-4,40 -dichlorodiphenyl sulfone
and 4,40 -dichlorodiphenyl sulfone for direct methanol fuel
cells (DMFCs). The carboxylic acid groups in the sulfonated
poly(arylene ether sulfone) copolymer side chain increased
the hydrophilicity of the copolymer and provided additional
acid sites. In addition, the carboxyl groups could be active
pendant groups that react with crosslinking agents to form
cross-linked membrane materials [27]. 1,2,3,4-butanetetracarboxylic acid was used to crosslink amorphous regions
of rayon fibers via anhydride-mediated esterification for the
purpose of enhanced mechanical properties [32]. The crosslinking was conducted under a series of curing temperatures
in the presence of sodium hypophosphite, which acted as an
effective catalyst for anhydride formation.
In this paper, we report on the preparation of sulfonated
poly(arylene ether sulfone) copolymers containing carboxylic
acid groups using phenolphthalin, 3,30 -disulfonated-4,40 dichlorodiphenyl sulfone, 4,40 -dichlorodiphenyl sulfone and
4,40 -bisphenol A. Crosslinked membranes were fabricated
with prepared copolymers containing carboxyl groups and
hydroquinone in the presence of the catalyst, sodium hypophosphite. The effect of crosslinked structure was also evaluated by varying the molar ratio of carboxylic acid groups in
the copolymers. The membranes were characterized by
Fourier transform infrared spectroscopy, thermo-gravimetric
analysis, water uptake and proton conductivity measurements. In addition, the membranes were tested in
a commercial single cell to investigate the ability of the
membranes to perform PEMFCs operations under high
temperatures and low humidity conditions.

2.

Experimental

2.1.

Materials

4,40 -dichlorodiphenyl sulfone (DCDPS), 4,40 -bisphenol A (BPA),


phenolphthalin (PP), anhydrous potassium carbonate, sodium
hypophosphite (SHP) and hydroquinone (HQ) were obtained
from Aldrich. DCDPS and BPA were dried under vacuum at
100  C for 24 h prior to use. PP, anhydrous potassium
carbonate, SHP and HQ were used as received. 3,30 -disulfonated-4,40 -dichlorodiphenyl sulfone (SDCDPS) was synthesized from DCDPS according to a procedure reported by
Harrion et al. [20] and dried under vacuum at 100  C for 24 h
before use. N-methyl-2-pyrrolidone (NMP), N,N-dimethylacetamide (DMAc) (Aldrich), toluene, methanol (J.T. Baker) and
sulfuric acid (Daejung reagents and chemical) were used as
received.

2.2.
Synthesis of sulfonated poly(arylene ether sulfone)
copolymer containing carboxylic acid group
The sulfonated poly(arylene ether sulfone) copolymers containing carboxylic acid groups were synthesized at different
molar ratios of PP to the total monomer content. A typical
copolymerization procedure to prepare the sulfonated copolymer (SDCDPS/DCDPS 60/40, PP/BPA 10/90) was as follows.
First, 7.3677 g (15 mmol) of SDCDPS, 2.8717 g (10 mmol) of
DCDPS, 0.8010 g (2.5 mmol) of PP and 5.1644 g (22.5 mmol) of
BPA were added in a three neck flask equipped with a nitrogen
inlet, Dean-Stark trap and magnetic stirrer. 90 mL of NMP was
added into the flask and stirred until the monomers were
dissolved. After toluene was added to the reaction flask
(usually, NMP/toluene 2/1 v/v), 4.1463 g (30 mmol) of anhydrous potassium carbonate was added. The reaction mixture
was refluxed at 150  C for 4 h to dehydrate the system. The
temperature was slowly raised to 190  C by controlled removal
of toluene from the mixture. The mixture was further reacted
for 24 h, during which the solution became very viscous. The
solution was cooled to room temperature and diluted with
DMAc to allow easier filtering. The solution was isolated by
coagulation in excess methanol after filtration with filter
paper of 100 mm pore size to remove most of the salts. Finally,
the precipitated copolymer was washed several times with
ethanol and dried under vacuum at 100  C for 24 h.

2.3.

Fabrication of crosslinked membrane

Membranes were prepared by dissolution of the copolymers in


DMAc to create a 10 % (w/w) polymer solution. An aqueous
solution containing 5 %(w/w) HQ and 5 %(w/w) SHP was added
to the polymer solution. The solution was then cast on a glass
substrate and maintained at 60  C for 12 h under a nitrogen
atmosphere. The cast membranes were then heated to 180  C
under vacuum for 12 h. The dried membranes were detached
from the glass substrate and washed several times with deionized water. Finally, acidification was carried out by
immersion in a 2 N sulfuric acid (H2SO4) for 24 h. The acidified
membranes were stored in de-ionized water at room
temperature.

international journal of hydrogen energy 36 (2011) 18131819

2.4.

