You are on page 1of 14

Journal of Membrane Science 225 (2003) 63–76

Synthesis and characterization of proton conducting


polymer membranes for fuel cells
B. Smitha, S. Sridhar, A.A. Khan∗
Membrane Separations Group, Chemical Engineering Division, Indian Institute of Chemical Technology,1 Hyderabad 500 007, India
Received 15 January 2003; received in revised form 15 July 2003; accepted 18 July 2003

Abstract
Commercial polymers with aryl backbones such as polystyrene (PS), polycarbonate (PC), polysulfone (PSf) and poly
(phenylene oxide) (PPO) were sulfonated using suitable reagents, and assessed for their potential to serve as proton exchange
membranes (PEM) in fuel cells (FCs). The membranes thus synthesized were characterized by Fourier transform infra-red
(FTIR) and 1 H NMR to verify sulfonation and to identify the sites available for proton conduction. Differential scanning
calorimetry (DSC) and thermogravimetric analysis (TGA) studies were carried out to investigate the thermal stability of the
sulfonated membranes. Surface morphology and tensile strength were evaluated by scanning electron microscopy (SEM) and
UTM, respectively. Sorption experiments were conducted to observe the interaction of sulfonated polymers with water and
methanol. The ion exchange capacity (IEC), which is a measure of proton conductivity, and the degree of substitution (DS) were
evaluated and found to be comparable with the commercially available Nafion membranes. Sulfonated polycarbonate (SPC)
was found to possess all the requisite properties of a PEM; namely, high IEC (0.57 meq./g), tensile strength (157 N/mm2 ) and
thermal stability (Tg = 120 ◦ C) besides low affinity towards methanol (%sorption = 2.97). The study reveals the possibility of
developing inexpensive and sturdy fuel cell membranes, which could provide attractive alternatives to substitute the expensive
commercially available membranes.
© 2003 Elsevier B.V. All rights reserved.
Keywords: Polymer electrolyte membranes; Aryl backbone; Ion exchange capacity; Degree of substitution; Membrane characterization

1. Introduction

Fuel cells (FCs) have emerged strongly as a viable


alternative source of power owing to their high-energy
Abbreviations: PEM, polymer electrolyte membrane; PPO, poly efficiency and eco-friendly nature [1,2]. The proton
(phenylene oxide); PSf, polysulfone; PS, polystyrene; PC, poly- exchange membrane (PEM) in fuel cells have been
carbonate; FTIR, Fourier transform infra-red; NMR, nuclear mag-
netic resonance spectroscopy; SEM, scanning electron microscopy;
well established for over five decades and are suc-
DSC, disc scanning calorimetry; TGA, thermogravimetric analy- cessfully commercialized as electrical power sources
sis; SPPO, sulfonated PPO; SPSf, sulfonated PSf; SPS, sulfonated in spacecrafts and submarines [3]. During the last
PS; SPC, sulfonated PC; IEC, ion exchange capacity; DS, degree decade the compatibility of PEM-FCs as power source
of substitution; DMAc, N,N-dimethyl acetamide in mass production of automobiles and portable elec-
∗ Corresponding author. Tel.: +91-40-719-3139/3626;

fax: +91-40-719-0387/0757.
trical devices, is well fathomed [1]. This has lead to
E-mail address: aakhan iict@rediffmail.com (A.A. Khan). an increasing interest in materials being used as elec-
1 IICT Communication No. 030613. trolytes. Currently, the choice is confined to hydrated

0376-7388/$ – see front matter © 2003 Elsevier B.V. All rights reserved.
doi:10.1016/S0376-7388(03)00343-0
64 B. Smitha et al. / Journal of Membrane Science 225 (2003) 63–76

