You are on page 1of 12

Science of the Total Environment 856 (2023) 159105

Contents lists available at ScienceDirect

Science of the Total Environment


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

Tailoring a highly conductive and super-hydrophilic electrode for


biocatalytic performance of microbial electrolysis cells
Sung-Gwan Park a,b, Chaeyoung Rhee c, Dipak A. Jadhav a, Tasnim Eisa a,b, Riyam B. Al-Mayyahi a,b, Seung Gu Shin c,
⁎ ⁎⁎
Mohammad Ali Abdelkareem d,e,f, , Kyu-Jung Chae a,b,
a
Department of Environmental Engineering, College of Ocean Science and Engineering, Korea Maritime and Ocean University, 727 Taejong-ro, Yeongdo-gu, Busan 49112, Republic of Korea
b
Interdisciplinary Major of Ocean Renewable Energy Engineering, Korea Maritime and Ocean University, 727 Taejong-ro, Yeongdo-gu, Busan 49112, Republic of Korea
c
Department of Energy Engineering, Future Convergence Technology Research Institute, Gyeongsang National University, 501 Jinju-daero, Jinju, Gyeongnam 52828, Republic of Korea
d
Chemical Engineering Department, Faculty of Engineering, Minia University, Minia, Egypt
e
Center of Advanced Materials Research, Research Institute of Science and Engineering, University of Sharjah, P.O. Box 27272, Sharjah, United Arab Emirates
f
Department of Sustainable and Renewable Energy Engineering, University of Sharjah, PO Box 27272, Sharjah, United Arab Emirates

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• The highly conductive and hydrophilic


properties were applied to carbon mate-
rial.
• The Coulombic efficiency was increased
by two-fold after functionalization of
anode.
• Increased internal mass transfer promoted
microbial growth on the anode surface.
• The bifunctional anode plays a role as an
electron-hive for efficient H2 production.
• Biofilm and microbial dynamics in 3D
structural anodes were first profiled in
depth.

A R T I C L E I N F O A B S T R A C T

Editor: Huu Hao Ngo Bioelectrochemical hydrogen production via microbial electrolysis cells (MECs) has attracted attention as the next gen-
eration of technology for the hydrogen economy. MECs work by electrochemically active bacteria reducing organic
Keywords: compounds at the anode. However, the hydrophobic nature of carbon-based anodes suppresses the release of the pro-
Bioelectrochemical system duced gas and water penetration, which significantly reduces the possibility of microbial attachment. Consequently, a
Hydrogen
limited surface area of the anode is used, which decreases hydrogen production efficiency. In this study, the bifunc-
Biocompatibility
Polymerized anode material
tional material poly(3,4-ethylenedioxythiophene):polystyrene sulfonate (PEDOT:PSS) was applied to the surface of a
Electrochemically active bacteria three-dimensional carbon felt anode to enhance the hydrogen production efficiency of an MEC owing to the high con-
Microbial electrolysis cells ductivity of PEDOT and super-hydrophilicity of PSS. In experiments, the PEDOT:PSS-modified anode almost doubled
the hydrogen production efficiency of the MEC compared with the control anode owing to the increased capacitance
current (239.3 %) and biofilm formation (220.7 %). The modified anode reduced the time required for the MEC to
reach a steady state of hydrogen production by 14 days compared to the control anode. Microbial community profiles
demonstrated that the modified anode had a greater abundance of electrochemically active bacteria than the control
anode. This simple method could be widely applied to various bioelectrochemical systems (e.g., microbial fuel cells
and solar cells) and to scaling up MECs.

Abbreviations: CE, Coulombic efficiency; CF, carbon felt; COD, chemical oxygen demand; CV, cyclic voltammetry; EAB, electrochemically active bacteria; EDLC, electric double-layer capaci-
tance; EDX, energy-dispersive X-ray; EIS, electrochemical impedance spectroscopy; GF, graphite felt.
⁎ Correspondence to: M.A. Abdelkareem, Chemical Engineering Department, Faculty of Engineering, Minia University, Minia, Egypt.
⁎⁎ Correspondence to: K.J. Chae, Department of Environmental Engineering, College of Ocean Science and Engineering, Korea Maritime and Ocean University, 727 Taejong-ro, Yeongdo-gu,
Busan 49112, Republic of Korea.
E-mail addresses: mabdulkareem@sharjah.ac.ae (M.A. Abdelkareem), ckjdream@kmou.ac.kr (K.-J. Chae).

