You are on page 1of 9

Journal of Polymers and the Environment

https://doi.org/10.1007/s10924-017-1146-7

ORIGINAL PAPER

Development of Chitosan Membranes as a Potential PEMFC Electrolyte


Karine N. Lupatini1,2,4 · Jéssica V. Schaffer1 · Bruna Machado1 · Eliane S. Silva1 · Luciana S. N. Ellendersen1 ·
Graciela I. B. Muniz1 · Ricardo J. Ferracin1,3 · Helton J. Alves1,2

© Springer Science+Business Media, LLC, part of Springer Nature 2018

Abstract
Commercial chitosan and chitosan extracted from shrimp shells are being used to design membranes to be tested as low cost
electrolyte in PEM fuel cells. This study investigated the influence of the deacetylation degree (DD) and molar mass (MV) of
the chitosans used in the composition of membranes on its performance regarding to proton conductivity and other properties.
Preliminary results indicate that the chitosan extracted from shrimp shells generated membranes with promising properties
such as proton conductivity, which demonstrated to be even a 100 times higher than those shown by commercial chitosan
membranes. The significant increase in proton conductivity can be associated with the higher number and availability of
amino groups (–NH2) in the chitosan produced in the laboratory, which presents higher DD and lower MV. It is believed that
the properties of chitosan can be manipulated in such a way that it would be possible to obtain proton conductivity values
closer to that presented by Nafion®.

Keywords Proton conductivity · Biopolymers · Hydrogen · Renewable energy · PEMFC

Introduction Proton exchange membrane fuel cell are the most efficient
power generation devices, once the electrochemical pro-
Fuel cells (FC) stand out among different emerging technol- cesses involved in PEM are not governed by Carnot Cycle.
ogies for sustainable energy generation by converting chemi- Therefore, the efficiency of these cells does not depend
cal energy directly into electrical energy [1]. Among the cell exclusively on the operating temperature of the system. High
types, the proton exchange membrane fuel cell (PEMFC) has efficiency can be achieved at relatively lower temperatures
gained considerable importance due to its great possibility of [3]. The commercial polymer electrolyte membrane cur-
applications in mobile. It has also been considered a solution rently used on this type of cell is Nafion® membrane which
for rural electrification, which enables the generation and presents proton conductivity (σ) of 9.0 × 10−2 S cm−1 at
isolated loads supply in areas where the electrical grid facili- 80 °C [4, 5]. Conductivity of Nafion® is originated from
ties are unviable by the distance from the consumer unit [2]. protons of sulfonic acid groups present at the end of its
chain. Then the water absorbed by the membrane allows the
protons transportation [6].
* Karine N. Lupatini
karinelupatini@gmail.com The most accepted and illustrative theoretical model is
the “cluster network model”, in which sulfonate groups
1
Laboratory of Catalysis and Biofuel Production (SO3−) fixed on the Nafion® structure form ionic clusters
(LabCatProBio), Postgraduate Program in Bioenergy - approximately spherical. During the transportation by
Federal University of Paraná (UFPR - Setor Palotina), R.
Pioneiro, 2153, Jardim Dallas, Palotina, PR 85950-000, structural diffusion or Grotthus mechanism, protons jump
Brazil from a hydrated ionic site (SO3− H3O+) to another through
2
Postgraduation Program in Engineering of Energy
the membrane. It occurs a constant formation and cleavage
in Agriculture, Western Paraná State University, Cascavel, of hydrogen bonds between the proton, the water mol-
PR, Brazil ecules and the SO3− of the structure. Thus, the proton is
3
Itaipu Technological Park (PTI), Itaipu Binational, transported migrating through the water cluster structures
Foz do Iguacu, PR, Brazil in equilibrium towards the cathode, due to the electric field
4
R. Dr. João Gonçalves Padilha, 811, Centro, Pitanga, effect. Whereas in the vehicular transport mechanism, the
PR 85200-000, Brazil

