Uploaded by endahrd

Ameliorated polyvinylidene fluoride based proton exchange membrane impregnated with graphene oxide, and cellulose acetate obtained from sugarcane bagasse for application in microbial fuel cell

advertisement
Journal of Environmental Chemical Engineering 9 (2021) 106681
Contents lists available at ScienceDirect
Journal of Environmental Chemical Engineering
journal homepage: www.elsevier.com/locate/jece
Ameliorated polyvinylidene fluoride based proton exchange membrane
impregnated with graphene oxide, and cellulose acetate obtained from
sugarcane bagasse for application in microbial fuel cell
Mukesh Sharma a, Pranjal P. Das b, Trishla Sood b, Arun Chakraborty a, Mihir K. Purkait b, *
a
b
Centre for Oceans, Rivers, Atmosphere and Land Sciences, Indian Institute of Technology Kharagpur, Kharagpur, West Bengal, India 721302
Department of Chemical Engineering, Indian Institute of Technology Guwahati, Guwahati, Assam, India 781039
A R T I C L E I N F O
A B S T R A C T
Editor: Dr. G. Palmisano
Polyvinylidene fluoride (PVDF) based, Cellulose acetate (CA), and Graphene oxide (GO) doped membrane acts as
a superior cost-effective cation exchange membrane (CEM) for the energy extraction application using microbial
fuel cells (MFCs). In this study, the waste sugarcane bagasse was used to prepare CA, and the Tour method was
used to prepare GO. CA and GO compositions were varied to prepare CEMs WCA-1, WCA-2, WCA-3, and WCA-4.
Doping the functionalized GO material in the hydrophilic membranes enhances the contribution of conducting
phase concerning its hydrophilic nature and facilitates its functional characteristics at the surface. Further,
adding an adequate amount of GO improves the physicochemical properties of the membrane and its mechanical
stability. Also, with the fusion of GO in the membrane, the trend for water uptake ability significantly increases.
Post sulfonating the prepared membranes the enhancement in the attained proton conductivity values from
0.025 Scm-1 for WCA-1 to 0.4 Scm-1 for WCA-4 was observed, which may be attributed to the increase in the GO
content in the prepared membranes. Among various prepared membranes, WCA-3 was considered for MFC
application. The WCA-3 membrane outperformed several membranes for wastewater treatment and energy
extraction using MFC. Columbic efficiency (CE) of 7.1%, chemical oxygen demand (COD) removal efficiency of
97.5 ± 0.8%, and power density of more than 150.22 mW m− 2 were achieved without any electrode modifi­
cation. Thus, the prepared low-cost novel PVDF-based, GO and CA membrane (WCA-3) elucidated its appro­
priateness for proliferating the efficacy of MFC and is recommended for scaling up of MFCs.
Keywords:
Cation exchange membrane
Microbial fuel cell
Wastewater treatment
Proton conductivity
1. Introduction
Rapid population growth has led to rising energy demands and
environmental concerns, necessitating the development of renewable
and sustainable energy production techniques. Advances in microbial
fuel cell (MFC) technology have been the subject of extensive investi­
gation, with promising outcomes. MFC is a bio-electrochemical reactor
system that utilizes electrons liberated in the biochemical oxidation of
organic substrates catalyzed by anaerobic microbes [1,2]. A conven­
tional MFC reactor comprises an anaerobic biotic anode chamber, an
aerobic biotic or abiotic cathode chamber, and a separator (such as a
proton exchange membrane (PEM)) [3]. The active biocatalyst in the
anodic chamber anaerobically oxidizes organic matter in wastewater to
produce electrons and protons. Protons are transported to the cathodic
chamber through the PEM. The external circuit conducts the electrons to
the cathode, completing the electrical circuit. At the cathode, electrons
and protons react in the presence of oxygen (or another electron
acceptor), which gets reduced to water [4]. MFCs use a variety of
inorganic compounds in addition to the fuel to enhance microbial
metabolism, which dissociates in the aqueous phase to create cations
such as K+, Na+, and NH4+ [5,6]. PEMs have a crucial role in selectively
allowing proton flow to the cathode while resisting the migration of
other ions [7–10]. Moreover, they must restrict oxygen flow from the
cathodic to the anodic chamber, ensuring that anaerobicity in the anodic
chamber is undiminished [7,8,13]. Poor power generation, low
coulombic efficiency (CE), and high cost of materials have hindered
practical applications of MFCs [14]. Nafion, the most common PEM used
for MFC applications, is permeable to cations such as Na+ and K+
leading to pH gradients, susceptible to oxygen transfer from anode to
cathode, expensive, and physically unstable at temperatures above
* Corresponding author.
E-mail address: mihir@iitg.ac.in (M.K. Purkait).
https://doi.org/10.1016/j.jece.2021.106681
Received 13 September 2021; Received in revised form 14 October 2021; Accepted 25 October 2021
Available online 28 October 2021
2213-3437/© 2021 Elsevier Ltd. All rights reserved.
M. Sharma et al.
Journal of Environmental Chemical Engineering 9 (2021) 106681
100 ◦ C [1–4,6,7,9–11]. Thus, designing low-cost, high-selectivity sepa­
rators with low oxygen transfer coefficients and simple fabrication
procedures is imperative. Nylon and glass fiber filter separators were
found to exhibit higher power density but decreasing CE with increasing
pore size [17]. Zirfon membranes have also been examined as a low-cost
alternative to Nafion separators [11]. Other materials such as cellulose,
polycarbonate, sulfonated polystyrene ethylene butylenes polystyrene
membrane (SPSEBS), clay ceramic, sulfonated polyether ether ketone
membrane (SPEEK), polyvinyl alcohol-Nafion-borosilicate have also
been explored [11,13,14]. High-performance PEMs must possess high
ion-exchange capacity, high perm-selectivity, good mechanical and
chemical stability, and resistance to organic fouling for applications in
MFCs [6,9]. Such desirable properties are developed by regulating the
composition of the polymer matrix and introducing fillers [8,15,17].
Ion-functionalized polymeric materials promote the improvement of the
conductivity of ions through the construction of water channels [21].
However, for PEMs containing sulfonic acid groups (~HSO3-), a high
degree of sulphonation can lower mechanical stability due to water
swelling [18,19]. PVDF, a hydrophobic polymer that is thermally stable
and possesses good chemical resistance, can be introduced to counteract
this [23]. Another effective way of tuning the properties of CEMs is
polymer blending. By introducing two or more polymers and controlling
their composition, properties such as ion conductivity, chemical stabil­
ity, and swelling degree can be tuned [13–15]. Similarly, various
advanced membranes incorporating crucial membrane modifications
using nanoparticles like GO have been used in several environmental
remediation applications [20–26,29]. For instance, Sharma et al., 2021
investigated the acid recovery analysis and dye rejection using a
CA-based membrane doped with varying concentrations of GO [22,35].
As sugarcane bagasse is a major waste product from sugarcane mill, and
hence using it for preparing membrane modifier material would pro­
mote utilization of the waste for preparing a value added product. In this
regard, cellulose has promising applications in membrane technology. In
particular, cellulose acetate (CA) is cost-effective, possesses neutral
properties, and can be cast into transparent films [18]. Various studies
have demonstrated that CA, on account of its hydrophilic nature, en­
hances membrane hydrophilicity, porosity, proton conductivity, and
antifouling property [16,19,27,28]. Moreover, agricultural wastes such
as sugarcane bagasse (SCB) can be utilized to synthesize CA [34].
Membrane performance in terms of ion exchange capacity and ion
conductivity can be further optimized by dispersing inorganic fillers in
the polymer matrix to form mixed matrix membranes (MMMs). A wide
range of inorganic materials like metal oxides and graphene oxide (GO)
have been investigated as additives [13,30]. GO has been extensively
used in membrane-based applications because of its unique properties
such as high mechanical and thermal stability, high surface area [36].