perfluorosulfonic polymers such as Nafion [4]. The In the present study, inexpensive and abundantly
perfluorosulfonic acid membranes cost US$ 800 m−2 available commercial polymers namely polycarbonate
and the power generated costs approximate to US$ and polystyrene (PS) have been investigated after sul-
150 kW−1 [4]. Since the technical feasibility of fuel fonation for their resistance to high-temperature and
cells has been to a large extent successfully demon- chemicals. Poly (phenylene oxide) (PPO) and polysul-
strated, the industry is currently engaged in field trials, fone were also sulfonated as these are polymers with
lowering costs and determining manufacturing strate- rigid aryl backbones and completely hydrophobic in
gies [5]. Hence, despite the superior performance of nature [17]. These polymers upon suitable structural
the membranes currently used in fuel cells, their inher- modifications tend to exhibit a number of desirable
ently high cost of production limits their usage in fu- properties including high ionic conductivity, good
ture mass-produced appliances including automobiles. mechanical strength and reasonably low affinity to
From the application point of view, any membrane methanol and water. This work is therefore an attempt
even the one without fluorine chemistry, but hav- to synthesize sulfonated membranes, with hydrophilic
ing high temperature resistance, greater mechanical regions around the cluster of side chains in a generally
strength and high proton conductivity would be useful. hydrophobic polymer for promoting proton conduc-
Currently there is an increasing interest in a va- tivity. This modification manifests itself in many ways
riety of polymers having fluorinated alkyl and aryl such as thermal transitional behavior, mechanical
backbones to serve as suitable membranes for fuel properties, morphology, and above all the crucial prop-
cell applications [6]. Sulfonation is a powerful and erty in the context of fuel cells namely ion exchange
versatile process, which can be used to simultane- capacity. Some of the polymers tested in the study, i.e.
ously render these polymers proton conductive as PPO, PSf and PS have been sulfonated earlier and eval-
well as hydrophilic in nature. Sulfonated polymers uated for their ability to behave as solid polymer elec-
can be prepared in the form of free acid (–SO3 H), trolytes in fuel cells but extensive characterization by
a salt (e.g. –SO3 − Na+ ) or an ester (–SO3 R) [7]. different methods has been lacking which is what this
The degree of sulfonation can be controlled as de- paper aims to report. Moreover, there is very little lit-
sired. The polymers can be sulfonated in the initial erature on feasibility of using alternatives such as sul-
stages of synthesis or can be sulfonated in their final fonated polycarbonate (SPC) for fuel cell applications.
form. Any of the homopolymers, random copolymers
and block and graft copolymers containing aromatic
rings or double bonds are suitable as building blocks 2. Experimental
for this application. Extensive sulfonation leads to
water-soluble polymers, which find use as thickeners 2.1. Materials
and flocculants. Sulfonation improves the dye-ability
of propylenestyrene copolymer fibers [8]. The sodium Poly (phenylene oxide) was synthesized in a sim-
salt of sulfonated butyl rubber is reportedly stronger ilar manner to that described in the literature [18].
than the unsulfonated polymer [9,10]. Sulfonated Polysulfone, polycarbonate and polystyrene of num-
polyesters and polyimides are also mentioned in the ber average molecular weight of 22,000, 64,000 and
literature [11–13]. Sulfonated polysulfone (SPSf) 430,000, respectively, were purchased from Aldrich.
membranes are reported to be useful in desalination Chloroform, 1,2-dichloroethane, acetic anhydride, sul-
applications [14,15]. furic acid and chlorosulfonic acid of AR grade, were
Polysulfone (PSf) has been sulfonated [7] using purchased from S.D. Fine Chem. Ltd., Mumbai.
sulfur trioxide–triethyl phosphate complex as sul-
fonating agent and the effect of sulfonation on the 2.2. Sulfonation of polymers
glass transition temperature, water absorption and gas
permeability was studied. The synthesis of sulfonated 2.2.1. Choice of sulfonating agent
poly (phenylene oxide) (SPPO) has been investigated Three different reagents, viz. sulfuric acid, acetyl
[16] and the effects of catalyst concentration, solvent sulfate and chlorosulfonic acid were used to sulfonate
and kinetics of sulfonation have been reported. the four polymers used in the study. The criteria for
B. Smitha et al. / Journal of Membrane Science 225 (2003) 63–76 65