13
Vol.:(0123456789)
Journal of Polymers and the Environment

proton is carried by a “vehicle” which leads it from one Few studies employ chitosan membranes in protons con-
point to another within a continuous medium. Considering duction due to the low values demonstrated (σ between 10−7
proton in an aqueous medium, the transporters are water and 10−3 S cm−1). Accordingly, it is necessary to obtain
molecules that, when diffused carry the protons that are composites from the matrices of other polymers, in order
solvating [7]. to achieve the most suitable properties for this purpose [5,
Although in the recent years fuel cells (FC) have wide- 8, 9]. Commonly, studies use membranes made from low
spread and emerged as high-efficiency devices in electric DD and high MV commercial chitosan, which serves exclu-
power generation, it still has a high production and mainte- sively as parameters to evaluate the behavior of the compos-
nance cost, and it also employs materials that originate from ite biopolymer.
non-renewable sources. In this way, Nafion® becomes a great A research conducted by Osifo and Masala [15] aimed
example of this conception [4, 5]. at evaluating the influence of different DD on pure chitosan
It is important to think about actions to reduce the cost membranes with high MV used in direct methanol fuel cell
of FC production making it more efficient or sustainable. In (DMFC). The authors suggested that chitosan membranes
this regard, the use of alternative low cost raw materials in with high DD must be used in high operating temperatures
PEMFC electrolyte production which are able to conduct of the cell. In the other hand, chitosan with lower DD should
protons efficiently at a temperature of 80 °C (temperature be applied for low operating temperatures, since the result-
limit for Nafion®) or higher than this has been the subject ing membranes presented different thermal properties and
of several studies, specially in the obtaining of polymeric flux of protons. However, there are few studies in the litera-
composite materials [5, 8, 9]. ture that allow a better analysis of the effects and advantages
A promising alternative is the chitosan, a natural biopol- of using chitosan in obtaining membranes for electrolyte use
ymer predominantly consisting of 2-amino-2-deoxy-D-glu- in PEMFC.
cose and 2-acetamido-D-glucose (copolymer), derived from The aim of the present study was to develop chitosan
the N-deacetylation of chitin one of the most abundant poly- membranes extracted from shrimp farming waste to be used
saccharides in nature. It is mainly found in the crustacean as an electrolyte in PEMFC. It was investigated the influ-
shells, invertebrate skeletal structures, diatom algae and cell ence of the DD and MV of the biopolymer on its perfor-
walls of some fungi [10]. About 4.1 million tons/year of mance regarding to the proton conductivity besides other
shrimp are produced worldwide generating about 1.2 ton/ parameters.
year of shells rich in chitin, which are considered an envi-
ronmental liability once it is frequently discarded in landfills
or in other inappropriate ways [11]. Materials and Methods
Amino (–NH2) and hydroxyl (–OH−) groups present in
chitosan structure facilitates its interaction with other mate- Materials
rials/hydrophilic groups, and thus it is possible to make
structural changes to diversify the possible chemical inter- Commercial chitosan with an average molar mass (Sigma-
actions [9]. Aldrich, > 99.8%); hydrochloric acid (Isofar, 36.5–38%);
In acid medium, chitosan is soluble and becomes a type sodium hydroxide (Synth, 98.0%); glacial acetic acid
of polyelectrolyte positive charge carrier due to the presence (Fmaia, 99.5–100.5%); sodium acetate (Isfar, 99.0%); sul-
of amino groups which act as a Lewis base, being easily furic acid (Synth, 95.0–98.0%); distilled water (1.99 μs).
protonated (–NH3+). In addition, the great availability, low
cost, possibility of structural modification, thermal stability Commercial Chitosan (CA) and the Chitosan
at temperatures below 150 °C, and easy manufacturing of Produced in the Laboratory (CB)
thin and resistant membranes make chitosan an interesting
polymeric material to be used as an electrolyte in H2–O2 A sample of commercial chitosan (CA) with an average
FC [10, 12]. molecular mass was used as reference.
Physico-chemical properties of the biopolymer mem- Aiming to obtain chitosan in laboratory (CB), first of all
branes can be affected by parameter such as the degree of chitin was extracted from freshwater shrimp’s exoskeletons
deacetylation (DD), the chitosan molar mass (MV) the semi- (shells). It was utilized the species known as Macrobrachium
crystalline character and the reticulation level resulting from rosenbergii from a shrimp farming in western Parana State,
the conditions employed in the membranes preparation. A in Brazil. The shells were washed in running water, dried in
high DD means that a larger quantity of amino groups is an electrical oven at 60 °C for 24 h and then milled in a dry
available, and the chains are more extended favoring the grinding ball mill. It was ground until it is completely sieved
protons conduction. The same phenomena can be observed in a 63 µm opening sieve. Obtained materials was subjected
with the decrease in crystallinity and MV [13, 14]. to demineralization and deproteinization according to the