Polymer and membrane characterization

2.6.

H NMR (400 MHz) spectra were recorded on a Varian instrument using dimethyl sulfoxide-d6 (DMSO-d6) as a solvent.
Fourier transform-infrared (FT-IR) spectra were recorded
with a Bomen DA-8 spectrometer.
The ion exchange capacity (IEC) of the copolymers was
measured using the following titration method. The dried
polymer powder was soaked in a 1 M NaCl solution for 12 h at
ambient temperature. This solution was titrated with a 0.01 M
NaOH solution to neutralize the exchanged proton (H) using
an automatic titrator (Metrohm). The IEC was calculated using
the following equation:
IEC

VM
m

mwet  mdry
 100
mdry

(2)

where mwet and mdry are the weights of the wet and dry
samples (g), respectively.
The proton conductivities at 25  C under fully hydrated
conditions were evaluated in liquid water by electrochemical
impedance analyzer (IM 6ex, Zahner) over the frequency
range of 101 MHz using the following equation:
s l=RS

(3)

where s, l, and S are the proton conductivity (S cm1), the


distance between electrodes used to measure the potential
(cm), and area for proton transport (cm2), respectively. R, the
resistance of membrane (Ohm), was derived from the
minimum imaginary response.

2.5.

Electrochemical measurement of MEAs

The electrochemical measurements were carried out using


a 9 cm2 single cell in humidified H2/air gases. The single cell
was operated at 80  C and 120  C under 100 %RH and 30 %RH
conditions, respectively. When operated at 120  C, the inlet
gases were at a pressure of 2 atm in order to prevent the
membranes from dehydrating. The flow rates of the gases
were fixed at 1.5 times the stoichiometry of the fuel and twice
the stoichiometric value of the oxidant. The polarization
curves were measured by applying a constant current for
3 min at each point using an electronic loader (EP-1000, Deagil
Electronics).

(1)

where IEC is the ion exchange capacity (meq g1), V is the


added titrant volume at the equivalent point (mL), M is the
molar concentration of the titrant, and m is the dry sample
weight (g).
Thermo-gravimetric analysis of the membranes was
carried out using a thermal analysis system (TGA 2050, TA
instruments) at temperatures ranging from room temperature
to 700  C at a heating rate of 10  C min1 in a N2 atmosphere.
The water uptake of the membranes was calculated by the
difference between the wet and dry weight of the membranes.
The wet weight was measured after immersing the
membranes in distilled water for 24 h at room temperature.
The dry weight (mdry) was measured after drying the
membranes at 100  C for 2 h. The water uptake (Wup%) was
calculated using the following equation:
Wup%

1815

Preparation of membrane electrode assembly (MEA)

The MEAs were prepared by directly spraying the catalyst


slurries on the membranes and pressing between the gas
diffusion layers (GDLs). The catalyst slurries were prepared by
mixing a 40 % (w/w) Pt/C catalyst (Johnson Matthey) with IPA
(J.T. Baker HPLC grade) and 5 % (w/w) Nafion solution
(DuPont) followed by ultrasonication for 15 min. The platinum
(Pt) loading on both electrodes was 0.4 mg Pt cm2. SGL carbon
papers, which were 400 mm in thickness, were used as the
GDLs.

3.

Results and discussion

3.1.

Synthesis and characterization of the copolymers

A series of copolymers were prepared by an aromatic substitution polymerization reaction using phenolphthalin (PP),
3,30 -disulfonated-4,40 -dichlorodiphenyl sulfone (SDCDPS),
4,40 -dichlorodiphenyl sulfone (DCDPS) and 4,40 -bisphenol A
(BPA) in the presence of potassium carbonate in NMP with
different compositions by varying the molar ratio of PP to BPA.
Toluene was used for azeotropic removal of water during the
reaction. The general reaction sequence is depicted in Scheme 1.
We synthesized CPS-PPxx, which has 60 mol% disulfonated
units in the polymer backbone, where xx means the molar ratio
of PP to BPA.
1
H NMR was used to identify and characterize the
sulfonated copolymers. The 1H NMR spectrum of CPS-PP10 is
shown in Fig. 1. The integration and appropriate analysis of
known reference protons of the copolymers allowed for the
relative compositions of the copolymers to be determined
[20,27,33]. A peak at 8.25 ppm was separated from the other
aromatic protons, and assigned to the protons adjacent to the
sulfonate group. Integration of the peak was used to calculate
the actual mol % of SDCDPS in the copolymers, which was
determined to be approximately 58.4 mol%.