The presence of different oxygen-containing functional groups entails
higher solubility and presents opportunities for surface functionalization
[31,32]. Incorporating GO in membranes has been shown to increase
membrane hydrophilicity and permeability [20]. However, no such
low-cost, high-performance proton exchange membrane integrating
cellulose acetate produced from sugarcane bagasse and GO with PVDF
as a base polymer has been used in MFC. Although multiple studies on
COD removal and energy extraction utilising microbial fuel cells with
various electrode modifications and catalysts have been published in the
past [50,51], only a handful have achieved significant COD removal
efficiency and power density. Moreover, the investigation has been done
without any catalyst involved and modification of the electrode. Studies
on wastewater treatment for COD removal have been previously re­
ported incorporating both individual and hybrid processes [33,34,38,
39].
The present work focuses on synthesizing a novel high-performance
CEM comprising a polymeric blend of PVDF and CA and impregnating
GO to obtain membranes with high ion exchange capacity (IEC) and
proton conductivity. Dual-chamber MFC design was selected as it is the
most widely used design owing to its usefulness in wastewater treatment
and energy generation. In batch mode, a dual-chamber is generally used.
It comprises two electrodes separated by a membrane that transfers
electrons between the chambers, while the proton exchange membrane
can be a porous polymeric or a porous ceramics [40]. The membrane
permits protons to flow from the anode to the cathode and O2 to diffuse
in the anode However, it does not allow substrates or microbes to pass
between the electrodes. Previous studies also reported the fabrication of
the functionalized zeolite for applications in water treatment [41]. In
this work, CA has been synthesized from waste sugar cane bagasse. GO
was synthesized by the Tour method [37]. The effect of different blend
ratios of GO and CA on physicochemical properties such as water up­
take, proton conductivity, and ion exchange capacity was examined. The
power density and polarization curves were analyzed, and the perfor­
mance of the membrane was compared to that of other separators in the
literature. Finally, an investigation for cost analysis has been done with
the previously reported high-performance membranes for application in
extracting clean energy.
2. Materials and methodology
2.1. Materials required
Sulfuric acid (95–97%), ortho-phosphoric acid (≥85% MSDS), po­
tassium permanganate (Grade value: ACS,Reag. Ph Eur), hydrogen
peroxide (30%), hydrochloric acid (≥37%), PVDF, Dimethylformamide
(DMF) (Grade:ACS,ISO,Reag. Ph Eur), sodium chloride (NaCl) (Grade:
ACS,ISO,Reag. Ph Eur), polyethylene glycol (PEG) (Mw=6000 g/mol),
NaOH (MFCD00003548), MgSO4, EDTA, glacial acetic acid (≥99%)
were obtained from Merck specialties Private Ltd. Germany. Deionized
Millipore water has been used throughout the experiments.
2.2. Graphene oxide synthesis using Tour method
GO was prepared according to the Tour method, a more efficient and
greener alternative to Hummer’s method [37]. A mixture of 4.5 g
KMnO4 and 0.5 g graphite was taken and ground to a fine powder using
pestle and mortar. To this, a mixture of 90 ml of conc. H2SO4 and 10 ml
of conc. H3PO4 (9:1 v/v solution) was added. The resulting mass was
heated to a temperature of 50 ℃ in a water bath, with continuous stir­
ring for 12 hrs. The mixture was observed to turn into a paste as the
reaction progressed. This paste was then cooled to room temperature,
and the reaction was arrested by adding 250 ml of distilled water. Re­
sidual KMnO4 was reduced to soluble MnSO4 by adding 10 ml of H2SO4.
The resulting reaction precipitated bubbling, and yellow color was
observed. MnSO4 was removed using filter paper. The graphite oxide
(GTO) cake formed was washed with 5% HCl solution repeatedly and
centrifuged at 4000 RPM until the removal of sulfate ions was confirmed
using BaCl2 solution. The GTO thus obtained was dispersed in distilled
water and stirred at 60 ◦ C for 12 h to form a single-layer GO. The
resulting dark brown paste was collected using a freeze drier (Lyophi­
lizer) for 24 h.
2.3. Cellulose acetate extraction from sugarcane bagasse
Cellulose acetate used in this work was synthesized from waste
sugarcane bagasse obtained from local vendors. The initial steps
involved the washing of waste sugarcane bagasse (SCB) followed by
cutting the same into small pieces after drying [34]. The material was
first dried for 12 h in a hot air oven at 80 ᵒC. Acid pre-treatment was
carried out with 10% (v/v) H2SO4 and 1/10 (w/v) solid/liquid ratio of
dry SCB. The resulting mixture was stirred at 250 RPM for 16 h at 60 ᵒC.
Further, the solid fraction obtained was repeatedly washed with
deionized water till it achieved a neutral pH and dried. The alkaline
pretreatment was carried out with 2 M NaOH and 1/10 (w/v) solid/­
liquid ratio of acid-treated material loading with stirring at 250 RPM
and heating at 70 ᵒC. The resulting mass was washed with deionized
2
M. Sharma et al.
Journal of Environmental Chemical Engineering 9 (2021) 106681
water till it achieved a neutral pH. Further, the chelation was carried out
with 0.5% EDTA at 70 ᵒC for 30 min. The material was then washed with
deionized water at 70 ᵒC till neutral pH. Bleaching was performed with
5% (v/v) H2O2 and 0.1% MgSO4. The material was again washed with
deionized water at 70 ᵒC till a pH of 7.0 was achieved.
Finally, the obtained cellulose was converted to cellulose acetate by
adding 24 ml glacial acetic acid to 10 g of material. The obtained
mixture was stirred at 37.8 ᵒC for 2 h. Later, 40 ml acetic acid and 0.08
ml H2SO4 were added to the product obtained. The material was later
cooled to room temperature. Thereafter, 28 ml of acetic acid anhydride
and 0.6 ml H2SO4 were added to it and the mixture obtained was stirred
for 2 h at 40 ᵒC. Subsequently, 10 ml of deionized water and 20 ml of
glacial acetic acid were added to the mixture. The material obtained was
washed with deionized water till a pH of 7.0 was realized.
solution and placed in an orbital shaker for 24 h. The samples were then
removed from the solutions, dried in a hot air oven, and weighed. The
solutions were titrated against 0.01 M NaOH solution using phenol­
phthalein as the endpoint indicator. The IEC was estimated using the
following Eq.:
IEC =
FIC =
σ=
2.4.3. Ion exchange capacity
The amount and species of ion-exchange groups are critical de­
terminants of CEM performance. Therefore, the IEC is an integral part of
membrane characterization [43]. The titration method was employed to
determine membrane IEC. Membrane samples of dimensions
2 cm × 2 cm were immersed in 50 ml each of freshly prepared 2 M NaCl
Table 1
Membrane composition.
GO (%)
DMF (%)
CA (%)
PEG (%)
18
15
13
11
0
3
5
7
80
80
80
80
1
1
1
1
1
1
1
1
(4)
2.4.6. MFC fabrication and operation
A dual-chamber aqueous-cathode MFC was designed and fabricated
for evaluating the performance of the as-synthesized membrane. The
MFC setup comprised of membrane synthesized using PVDF as a base
polymer and, graphene oxide along with cellulose acetate extracted
from sugarcane bagasse as the membrane modifier materials. The
membrane was pre-treated for removal of impurities before using in
fabricated MFC. The pre-treatment process involved the treatment of
membranes using alkali and DI water and further followed by the sul­
fonating process to attach the necessary functional groups in the mem­
branes. Carbon cloth was used as the electrode material for the prepared
MFC assembly. The involved projected electrode surface area is similar
to the anode surface area (16 cm2), and the internal volume of the
anodic and cathode chamber of the MFC was100 ml each. The stainlesssteel wires were used for all the necessary connections, and the external
resistance of 100 Ω was connected across the two electrodes. The anodic
chamber of the MFC was inoculated with the anaerobic mixed sludge
collected from the sewage tank at IIT Guwahati, India. Before inocula­
tion of the MFC, the methanogenesis suppression in the anaerobic sludge
was done with Chloroform [44]. In order to maintain the anaerobic
environment in the anodic chamber, the MFC was hermetically sealed at
the anodic sides. To ensure continuous flow of oxygen in the cathodic
chamber a hole of diameter 1.5 in. was made, and the aeration pump
was used to supply continuous flow of air into it. Sucrose-based synthetic
wastewater having chemical oxygen demand (COD) of 3 g L-1 was used
in the anodic chamber of the MFC [45]. The MFC was operated in the
Where Wd is the dry and Ww is the wet weight of membrane samples.