13
Journal of Polymers and the Environment

method described by Tolaimate et al. [16]. To ensure the 4.0 M sulfuric acid solution for 15 h. The reticulated mem-
obtaining of a high DD and low MV chitosan, the resulting branes were neutralized again, washed and dried under the
chitin was deacetylated in a 50% NaOH solution (m V−1) in same conditions mentioned above.
a proportion of 2.5% (m V−1) under reflux for 10 h at 120 °C. This study evaluated four types of membranes, with avail-
The reaction product was washed in a neutral pH and dried able area of 5 cm × 5 cm, differing from each other with
at 60 °C electric oven for 24 h [13]. respect to the chitosan used (CA or CB) and the reticulation
A triplicate conductometric titration was performed to reaction applied. Hence, the tested membranes were termed
determine the DD percentage of the chitosan samples. It was as: MCAN (commercial/neutralized); MCAR (commercial/
used 200 milligrams of biopolymer in 40 mL of 0.05 mol reticulated); MCBN (laboratory/neutralized); and MCBR
L−1 hydrochloric acid solution, and it was kept under orbital (laboratory/reticulated).
shaking for 18 h at 25 ± 0.1 °C. The samples were titrated
in 0.17 mol L−1 sodium hydroxide solution. Thereafter the
Eq. 1 was used for DD determination (%): Description of MCA and MCB Chitosan Membranes
16.1 × [base] × (V2 − V1 ) X-ray diffraction (XRD) analyzes were conducted on a Shi-
DD = (1)
m madzu diffractometer (Model XRD-6000) in the range of
where V1 is the volume of NaOH (in mL) used for neutral- 5° < 2θ < 70°, with CuKα radiation (λ = 1.5406 Å; 40 kV;
izing the excess of HCl, V2 is the volume of NaOH (in mL) 30 mA) and continuous scanning speed of 1.5° min−1. In
required for neutralization of the protonated chitosan acid order to improve the results interpretation, the analysis was
groups (–NH3+), base is the used concentration of NaOH also performed on CA and CB powder samples [18].
(in M), and m is the mass of the chitosan sample added to Infrared spectroscopy (FTIR) the identification of func-
the acidic solution. V1 and V2 were determined based on tional groups and their possible displacements were evalu-
inflections observed in the conductivity graphs (µS cm−1) × ated in the spectral range between 600 and 4000 cm−1 and
NaOH (mL) consumed [10]. resolution of 4 cm−1 in a Vertex 70 spectrophotometer
The average viscosimetric molar mass was obtained by (Bruker, USA) with attenuated total reflectance (ATR), and
determining the flow time of the solvent and dilute chitosan equipped with a carbon crystal at 25 ± 2 °C. The membranes
solutions (0.31–0.73 mg mL−1) in 0.3 M HAc/NaAc 0.2 M. were placed in a holder and it was compressed at high and
Tests were conducted in triplicate using a Ubbelohde vis- controlled pressure. Each graph produced represents the
cometer with a glass capillary (φ = 0.44 mm) immersed in a result of the average of sixty-four (64) points collected from
thermostated bath with temperature kept at 25 ± 0.1 °C. The the same sample [9].
relationship between the intrinsic viscosity [η] (dL g−1) and Water absorption the coefficient of water absorption was
the average viscosimetric molar mass of the polymer MV determined by the ratio between the mass of dry and wet
(Da) was established by Mark–Houwink–Sakurada’s equa- membrane according to Eq. 3, in which ΔM is the coef-
tion below [17]: ficient of water absorption, Mw is the mass of the wet mem-
brane, and Md is the mass of the dry membrane. The initial
[ ]
̄𝛂
η = KM (2) dry weight was determined by thermal treatment in electric
V
oven at 60 °C for 24 h. Subsequently, the membranes were
where K and α are constants for a given polymer–solvent hydrated at temperatures of 40–70 °C for 1 h followed by
system which in the case of chitosan varies according to the excess water removal and weighing the membranes [9].
DD as well as with the molar mass range [12].
Mw − Md
ΔM = × 100 (3)
Md
Preparation of Chitosan Membranes MCA and MCB
Proton conductivity σ (S cm−1) it was determined by
To obtain each membrane 2.0 g of chitosan were dissolved electrochemical impedance analysis using a testing station
in 150 mL of 15% acetic acid (m V−1) in a thermostated (RSI), described by Paganin et al. [19], and a PGSTAT 30
bath at 50 °C. After that it was filtered to remove undis- Autolab potentiostat (Eco Chemie) with associated Fre-
solved material, and then the solution was transferred into quency Response Analyzer (FRA). The measurements were
a silicone mold and dried in an electric oven at 60 °C for performed in open circuit voltage (OCV) with scan rate of
24 h. The resulting membranes were neutralized in a 0.2 M 1 mV s−1 in a frequency range of 1.0 MHz–0.1 Hz at the
NaOH solution for 30 min, washed with distilled water until temperature of 25, 40, 60 and 80 °C, with humidified hydro-
it reached neutral pH (7) and then it was stretched for drying gen gas (prepurified > 99.95%) and oxygen (> 90%). Gas dif-
at ambient temperature of 25–30 °C for 24 h [5]. After the fusion layer (GDL) was used as electrodes along with a com-
neutralization, some of the membranes were reticulated in mercial catalyst (Vulcan XC 72) (0.4 mg cm−2 of Pt, 30 wt%