http://dx.doi.org/10.1016/j.scitotenv.2022.159105
Received 20 July 2022; Received in revised form 14 September 2022; Accepted 24 September 2022
Available online 29 September 2022
0048-9697/© 2022 Elsevier B.V. All rights reserved.
S.-G. Park et al. Science of the Total Environment 856 (2023) 159105

1. Introduction Thus, most studies on polymer-modified anodes mainly focused on in-


creasing electron transfer efficiency by enhancing conductivity. However,
As the energy sector moves toward the era of the hydrogen (H2) econ- few studies have considered improving the hydrophilicity. The uniform
omy, sustainable and efficient H2 production has become a critical issue. formation of microbial communities in 3D anodes is crucial for the rapid
Microbial electrolysis cells (MECs) represent an ecofriendly technology decomposition of organic compounds and the promotion of extracellular elec-
that recovers clean energy such as H2 from organic waste via tron transfer. In this context, poly(3, 4-ethylenedioxythiophene) (PEDOT)
bioelectrochemical activity (Yang et al., 2021). In MECs, the organic com- and polystyrene sulfonate (PSS) can be combined into a dual-functional ma-
pounds in wastewater are initially degraded into CO2 and biomass, while terial as a unique potential candidate for improving the MEC performance.
the electrochemically active bacteria (EAB) generate protons and electrons PEDOT is a positively charged ionic polymer that is highly conductive but
at the anode (Jadhav et al., 2020; Santoro et al., 2022). Then, protons are is naturally hydrophobic. However, when it is electropolymerized with PSS,
reduced by electrons at the cathode to produce H2. However, this reduction which is a negatively charged ionic polymer with SO− 3 groups and thus hy-
at the cathode may be hindered by insufficient generation of protons and drophilic, PSS surrounds the outside of PEDOT, so PEDOT:PSS is hydrophilic
electrons at the anode. Therefore, improving the bioelectrocatalytic activity (Lu et al., 2019). Furthermore, the strong interaction between PEDOT and
at the anode is crucial for the commercialization of MECs and realizing sus- PSS can result in anodes with high stability and easy modification. Zajdel
tainable and efficient H2 production (Jadhav et al., 2022). To increase the et al. (2018) developed a bioanode covered with a multilayer conductive bac-
concentrations of protons and electrons at the cathode, two essential pro- terial composite film using PEDOT:PSS that improved the MFC performance.
cesses must be achieved: the rapid degradation of organic compounds Despite its effectiveness, the enrichment of specific microbial cultures for
and efficient electron transfer to the anode with minimal electron loss. In industrial-scale plants is technically challenging. PEDOT:PSS increases both
particular, EAB, which are also known as exoelectrogens, play a crucial the hydrophilicity and electron transfer of carbon-based anodes and allows
role in both processes (Pawar et al., 2022). The anode should have high for easy mass transfer within. The hydrophilicity of the anode is important be-
conductivity and biocompatibility to allow for vigorous EAB proliferation. cause it accelerates the biofilm formation. The hydrophobic nature of CF can
In this context, several types of carbon-based materials such as carbon felt increase the resistance of the interface between the solution and anode by
(CF), carbon nanotubes (CNTs), carbon cloth, and carbon paper have preventing surface wetting. Additionally, this hydrophobicity can induce
been widely used as anodes because they are cheaper than other conductive the capture of produced gases, which interferes with internal mass transfer.
materials and have a three-dimensional (3D) structure comprising multiple Therefore, the hydrophilic surface caused by the SO− 3 groups of PEDOT:PSS
pores that is suitable for microbial community growth (Park et al., 2022). smoothens the flow of the internal solution, hence increasing the microorgan-
However, the hydrophobicity of carbon-based materials hinders the mass ism attachment possibility and facilitating the natural release of produced
transfer of the solution and substrates, which forms dead space in the elec- gases. Fig. 1 shows the concept of functionalizing the surface of a CF anode
trode. In other words, only the outer surface of the electrode is available for with PEDOT:PSS to improve the MEC performance.
microbial growth, so only a 2D surface is used despite the electrode having In this study, a simple method was used to apply a highly conductive
a 3D structure. and super-hydrophilic electrode to an MEC for enhanced H2 production.