3.2.
Crosslinking of sulfonated poly(arylene ether
sulfone) copolymers
Crosslinked
sulfonated
poly(arylene
ether
sulfone)
membranes were prepared by casting their 10 % (w/w) solutions in DMAc with an aqueous solution containing 5 % (w/w)
HQ and 5 % (w/w) SHP. In the case of CPS-PP00, the polymer
solutions were prepared without HQ and SHP because the
CPS-PP00 contains no functional groups for esterification
reactions. The cast membranes were dried at 60  C for 12 h
under a nitrogen atmosphere and heated to 180  C under
vacuum for 12 h. The dried membranes were washed several
times with de-ionized water.
The crosslinked structure was verified by comparing the
FTIR spectra of the membranes with different molar ratios of
PP. FTIR spectra of the membranes with different molar ratios
of PP were shown in Fig. 2. The absorption bands at 1745 and
1629 cm1 were assigned to the ester and carboxylate
carbonyl, respectively [32]. It should be noted that there was

1816

international journal of hydrogen energy 36 (2011) 18131819

Scheme 1 Synthesis scheme of the sulfonated poly(arylene ether sulfone) copolymers.

no absorption corresponding to the ester or carboxylate


carbonyl for the uncrosslinked membrane (CPS-PP00
membrane). The degree of crosslinking can be evaluated by
comparing the absorption intensity of the ester linkages of the
membranes. The intensity of the absorption bands at
1745 cm1 increased when the molar ratio of PP in the
copolymers was increased (Fig. 2). Consequently, the crosslinked structures of membranes increase as the contents of
carboxyl groups in the copolymers increases.

3.3.

Characterization of the membranes

The content of PP, calculated ion exchange capacities (IEC),


measured IEC and thickness of the obtained membranes are
summarized in Table 1. The CPS-PP10(p) was also prepared
without esterification for comparison with CPS-PP10 to evaluate the effect of crosslinked structure. The measured IEC
1
mmol-COO g1) were lower than the
values (mmol-SO
3 g

1
calculated IEC values (mmol-SO
mmol-COO g1)
3 g
except for CPS-PP00 due to the consumption of carboxylic acid
groups during membrane preparation. The CPS-PP10(p) had
measured IEC values of 2.35 meq g1 which is similar to
calculated IEC values of 2.37 meq g1 and according to
60 mol% of disulfonated groups (2.19 meq g1) and 10 mol% of
carboxylic acid groups (0.18 meq g1). This result shows that
the most of carboxylic acid groups in the copolymer were not
consumed during membrane preparation in the case of CPSPP10(p). However, the CPS-PP10 which is esterified
membrane had measured IEC values between 2.19 and
2.37 meq g1 because carboxylic acid groups were consumed
by esterification reaction and unreacted free carboxylic acid
groups also exist.
The thermal stability of the membranes was investigated
by thermo-gravimetric analysis (TGA). Three consecutive
mass loss steps were seen in the TGA thermograms, which
resulted from the processes of thermal salvation, thermal

Fig. 1 1H NMR spectroscopy of CPS-PP10 in DMSO-d6.

1817

international journal of hydrogen energy 36 (2011) 18131819

Fig. 3 TGA thermograms of the membranes.


Fig. 2 FTIR spectra of the membranes containing different
molar ratios of PP.

desulfonation and thermal oxidation of the polymer matrix


(Fig. 3). In addition, the temperature at which these decompositions occurred was slightly shifted in the crosslinked
membranes and no significant difference in thermal stability
among the crosslinked membranes was observed. The loss of
mass at approximately 100  C corresponds to water desorption. The de-sulfonation process begins at approximately
280  C, and the polymer backbone chain begins to decompose
after 450  C. The prepared copolymers were found to be
thermally stable up to 250  C.
Table 2 shows the results of water uptake and proton
conductivity measurements of the membranes. The water
uptake and proton conductivity of the membranes were
decreased as the PP content increased due to the crosslinked
structure. Generally, the water uptake of sulfonated polymers
is known to have profound effect on proton conductivity and
mechanical properties [34]. High water retention capacity
leads to higher proton conductivity because the water in the
membrane provides a carrier for proton transport. However,
excessive water uptake causes undesirable dimensional
changes and the membranes can become highly swollen and/
or soluble in water. From the results of water uptake
measurement of the membranes, we can see that the highly

crosslinked membrane has low water retention capacity, that


is low swelling ratio. Although the membrane properties of
water retention capacity and poton conductivity were
decreased with PP content, the crosslinked structure of
the membranes allows higher dimensional stability and the
unreacted free carboxylic acid groups may contribute to the
increased proton conductivity and water retention capacity
under high temperature and low humidity conditions.