13
Journal of Polymers and the Environment

Pt/C), and a 0.87 mg cm−2 Nafion®. Membrane resistance


values were determined directly from Nyquist diagrams. The
imaginary component (Z″, y axis) is plotted versus the real
component (Z′, x axis) for each frequency, in which Z′ is the
resistivity of the material (R). The proton conductivity (in S
cm−1), was determined according to Eq. 4:
1 L
σ= . (4)
R A
where L is the distance between the electrodes (cm), A is the
membrane area (cm2) and R is the electric resistance (Ω).

Results and Discussion

DD and MV of the CA and CB Chitosans Fig. 1 MCB membranes in drying process at ambient temperature

According to the conductometric titration curves of the


CA and CB chitosans, the DD calculated from Eq. 1 was
64.3 ± 1.4% for CA and 76.0 ± 0.8% for CB. These results
confirm the efficiency of the chosen method to obtain CB,
once the employed conditions in the deacetylation reaction
provided greater DD than the CA chitosan.
Viscosity measurements were based on solvent flow time
and diluted chitosan solutions. For CA and CB, we adopted
K = 0.074 and α = 0.76, based on obtained DD values [17].
The values of η were determined by the linear slope of the
curve established between samples concentration (g mL−1)
and reduced viscosity (mL g−1) reaching the following val-
ues for CA and CB, respectively: 825.56 and 207.46. There-
fore, MV values for CA and CB were 220.8 and 63.07 kDa,
respectively, indicating that besides CB has a higher DD it
also demonstrates a low MV (more than 3 times lower than in
CA). It is worth pointing out that these results were consist-
ent with the initial aim since larger DD values and smaller Fig. 2 Powder chitosan diffractograms (CA and CB) and chitosan
MV for CB should result in an increase in the number of membranes diffractograms (MCAN, MCAR, MCBN e MCBR)
amino groups available for conducting protons.

Description of the MCA and MCB Chitosan to its peak corresponding to plan (010) presents a lower
Membranes intensity [10, 13]. In general, it is observed that the dif-
fractograms of chitosan membranes are analogous to each
Regarding to appearance, the MCA and MCB chitosan other and quite different from those of the powder samples.
membranes showed an excellent aspect: transparent and No clear crystalline peaks were observed for the mem-
free of bubbles and particles. Figure 1 shows a picture of branes, only large amorphous halos with main reflections
the obtained membranes. in two areas of the diffractograms (at 2θ = 19°–21° and
X-ray diffraction (XRD) the main crystalline CA and 27°–31°). It showed the reduction of the original semi-
CB peaks are located at 2θ = 9.42° and 20° (Fig. 2). The crystalline character of CA and CB. MCB membranes
observed peak at approximately 10° corresponds to the have diffractograms with faster response, indicating a
crystal planes (010), attributed to orthorhombic crystals more amorphous character than MCA. According to Agu-
derived from α-chitin semi-crystalline structure found in iar et al. [20], it favors its application as electrolyte in FC
shrimp shells. The crystalline high intensity peak observed once amorphous structures tend to have superior proton
at about 20° corresponds to planes (110) and (020). Thus, conductivity. Regarding to reticulation, it couldn’t be said
it can be stated that CB is more amorphous than CA due that it has changed biopolymer’s chains arrangement in