Acid treatment is typically used to increase the surface area and polarity The optimal electropolymerization conditions of the PEDOT:PSS anode
of carbon-based materials by damaging the carbon fibers (Zhang et al., were determined based on the physical and chemical characteristics. The
2009). However, physically pitted and fragmented carbon fibers become long-term effects of PEDOT:PSS application on the MEC performance
extremely fragile and have been shown to lose 28.4 % of their weight were investigated over a period of >4 months. The biofilm formation and
when treated with 70 % nitric acid for 240 min (Wu et al., 1995). Hydro- microbial dynamics in 3D-structure anodes were profiled in depth to pro-
philic conductive polymers are a sustainable alternative because they vide insight into the relationship between the microbial community struc-
enable super-hydrophilicity without any damage or weight loss of the car- ture and MEC performance. The findings of this study may contribute to
bon fibers. This enables a rapid exchange of solution across carbon fibers the potential application of PEDOT:PSS in biocatalytic materials, particu-
for active microbial growth even inside the anode. In addition, the electrical larly for bioelectrochemical systems.
conductivity of the polymer increases the conductivity of the treated carbon
fibers while acid treatment cannot enhance the conductivity beyond the 2. Materials and methods
original level.
Most of the recent attempts to modify carbon-based anodes have used 2.1. Preparation of functionalized anodes
conductive polymers to enhance the electron transfer efficiency. For exam-
ple, Cui et al. (2015) electropolymerized polyaniline (PANI) on the surface Anodes were manufactured in accordance with the design of the MEC
of large porous graphite felt (GF) followed by the electrophoretic deposi- reactor used in a previous study (Park et al., 2017, 2021). Rinsed CF (5 ×
tion of CNTs to improve the performance of microbial fuel cells (MFCs). 5 cm2, thickness 5 mm) was attached to stainless steel as an electron collec-
MFCs are similar to MECs within the framework of bioelectrochemical tor, which was used as a control anode for the experiment (Fig. 2).
systems except that the former outputs electricity while the latter outputs Electropolymerization solutions (EPSs) were prepared by mixing 3,4-
hydrogen (Sunghoon et al., 2021). With the CNT-125/PANI/GF anode, ethylenedioxythiophene (EDOT, Sigma–Aldrich), poly(sodium 4-styrene
they realized a maximum power density (308 mW/m2) that was 6.2- and sulfonate) (NaPSS, Sigma–Aldrich), lithium chloride (LiClO4, Samchun
3.9-fold higher than those with the original GF (49 mW/m2) and PANI/GF Chemicals), and deionized (DI) water at five concentrations. LiClO4 was
(80 mW/m2) anodes, respectively. Feng et al. (2010) modified an anode used as the electrolyte in a mixed solution. Concentration 8 (CONC.8)
by using polypyrrole/anthraquinone-2,6-disulphonic disodium salt (PPy/ was obtained by referring to Benoudjit et al. (2018), who successfully
AQDS) and demonstrated a dual-chamber MFC with a maximum power den- electropolymerized PEDOT:PSS on electrochemical sensors, and four addi-
sity of 1303 mW/m2, which was 13-fold higher than that with an unmodi- tional concentrations (i.e., CONC.1, 2, 4, and 16) were added to identify the
fied anode. Li et al. (2014) prepared PANI/multiwalled carbon nanotube effects of different PEDOT:PSS concentrations (Table 1).
(MWCNT) and PPy/MWCNT composite anodes and revealed that the Electropolymerization was conducted to functionalize the CF surface by
PPy/MWCNT anode (34.95 Ω) exhibited better electrical conductivity using a three-electrode system comprising a control anode as the working
than the PANI/MWCNT (87.93 Ω) and unmodified (226.20 Ω) anodes. Fur- electrode, a graphite counter electrode, and an Ag/AgCl reference electrode
thermore, some researchers fabricated magnetic and conductive anodes by (Fig. S1a). A galvanostat (ZIVE SP1, WonATech, Korea) was operated in
encapsulating uniformly dispersed SrFe12O19 nanoparticles into poly(vinyl galvanic mode at 100 mA for 12 h with each EPS concentration to fabricate
alcohol-co-ethylene) nanofibers to form a 3D PPy network and realized a the PEDOT:PSS-functionalized CFs (PEDOT:PSS@CF). The three-electrode
maximum power density of 3317 mW/m2, which was significantly greater cell was composed of quartz instead of the commonly used acrylic material
than that of the control anode (2471 mW/m2) (Li et al., 2019). to prevent undesired reactions such as deformation of the surface into deep

2
S.-G. Park et al. Science of the Total Environment 856 (2023) 159105

Fig. 1. Conceptual diagram of the advantage from functionalizing the MEC anode surface with PEDOT:PSS.