PVDF (%)
l
R×d×w
Here, l, d, w, and R are the distance between the platinum rods,
membrane thickness, width of the membrane, and low intersect of highfrequency semicircle obtained in the Nyquist plot on complex imped­
ance plane with the real (Z) axis.
(1)
WCA-1
WCA-2
WCA-3
WCA-4
(3)
2.4.5. Proton conductivity
The proton conductivity of the prepared membranes was calculated
using electrochemical impedance spectroscopy (EIS) with frequency in
the range of 100 Hz to 2 MHz. The membranes (2 cm × 2 cm) were well
hydrated by immersing in water before testing to ensure complete
saturation. The period of immersion was varied to confirm that inter­
calated sulfuric acid did not leach out during soaking in water and the
proton conduction was effectively due to the functional groups present
in GO. The proton conductivity was calculated using Eq. (4) [42].
2.4.2. Water uptake
Water uptake (WU) was calculated by measuring the difference in
weight of the membranes before and after hydration. 2 cm × 2 cm
samples were immersed in water in a glass dish at ambient temperature
overnight. Excess water on the samples was absorbed using tissue paper,
and the samples were then weighed. Water uptake was estimated using
the following Eq. [22]:
Sample
IEC
S
Where FIC is expressed in meq L-1.
2.4.1. Membrane pre-treatment
Deionized Millipore water was used throughout the sulphonation
process. A 3% (v/v) solution of H2O2 was prepared, and the membranes
were immersed in it for 1 h at a temperature of 65ᵒC. The membranes
were then immersed in deionized Millipore water for 2 h at 65ᵒC to
remove excess H2O2. Afterward, the membranes were immersed in
freshly prepared 0.5 M H2SO4 for 4 hr at 65ᵒC. The membranes were
again rinsed with deionized millipore water to remove excess acid [42].
× 100
(2)
2.4.4. Fixed ion concentration
IEMs consist of fixed-charged groups connected to the polymer
backbone. Fixed ion concentration (FIC) of a membrane defines the
counter ion transportation perm-selectivity across the membrane. The
following Eq. can be used to determine the FIC for the CEM [12].
Dispersion of graphene oxide (GO) in DMF was prepared by adding
0.3 gm GO to 15 ml DMF, followed by sonication for 10 min. The base
polymer, PVDF, along with CA, was dissolved in DMF. Polyethylene
glycol (average molecular weight 6000 g/mol) (PEG-6000) was added
as the pore-forming agent. The compositions of the casting solutions of
the four membranes prepared are given in Table 1. The casting solutions
were mixed using a magnetic stirrer operating at 300 RPM at 65ᵒC for
16 h. The solutions obtained were then degassed in a hot air oven at 50ᵒC
for 12 h. The solutions were then cast on a dry glass plate. A thickness of
120 µm was obtained with the help of a casting knife. Deionized water
was used as the non-solvent for immersion precipitation and the mem­
branes prepared were kept in deionized water for 24 h.
Ww −
Wd
Wd
× NNaOH
Wdry
Where VNa0H is the volume of NaOH consumed in the titration, NNaOH is
the normality of the NaOH solution used and Wdry is the mass of the
dried membrane samples.
2.4. Membrane preparation and characterization
WU(%) =
VNaOH
3
M. Sharma et al.
Journal of Environmental Chemical Engineering 9 (2021) 106681
batch mode and after each three days interval, the synthetic wastewater
was replaced with freshly prepared synthetic wastewater. The MFC was
operated for 30 days and the performance of the MFC was evaluated
during this period.
Further, the measurements for the current density and power density
were obtained by dividing the current and power with the projected
anode surface area. The polarization was performed after the MFC
attained steady-state performance. During polarization, the feed of the
MFC was replaced by freshly prepared synthetic wastewater and after
that, the MFC was kept in open circuit condition for about 30 min to 1 h
till the MFC attain a stable voltage. After that, a very high external
resistance was connected between cathode and anode and after getting
stable reading the voltage reading was measured, and the current value
was evaluated using Ohm’s law. After the external resistance was
gradually decreased till 10 Ω and stable voltage readings were noted
down and corresponding current readings were calculated current
density and power density were further calculated by using Eq. 1 and
dividing the current and power value with projected electrode surface
area. Further, the internal resistance was evaluated by measuring the
slope of the Current vs. Voltage graph of polarization. Finally, the
columbic efficiency (CE) of the MFC was calculated by using Eq. (6)
2.4.7. Electricity recovery from MFC
Electricity recovery in terms of voltage and current was also evalu­
ated. The voltage of the MFC corresponding to an eternal resistance of
100 Ω was measured daily and the corresponding current (I) was
calculated by dividing the voltage (V) by external resistance (R) as per
the Ohms law, V=IR. The effluent and influent concentrations of COD
were measured after every three days batch cycle with the help of a
spectrophotometer (Hach; Model DR-900). Regular measurement of
open-circuit voltage (OCV) was also conducted by removing external
resistance. HTC instrument DM-97, a digital multimeter was used for
measuring all electrical parameters. The power recovery was calculated
as the product of voltage ‘V’ and the current produced ‘I’ [42].
Fig. 1. a. FTIR spectrum for prepared cellulose acetate. Inset: FTIR spectrum for commercial CA 1b. Raman shift for CA (Inset: FESEM (CA)). 1c. XRD analysis; outset:
prepared CA; inset: XRD for commercial CA.
4
M. Sharma et al.
Journal of Environmental Chemical Engineering 9 (2021) 106681
[44].
∫t
CE =
M 0 Idt
× 100
FbvΔCOD
– C bonding
stretching of the carboxyl-carbonyl group and unchanged C–
were observed for before and after oxidation of graphite by two familiar
peaks at 1735 cm-1 and 1617 cm-1, respectively. The availability of the
epoxy group in the basal plane of GO is identified at a peak near
1060 cm-1. Additionally, the oxide functional group with lower trans­
mittance % was observed. Therefore, it can be concluded that the oxy­
gen functional groups with stretching vibrations of GO can be located at
– O), and 1060 cm-1 (C-O)
about 3000–3700 cm-1 (OH), 1735 cm-1 (C–
which can overlap the CA peaks. The obtained results were further
supported by the respective Raman spectra (Fig. 2c), where several new
peaks attributed to the availability of various functional groups. Further,
the morphological analysis of the synthesized GO was also done using a
field emission scanning electron microscope (FESEM) to confirm the
exfoliation process of the obtained product (Fig. 2d). Hence it can be
inferred that the quality of the GO and, CA obtained from the sugarcane
bagasse is superior, which is further reflected by the enhanced values of
the IEC and proton conductivity post incorporating them.
(6)
Here, I, M, v, b, F and ΔCOD represents the, generated electric cur­
rent, Oxygen’s molecular weight, the volume of the anodic chamber,
number of electrons available in a mole of oxygen (4), Faraday’s con­
stant (96,485 C mol-1) and variation in the effluent and influent COD
concentration in gL-1 in time t [45].
3. Results and discussion
3.1. Modifiers characterization
3.1.1. FTIR study of GO and CA
The FTIR study is a crucial method for a qualitative investigation of
the synthesized materials. As depicted in Fig. 1a the comparative anal­
ysis of the extracted CA and the commercially available CA has been
done. As can be observed clearly, the obtained CA as compared to the
commercially available CA exhibited functional groups at nearly the
same intensity, which describes the quality of the product synthesized.