3.4.

Single cell performances

Fig. 4 shows the polarization curves of the single cells for the
membranes at 80  C and 120  C. All the polarization curves
were obtained after single cell operation for 72 h. As shown in
Fig. 4 (a), the uncrosslinked membrane (CPS-PP00) performed
better than the crosslinked membranes and the cell performance was decreased with PP content at 80  C and 100 % RH.
By contrast, crosslinked membranes showed better performances than the uncrosslinked membrane at 120  C and 30 %
RH. When comparing Fig. 4(a) and (b), it is quite apparent that
the composite membranes are more effective under high
temperature and low humidification conditions. In addition,
the CPS-PP10 showed the highest current density of 600
mA/cm2 at 0.6 V and better performance than the commercial
Nafion 112 at 120  C and 30 % RH. The cell performances of
CPS-PP00 and CPS-PP10(p) drastically decreased with operation time at 120  C due to dimensional deformation and some

Table 1 Properties of prepared membranes.


Samples

Nafion 112
CPS-PP00
CPS-PP05
CPS-PP10
CPS-PP10(p)
CPS-PP15

PP
Calculated Measured Thickness
IECb
(mm)
Content
IECa
(meq/g)
(meq/g)
(mol %)

0
5
10
10
15

2.22
2.30
2.37
2.37
2.44

0.90
2.22
2.22
2.24
2.35
2.25

51
50
51
51
50
51

1
a Theoretical calculated (mmol-SO
mmol-COO g1).
3 g
1
b Measured (mmol-SO
mmol-COO g 1).
3 g

Table 2 Water uptake and proton conductivity


measurements of the membranes.
Samples

Nafion 112
CPS-PP00
CPS-PP05
CPS-PP10
CPS-PP10(p)
CPS-PP15

Water
uptake
(w/w %)

Proton
conductivity
(mS/cm)

36.8
65.7
64.8
62.4
66.1
59.2

94  6
88  8
85  4
84  4
89  7
82  6

1818

international journal of hydrogen energy 36 (2011) 18131819

ratios of PP and characterized by 1H NMR spectrum. Crosslinked membranes were prepared by thermal treatment in the
presence of the catalyst, SHP, which catalyzed the esterification reaction, and the structure was verified by FTIR. The
prepared membranes were thermally stable up to 250  C. An
increase in the PP content of the copolymer was associated
with the degree of membrane crosslinking. The crosslinked
structure of the membranes allowed for higher dimensional
stability. The uncrosslinked membrane performed better than
the crosslinked membranes at 80  C and 100 % RH. By contrast,
crosslinked membranes performed better at 120  C and 30 %
RH. Although the crosslinked membrane structure increases
dimensional stability at high temperature, the cell performance was decreased due to low water uptake and proton
conductivity when the degree of crosslinking exceeded
a certain level. The CPS-PP10 showed the best performance of
620 mA/cm2 at 0.6 V and better performance than the
commercial Nafion 112 at 120  C and 30 % RH.

Acknowledgement
This work was financially supported by the Seoul research and
business development Program.

references

Fig. 4 Polarization curves for the tested membranes; (a) at


80 8C, 100% RH and (b) 120 8C, 30% RH.

dissolution to water. The dissolution of the membrane caused


serious diminish of the cell performance, because the dissolved polymer hindered the active sites of catalyst and
blocked the pore of gas diffusion layer. These results show the
effect of crosslinked structure which improved the dimensional stability of the crosslinked membrane at high temperature. Although the crosslinked membrane structure
increases dimensional stability at high temperature, the
crosslinked structure also causes low water uptake and low
proton conductivity. The CPS-PP15 showed lower performance than the CPS-PP10. Therefore, the appropriate degree
of crosslinking of membranes is required for PEMFC operations under high temperature and low RH conditions.

4.