white during electropolymerization. The quartz reactor allowed identical spectroscopy (EDX) to confirm that PEDOT:PSS covered the CF. The ele-
conditions to be maintained during each electropolymerization, such as mental structure and functional groups of the anodes were analyzed by
the distances between the three electrodes. Finally, the anodes were heat- using Raman spectroscopy (JASCO, NRS-5100, Japan) with laser excitation
treated at 110 °C for 12 h to minimize the release of PEDOT:PSS into the at 532 nm (range of 700–4200 cm−1) and Fourier-transform infrared spec-
MEC anolyte. The prepared anodes were then equipped in the anode cham- troscopy (FT-IR; Spectrum GX, Perkin Elmer, USA) in attenuated total re-
ber of the H2-producing MEC for performance evaluation (Fig. S1b). flectance mode (650–4000 cm−1).
The water contact angle (WCA) test (DSA25, Kruss, Germany) was con-
2.2. Characterization of functionalized anodes ducted to evaluate the effect of different PEDOT:PSS concentrations on the
hydrophilicity of the anode. During the WCA analysis, 6 μL of DI water was
The surface morphology and elemental composition of the anodes were dropped on the surface and inside of the anodes, and the angle between the
characterized by field emission scanning electron microscopy (FE-SEM, droplet and sample was measured at 21° ± 1°. The immediate water ab-
TESCAN-MIRA 3 LMU, Czech) coupled with energy-dispersive X-ray sorption performance was filmed and can be seen in Video S1.

Fig. 2. Preparation of control and PEDOT:PSS anodes.