– O (1740.15 cm-1), CThe characteristics peak pf OH (3475.86 cm-1), C–
CH3 (1805.00 cm-1), CH2 (1369.28 cm-1), C-O (1229.23 cm-1), C-O-C
(1040.13 cm-1) and CH (902.93 cm-1) were obtained nearly at the same
location for both the cases. Further, the results were supported with the
respective Raman spectra (Fig. 1b), where several new peaks correspond
to the presence of several functional groups. The analysis facilitates the
identification of the organic/ inorganic functional groups in the ob­
tained material. The modifiers, GO and, CA obtained from sugar bagasse
were observed via the FTIR technique. As illustrated in Fig. 2a the
specific and intense peak at 3421 cm-1 with lower transmittances states
–O
the presence of the hydroxyl groups in the prepared GO. Further, C–
3.1.2. XRD analysis of GO and CA
The X-Ray diffraction plots for the CA synthesized from the waste
sugarcane bagasse, and the CA purchased commercially are illustrated
in Fig. 1c. The obtained peaks reveal the interlayer distance variation in
the two nanofibers. As can be clearly seen from Fig. 1c the diffraction
peaks for the synthesized CA shows two major peaks at 2θ = ~15o and
~24o whereas, for the commercially available CA the peaks were ob­
tained at 2θ= ~11o, ~13o, ~15o, and ~18o. The absence of the Peak at
24o at the commercially available CA may be attributed to the fact that it
is purer than the CA extracted from sugarcane bagasse. The crystal
structure study, D-spacing, and phase identification of GO were carried
out using the X-ray diffraction technique after the GO powder had been
properly dried. The investigation of crystal structure, D-spacing, and
phase identification of graphene oxide (GO) obtained after drying was
Fig. 2. a. FTIR analysis for GO 2b. XRD plot for GO 2c. Raman shift for GO 2d. FESEM analysis for GO.
5
M. Sharma et al.
Journal of Environmental Chemical Engineering 9 (2021) 106681
done by the powder X-ray diffraction technique. The XRD spectra of the
prepared GO using the Tours method are depicted in Fig. 2b, with
several needed modifications. From figure Fig. 2c it can be observed that
post oxidation the peaks have shifted to smaller angles. The obtained
sharp diffraction peak at 2θ = ~8o (d = 0.832 nm) confirms the
oxidation reactions efficiency by enhancing the interplanar distance of
the graphene plates with the inclusion of the oxygen groups [46].
uptake values were observed with an increment in the GO content for
the membranes incorporated with CA extract, and PVDF as the base
polymer, however, the WCA-3 was found to be superior considering the
other relevant parameters. However, it was seen that any further in­
crease in the GO content for the membrane preparation may lead to
detrimental effect on the membrane properties. The reason may be
attributed to the incorporation of a large quantity of GO in the mem­
brane which behaves as a filler material, thereby leading to a decrease in
the water uptake ability up to a certain level. Also, regardless of a more
hydrophilic surface, the depleted value of S refers to an insubstantial
holding capability of the water molecules in the membrane matrix of the
fabricated membranes. The value of IEC (0.169 m. mol./g) for pure
PVDF incorporated CA membrane is because of the availability of the
functional groups available in the CA.
The value of IEC obtained in this study can be compared with the
most widely used separators. The water uptake ability of the membrane
significantly influences its IEC value due to the permeation of both the
ions and water molecules through the membrane. Moreover, the proton
mobility through the membrane is highly influenced by its water con­
tent. The water uptake ability of the considered membrane for the MFC
study was found to be as high as 54.5%, which is comparable to the
water uptake ability of several standard reported membranes [42]. In
fact, the values were even better than that of the widely recognized
Nafion 117 membrane (29%). As such, the water absorption capacity
can be considered as the performance evaluating factor in WCA-3, while
the sulfonated groups are responsible for the higher IEC in Nafion 117
membrane. However, excess addition of the GO as a filler material de­
creases the uniformly formed pores. The ion exchange groups present in
a membrane determine the ion exchange capacity of the membrane. The
involved separator showed an ion exchange capacity of 0.7 m. mol. /g,
while in the case of commercially available Nafion 117 membrane an ion
exchange capacity of 0.95 m. mol./g was reported [42]. As seen from the
XRD and FTIR studies, groups which are competent in the transportation
of cations in WCA-3 are accountable for their higher IEC and proton
conductivity.
3.2. Water uptake and ion exchange capacity (IEC)
The water uptake ability of the WCA-3 membrane was found to be
54.5% (Fig. 3a). As seen in the figure, there was a steep inclination in the
water uptake with the addition of the CA in the membrane. However,
with the increase in the GO content the water uptake value was moni­
tored and was controlled to the desired amount. This is because GO acts
as a filler material, and the hydrophilic characteristics of CA extracted
from the waste sugarcane bagasse assists in retaining the water mole­
cules in the prepared membrane matrix. Thereby, a thin film of desired
properties (hydrophilicity and strength) is developed for use in the MFC
application. The water uptake ability of the CA plays a crucial part in the
transportation of protons through them. The impregnation of CA in the
separator not only aids in the improvement of water uptake ability but
also enhances the IEC and proton conductivity of the membrane matrix
due to the availability of the hydroxyl and methyl functional groups. The
presence of ion-exchange functional groups within a CEM could be used
to determine the IEC of the membrane. The negatively charged groups
present on the surface of CEM allow the protons transportation. This
privileged uptake of counterions and the capacity to discard similarly
charged ions with the help of a proton exchange separator is designated
as Donnan equilibrium. In the present study, WCA-3 showed an IEC
value of 0.7 m. mol./g, which is found to be closer to the IEC value of
0.95 m. mol./g for the commercially available and widely used Nafion
membrane. Similar trends in the fixed ion concentration (FIC) values
were also observed in the prepared membranes. It was observed that the
FIC value increased steeply, and later got almost uniform with the
appropriate GO loading (Fig. 3a (Inset)). The variation of chemical
bonds on the membrane surface as well as between the GO layers is
shown in the FTIR analysis. Integration of functionalized GO material in
hydrophilic WCA-3 further intensifies the contribution of conducting
phase about its hydrophilic nature and its functional characteristics at
the fabricated membrane surface. The presence of GO improves the
physicochemical properties of the membrane and its mechanical sta­
bility. Membrane thickness, fixed charge density, and ion exchange ca­
pacity are evaluated to assess the physicochemical properties of the
fabricated membranes. Also, with the fusion of GO in the membrane, the
trend for water uptake ability significantly increases. Better water
3.3. Swelling degree and proton conductivity
The swelling degrees for the fabricated membranes viz. WCA-1,
WCA-2, WCA-3, and WCA-4 were found to be ~73.39, 73.74, 73.82
and 73.84, respectively. The inclination in the swelling values indicates
a higher water uptake ability, which was already conferred in the pre­
vious section. The addition of GO as filler material was found to improve
the membrane stability, while the incorporation of a higher quantity of
GO nano-sheet in the WCA-3 membrane considerably minimizes the
swelling index of the membrane. The value of FCD was enhanced from
Fig. 3. a. Change in EWC of the prepared membranes with different compositions of GO, and the prepared CA. Inset: FIC comparison of different membranes 3b.
Proton conductivity and IEC variation of the prepared membranes.
6
M. Sharma et al.
Journal of Environmental Chemical Engineering 9 (2021) 106681
0.002 meq/L for WCA-1 membrane to 0.01 meq/L for WCA-4 mem­
brane, owing to the addition of GO nano-sheets. Likewise, the proton
conductivity value also increased from 0.025 Scm-1 for the WCA-1
membrane to 0.4 Scm-1 for the WCA-4 membrane as shown in Fig. 3b.
It was observed that an increasing amount of GO in the membrane
matrix led to a considerable decrease in the swelling degree values,
while both the values for IEC and proton conductivity gets enhanced
simultaneously. Such occurrence was due to the availability of func­
tional groups in the GO nano-sheets.