Conclusions

A sulfonated poly(arylene ether sulfone) copolymer was


successfully synthesized via an aromatic substitution
polymerization reaction from phenolphthalin (PP), 3,30 -disulfonated-4,40 -dichlorodiphenyl sulfone (SDCDPS), 4,40 -dichlorodiphenyl sulfone (DCDPS) and 4,40 -bisphenol A (BPA) in the
presence of potassium carbonate in NMP with different molar

[1] Sacca A, Gatto I, Carbone A, Pedicini R, Passalacqua E. ZrO2


Nafion composite membranes for polymer electrolyte fuel
cells (PEFCs) at intermediate temperature. J Power Sources
2006;163:47.
[2] Zhai Y, Zhang H, Hu J, Yi B. Preparation and characterization
of sulfated zirconia (SO2
4 /ZrO2)/Nafion composite
membranes for PEMFC operation at high temperature/low
humidity. J Membr Sci 2006;280:148.
[3] Jalani NH, Dunn K, Datta R. Synthesis and characterization of
Nafion-MO2 (M Zr, Si, Ti) nanocomposite membranes for
higher temperature PEM fuel cells. Electrochim Acta 2005;51:553.
[4] Ren S, Sun G, Li C, Song S, Xin Q, Yang X. Sulfated zirconia
Nafion composite membranes for higher temperature direct
methanol fuel cells. J Power Sources 2006;157:724.
[5] Zeng R, Wang Y, Wang S, Shen PK. Homogeneous synthesis
of PFSI/silica composite membranes for PEMFC operating at
low humidity. Electrochim Acta 2007;52:3895.
[6] Ramani V, Kunz HR, Fenton JM. Stabilized composite
membranes and membrane electrode assemblies for
elevated temperature/low relative humidity PEFC operation.
J Power Sources 2005;152:182.
[7] Jung UH, Park KT, Park EH, Kim SH. Improvement of lowhumidity performance of PEMFC by addition of hydrophilic
SiO2 particles to catalyst layer. J Power Sources 2006;159:529.
[8] Park KT, Jung UH, Choi DW, Chun K, Lee HM, Kim SH. ZrO2
SiO2/Nafion composite membrane for polymer electrolyte
membrane fuel cells operation at high temperature and low
humidity. J Power Sources 2008;177:247.
[9] Hagihara Hiroki, Uchida Hiroyuki, Watanabe Masahiro.
Preparation of highly dispersed SiO2 and Pt particles in
Nafion112 for self-humidifying electrolyte membranes in
fuel cells. Electrochim Acta 2006;51:3979.
[10] Zhang YF, Wang SJ, Xiao M, Bian SG, Meng YZ. The silicadoped sulfonated poly(fluorenyl ether ketone)s membrane
using hydroxypropyl methyl cellulose as dispersant for high

international journal of hydrogen energy 36 (2011) 18131819

[11]

[12]

[13]

[14]

[15]

[16]

[17]

[18]

[19]

[20]

[21]

[22]

temperature proton exchange membrane fuel cells. Int J


Hydrogen Energy 2009;34:4379.
Wen Sheng, Gong Chunli, Tsen Wen-Chin, Shu Yao-Chi,
Tsai Fang-Chang. Sulfonated poly(ether sulfone) (SPES)/
boron phosphate (BPO4) composite membranes for hightemperature proton-exchange membrane fuel cells. Int J
Hydrogen Energy 2009;34:8982.
Jung Guo-Bin, Weng Fang-Bor, Su Ay, Wang Jiun-Sheng, Leon
Yu T, Lin Hsiu-Li, et al. Nafion/PTFE/silicate membranes for
high-temperature proton exchange membrane fuel cells. Int
J Hydrogen Energy 2008;33:2413.
Li QF, He RH, Jensen JO, Bjerrum NJ. Approaches and recent
development of polymer electrolyte membranes for fuel cells
operating above 100  C. Chem Mater 2003;15:4896.
Hickner MA, Ghassemi H, Kim YS, Einsla BR, McGrath JE.
Alternative polymer systems for proton exchange
membranes (PEMs). Chem Rev 2004;104:4587.
Lee HS, Roy A, Lane O, Dunn S, McGrath JE. Hydrophilic
hydrophobic multiblock copolymers based on poly(arylene
ether sulfone) via low-temperature coupling reactions for
proton exchange membrane fuel cells. Polymer 2008;49:715.
Lee JK, Li W, Manthiram A. Poly(arylene ether sulfone)s
containing pendant sulfonic acid groups as membrane
materials for direct methanol fuel cells. J Membr Sci 2009;
330:73.
Badami AS, Lane O, Lee HS, Roy A, McGrath JE. Fundamental
investigations of the effect of the linkage group on the
behavior of hydrophilichydrophobic poly(arylene ether
sulfone) multiblock copolymers for proton exchange
membrane fuel cells. J Membr Sci 2009;333:1.
Roy A, Lee HS, McGrath JE. Hydrophilic-hydrophobic
multiblock copolymers based on poly(arylene ether sulfone)s
as novel proton exchange membranes - Part B. Polymer 2008;
49:5037.
Oh YS, Lee HJ, Yoo M, Kim HJ, Han J, Kim TH. synthesis of
novel crosslinked sulfonated poly(arylene ether sulfone)s
using bisazide and their properties for fuel cell application.
J Membr Sci 2008;323:309.
Harrison WL, Wang F, Mecham JB, Bhanu VA, Hill M, Kim YS,
et al. Influence of the bisphenol structure on the direct
synthesis of sulfonated poly(arylene ether) copolymers. I.
J Polym Sci A Polym Chem 2003;41:2264.
Im SJ, Patel R, Shin SJ, Kim JH, Min BR. Korean. Sulfonated
poly(arylene ether sulfone) membranes based on biphenol
for direct methanol fuel cells. J Chem Eng 2008;25(4):732.
Goh YT, Patel R, Im SJ, Kim JH, Min BR. Korean. Synthesis and
characterization of poly(ether sulfone) grafted poly(styrene