3
S.-G. Park et al. Science of the Total Environment 856 (2023) 159105

Table 1 The gas production rate was automatically logged in real-time by using
EPS concentrations for PEDOT:PSS synthesis. a respirometer (BRS-110, EETech, South Korea), and its composition was
EPS concentrations DI water (mL) LiClO4 (g) EDOT (mL) NaPSS (g) analyzed by gas chromatography (Clarus 580, Perkin Elmer, USA)
CONC.1 20 0.0112 0.06 0.1
equipped with a thermal conductivity detector and Elite-PLOT Q Column
CONC.2 20 0.0112 0.12 0.2 (6 f. × 1/8″, SS) with nitrogen gas (99.999 %) as the carrier gas. The gas
CONC.4 20 0.0112 0.25 0.4 for gas chromatography analysis was sampled by using a gastight syringe
CONC.8 20 0.0112 0.5 0.8 (2.5 mL, Hamilton SampleLock Syringe # 1002, USA) from the headspace
CONC.16 20 0.0112 1 1.6
of each chamber. The amount of H2 production was calculated based on
the gas production rate and gas composition results. The overall H2 effi-
ciency (OHE) from the substrate to H2 was calculated as the ratio of the ac-
The electrocatalytic activity of the samples was examined by using cy- tual amount of H2 produced and theoretical amount of H2 in the substrate.
clic voltammetry (CV) and electrochemical impedance spectroscopy (EIS; To compare the characteristics of H2 production in the MECs, the maximum
ZIVE SP1, WonATech, South Korea) on a three-electrode cell (Fig. S1a). gas production rate, saturation constant, and reaction immediacy after sub-
The prepared anodes were vertically sliced into three layers to obtain the strate injection were determined by using the modified Gompertz model,
inside (first and third layers) and outside (second layer) layers of the elec- which Ware and Power applied to the degradation of simple organic sub-
trode. Additionally, the electrochemical behaviors of the first and second strates such as acetate (Ware and Power, 2017):
layers were investigated. Each sample was connected to ZIVE SP1 to act   
as the working electrode, and Pt wire and Ag/AgCl were used as the counter Rmax  e
M ¼ Pmax  exp  exp  ðλ  tÞ þ 1 (2)
and reference electrodes, respectively. CV tests were conducted at −1 to Pmax
1 V at 50 mV/s in a phosphate buffer solution (pH 7) under anoxic condi-
tions, which were achieved by nitrogen gas purging. EIS measurements where M is the gas production (mL), Pmax is the gas production constant,
were performed under the same conditions as CV, but the frequency Rmax is the maximum production rate constant, λ is the lag-phase time
range, potential, and amplitude were fixed at 0.01–100 kHz, 0 V (vs. (h), and t is the operating time (h). The software program MATLAB
open-circuit potentiometry), and 10 mV, respectively. R2015a was used to fit the experimental gas production data to the model-
ing algorithm.
2.3. Microbial electrolysis cell setup and operation
2.5. Microbial attachment and community analysis according to anode depth
H2-producing MECs were operated with two different anodes (CF and
PEDOT:PSS@CF) to compare their performances (Fig. S1c). MEC opera- 2.5.1. Sample preparation
tions with each anode were conducted in triplicate: CF-MEC1, CF-MEC2, Biofilm formation on the anodes of all experimental reactors was ob-
and CF-MEC3 for CF and PEDOT:PSS-MEC1, PEDOT:PSS-MEC2, and served after four months of MEC operation (Fig. S2). To identify the micro-
PEDOT:PSS-MEC3 for PEDOT:PSS@CF. The MEC reactors comprised an- bial growth distribution and differences in microbial diversity depending
odic and cathodic chambers (200 mL; working volume = 160 mL in each on the depth of the anode, a 1 × 1 cm2 piece was obtained from the anodes
chamber) separated by a Nafion-211 membrane (5 × 5 cm2, DuPont, and sliced into five layers, each with a thickness of 1 mm (Fig. S3). FE-SEM,
USA). Identical Pt-coated perforated Ti plates (0.5 mg Pt/cm2) were used next generation sequencing (NGS), and quantitative polymerase chain reac-
as cathodes. The anodic chamber was filled with a mixture of nutrient min- tion (qPCR) were conducted as follows.
eral buffer solution and anaerobic digestion sludge (Soo-young Wastewater
Treatment Plant, South Korea) to inoculate microorganisms on the anode 2.5.2. Field emission scanning electron microscopy
surface while the cathodic chamber was filled with phosphate buffer To observe the shape of microbes on both the surface and inside of the
(electrolyte, pH 7). The substrate (0.5 mL of 500 mM acetic acid) was anode using FE-SEM, the first and third anode layers were carefully
injected into the anodic chamber for every batch after purging with nitro- dehydrated, which was followed by the fixation process. The samples
gen gas. In total, 30 batches were tested for 4 months. To supply an energy were soaked in 2.5 % glutaraldehyde (primary fix, Sigma–Aldrich, USA)
deficit, an external voltage of 0.5 V was applied to the MEC reactors by and 1 % osmium tetroxide (secondary fix, Sigma–Aldrich, USA) for 1 h
using a DC power supply (OPM\\303D, ODA, South Korea). A multimeter and were washed three times with DI water after each step. Then, the dehy-
(Keithley 2700, USA) was used to monitor the voltage and calculate the cur- dration process was performed by using increasing concentrations (50 %,
rent through the external wire. All experiments were performed at 30 °C 70 %, 90 %, 95 %, and twice 100 %) of ethanol, and the solution was
while using an incubator with a magnetic stirrer (350 rpm). changed twice to 99.9 % hexamethyldisilazane (Sigma–Aldrich, USA) for
10 min per step. Before observation, the samples were dried overnight at
2.4. Electron transfer and gas production evaluation 25 °C ± 1 °C, and then the sample surfaces were coated with Pt by ion
sputtering (E-1030, Hitachi Ltd., Japan).
The substrate-to-electron conversion efficiency can be expressed
in terms of the Coulombic efficiency (CE) from the measured current 2.5.3. 16S rRNA gene sequencing and bacterial taxonomic assignment
flowing from the anode to the cathode through an external wire using a The five layers of each anode were sonicated with distilled water and
fixed resistance: centrifuged to obtain cell pellets. DNA was extracted by using the AccuPrep
Genomic DNA Extraction Kit (Bioneer, South Korea). To normalize the DNA
Rt extraction yield, a self-ligated pGEM-T easy vector (3015 base pairs,
M 0 I dt
CE ð%Þ ¼ (1) Promega, USA) was added during the cell lysis step. In total, 108 copies
FbV an ΔCOD
of the internal standard were co-extracted with sample DNA and purified
in 100 μL of elution buffer. The amount of eluted internal standard was
where M is the molecular weight of O2 (32 g/mol), F is Faraday's constant quantified by real-time PCR.
(96,485 C/mol), b is the number of moles of electrons produced per mole Amplicon sequencing was performed according to the manufacturer's
of substrate (4), and Van is the volume of the reactor. The current (I) flowed instructions (Illumina, CA, USA). Oligomers containing the specific se-
through a fixed resistance (10 Ω) on the cathode side and was automatically quence of the V3–4 region of bacterial 16S rDNA (518F: 5′-CCAGCAGCC
monitored by a Keithley multimeter. The chemical oxygen demand (COD) GCGGTAATACG-3′ and 805R: 5′-GACTACCAGGGTATCTAATCC-3′)
was measured before and after each batch by using an HS COD-LR kit (Lane, 1991; Yu et al., 2005) and the Illumina overhang adapter sequence
(Humas, HS 2300 Plus, Korea). were used as amplicon primers. Sample DNA was PCR-amplified through