The proton conductivity of the fabricated membrane was evaluated
by determining the charge transfer resistance, achieved from EIS anal­
ysis. The addition of both CA (1% by weight) and GO (5% by weight) in
the membrane matrix enhances the proton conductivity of the WCA-3
membranes (~0.3 S cm-1) by several folds, which was way cheaper
than the Nafion 117 membrane. Moreover, the proton conductivity of
the WCA-3 membrane can be further improved by increasing the dosage
of CA and GO in the membrane matrix, thereby providing a proton
conductivity similar to that of Nafion 117. Fig. 3b depicts the improved
performance of the fabricated membrane in terms of proton conductivity
and IEC. The presence of several functional groups is responsible for the
improved proton conductivity and IEC of the separators. The formation
of single-layered GO nano-sheets minimizes the chance of sulphuric acid
molecules getting confined between the GO layers, thereby favoring the
role of the functional groups to control the improved proton conduc­
tivity. The hydrophilic properties of both GO and CA provide a high
adsorption capacity of the water molecules, which eventually increases
the swelling degree. However, the swelling degree for the WCA-3
membrane increased due to the improved filling property by GO nanosheets, and further incorporation of a substantial quantity of GO could
lead to a possible reduction in the mechanical strength of the membrane.
results were comparable to those reported for several other membranes
[48].
3.5. Electricity recovery from the MFC
The operating and open-circuit voltage was monitored regularly
during the operation of the MFC (Fig. 4a). Initially, a sharp increase in
the operating voltage was recorded for the MFC and, it took approxi­
mately 20 days to achieve stable voltage. The average stable operating
voltage of WCA-3 incorporated MFC was approximately 131.5 mV cor­
responding to 100 Ω external resistance. Further, during the steady-state
operation, the average open-circuit voltage for MFC was around
763 mV. The obtained results were comparable to the several reported
results without involving the catalysts and the electrode modifications
[49]. In fact, the obtained results were found to be superior to the
recently published investigation with similar structural configurations
involving comparative analysis of the MFCs with standard clayware
membrane and the commercially available Nafion 117 membrane [33].
3.6. MFC operation, polarization, and voltage generation
The MFC was operated in the batch mode for a time period of 30 days
and the power density and current density were measured (Fig. 4c). It
was observed that the power density values increased initially during
the lag phase however during steady-state operation of 48 h and a stable
power density and current density was observed after that period. The
current vs. voltage and current density vs. power density graphs of the
synthesized WCA-3 membrane are illustrated in Fig. 4b and c. As
illustrated in Fig. 4b, the polarization curve was found to be a steep
curve which signifies the internal resistance of the MFC 252.6 Ω. The
current density corresponding to the 100 Ω external resistance was
found to be 1004.1 mA/m2. As per the reported studies in the literature,
the unaltered MFC incorporating a cost-effective clayware membrane
demonstrated a current density of 103 mA/m2 [39,40]. In a similar
study conducted by Das et al., the clayware membrane was blended with
20% montmorillonite to enhance the performance output, and the
achieved current density (492 mA/m2) was much more than the previ­
ously reported study [42]. Hence, the WCA-3 incorporated MFC was
evidently superior in performance, which may be attributed to the
higher proton conductivity values, less movement of the cations, and
proton selective nature of the synthesized membrane.
Low power output is the biggest drawback of the MFCs, and re­
searchers are constantly working on designing better membranes and
configuration alterations to obtain an ameliorated setup for efficient
energy extraction. The previously reported studies have reported a high
power density of 1.67 W/cm2, but investigations suggest that the illus­
trations can only be limited to small-scale setups [52]. The obtained
power density value (150.22 Wm-2) was superior to the other widely
recognized costlier membranes incorporating electrode modifiers like
platinum. The obtained power density value was due to the availability
of abundant functional groups like epoxy and oxide functional groups in
the WCA-3 membrane, resulting in better transportation of ions and
enhanced proton conductivity (through the absorbed water molecules
added to the available functional groups in the GO) in the wet condi­
tions. The fabricated PEM may be considered as a better membrane
capable of reducing the COD to 97.5 ± 0.8% and obtaining superior
power density without any electrode modification. The obtained power
densities were correlated with the previous crucial investigations
(Table 2), and it can be concluded that the prepared separator (WCA-3)
outperformed other costlier membranes with optimized doping of GO.
This result was further verified with the proton conductivity values
obtained. For the polarization study, initially, the resistance was
disconnected from the MFC and the setups were kept in the open circuit
mode for a duration of around 12–14 h allowing no current flow through
the circuit. Further, the connections were made to close the circuit and
the external resistance of 10,000 Ω was connected, and the cell voltage
3.4. COD removal and columbic efficiency
The wastewater treatment efficiency of the MFC was evaluated by
measuring COD removal efficiency. The WCA-3 membrane demon­
strated higher COD removal efficiency (97.5 ± 0.8%,) as compared to
the several other reported membrane-based MFC [3]. The CE was also
calculated during the steady-state operation of MFC and it was also
observed that the average CE value for WCA-3 membrane was 7.1%,
which was comparable to the results obtained in several investigations
involving standard membranes [3,44]. During the utilization of
waste-activated sludge, the average value of CE and COD was found to
decrease marginally. A CE value was obtained for WCA-3 mem­
brane-based MFC (100 ml volume) without the use of catalyst and
involved no electrode modifications. Nevertheless, the columbic effi­
ciency of an MFC setup is highly dependent on its volume, shape, and
electrode materials and direct consideration of columbic efficiency of
different MFCs with non-identical volume, shape, and electrode material
that were run under different operational conditions in separate studies
may not provide uniform output. As such, CE acquired in the current
study may not be similar to the CE obtained in previous studies.
Furthermore, the improved CE and COD values of WAC-3 suggested
higher proton mobility and lesser substrate and oxygen diffusion ca­
pacity, which assisted in an improved electrogenic activity in the anodic
chamber, thereby resulting in an increased columbic efficiency and
wastewater pollutant removal. Further, the oxygen mass transfer coef­
ficient and oxygen diffusion coefficient of the membrane (WCA-3) were
determined using the methodology adopted by Neethu et al., 2019 [47].
The oxygen mass transfer coefficient and oxygen diffusion coefficient
are calculated to determine the membranes’ ability to keep oxygen out
of the anaerobic anodic chamber. The obtained value for the oxygen
mass transfer coefficients was (2.7 ± 0.4) × 10-4 cm/s, whereas the
oxygen diffusion coefficient was calculated as 4.05 × 10-6 cm2/s.
Furthermore, the substrate diffusion may also occur because of the
crossover of the substrate through the membrane. The substrate diffu­
sion of the WCA-3 membrane was 0.04 × 10-6 cm2/s. The achieved
7
M. Sharma et al.
Journal of Environmental Chemical Engineering 9 (2021) 106681
Fig. 4. a. Open circuit voltage and operating voltage for the MFC 4b. Polarization curves of the investigated MFC using sewage waste water obtained from waste
water treatment plant IIT Guwahati, and synthetic waste water as a feed 4c. Obtained power density curve of the MFC.
4. Cost efficiency
Table 2
Performance comparison of MFCs employing various separators.
Membrane
Cathode
Power
density
(mW m-2)
References
Textile separator was made from
46% cellulose and 54%
polyester
PVA-Nafionborosilicatemembrane
SPEEK+ 7.5% TiO2 composite
membrane
Nafion 117
Sulfonated polyphenylenesulfide
Carbon cloth
coated with
Pt/C
Pt/C
120
[53]
103.86
[2]
98.1
[54]
Pt/C
Carbon cloth
coated with
Pt/C
Carbon felt
Carbon felt
47.6
190
[54]
[55]
124.16
112.8
[56]
[3]
Carbon felt
coated with
Pt/C
Carbon felt
coated with
Pt/C
Carbon cloth
226
[57]
312.50
[58]
150.22
Present
study
NA
5% goethite blended clayware
membrane
PTDPBSH-70
PTPFBSH-90
WCA-3
Pt/C
The cost estimation was primarily done by considering the material
costs needed for the fabrication of the WCA-3 membrane (surface area of
100 cm2), which was then compared with the widely recognized and
high-performance Nafion 117 membrane having a surface area of
100 cm2 [42]. The cost-effectiveness of the fabricated 100 cm2 WCA-3
membrane with commercially available Nafion 117 and other in­
vestigations with similar setup configuration was done [13]. During the
membrane calcination, the energy consumption of 3.6 kWh with a
subsequent cost of 0.51 USD was estimated under the Indian scenario.