selection of the best reagent for a specific polymer transferred into a dropper and gradually added to the
was based on its compatibility with polymer and film PPO solution over a period of 20 min while stirring
forming property besides the mechanical strength of the solution vigorously at 25 ◦ C. Special precautions
the sulfonated polymer. had to be taken to carry out the reaction involving
the usage of chlorosulfonic acid. The precipitated
• Sulfuric acid: Sulfonation was done with concen- polymer, i.e. sulfonated PPO (SPPO), dark brown in
trated sulfuric acid (98%) as per the procedure re- color, was washed with distilled water and allowed to
ported in the literature [19]. Despite the addition dry in air for 24 h at room temperature [18].
of small quantities of acid, it was observed that the
polymer so obtained dissolved completely in water. 2.2.3. Sulfonation of polysulfone
Excess of the acid resulted in the degradation of the Polysulfone was sulfonated on similar lines as that
polymer which is undesirable. The degree of sul- of PPO.
fonation could not be controlled and hence sulfuric
acid was found unsuitable as sulfonating agent. 2.2.4. Sulfonation of polystyrene
• Chlorosulfonic acid: PPO and PSf were sulfonated Sulfonated polystyrene (SPS) was prepared by us-
by this reagent. As the extent of sulfonation could be ing acetyl sulfate as the sulfonating agent [21] which
controlled, it was observed that the resultant poly- was freshly prepared by mixing 15.3 ml of acetic
mers swelled in water but did not dissolve com- anhydride with 5.6 ml of sulfuric acid in 79 ml of
pletely. The mechanical strength and film forming 1,2-dichloroethane. The desired quantity of acetyl
capabilities of the polymer thus formed were rea- sulfate solution was added to a reaction flask con-
sonably good. PC and PS could not be sulfonated taining a solution of 20.8 g of polystyrene in 98 ml
using this sulfonating agent as the polymer pre- of 1,2-dichloroethane. The sulfonation reaction was
cipitated as such without undergoing any change. allowed to proceed for 1 h at 50 ◦ C and was then ter-
This was confirmed by Fourier transform infra-red minated by adding methanol. The resulting sulfonated
(FTIR). polystyrene was isolated by introducing steam. The
• Acetyl sulfate: It is reported that a homogenous dis- polymer was then collected and recovered by filtration.
tribution of sulfonic acid groups in the resultant
polymer is obtained when PS was sulfonated using 2.2.5. Sulfonation of polycarbonate
this reagent [12]. Apart from PS, PC too reacted Polycarbonate was sulfonated in a manner analo-
readily with acetyl sulfate and the properties of re- gous to that of polystyrene.
sultant product are encouraging. However, PPO and
PSf could not be sulfonated using this method be- 2.2.6. Effect of concentration of sulfonating agent
cause of the lack of their compatibility with the Experiments were carried out by adding varying
reagent. quantities of sulfonating agents to the same quantity
of the polymer. The effect is best illustrated by citing
2.2.2. Sulfonation of poly (phenylene oxide) examples of sulfonation of PPO and PSf with chloro-
Sulfonation was carried out in a chloroform solvent sulfonic acid.
system at ambient conditions using chlorosulfonic
acid as the sulfonating agent as described by Plum- (a) 2.5% of chlorosulfonic acid: Addition of entire
mer et al. [20]. Prior to the sulfonation step, the small quantity did not result in the formation of any pre-
quantity of sulfonable material (ethanol and water) cipitate and hence the polymer had to be precip-
in chloroform, used as a solvent for PPO, was deter- itated in methanol. The polymer so obtained did
mined and neutralized with chlorosulfonic acid. Ten not exhibit film forming property as it was not sol-
grams of PPO was then added to 100 ml of neutral- uble in any solvent after drying.
ized chloroform, in a three-neck reaction flask, and (b) 5% of chlorosulfonic acid: Addition of 70–75 ml
dissolved by stirring for about 30 min at room temper- of chlorosulfonic acid and chloroform mixture into
ature to form 3–6 wt.% solution. A 5% (v/v) solution the reaction vessel resulted in the precipitation
of chlorosulfonic acid in 100 ml of chloroform was of the polymer. The sulfonated polymer did not
66 B. Smitha et al. / Journal of Membrane Science 225 (2003) 63–76

exhibit excessive swelling and hence membranes clean glass plate to obtain dense membranes. The cast
cast out of this method were chosen for character- films were evaporated to dryness in open air at room
ization. temperature (30–35 ◦ C) and were later vacuum dried
(c) 10% of chlorosulfonic acid: In this case, it was for a period of 12 h at ambient temperature to expel
observed that addition of 25–35 ml of the mixture traces of solvent. Membranes were also prepared from
resulted in the precipitation of the polymer. The respective unsulfonated polymers, using a similar pro-
polymer obtained was in the degraded state and cedure, for comparison of various characteristics.
the membrane so formed was completely soluble Fig. 1 shows the chemical structures of the sul-
in water. fonated membranes prepared in the study.
Thus, based on the trials, addition of 5% of chloro-
sulfonic acid to chloroform was chosen as the opti- 2.4. FTIR studies
mum quantity of sulfonating agent.
The FTIR spectra of unsulfonated and sulfonated
membranes were scanned using Nicolet-740, Perkin-
2.3. Preparation of membranes Elmer-283B FTIR Spectrometer. These spectra are
shown in Figs. 2 and 3.
Membranes synthesized for the study were prepared
by solution casting and solvent evaporation technique. 2.5. NMR studies
Approximately 5 wt.% of SPPO in DMAc, 15 wt.%
of SPSf in N,N-dimethyl formamide and 12 wt.% of The 1 H NMR spectra of unsulfonated and sul-
SPC in 1,4-dioxane were dissolved in the respective fonated membranes were scanned to detect the occur-
solvents. rence of sulfonation. The solvent used to dissolve the
A solution (32 g/l) of SPS in toluene/n-butanol mix- polymer samples was deuterated chloroform (CDCl3 ).
ture (7:3, v/v) is prepared, filtered and spread on a Observed spectrum is recorded in Fig. 4.