4
S.-G. Park et al. Science of the Total Environment 856 (2023) 159105

28 cycles at an annealing temperature of 58 °C. The purified amplicons were Biosystems, USA). The qPCR assay was conducted by using the PowerUp
PCR-indexed by using the Illumina Nextera XT index kit. A TaqMan DNA po- SYBR Green Master Mix (Applied Biosystems) with the bacteria-specific
lymerase kit (Solgent, South Korea) was used for library construction. The pu- primers and probe set (338F; 5′-ACTCCTACGGGAGGCAG-3′, 516F; 5′-
rified library was quantified, pooled, and combined with PhiX control TGCCAGCAGCCGCGGTAATAC-3′, and 805R; 5′-GACTACCAGGGTATCT
(Illumina). The amplicon sequencing was paired-end (150 bp × 2) sequenced AATCC-3′) (Shin et al., 2010). The cycling conditions were as follows: ini-
by using the iSeq 100 platform. Sequence reads were first merged and then tial denaturation at 95 °C for 2 min followed by 40 cycles of denaturation
processed to remove short-or low-quality sequences and potential chimeras. at 95 °C for 6 s and primer annealing and elongation at 60 °C for 45 s.
More than 43,000 filtered reads (average: 76,000 reads) were obtained The DNA standard was previously constructed as a purified plasmid derived
from each sample. Operational taxonomic units (OTUs) were defined at a from the bacterial species Escherichia coli k12 (Shin et al., 2010). To normal-
97 % sequence identity cutoff by using the VSEARCH algorithm (Rognes ize the DNA extraction yield, the internal standard (pGEM-T easy plasmid)
et al., 2016). Taxonomic assignment was conducted by using the online was amplified with T7 (5′-TAATACGACTCACTATAGGGCGA-3′) and SP6
RDP classifier (https://rdp.cme.msu.edu/classifier/). (5′-ATTTAGGTGACACTATAGAATACTCAAGCT-3′) primers by using the
PowerUp SYBR Green Master Mix. The cycling conditions were as follows:
2.5.4. Real-time quantitative polymerase chain reaction and quantification of initial denaturation at 95 °C for 2 min followed by 40 cycles of denaturation
bacterial population at 95 °C for 5 s and primer annealing and elongation at 60 °C for 30 s. The
To measure the population size of bacteria, real-time qPCR was DNA extraction yield (%) was determined by dividing the plasmid copy
performed by using the StepOnePlus Real-Time PCR system (Applied number in the extracted DNA by the input copy number of the plasmid.

Fig. 3. FE-SEM and EDX results for the surfaces of (a) CF and (b) PEDOT:PSS@CF with heat treatment. (c) Chemical structure of PEDOT:PSS. (d) Raman spectra and (e) FT-IR
results for CF and PEDOT:PSS@CF, where the rhombuses and stars indicate the thiophene ring and SO− 3 group, respectively.