The total fabrication cost per square meter of the WCA-3 membrane was
found to be 475 USD. The cost comparison of the prepared membrane in
terms of price per square meter of the widely recognized and
high-performance membranes such as Nafion, PDMS cast on carbon
cloth, Nafion, and G5 membrane was done (Table 3), and the reported
values were 1000 USD, 700 USD, 3300 USD, and 607 USD, respectively.
Moreover, the cost incurred in generating per unit of electrical power
(USD/W) was evaluated by including both the membrane cost and the
fabrication cost of the microbial fuel cell. Implementing WCA-3 mem­
brane for energy extraction using MFC in place of costly proton ex­
change separators will lead to cost reduction by several folds. In
addition, the cost of the WCA-3 membrane can be decreased consider­
ably if the utilized CA during the membrane fabrication could be
was recorded as 561 ± 11 mV for WCA-3 incorporated MFC. Further,
the external resistances were reduced to 1.5 Ω and, current vs. voltage
and current density vs. power density curves were obtained (Fig. 4b and
c). The internal resistance of the MFC incorporating WCA-3 membrane
was 253.6 Ω, which was lower than several reported MFC incorporating
standard membranes like clayware and Nafion membranes [3].
Table 3
Cost comparison of the various membranes.
8
Membrane
Price (USD)
Reference
Nafion
PDMS cast on carbon cloth
Nafion 117
G5
WCA-3
1000 USD/ m2
700 USD/ cm2
3300 USD/ m2
607 USD/ m2
475 USD/ m2
[59]
[59]
[3]
[3]
Present work
M. Sharma et al.
Journal of Environmental Chemical Engineering 9 (2021) 106681
retrieved from several other wastes sources. As such, the WCA-3 mem­
brane can easily replace the illustrated membranes shown in Table 3 for
the MFC application due to its high cost-effectiveness. Further, the cost
of the prepared membrane was estimated as 476.72 $ m− 2 (Table S1(c)),
which is much less than the reported commercially available Nafion 117
membranes (2860 $ m− 2) [13]. It can be inferred that, the prepared
membrane is approximately 6 times cheaper than the Nafion 117
membrane. Further, the power density obtained per unit cost of mem­
brane synthesized was 0.0344 W $− 1 (Table S3) which is 8 folds lesser
than the reported value for Nafion 117 membrane (0.278 W $− 1). The
obtained values for power density per unit cost of membrane are how­
ever more than the reported MFC-SBC membranes, which can be further
reduced by implementing the crucial electrode modifications and cata­
lysts to enhance the power density values. Therefore, lowering the cost
of fabrication during field-scale setup would not only reclaim additional
power but also simultaneously provides efficient treatment to the
wastewater.
References
[1] M. Rahimnejad, A. Adhami, S. Darvari, A. Zirepour, S.-E. Oh, Microbial fuel cell as
new technology for bioelectricity generation: a review, Alex. Eng. J. vol. 54 (3)
(2015) 745–756, https://doi.org/10.1016/j.aej.2015.03.031.
[2] B.R. Tiwari, M.T. Noori, M.M. Ghangrekar, A novel low cost polyvinyl alcoholNafion-borosilicate membrane separator for microbial fuel cell, Mater. Chem. Phys.
vol. 182 (2016) 86–93, https://doi.org/10.1016/j.matchemphys.2016.07.008.
[3] I. Das, S. Das, R. Dixit, M.M. Ghangrekar, Goethite supplemented natural clay
ceramic as an alternative proton exchange membrane and its application in
microbial fuel cell, Ion. (Kiel. ) vol. 26 (6) (2020) 3061–3072, https://doi.org/
10.1007/s11581-020-03472-1.
[4] B.H. Kim, I.S. Chang, G.M. Gadd, Challenges in microbial fuel cell development and
operation, Appl. Microbiol. Biotechnol. vol. 76 (3) (2007) 485–494, https://doi.
org/10.1007/s00253-007-1027-4.
[5] Y. Ye, H.H. Ngo, W. Guo, S.W. Chang, D.D. Nguyen, Y. Liu, B.J. Ni, X. Zhang,
Microbial fuel cell for nutrient recovery and electricity generation from municipal
wastewater under different ammonium concentrations, Bioresour. Technol. vol.
292 (2019), 121992, https://doi.org/10.1016/j.biortech.2019.121992.
[6] B. Min, B.E. Logan, Continuous electricity generation from domestic wastewater
and organic substrates in a flat plate microbial fuel cell, Environ. Sci. Technol. vol.
38 (21) (. 2004) 5809–5814, https://doi.org/10.1021/es0491026.
[7] H. Regassa, D. Bose, A. Mukherjee, Review of microorganisms and their enzymatic
products for industrial bioprocesses, Ind. Biotechnol. vol. 17 (4) (. 2021) 214–226,
https://doi.org/10.1089/ind.2021.0002.
[8] D. Bose, M. Santra, R.V.S.P. Sanka, B. Krishnakumar, Bioremediation analysis of
sediment-microbial fuel cells for energy recovery from microbial activity in soil,
Int. J. Energy Res vol. 45 (4) (. 2021) 6436–6445, https://doi.org/10.1002/
er.6163.
[9] D. Bose, R. Rawat, S. Sridharan, M. Gopinath, P. Vijay, in: V. Kumar, G. Saxena, M.
P.B.T.-B. for, E.S. Shah (Eds.), Chapter 8 - Aspects of microbial fuel cell technology
for wastewater treatment and bioelectricity generation, Elsevier, 2021,
pp. 167–190.
[10] D. Bose, S. Sridharan, H. Dhawan, P. Vijay, M. Gopinath, Biomass derived activated
carbon cathode performance for sustainable power generation from Microbial Fuel
Cells, Fuel vol. 236 (2019) 325–337, https://doi.org/10.1016/j.fuel.2018.09.002.
[11] D. Pant, G. Van Bogaert, M. De Smet, L. Diels, K. Vanbroekhoven, Use of novel
permeable membrane and air cathodes in acetate microbial fuel cells, Electrochim.
Acta vol. 55 (26) (2010) 7710–7716, https://doi.org/10.1016/j.
electacta.2009.11.086.
[12] J. Ran, L. Wu, Y. He, Z. Yang, Y. Wang, C. Jiang, L. Ge, E. Bakangura, T. Xu, Ion
exchange membranes: New developments and applications, J. Memb. Sci. vol. 522
(2017) 267–291, https://doi.org/10.1016/j.memsci.2016.09.033.
[13] I. Chakraborty, S. Das, B.K. Dubey, M.M. Ghangrekar, Novel low cost proton
exchange membrane made from sulphonated biochar for application in microbial
fuel cells, Mater. Chem. Phys. vol. 239 (2020), 122025, https://doi.org/10.1016/j.
matchemphys.2019.122025.
[14] K. Watanabe, Recent developments in microbial fuel cell technologies for
sustainable bioenergy, J. Biosci. Bioeng. vol. 106 (6) (2008) 528–536, https://doi.
org/10.1263/jbb.106.528.
[15] R.E. Khalifa, A.M. Omer, M.H. Abd Elmageed, M.S. Mohy Eldin, Titanium Dioxide/
phosphorous-functionalized cellulose acetate nanocomposite membranes for DMFC
applications: enhancing properties and performance, ACS Omega vol. 6 (27) (.
2021) 17194–17202, https://doi.org/10.1021/acsomega.1c00568.
[16] X. Zhang, S. Cheng, X. Wang, X. Huang, B.E. Logan, Separator characteristics for
increasing performance of microbial fuel cells, Environ. Sci. Technol. vol. 43 (21) (.