Fig. 1. Chemical structure of: (a) SPSf, (b) SPC, (c) SPS and (d) SPPO.
B. Smitha et al. / Journal of Membrane Science 225 (2003) 63–76 67

Fig. 2. FTIR spectra of: (a) polystyrene and (b) sulfonated polystyrene.

2.6. Thermal analysis 700 ◦ C heated at 10 ◦ C/min in nitrogen gas flushed at


200 ml/min. The samples were subjected to thermo-
2.6.1. DSC studies gravimetric analysis (TGA) both before and after sul-
The differential scanning calorimetry (DSC) spectra fonation to determine the decomposition temperatures.
of sulfonated and unsulfonated polymers were obtai- Results are shown in Fig. 7.
ned on Perkin-Elmer DSC Model 7. Measurements
were performed over the temperature range of
30–200 ◦ C at the heating rate of 5 ◦ C/min in herme- 2.7. Determination of the ion exchange capacity
tically sealed aluminium pans. Membrane samples and degree of substitution
were allowed to attain steady state with the solvents
and the sample pan conditioned in the instrument The ion exchange capacity (IEC) indicates the
before running the experiment. Results are shown in number of milli-equivalents of ions in 1 g of the dry
Figs. 5 and 6. polymer. The degree of substitution (DS) indicates
the average number of sulfonic groups present in the
2.6.2. TGA sulfonated polymer. To determine the degree of sub-
Thermal stability of the polymer films was exam- stitution by acid groups, the sulfonated membranes
ined, using Seiko 220TG/DTA analyzer, from 25 to and unsulfonated specimens of similar weight were
68 B. Smitha et al. / Journal of Membrane Science 225 (2003) 63–76

Fig. 3. FTIR spectra of: (a) polycarbonate and (b) sulfonated polycarbonate.

soaked in 50 ml of 0.01 N sodium hydroxide solution Relationship between DS and IEC is


for 12 h at ambient temperature. Then, 10 ml of the
120 × IEC
solution was titrated with 0.01 N sulfuric acid [5]. The DS =
sample was regenerated with 1 M hydrochloric acid, 1000 + 120 × IEC − 200 × IEC
washed free of acid with water and dried to a constant
weight. The IEC was calculated according to: 2.8. Swelling

B − P × 0.01 × 5 In order to determine their interaction, weighed


IEC =
m samples of circular pieces of sulfonated polymer films
where IEC is the ion exchange capacity (meq./g), (3 cm diameter) were soaked in water and methanol.
B the sulfuric acid used to neutralize blind sample The films were taken out after different soaking pe-
soaked in NaOH (ml), P the sulfuric acid used to riods and quickly weighed after carefully wiping out
neutralize the sulfonated membranes soaked in NaOH excess water/methanol to estimate the amount ab-
(ml), 0.01 the normality of the sulfuric acid, 5 the sorbed at the particular time “t”. The film was then
factor corresponding to the ratio of the amount of quickly placed back in the solvent. The process was
NaOH taken to dissolve the polymer to the amount repeated until the films attained steady state as in-
used for titration, and m the sample mass (g). dicated by constant weight after a certain period of
B. Smitha et al. / Journal of Membrane Science 225 (2003) 63–76 69

Fig. 4. 1 H NMR spectra of: (a) polystyrene and (b) sulfonated polystyrene.

soaking time. The degree of swelling was calculated between the grips of the testing machine. The grip
from: length was 5 cm and the speed of testing was set at the
Ms − Md rate of 12.5 mm/min. Tensile strength was calculated
swelling (%) = × 100 using:
Md
where Ms is the mass of the swollen polymer in g and max load
tensile strength (N/mm2 ) = .
Md the mass of the dry polymer in g. cross-sectional area

2.9. Mechanical properties 2.10. Scanning electron microscopy

The equipment used for carrying out the test was The films were thoroughly dried and the surface
Universal Testing Machine (Shimadzu make, model morphology was studied by scanning electron mi-
AGS-10kNG) with an operating head load of 5 kN. croscopy (SEM) using a Hitachi S2150 microscope.
Cross-sectional area of the sample of known width and The SEM images for polysulfone and polystyrene
thickness were calculated. The films were then placed before and after sulfonation are shown in Fig. 8.
70 B. Smitha et al. / Journal of Membrane Science 225 (2003) 63–76

3. Results and discussion

The chemical structures of the sulfonated polymers


prepared in this study are presented in Fig. 1. The
membranes made from the polymers were charac-
terized extensively for their thermal stability, tensile
strength, swelling and finally for IEC.