5
S.-G. Park et al. Science of the Total Environment 856 (2023) 159105

3. Results and discussion stretching bend (Mengistie et al., 2013). Therefore, these results indicate
that the CF surface was successfully functionalized by PEDOT:PSS.
3.1. Confirmation of the functionalized anode surface
3.2. Characterization of the anodes
The surface morphology of the prepared anodes was confirmed by FE-
SEM (Fig. 3a and b). The CF anode had a clean surface, whereas the surface 3.2.1. Wettability
of the PEDOT:PSS@CF anodes was overlaid by a thin layer with a rough Fig. 4a shows the wettability, represented by the WCA, of the samples
texture. The heat treatment, which was intended to increase the durability according to the PEDOT:PSS concentration. Pristine CF had the largest
of the PEDOT:PSS coating, also seemed to increase the surface roughness. WCA at 136.5° ± 12.3°, and the hydrophilicity increased with the
The functionalization of the CF surface with PEDOT:PSS was confirmed PEDOT:PSS concentration. Super-hydrophilicity (i.e., WCA = 0°) was ob-
by EDX. The original CF comprised 100 % elemental C, whereas the surface tained at PEDOT:PSS@CF_CONC.8. For further investigation, the prepared
of the PEDOT:PSS@CF showed a lower C content of 77.05 wt% with S and anodes were cut in half, and the wettability inside and outside the anodes
O contents of 17.09 and 5.86 wt%, respectively. This supports the presence was measured (Fig. 4b and c). Pristine CF was hydrophobic both outside
of PEDOT:PSS, whose structure comprises elemental S and O from PEDOT (136.5° ± 12.3°) and inside (143.2° ± 4.3°). PEDOT:PSS@CF_CONC.8
and the SO− 3 groups of PSS on its surface (Fig. 3c). was super-hydrophilic both outside (0°) and inside (0°), which indicates
The chemical structure of the prepared anodes was examined by using that the coating was uniform across the entire anode (Video S1).
Raman spectroscopy (Fig. 3d). The presence of the precursors EDOT and Sharma et al. (2019a) studied the physicochemical properties of carbon-
NaPSS was observed in the Raman peaks of PEDOT:PSS@CF, which indi- based materials. These materials showed >100° of WCA regardless of sur-
cates that the CF surface was successfully functionalized by PEDOT:PSS. face area and porosity, causing low mass transfer into the electrode. How-
The Raman spectrum of CF showed only two peaks (D band: 1358 cm−1 ever, the superhydrophilic property allowed the solution to penetrate
and G band: 1589 cm−1), which are typical of carbon-based materials. deeply into the electrode and completely utilize the 3D porous structure
The peaks at 1003, 1269, 1361, and 1436 cm−1 in PEDOT:PSS@CF repre- of CF. This should increase the mass transfer, electron transfer efficiency,
sent Cβ–Cβ′ asymmetric, Cα–Cα′ inter-ring stretching, Cβ–Cβ′ stretching, and and biocompatibility of the anode. Based on the previous results, any fur-
Cα = Cβ symmetric stretching vibrations, respectively (Chang et al., 2014; ther discussion of the PEDOT:PSS@CF results is referring to PEDOT:PSS@
Xiong et al., 2013). The peak at 1507 cm−1 corresponds to the Cα = Cβ CF_CONC.8.
asymmetric elongation vibration associated with the thiophene ring Guselnikova et al. (2017) also performed wettability studies using
in the PEDOT chain (Lindfors et al., 2014). In addition, a peak at PEDOT:PSS. They observed that pristine PEDOT:PSS films exhibited high
1565 cm−1 was observed indicating an obvious quinoid structure (Cα–Cβ) hydrophilicity owing to the presence of polar groups in the polymer
(Funda et al., 2016). These peaks have also been identified in commercial blend. Moreover, when PEDOT:PSS was modified by different substituents
PEDOT:PSS polymers, which indicates that PEDOT:PSS thoroughly coated (NO2, NH2, and CF3), the WCA increased because the hydrophilic SO− 3
the CF surface. groups turned into nonpolar esters. This implies that the SO− 3 groups of
FT-IR analysis was used to evaluate the surface compounds of the pre- PSS clearly contribute to the hydrophilicity of PEDOT:PSS. Lu et al.
pared anodes with IR absorption properties that could not be identified (2019) developed pure PEDOT:PSS hydrogels that are highly desirable for
by Raman analysis (Fig. 3e). PEDOT:PSS@CF demonstrated the asymmetric bioelectronic applications. They noted that a hydrophobic PEDOT-rich
stretching modes of Cα = Cβ, symmetric Cα = Cβ, and asymmetric Cα–Cβ core covered by a hydrophilic PSS-rich shell can improve the intercon-
indicating the quinoid structure of the thiophene ring by peaks at 1716, nected networks of PEDOT:PSS due to the water-swelling characteristics
1654, and 1427 cm−1, respectively (Han et al., 2014; Xu et al., 2017). of the PSS-rich domain, which agrees with the results of the present
The bands at 1334 cm−1 correspond to ring strains in the stretching study. Biessmann et al. (2018) investigated the swelling behavior of
mode of ethylenedioxy (C–O–C) group (Susanti et al., 2018). A band ap- PEDOT:PSS electrodes under humid conditions and observed the highest
peared at approximately 1137 cm−1 because of the stretching vibrations swelling ratio and water content in pristine PEDOT:PSS thin films. They
of the SO− 3 groups of PSS (Xu et al., 2017). The observed peak at attempted to decrease the swelling of PEDOT:PSS by treatment with
998 cm−1 (C\\S) can be attributed to the stretching vibrations of the thio- Zonyl and post-treatment with ethylene glycol to stabilize the PEDOT:PSS
phene ring (Olivares et al., 2019). The peak at 2875 cm−1 indicates a C\\H thin films. In the present study, however, the swelling of PEDOT:PSS was

Fig. 4. (a) WCA of prepared samples according to the PEDOT:PSS concentrations given in Table 1. Inside and outside WCAs of (b) CF and (c) PEDOT:PSS@CF.