13
Journal of Polymers and the Environment

Fig. 4 MCA and MCB membrane’s water absorption at different tem-


peratures
Fig. 3 FTIR of MCA and MCB membranes

the membranes, which remain similar to neutralized mem- the reticulated membranes with H2SO4 (MCAR and MCBR).
branes (without reticulation). In other words, the promoted reaction between the reticulat-
Infrared spectroscopy (FTIR) infrared spectrum (Fig. 3) ing agent and the amino groups of these membranes would
shows the main absorption bands of MCA and MCB mem- benefit cross-chain formation and consequently reduce water
branes, which are located in: 3500–3300 cm −1, attrib- absorption [5]. However, there is an opposite behavior in
uted to axial deformation of hydroxyl and amine groups; which the reticulated membranes had higher water absorp-
2940–2867 cm−1, corresponding to the C–H stretching; tion values than the non-reticulated (MCAN and MCBN).
1641 cm−1, typical amide I band due to carbonyl stretching; This occurrence can be explained by the fact that cross-chain
1556 cm−1, related to amide II due to the vibrations in the formations resulting from the interaction between the proto-
NH bond plan and to the CN stretching [21–26]; 1411 cm−1, nated amino groups (–NH3+) of chitosan chains (in contact
corresponding to the CH2 and/or CH3 stretching; next to with acid) and sulfate group SO42−, change the hydrophilic
1326, 1256 and 1374 cm−1, attributed to the amine I, II characteristics of chitosan, attracting water molecules and
and III stretching, respectively; 1089 and 1035 cm−1, cor- promoting membrane swelling. It is important to mention
responding to the C–O–C or COH stretching; and finally that since MCB membranes have a higher DD it makes the
895 cm−1, attributed to the CH deformation out of the plan recirculation occurs more effectively leading to the observed
[9]. The spectrums of the membranes show similarities difference between the water absorption values of MCBN
between each other, and all of the observed bands were simi- and MCBR membranes. Another issue that should be stated
lar to those reported in the literature for chitosan membranes is the fact that reticulation reduces the crystalline domains
[9, 26]. In this way, it is possible to assert that the reticula- of the semi-crystalline structure of chitosan, which can
tion did not change the functional groups which are chitosan make the membrane more receptive to water molecules
membrane characteristics. However, both MCA and MCB capture. Furthermore, all membranes showed greater water
have a very low transmittance. absorption rate at the higher temperatures tested (70 °C).
Water absorption MCB chitosan membranes had lower The increase in cell operating temperature directly affects
percentage of water absorption (80–160%) than MCA mem- the electro-osmotic flow mechanism, since with increasing
branes (160–180%). The same effect was observed by Osifo temperature the pores and channels present in the polymeric
and Masala [15] and it may be attributed to the greater DD structure of the chitosan membrane expand allowing greater
of the used CB chitosan. This fact can be explained by the absorption of water. Therefore, the temperature acts directly
greater number of available amino groups, which promote a on membrane hydration [5, 7, 27]. Nonetheless, as reported
greater occurrence of intermolecular forces existing between in the literature, a cell operation at close to 100 °C may
the biopolymer chains, mainly hydrogen bonds between cause dehydration of the membrane making it difficult to
–OH− and –NH 2 groups, making its three-dimensional form channels through which the protons would be trans-
structure more compact. The higher compression hinders ported. Consequently, it might cause the loss of conductivity
the entry of water molecules in these membranes (Fig. 4). It [28, 29]. In all assays, the standard deviation values were
might be expected this behavior was even more enhanced in below 3%.

13
Journal of Polymers and the Environment

Proton conductivity proton conductivity was determined -1


10
according to Eq. 4, and it is correlated to the resistance of the MCAR (32
solid electrolyte R. In the Nyquist diagram the point used in MCAN (30
MCBR (47
each graph (Fig. 5) is the one that appears on the real axis (x MCBN (25
axis), in which the resistor is not conditioned to frequency. In