3.1. FTIR analysis

Two sets of FTIR spectra comparing sulfonated


and unsulfonated films pertaining to polystyrene and
polycarbonate are presented in Figs. 2 and 3, respec-
tively, and the spectra relating to polysulfone and
poly (phenylene oxide) are only presented in the text
in order to save the manuscript space. Referring to
Fig. 2 it can be seen that the sharp peaks at 700
and 780 cm−1 in unsulfonated polystyrene are due
to –C–H out of plane deformation which represent
mono substitution. The disappearance of these peaks
after sulfonation followed by the occurrence of a new
Fig. 5. DSC spectra of: (a) polycarbonate and (b) sulfonated
peak at 520 cm−1 representing para substitution is
polycarbonate.
indicative of the attachment of sulfonic acid groups
at para position. The peak identified in the spectra
at 1360 cm−1 is due to the asymmetric stretching
of S=O bond. The symmetric vibration of this bond
produces the characteristic split band of absorbance
at 1150–1185 cm−1 . There was no significant change
observed in the peaks at 2925 and 3000 cm−1 , which
represent C–C and C–H bond, respectively.
Sulfonated polycarbonate spectra in Fig. 3 shows
sharp peaks at 1160 and 1340 cm−1 which are not
present in the broad band peaks at 1150–1220 cm−1
observed for unmodified polycarbonate. As in the case
of sulfonated polystyrene, the peak at 1360 cm−1 iden-
tified in the spectra is due to the asymmetric stretching
of S=O bond and the symmetric vibration of this bond
produces the characteristic split band of absorbance at
1150–1185 cm−1 .
Sulfonated polysulfone shows a peak represent-
ing a broad H-bonded O–H stretch in the range
of 1175–1150 cm−1 . The asymmetric stretching of
S=O bond occurs at 1350–1340 cm−1 with an O–H
stretching around 1165–1150 cm−1 .
Spectra of sulfonated PPO showed the aromatic
group at 1600 and 1490 cm−1 , sulfonic acid at 1070
Fig. 6. DSC spectra of: (a) polystyrene and (b) sulfonated and 675 cm−1 and S=O stretching vibration at 1369
polystyrene. and 1163 cm−1 .
B. Smitha et al. / Journal of Membrane Science 225 (2003) 63–76 71

Fig. 7. TGA of: (a) sulfonated PPO and (b) sulfonated polystyrene.

Thus, the results of FTIR analysis clearly prove the served in the case of sulfonated polysulfone. This is
occurrence of sulfonation by the presence of sulfonate due to the stronger electron attracting force of the sul-
groups after the reaction in all the four polymers. fonic acid group. Similar indications of sulfonation
were observed in the 1 H NMR spectra of sulfonated
3.2. NMR analysis PPO.
1 H NMR spectra before and after sulfonation is pre- 3.3. Tensile strength and elongation
sented for polystyrene only (Fig. 4) whereas the in-
terpretation is reported in case of all the rest of the The tensile strength at break of the polymers before
polymers. In the case of sulfonated polystyrene, no and after sulfonation is given in Table 1. From the re-
significant change is noticed in the signals at 7.3, 6.6 sults, it can be observed that there is a decrease in the
and 7 ppm. A hump between 5 and 6 ppm indicates tensile strength and elongation of all the polymers af-
the presence of sulfonic acid linkage on the aromatic ter sulfonation. This reduction may be attributed to the
benzene ring. degree of substitution. Sulfonation causes the polymer
The aromatic portion of the 1 H NMR spectra of matrix to expand. This expansion permits an increase
sulfonated polycarbonate is diminished on account of in the free volume thereby increasing the chain move-
sulfonation. This depicts replacement of aromatic pro- ments which makes the plastic material softer and
tons by sulfonic acid groups. New signals at 3.2, 3.4 flexible. This greater chain movement implies that the
and 3.6 ppm portray the change of field around the material changes partially from glassy state (hard and
protons on the methyl groups. brittle) to rubbery state (flexible and soft), thereby
A shift in the range of protons, from 8 to 8.3 ppm causing a reduction in the tensile strength at break
next to the sulfone bridge of the polysulfone, is ob- and a corresponding reduction in the elongation [23].
72 B. Smitha et al. / Journal of Membrane Science 225 (2003) 63–76

Fig. 8. SEM picture of: (a) PS, (b) SPS, (c) PSf and (d) SPSf.