2009) 8456–8461, doi: 10.1021/es901631p.
[17] X. Zhang, S. Cheng, X. Huang, B.E. Logan, The use of nylon and glass fiber filter
separators with different pore sizes in air-cathode single-chamber microbial fuel
cells, Energy Environ. Sci. vol. 3 (5) (2010) 659–664, https://doi.org/10.1039/
B927151A.
[18] X. Zhang, J. Zhou, X. Zou, Z. Wang, Y. Chu, S. Wang, Preparation of nano-SiO2/
Al2O3/ZnO-blended PVDF cation-exchange membranes with improved membrane
permselectivity and oxidation stability, Materials vol. 11 (12) (2018) 2465,
https://doi.org/10.3390/ma11122465.
[19] V. Kugarajah, S. Dharmalingam, Investigation of a cation exchange membrane
comprising sulphonated poly ether ether ketone and sulphonated titanium
nanotubes in microbial fuel cell and preliminary insights on microbial adhesion,
Chem. Eng. J. vol. 398 (2020), 125558, https://doi.org/10.1016/j.
cej.2020.125558.
[20] A. Alabi, L. Cseri, A. Al Hajaj, G. Szekely, P. Budd, L. Zou, Electrostatically-coupled
graphene oxide nanocomposite cation exchange membrane, J. Memb. Sci. vol. 594
(2020), 117457, https://doi.org/10.1016/j.memsci.2019.117457.
[21] M. Sharma, P. Mondal, A. Chakraborty, J. Kuttippurath, M. Purkait, Effect of
different molecular weight polyethylene glycol on flat sheet cellulose acetate
membranes for evaluating power density performance in pressure retarded osmosis
study, J. Water Process Eng. vol. 30 (2019), 100632, https://doi.org/10.1016/j.
jwpe.2018.05.011.
[22] M. Sharma, P. Mondal, A. Sontakke, A. Chakraborty, and M.K. Purkait, High
Performance Graphene-oxide Doped Cellulose Acetate based Ion Exchange
Membrane for Environmental Remediation Applications, Int. J. Environ. Anal.
Chem., doi: 10.1080/03067319.2021.1975276.
[23] C. Mu, Y. Su, M. Sun, W. Chen, Z. Jiang, Remarkable improvement of the
performance of poly(vinylidene fluoride) microfiltration membranes by the
additive of cellulose acetate, J. Memb. Sci. vol. 350 (1) (2010) 293–300, https://
doi.org/10.1016/j.memsci.2010.01.004.
5. Conclusion
The cation exchange membrane is a unique and essential barrier in
MFC assisting in separation of charges and potential development. The
developed PVDF-based GO and, CA obtained from sugarcane bagasse
were used as modifiers for obtaining a cost effective CEM, and the MFC
operation incorporating the WCA-3 was done. Considering the mem­
brane characterization results, like IEC, proton conductivity, water up­
take, and other crucial membrane parameters the WCA-3 membrane was
used for application in MFC, and the obtained COD removal efficiency of
97.5 ± 0.8% for the WCA-3 was observed. The obtained results were
compared with several previous studies involving no catalysts and
electrode modification. WCA-3 membrane demonstrated superior elec­
tricity recovery and wastewater treatment. The investigation approves
the applicability of WCA-3 membrane as a suitable PEM for application
in MFC. Incorporating electrode modification like Pt may surely enhance
the process performance of the cell, and can be explored further for
successful field-scale application of MFCs.
CRediT authorship contribution statement
Mukesh Sharma: Writing original draft, Conceptualization, Meth­
odology, Software; Pranjal P. Das: Review and editing, Conceptuali­
zation, Analysis, Software; Trishla Sood: Review and editing; Arun
Chakraborty: Supervision; Mihir K Purkait: Supervision,
Visualization;.
Declaration of Competing Interest
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
the work reported in this paper.
Acknowledgment
This work is supported by a grant (DST/TM/WTI/WIC/2K17/84(G))
from the DST (Department of Science and Technology) New Delhi. Any
opinions, findings, and conclusions expressed in this paper are those of
the authors and do not necessarily reflect the views of DST, New Delhi.
Appendix A. Supporting information
Supplementary data associated with this article can be found in the
online version at doi:10.1016/j.jece.2021.106681.
9
M. Sharma et al.
Journal of Environmental Chemical Engineering 9 (2021) 106681
[43] S.M. Hosseini, A. Gholami, S.S. Madaeni, A.R. Moghadassi, A.R. Hamidi,
Fabrication of (polyvinyl chloride/cellulose acetate) electrodialysis heterogeneous
cation exchange membrane: Characterization and performance in desalination
process, Desalination vol. 306 (2012) 51–59, https://doi.org/10.1016/j.
desal.2012.07.028.
[44] G.D. Bhowmick, S. Das, M.M. Ghangrekar, A. Mitra, R. Banerjee, Improved
Wastewater Treatment by Combined System of Microbial Fuel Cell with Activated
Carbon/TiO2 Cathode Catalyst and Membrane Bioreactor, J. Inst. Eng. Ser. A vol.
100 (4) (2019) 675–682, https://doi.org/10.1007/s40030-019-00406-7.
[45] G.S. Jadhav, M.M. Ghangrekar, Performance of microbial fuel cell subjected to
variation in pH, temperature, external load and substrate concentration, Bioresour.
Technol. vol. 100 (2) (2009) 717–723, https://doi.org/10.1016/j.
biortech.2008.07.041.
[46] J.L.S. Gascho, S.F. Costa, A.A.C. Recco, S.H. Pezzin, Graphene oxide films obtained
by vacuum filtration: x-ray diffraction evidence of crystalline reorganization,
5963148-12, J. Nanomater. vol. 2019 (2019), https://doi.org/10.1155/2019/
5963148.
[47] B. Neethu, G.D. Bhowmick, M.M. Ghangrekar, A novel proton exchange membrane
developed from clay and activated carbon derived from coconut shell for
application in microbial fuel cell, Biochem. Eng. J. vol. 148 (2019) 170–177,
https://doi.org/10.1016/j.bej.2019.05.011.
[48] A. Asghar, A.A. Abdul Raman, W.M.A. Wan Daud, Challenges and
recommendations for using membranes in wastewater-based microbial fuel cells
for in situ Fenton oxidation for textile wastewater treatment, Rev. Chem. Eng. vol.
31 (1) (2015) 45–67, https://doi.org/10.1515/revce-2014-0030.
[49] O. Prakash, A. Mungray, S.K. Kailasa, S. Chongdar, A.K. Mungray, Comparison of
different electrode materials and modification for power enhancement in benthic
microbial fuel cells (BMFCs), Process Saf. Environ. Prot. vol. 117 (2018) 11–21,
https://doi.org/10.1016/j.psep.2018.04.009.
[50] M. Behera, P.S. Jana, M.M. Ghangrekar, Performance evaluation of low cost
microbial fuel cell fabricated using earthen pot with biotic and abiotic cathode,
Bioresour. Technol. vol. 101 (4) (2010) 1183–1189, https://doi.org/10.1016/j.
biortech.2009.07.089.
[51] A.N. Ghadge, M.M. Ghangrekar, Development of low cost ceramic separator using
mineral cation exchanger to enhance performance of microbial fuel cells,
Electrochim. Acta vol. 166 (2015) 320–328, https://doi.org/10.1016/j.
electacta.2015.03.105.
[52] M. Ramya, E. . Senthilkumar, G. . Sivagaami Sundari, K. . Thileep Kumar, High
power and energy density of redox additive in microbial fuel cell, Rasayan J. Chem.
vol. 12 (1) (2019) 91–100, https://doi.org/10.31788/RJC.2019.1215071.
[53] Y. Ahn, B.E. Logan, Domestic wastewater treatment using multi-electrode
continuous flow MFCs with a separator electrode assembly design, Appl. Microbiol.