)
-2 - - - Nafion (51
10

-1
Conductivity (S cm
other words, the one that at high frequencies Z″ is equal to
zero and Z′ is the actual membrane resistance (R) [19]. The
resistivity of the material is inversely proportional to the pro-
ton conductivity; thus, it is noted that Z′ of MCBN and MCBR -3
10
membranes at 60 and 80 °C is between 0 and 1 Ω indicat-
ing higher conductivity values. On the other hand, MCAN e
MCAR membranes showed Z′ between 3 and 10 Ω, indicating
lower conductivity values. The proton conductivity (σ) and the -4
10
20 30 40 50 60 70 80
MCA e MCB membranes thickness at different temperatures
Temperature (°C)
are found in Fig. 6. The thickness of the chitosan membranes
ranged between 25 and 47 µm which are lower values if com-
pared to the ones typically found for Nafion®. For all tested Fig. 6 Proton conductivity (σ) at different temperatures and thickness
(measurement was carried out in seven random points of each mem-
membranes, the increase in temperature resulted in higher σ brane using Regmed (electronic equipment) and following the rec-
values, as also observed by Luo et al. [30]. The proton con- ommendations of Standard T 411 om-97. The results were expressed
ductivity of MCA membranes in general was extremely low as the arithmetic average of the measurements [33–35]) of chitosan
even at higher temperatures. The values were restricted to the membranes
magnitude order of 10−4 S cm−1, far below Nafion®’s conduc-
tivity. Similar results were found by Rahman et al. [24] (10−5 exchange is further reduced due to its reaction with reticulating
S cm−1 for chitosan and cellulose membranes), Wan et al. [26] agent, shown in Fig. 7. Better results of proton conductivity
(10−3–10−5 S cm−1 for pure chitosan membranes and phospho- for MCB membranes (up to a 100 times greater) was achieved
rylated chitosan membranes) and Vijayalekshmi and Khastgir with larger DD and smaller CB chitosan molar mass. It can be
[31] (10−4 S cm−1 for pure chitosan membrane and composite concluded that the greatest amount of –NH2 and its favorable
membranes of chitosan/methane sulfonic acid and chitosan/ arrangement in the polymer chain benefited protons conduc-
sodium salt of dodecylbenzene sulfonic acid). The reticula- tion in these membranes, reaching values up to to 1.9 × 10−2 S
tion effect in MCA membranes was not favorable for proton cm−1 at 80 °C (same magnitude order of Nafion®). Even con-
conductivity, which may be justified due to its smaller amount sidering the worst MCB membrane result for conductivity, it is
of –NH2 groups induced by the low degree of deacetylation of superior to the best MCA result. Unlike the MCA membrane,
chitosan CA. As noted by Bispo [32] after reticulation, the pro- reticulation did not negatively influence the conductivity of the
portion of the –NH2 groups available to carry out the proton material in MCB membrane due to the larger amount of amino

10
(a) 60 °C (b) 80 °C
MCAR MCAR
8 MCAN MCAN
MCBR MCBR
MCBN MCBN
6

0
0 2 4 6 8 10 0 2 4 6 8 10

Fig. 5 Electrochemical impedance spectroscopy for MCA and MCB membranes at 60 °C (a) and 80 °C (b)

13
Journal of Polymers and the Environment

Fig. 7 Representation of the


crosslinking reaction of chitosan
with sulfuric acid [5]

groups available. It is possible to emphasize that only MCBR proton conductive membranes with performance similar
membrane presented an increase in the σ value at temperatures or higher than Nafion®. The control of the conditions to
of 60–80 °C, which may be associated with a better stability obtaining chitosan from the chitin extracted from shrimp
at higher temperatures due to reticulation. Comparing Figs. 5 shells allowed membranes for PEMFC to be produced with
and 6, it is noted the increase material resistance (Z′) results superior characteristics compared to commercial. Chitosan
in lower proton conductivity values. Table 1 presents some membrane obtained from shrimp shells reached protonic
researches using chitosan membrane as polymer electrolyte. conductivity values of: 1.6 × 10−2 and 1.9 × 10−2 S cm−1
It is noteworthy, the results for pure chitosan membranes were at 60 and 80 °C, respectively. These results were the clos-
originated from studies with composite membranes of chitosan est to the performance achieved by Nafion ® (9.0 × 10 −2
considering that the preparation of pure chitosan membranes S cm−1), demonstrating potential use as an electrolyte in
was performed only for comparison, without exploring the PEMFC. It is important to highlight that further assays
individual characteristics of chitosan. should be conducted in future studies in order to deter-
mine the chemical, mechanical and thermal stability of
the membranes.
Conclusions
Acknowledgements The authors gratefully acknowledge PTI Foun-
dation (Brazil) for the financial support and for the granting scholar-
Considering the possibility of exploitation and chemi- ships (Notice 058/2014 N. 014/2014). We also gratefully Mr. Valdecir
cal modifications of its structure, the chitosan obtained Antônio Paganin from University of São Paulo (USP/IQSC) for provid-
from shrimp farming waste showed to be a very attrac- ing the structure and equipment for the research development.
tive alternative for researches seeking the development of