[23]

[24]

[25]

[26]
[27]

[28]

[29]

[30]

[31]

[32]

[33]

[34]

1819

sulfonic acid) for proton conducting membranes. J Chem Eng


2009;26(2):518.
Devrim Yilser, Erkan Serdar, Bac Nurcan, Eroglu Inci.
Preparation and characterization of sulfonated polysulfone/
titanium dioxide composite membranes for proton exchange
membrane fuel cells. Int J Hydrogen Energy 2009;34:3467.
Sx engul Erce, Erdener Hulya, Gultekin Akay R,
_ I.
_ Effects of
Yucel Hayrettin, Bac Nurcan, Eroglu Inc
sulfonated polyetheretherketone (SPEEK) and composite
membranes on the proton exchange membrane fuel cell
(PEMFC) performance. Int J Hydrogen Energy 2009;34:4645.
Rogers ME, Long TE. Synthetic methods in step-growth
polymers. In: Wang S, McGrath JE, editors. Synthesis of
poly(arylene ether)s. New York: John Wiley & Sons, Inc.;
2003. p. 327.
Rose JB. Preparation and properties of poly(arylene ether
sulphones)*. Polymer 1974;15:456.
Kim DS, Shin KH, Park HB, Chung YS, Nam SY, Lee YM.
Synthesis and characterization of sulfonated poly(arylene
ether sulfone) copolymers containing carboxyl groups for
direct methanol fuel cells. J Membr Sci 2006;278:428.
Kim YS, Hickner MA, Dong L, Pivovar BS, McGrath JE.
Sulfonated poly(arylene ether sulfone) copolymer proton
exchange membranes: composition and morphology effects
on the methanol permeability. J Membr Sci 2004;243:317.
Kreuer KD. On the development of proton conducting
polymer membranes for hydrogen and methanol fuel cells.
J Membr Sci 2001;185:29.
Heo KB, Lee HJ, Kim HJ, Kim BS, Lee SY, Cho E, et al. Synthesis
and characterization of cross-linked poly(ether sulfone) for
a fuel cell membrane. J Power Sources 2007;172:215.
Feng S, Shang Y, Xie X, Wang Y, Xu J. Synthesis and
characterization of crosslinked sulfonated poly(arylene ether
sulfone) membranes for DMFC applications. J Membr Sci
2009;335:13.
Li W, Xu X, Chen S, Zhou X, Li L, Chen D, et al. Esterification
crosslinking structures of rayon fibers with 1,2,3,4butanetetracarboxylic acid and their water-responsive
properties. Carbohydr Polym 2008;71:574.
Gao Y, Robertson GP, Guiver MD, Jian X. Synthesis and
characterization of sulfonated poly(phthalazinone ether
ketone) for proton exchange membrane materials. J Polym
Sci A Polym Chem 2003;41:497.
Zawodzinski TA, Springer TE, Davey J, Jestel R, Lopez C,
Valerio J, et al. A comparative study of water uptake by and
transport through ionomeric fuel cell membranes.
J Electrochem Soc 1993;140:1981.

Das könnte Ihnen auch gefallen