3.4. Thermal analysis tion from 147 to 120 ◦ C, poly (phenylene oxide) from
145 to 114 ◦ C and polysulfone from 185 to 140 ◦ C.
3.4.1. DSC spectra The unneutralized free acid forms (–SO3 H) of all the
The DSC spectra are shown in Figs. 5 and 6. In the sulfonated polymers used in the study display a signif-
case of polystyrene, the Tg value after sulfonation is re- icantly lower Tg than the unsulfonated ones. This may
duced from 87 to 30 ◦ C, polycarbonate shows a reduc- be ascribed to the structural changes introduced into
B. Smitha et al. / Journal of Membrane Science 225 (2003) 63–76 73

Table 1 Sulfonated polystyrene (Fig. 7) behaved differently


Tensile strength and elongation of all the polymers used in the with only one weight loss stage at 465–510 ◦ C. Weight
study
loss at this stage may be due to the splitting of the
Polymer Tensile strength (N/mm2 ) Elongation (%) main chains or it could also be related to the evapo-
SPPO 146.666 5.671 ration of residual solvent. Though the boiling point of
PPO 176.864 13.456 the solvent is much less than the temperature in the
SPSf 138.422 3.732 weight loss zone, the fact that the polymer is in a rub-
PSf 191.822 10.783 bery state cannot be overlooked. An increase in free
SPS 119.666 2.648
PS 140.321 8.965
volume in the rubbery state and the excessive sliding
SPC 157.277 7.952 of polymer chains one upon another with an increase
PC 182.333 19.734 in temperature would enhance evaporation of residual
solvent entrapped in the polymer at glassy state result-
ing in weight loss at a temperature much higher than
the polymer on account of sulfonation. The higher the its boiling point.
degree of substitution, the greater is the free volume
of a sulfonated product membrane enabling a change 3.5. Water and methanol sorption
in the state of polymer from more amorphous to more
crystalline resulting in the reduction of the glass tran- Table 2 shows the equilibrium percentage sorption
sition temperature. This effect is illustrated clearly by of water and methanol in the normal and sulfonated
the DSC spectra for polycarbonate and polystyrene polymers. Water sorption depends on the extent of
before and after sulfonation as shown in Figs. 5 and 6. sulfonation, hence higher the degree of substitution,
greater the water uptake. From the table it can be
3.4.2. Thermal stability by thermogravimetric observed that the water sorption prior to sulfonation,
analysis in the unmodified hydrophobic polymers was low and
Fig. 7 shows the thermogravimetric analysis spec- varied from as low a value as 0.01% in PSf to 0.27
tra of SPPO films prepared by using DMAc. The films in PPO. After sulfonation, the values obtained were
were dried at room temperature for several days and, considerably high and are in the range of 10.8–17.2%
before TGA analysis, the films were exposed to air indicating the presence of hydrophilic sites within the
at atmospheric conditions. Three weight loss stages at hydrophobic matrix. The swelling ratio for sulfonated
60–120, 220–270 and 420–460 ◦ C followed by the fi- polycarbonate was 1.17 followed by sulfonated poly
nal decomposition of polymer can be distinguished in (phenylene oxide) with 1.1.4 and sulfonated polysul-
the figure. The final decomposition begins at around fone with 1.138. An earlier study reported a swelling
460 ◦ C. The locations of the weight loss stages ob- ratio of 1.245 at 0.5 DS for sulfonated polysulfone
served here are almost similar to those reported by [7] which is comparable to the present case. For
Gilbert et al. [22]. For the films being exposed to air sulfonated polysulfone which are ideal for fuel cell
before subjecting to TGA, the weight loss in the first, membranes since excessive swelling can cause rup-
second and third stage can be attributed, respectively, ture of membranes losses of hydrogen/methanol by
to the loss of moisture absorbed from air, the decompo-
sition of sulfonic groups and the splitting of the main Table 2
Sorption values for water and methanol in the sulfonated polymers
chains before the final decomposition of the polymer.
In the case of SPC, the three weight loss stages were Polymer Water sorption (%) Methanol sorption (%)
found at 70–140, 260–410 and 470–500 ◦ C and the BSa ASb BS AS
final decomposition occurred at around 510 ◦ C. The
SPPO 0.27 14.37 1.12 1.14
causes for the weight loss are similar to the ones stated SPSf 0.01 13.84 0.23 4.62
earlier. SPS 0.16 10.78 0.44 1.08
Sulfonated polysulfone again shows three weight SPC 0.13 17.21 0.98 2.97
loss stages at 110–140, 195–250 and 325–360 ◦ C. Fi- a Before sulfonation (BS).
nal decomposition is observed at around 510 ◦ C. b After sulfonation (AS).
74 B. Smitha et al. / Journal of Membrane Science 225 (2003) 63–76