13
Journal of Polymers and the Environment

Table 1 Performance of chitosan membranes in fuel cells at 25 ± 2 °C

Membrane Degree of deacetyla- Conductivity (S cm−1) References


tion (%)

Chitosan 85 0.009 [5]


Chitosan/polysulfone 0.034
Chitosan/sulphonated polysulfone 0.046
Chitosan/phosphotungstic acid 81 0.00577 [25]
Chitosan/phosphotungstic acid/montmorillonite 0.00225
Chitosan/phosphotungstic acid/montmorillonite/silane 5% 0.00617
Chitosan/phosphotungstic acid/montmorillonite/silane 10% 0.00484
Chitosan/phosphotungstic acid/montmorillonite/silane 15% 0.00292
Chitosan 90 0.0238 [9]
Chitosan/molecular sieves 0.0144–0.0181
Chitosan/PVA 0.0169
Chitosan/PVA/molecular sieves 0.0127–0.0225
Chitosan/TiO2 0.0168
Chitosan/montmorillonites 0.0065–0.0187
Chitosan 95 0.0035 [36]
Chitosan/phosphomolybdic acid 0.015
Chitosan/phosphotungstic acid 0.0122
Chitosan/silicotungstic acid 0.01
Chitosan 80 0.021 [37]
Chitosan/zeolites 0.017–0.020
Chitosan/phosphotungstic acid – 0.018 [38]
Chitosan (2%)/phosphotungstic acid—crosslinked for 30 s – 0.00791 [39]
Chitosan (2%)/phosphotungstic acid—crosslinked for 1 min 0.00807
Chitosan (2%)/phosphotungstic acid—crosslinked for 5 min 0.00649
Chitosan (2%)/phosphotungstic acid—crosslinked for 10 min 0.00688
Chitosan (2%)/phosphotungstic acid—crosslinked for 20 min 0.00742
Chitosan (2%)/phosphotungstic acid—crosslinked for 40 min 0.00877
Chitosan (2%)/phosphotungstic acid—crosslinked for 60 min 0.0104
Chitosan 90 0.0029 [40]
Chitosan/silica supported silicotungstic acid—1 0.0034
Chitosan/silica supported silicotungstic acid—3 0.0039
Chitosan/silica supported silicotungstic acid—5 0.0058
Chitosan/silica supported silicotungstic acid—7 0.0015

– Not specified

References 6. Khoshkroodi LG (2010) Polymer electrolyte membrane degrada-


tion and mobility in fuel cells: a solid-state NMR investigation.
Doctoral Thesis, University of Stuttgart, Alemanha
1. Ma J, Sahai Y (2013) Carbohydr Polym 92:955
7. Perles CJ (2008) Rev Polímeros 4:281
2. Andrade SBM, Ladchumananandasivam R, Nascimento RM
8. Santamaria M, Pecoraro CM, Di Franco F, Di Quarto F, Gatto I,
(2010) Extração e caracterização de quitina e quitosana e a
Saccà A (2016) Int J Hydrog Energy 41:5389
sua utilização na fabricação de nanofibras. In: V Congresso
9. Vicentini DS (2009) Effect of inclusion of molecular sieves, pol-
Nacional de Engenharia Mecânica, Campina Grande, paraíba,
yvinyl alcohol, montmorillonites and titanium dioxide chitosan
18–21 August 2010
membranes. Doctoral Thesis, Federal University of Santa Cata-
3. Wang C, Wang S, Peng L, Zhang J, Shao Z, Huang J, Sun C,
rina, Florianópolis, Brazil
Ouyang M, He X (2016) Rev Energies 9:603
10. Campana-Filho SP, Britto D, Curti E, Abreu FR, Cardoso MB,
4. Malis J, Mazúr P, Paidar M, Bystron T, Bouzek K (2016) Int J
Battisti MV, Sim PC, Goy RC, Signini R, Lavall RL (2007) Quím
Hydrog Energy 41:2177
Nova 30:644
5. Smitha B, Devi A, Sridhar S (2008) Int J Hydrog Energy
11. Bessa-Junior AP, Gonçalves AA (2013) Actapesca 1:13
33:4138