Biotechnol. vol. 97 (1) (2013) 409–416, https://doi.org/10.1007/s00253-0124455-8.
[54] P.N. Venkatesan, S. Dharmalingam, Effect of cation transport of SPEEK – Rutile
TiO2 electrolyte on microbial fuel cell performance, J. Memb. Sci. vol. 492 (2015)
518–527, https://doi.org/10.1016/j.memsci.2015.06.025.
[55] J.M. Moon, S. Kondaveeti, B. Min, Evaluation of low-cost separators for increased
power generation in single chamber microbial fuel cells with membrane electrode
assembly, Fuel Cells vol. 15 (1) (. 2015) 230–238, https://doi.org/10.1002/
fuce.201400036.
[56] M.I. Simeon, F.U. Asoiro, M. Aliyu, O.A. Raji, R. Freitag, Polarization and power
density trends of a soil-based microbial fuel cell treated with human urine, Int. J.
Energy Res vol. 44 (7) (. 2020) 5968–5976, https://doi.org/10.1002/er.5391.
[57] A. Ghorai, S. Roy, S. Das, H. Komber, M.M. Ghangrekar, B. Voit, S. Banerjee,
Preparation of sulfonated polytriazoles with a phosphaphenanthrene unit via click
polymerization: fabrication of membranes and properties thereof, ACS Appl.
Polym. Mater. vol. 3 (8) (. 2021) 4127–4138, https://doi.org/10.1021/
acsapm.1c00600.
[58] A. Ghorai, S. Roy, S. Das, H. Komber, M.M. Ghangrekar, B. Voit, S. Banerjee,
Chemically Stable Sulfonated Polytriazoles Containing Trifluoromethyl and
Phosphine Oxide Moieties for Proton Exchange Membranes, ACS Appl. Polym.
Mater. vol. 2 (7) (. 2020) 2967–2979, https://doi.org/10.1021/acsapm.0c00443.
[59] J. Chouler, I. Bentley, F. Vaz, A. O’Fee, P.J. Cameron, M. Di Lorenzo, Exploring the
use of cost-effective membrane materials for Microbial Fuel Cell based sensors,
Electrochim. Acta vol. 231 (2017) 319–326, https://doi.org/10.1016/j.
electacta.2017.01.195.
[24] M. Hossein Razzaghi, A. Safekordi, M. Tavakolmoghadam, F. Rekabdar,
M. Hemmati, Morphological and separation performance study of PVDF/CA blend
membranes, J. Memb. Sci. vol. 470 (2014) 547–557, https://doi.org/10.1016/j.
memsci.2014.07.026.
[25] M.K. Sinha, M.K. Purkait, Preparation of fouling resistant PSF flat sheet UF
membrane using amphiphilic polyurethane macromolecules, Desalination vol. 355
(2015) 155–168, https://doi.org/10.1016/j.desal.2014.10.017.
[26] M.K. Purkait, P.K. Bhattacharya, S. De, Membrane filtration of leather plant
effluent: flux decline mechanism, J. Memb. Sci. vol. 258 (1–2) (2005) 85–96,
https://doi.org/10.1016/j.memsci.2005.02.029.
[27] M.K. Sinha, M.K. Purkait, Preparation and characterization of novel pegylated
hydrophilic pH responsive polysulfone ultrafiltration membrane, J. Memb. Sci. vol.
464 (2014) 20–32, https://doi.org/10.1016/j.memsci.2014.03.067.
[28] B.K. Nandi, R. Uppaluri, M.K. Purkait, Treatment of oily waste water using low-cost
ceramic membrane: flux decline mechanism and economic feasibility, Sep. Sci.
Technol. vol. 44 (12) (2009) 2840–2869, https://doi.org/10.1080/
01496390903136004.
[29] B.K. Nandi, B. Das, R. Uppaluri, M.K. Purkait, Microfiltration of mosambi juice
using low cost ceramic membrane, J. Food Eng. vol. 95 (4) (2009) 597–605,
https://doi.org/10.1016/j.jfoodeng.2009.06.024.
[30] S. Emani, R. Uppaluri, M.K. Purkait, Preparation and characterization of low cost
ceramic membranes for mosambi juice clarification, Desalination vol. 317 (2013)
32–40, https://doi.org/10.1016/j.desal.2013.02.024.
[31] V.K. Bulasara, H. Thakuria, R. Uppaluri, M.K. Purkait, Effect of process parameters
on electroless plating and nickel-ceramic composite membrane characteristics,
Desalination vol. 268 (1) (2011) 195–203, https://doi.org/10.1016/j.
desal.2010.10.025.
[32] A. Achilli, T.Y. Cath, A.E. Childress, Power generation with pressure retarded
osmosis: an experimental and theoretical investigation, J. Memb. Sci. vol. 343 (1)
(2009) 42–52, https://doi.org/10.1016/j.memsci.2009.07.006.
[33] S.W. Kim, S.O. Han, I.N. Sim, J.Y. Cheon, W.H. Park, Fabrication and
characterization of cellulose acetate/montmorillonite composite nanofibers by
electrospinning, J. Nanomater. vol. 2015 (2015) 275230–275238, https://doi.org/
10.1155/2015/275230.
[34] R.G. Candido, G.G. Godoy, A.R. Gonçalves, Characterization and application of
cellulose acetate synthesized from sugarcane bagasse, Carbohydr. Polym. vol. 167
(2017) 280–289, https://doi.org/10.1016/j.carbpol.2017.03.057.
[35] Y. Shi, C. Li, D. He, L. Shen, N. Bao, Preparation of graphene oxide–cellulose
acetate nanocomposite membrane for high-flux desalination, J. Mater. Sci. vol. 52
(22) (2017) 13296–13306, https://doi.org/10.1007/s10853-017-1403-0.
[36] S.M. Hosseini, E. Jashni, M. Habibi, M. Nemati, B. Van der Bruggen, Evaluating the
ion transport characteristics of novel graphene oxide nanoplates entrapped mixed
matrix cation exchange membranes in water deionization, J. Memb. Sci. vol. 541
(2017) 641–652, https://doi.org/10.1016/j.memsci.2017.07.022.
[37] A.T. Habte, D.W. Ayele, Synthesis and Characterization of Reduced Graphene
Oxide (rGO) Started from Graphene Oxide (GO) Using the Tour Method with
Different Parameters, Adv. Mater. Sci. Eng. vol. 2019 (2019) 5058163–5058169,
https://doi.org/10.1155/2019/5058163.
[38] P.P. Das, P. Mondal, Anweshan, A. Sinha, P. Biswas, S. Sarkar, M.K. Purkait,
Treatment of steel plant generated biological oxidation treated (BOT) wastewater
by hybrid process, Sep. Purif. Technol. vol. 258 (2021), 118013, https://doi.org/
10.1016/j.seppur.2020.118013.
[39] P.P. Das, Anweshan, M.K. Purkait, Treatment of cold rolling mill (CRM) effluent of
steel industry, Sep. Purif. Technol. vol. 274 (2021), 119083, https://doi.org/
10.1016/j.seppur.2021.119083.
[40] B.E. Logan, Simultaneous wastewater treatment and biological electricity
generation, Water Sci. Technol. vol. 52 (1–2) (. 2005) 31–37, https://doi.org/
10.2166/wst.2005.0495.
[41] N.S. Samanta, S. Banerjee, P. Mondal, Anweshan, U. Bora, M.K. Purkait,
Preparation and characterization of zeolite from waste Linz-Donawitz (LD) process
slag of steel industry for removal of Fe3+ from drinking water, Adv. Powder
Technol. vol. 32 (9) (2021) 3372–3387, https://doi.org/10.1016/j.
apt.2021.07.023.
[42] I. Das, S. Das, S. Sharma, M.M. Ghangrekar, Ameliorated performance of a
microbial fuel cell operated with an alkali pre-treated clayware ceramic
membrane, Int. J. Hydrog. Energy vol. 45 (33) (2020) 16787–16798, https://doi.
org/10.1016/j.ijhydene.2020.04.157.
10
Download