permeation through the barrier, and difficulties in gas value for SPC (0.5798) is highest and SPS (0.2799)
flow control through the fuel cell manifold/channels is the lowest. SPPO and SPSf showed moderate IEC
whereas very low swelling (<1.05) results in inef- values.
ficient proton transfer rates. Though the methanol Sulfonated polycarbonate appears to be a promising
sorption values increased considerably after sulfona- alternative proton exchange membrane since it shows
tion for polycarbonate from 0.98 to 2.975% and PSf a greater IEC value. Unlike Nafion which is expen-
from 0.23 to 4.62% compared to PPO and PS the sive, polycarbonate polymer is inexpensive and can
range for degree of swelling was still low enough be synthesized and sulfonated on a commercial scale
(1.01–1.046) to prevent excessive losses of methanol [13]. Characterization of SPC by UTM shows rea-
through the membrane in DMFC. sonable tensile strength (157 N/mm2 ) and flexibility
(percentage elongation ∼8). Additionally, the Tg of
3.6. SEM sulfonated polycarbonate is found to be much higher
than the operating temperature range (80 ◦ C) of a PEM
Fig. 8 shows the surface morphology of polystyrene fuel cell [24]. The crucial second stage weight loss
and polysulfone before and after sulfonation. The at 260–410 ◦ C caused by decomposition of sulfonic
enhancement in the size of pores after sulfonation of groups and depicted by TGA, proves the thermal sta-
the membranes can be observed clearly. Further, there bility of SPC.
appears to be a uniform distribution of pores through-
out the membranes after sulfonation. The number of
pores per sq.cm area for sulfonated polystyrene and 4. Conclusions
sulfonated polysulfone ranges from 2 to 3 and 3 to
4 millions, respectively. The pore size in sulfonated Introduction of hydrophilic sites in hydrophobic
polystyrene and sulfonated polysulfone are 2.5 and polymers with rigid aryl backbones appears to pro-
1.9 ␮m, respectively. Their uniform and adequate duce promising electrolyte membranes for fuel cell
size and even distribution ensure a desirable and ef- applications. Three of the four membranes tested,
ficient conductivity of protons besides a sufficiently except SPS, showed IEC values and other essential
large interfacial area between the hydrophobic and properties comparable to commercial Nafion mem-
hydrophilic interface [1]. brane. SPC appears to be a viable alternative owing
to its low cost, high IEC and DS, good mechanical
and thermal stability and low sorption for methanol.
3.7. IEC and DS
SPS showed a steep fall in glass transition tempera-
ture to a level which does not comply with requisite
Ion exchange capacity provides an indication of the
operating conditions in PEM fuel cells. However,
content of acid groups present in a polymer matrix [5]
this drawback can be overcome by crosslinking or
which are responsible for the conduction of protons
copolymerization with other suitable polymers. All
and thus is an indirect and reliable approximation of
the sulfonating agents produced membranes capable
the proton conductivity. Results are based on the sam-
of proton conduction but the best reagent–polymer
ple calculations reported in Appendix A [5]. The IEC
combinations appeared to be chlorosulfonic acid
and DS values of all the polymers used for the study
for PPO and PSf and acetyl sulfate for PS and PC.
are tabulated in Table 3. It can be seen that the IEC
Increase in concentration of sulfonating agents in-
creased the IEC and DS of the polymers but can-
Table 3
IEC and DS values of polymers prepared for the study
not ensure uniform distribution of sulfonic groups
within the polymer matrix. It may yield completely
Polymer IFC (meq./g) DS soluble membranes or polymer degradation may be
SPS 0.2799 0.03436 observed in some cases. Thus there is a general
SPSf 0.3976 0.04926 trend of reduction in Tg and mechanical strength.
SPPO 0.4217 0.05237
Extensive sulfonation leads to proportional increase
SPC 0.5798 0.07296
in interaction with water and ultimately at high
B. Smitha et al. / Journal of Membrane Science 225 (2003) 63–76 75