13
Journal of Polymers and the Environment

12. Santos JE, Soares JP, Dockal ER, Campana-Filho SP, Cavalheiro trocadora de prótons (PEMFC) por impressão à tela. Dissertation,
ETG (2003) Rev Polimeros 13:242 Master’s in Science, University of São Paulo, São Paulo, Brazil
13. Arantes MK, Kugelmeier CL, Cardoso-Filho L, Monteiro MR, 29. Matos BR (2008) Preparação e caracterização de eletrólitos com-
Oliveira CR, Alves HJ (2015) Polym Eng Sci 55:1969 pósitos Naion/TiO2 para aplicação em células a combustível de
14. Tavares IS (2011) Obtenção e caracterização de nanopartículas de membrana de troca protônica. Dissertation, Master’s in Science,
quitosana. Dissertation, Master’s in Chemical, Federal University University of São Paulo, São Paulo, Brazil
of Rio Grande do Norte, Natal, Brazil 30. Luo Z, Chang Z, Zhang Y, Liu Z, Li J (2010) Int J Hydrog Energy
15. Osifo PO, Masala A (2012) J Fuel Cell Sci Technol 9:1 35:3120
16. Tolaimate A, Desbrieresb J, Rhazia M, Alaguic A (2003) Rev 31. Vijayalekshmi V, Khastgir D (2017) J Membr Sci 523:45
Polym 44:7939 32. Bispo VM (2009) Estudo do Efeito da Reticulação por Genipin
17. Kassai MR (2007) Carbohydr Polym 68:477 em suportes biocompatíveis de Quitosana-PVA. Doctoral Thesis,
18. Ribeiro C, Scheufele FB, Espinoza-Quiñones FR, Modenes AN, Federal University of Minas Gerais, Minas Gerais, Brazil
Silva CMG, Vieira MGA, Borba CE (2015) Physicochem Eng Asp 33. Cardoso MT, Carneiro ACO, Oliveira RC, Carvalho AMML, Pat-
482:693 rício Júnior W, Martins MC, Santos RC, Silva JC (2012) Cienc
19. Paganin VA, Oliveira CLF, Ticianelli EA, Springer TE, Gonzales Florest 22:403
ER (1998) Electrochim Acta 43:3761 34. Carpiné D, Dagostin JLA, Bertan LC, Mafra MR (2015) Food
20. Aguiar KR, Batalha GP, Peixoto M, Ramos A, Pezzin SH (2012) Bioprocess Technol 8:1811
Rev Polimeros 22:453 35. TAPPI—Technical Association of the Pulp & Paper Industry
21. Amaral IF, Granja PL, Barbosa MA (2005) J Biomater Sci Polym (1997) T 411 om-97—Thickness (caliper) of paper, paperboard,
Ed 16:1575 and combined board
22. Oliveira PN, Mendes AMM (2016) Mater Res 19:954 36. Cui Z, Xing W, Liu C, Liao J, Zhang H (2008) J Power Sources
23. Witt MA, Barra GMO, Bertolino JR, Pires ATN (2010) J Braz 188:24
Chem Soc 21:1692 37. Wang J, Zheng X, Zheng HWB, Jiang Z, Hao X, Wang B (2008)
24. Rahman NFA, Loh KS, Mohamad AB, Kadhum AAH, Lim KL J Power Sources 178:9
(2016) Malays J Anal Sci 20:885 38. Santamaria M, Pecoraro CM, Di Quarto F, Bocchetta P (2015) J
25. Permana D, Purwanto M, Ramadhan LOAN, Atmaja L (2015) J Power Sources 276:189
Chem 15:218 39. Pecoraro CM, Santamaria M, Bocchetta P, Di Quarto F (2015) Int
26. Wan Y, Creber KAM, Peppley B, Bui VT (2003) Rev Polym J Hydrog Energy 40:14616
44:1057 40. Vijayalekshmi V, Khastgir D (2018) Energy 142:313
27. Liu L, Chen W, Li Y (2016) J Membr Sci 504:1
28. Andrade AB (2008) Desenvolvimento de conjuntos eletrodo-
membrana-eletrodo para células a combustível a membrana

13

You might also like