levels of sulfonation the polymer is soluble in wa- The degree of substitution can be calculated by the
ter. This is unacceptable for any proton exchange formula:
membrane as the fuel cell conditions demand a 120 × IEC
level of humidity to promote proton conduction and DS = = 0.03436
1000 + 120 × IEC − 200 × IEC
suitable modification in polymer structure are war-
ranted. IEC and DS values for sulfonated polycarbonate,
polysulfone and poly (phenylene oxide) were calcu-
lated on the same lines.
Acknowledgements
References
The authors would like to thank Council of
Scientific and Industrial Research, India, for funding [1] K.D. Kreuer, On the development of proton conducting
the Fuel Cell Project (ES/P 81-1-03, EMR Head) polymer membranes for hydrogen and methanol fuel cells, J.
Dr. Sridhar, Ms. Dhanuja, Ms. Bhavani, Membrane Membr. Sci. 185 (2001) 29–39.
[2] A.J. Appleby, F.R. Foulkes, Fuel Cell Handbook, Van
Separations Lab, Mr. Sai Babu, Ms. K. Uma Gaya-
Nostrand Reinhold, New York, 1989.
tri, Pheromones Lab and Mr. Tajuddin of Chemical [3] J. Larminie, A. Dicks, Fuel Cell System Explained, Wiley,
Engineering Division for membrane characterization. West Sussex, 2000.
The consistent support and encouragement of Dr. [4] W. Becker, G. Schmidt-Naake, Proton exchange membranes
M. Ramakrishna and Dr. K. Babu Rao of Chemical by irradiation grafting of styrene onto FEP and ETFE:
influences of the crosslinker N,N-methylene-bis-acrylamide,
Engineering Division is gratefully acknowledged. Chem. Eng. Technol. 25 (2002) 364–370.
[5] J.St. Pierre, D.P. Wilkinson, Fuel cells: new, efficient and
cleaner power source, AIChE J. 47 (1998) 1482–1486.
Nomenclature [6] R.J. Bellows, M.Y. Lin, M. Arif, A.K. Thompson, D.
Tg glass transition temperature Jacobson, Neutral imaging technique for in situ measurement
of water transport gradients within Nafion in polymer
electrolyte fuel cell, J. Electrochem. Soc. 146 (3) (1999)
1099–1103.
[7] A. Noshay, L.M. Robenson, Sulfonated polysulfone, J. Appl.
Polym. Sci. 20 (1976) 1885–1903.
Appendix A. Calculations DS and IEC [5] [8] C. Lopatin, H.A. Newey, Am. Chem. Soc., Div. Polym.
Chem., Polym. Prepr. 12 (2) (1971) 230.
Consider the case of sulfonated polystyrene: [9] N.H. Canter, Ger. Offen. 1 (915) (1969) 236.
[10] C.P.O. Farrell, G.E. Serniuk, US Patent No. 3,836,511
(1974).
• Initial conditions: [11] J. Veno, M. Watanabe, Y. Tanaka, Japanese Patent No. 24,998
◦ Mass of dry sulfonated polystyrene = 0.01429 g (1968).
◦ Normality of H2 SO4 = 0.01 N (N1 ) [12] C.J. Kibler, G.R. Lappin, US Patent No. 3,734,874
◦ Normality of NaOH = 0.01 N (N2 ) (1973).
[13] R. Salle, B. Sillion, French Patent No. 2,212,356 (1974).
• After 12 h: [14] J.P. Quentin, US Patent No. 3,709,841 (1974).
◦ Blank titration = 7.0 ml [15] J. Bourganel, US Patent No. 3,855,122 (1974).
◦ Sample titration = 6.2 ml [16] R.Y.M. Huang, J.J. Kim, Synthesis and transport properties
◦ Volume of NaOH neutralized = 10 ml of thin film composite membranes, J. Appl. Polym. Sci. 29
(1984) 4017–4027.
[17] Mitchell, Will, Fuel Cells, vol. 1, Academic Press, New York,
The ion exchange capacity of the sulfonated mem- 1963.
brane is calculated as follows: [18] B. Kruczek, T. Matsurra, Development and characterization
of homogeneous membranes to form high molecular weight
(blank titration − sample titration) SPPO, J. Membr. Sci. 146 (1998) 263–275.
× normality × 5 [19] U. Pedretti, A. Gandini, A. Roggero, C. Sisto, C. Valentini,
IEC = A. Stopponi, US Patent No. 5,169,416 (1992).
membrane weight [20] C.W. Plummer, G. Kimura, A.B. LaConti, Development of
(7 − 6.2) × 0.01 × 5 Sulfonated Polyphenylene Oxide Membranes for Reverse
= = 0.2799 Osmosis, General Electric Co., Lynn, MA, 1970.
0.1429
76 B. Smitha et al. / Journal of Membrane Science 225 (2003) 63–76

[21] N. Carretta, V. Tricoli, F. Picchion, Ionomeric membranes [23] Brandrup, Encyclopedia of Polymer Science and Engineering,
based on partially sulfonated poly(styrene), J. Membr. Sci. Wiley/Interscience, New York, 1985.
166 (2000) 189–197. [24] L. James, A. Dicks, Fuel Cell Systems Explained, Wiley,
[22] E.E. Gilbert, Sulfonation and Related Reactions, West Sussex, 2000.
Wiley/Interscience, New York, 1965, pp. 62–92.

You might also like