6
S.-G. Park et al. Science of the Total Environment 856 (2023) 159105

used to allow the solution to penetrate the anode, which increased the mass be attributed to the high scan rate (50 mV/s), which exceeded the charging
transfer and fully exploited the 3D structure. speed of the electrochemical capacitor, as well as the pore resistance for
mass transfer (Lee et al., 2015). Neither electrode showed any irreversible
or faradic response, which indicated their inactivity in phosphate buffer
3.2.2. Electrochemical properties and consequently their potential to promote stable and direct contact for in-
The cyclic voltammograms showed mixed capacitive behavior along the terspecies electron transfer. However, the PEDOT:PSS@CF electrode exhib-
tested potential for both CF and PEDOT:PSS@CF (Fig. 5a). The current in- ited a considerably higher capacitance current (9.90 mA) than the CF
creased positively with the charging forward scan, followed by a similar electrode (4.17 mA) at the same specific potential (0.37 V) owing to its in-
smooth increase in negative current during the reverse scan. In both for- herent capacitive properties (Volkov et al., 2017). This increase in the redox
ward and reverse scans, the change in current showed no peaks or a sudden current indicates that the modified anode demonstrated better electrocata-
onset of rapid current increase. Therefore, the response was inferred to ex- lytic behavior than the unmodified anode. The enhanced EDLC of the mod-
hibit capacitive behavior. Electric double-layer capacitance (EDLC) domi- ified anode may also be a result of the larger electrochemically active
nated the response, whereas pseudocapacitance was closer to the surface area and more accessible pores because of the higher hydrophilicity
maximum and minimum in the tested potential window (−1 and 1 V). (Barbieri et al., 2005; Chmiola et al., 2006; Hryniewicz et al., 2018). More-
The noticeable deviation from the rectangular response of the EDLC can over, the current densities were similar inside and outside PEDOT:PSS@CF.
Thus, PEDOT:PSS thoroughly coated even inside the CF, which increased
the active surface area and promoted electrical conductivity.
EIS was performed to obtain a better understanding of the electrochem-
ical behavior of the electrodes in the phosphate buffer. The EIS data before
and after electropolymerization showed a clear difference between CF and
PEDOT:PSS@CF. Overall, the Nyquist plot showed a decrease in the imped-
ance magnitude of the modified electrode (Fig. 5b and c). The impedance
spectrum showed three distinct regions depending on the dominant
limiting process in each frequency range: physical, electrochemical, or
mass transfer.
At high frequencies, the physical impedance had a close encounter with
the x-axis (i.e., real impedance), which represented the electrode resistance,
electrode–current collector contact resistance, and bulk electrolyte resis-
tance (Mei et al., 2018). The first minimum at the starting point for
PEDOT:PSS@CF ranged between 63 and 66 Ω, which is considerably
lower than that of CF, ranging between 829 and 895 Ω. The lower intercept
indicates that the introduction of PEDOT:PSS improved the conductivity of
the original electrode material. As the frequency decreased, the impedance
manifests as a distorted arc. The impedance varied between CF and PEDOT:
PSS. This represents a unique interaction at the electrode surface and is not
a result of a common feature such as electrolyte resistance. Consequently,
the phenomena were attributed to electrochemical interfacial surface inter-
actions between the electrolyte and electrode as well as the pore resistance
(Bastidas, 2007; Lee et al., 2015). Such impedance was high for CF because
the hydrophobic pores promoted the charging resistance at a lower capac-
itance (Choi, 2010). In contrast, the resistance of PEDOT:PSS@CF was
only ~16.2 Ω (Jones et al., 2020; Nuramdhani et al., 2018; Stöcker et al.,
2012). The distortion in the arc indicates non-ideal capacitance for the
EDLC (constant-phase element) (Chen et al., 2015; Mei et al., 2018;
Zajdel et al., 2018).
Finally, at lower frequencies, ions started to move a considerable dis-
tance demonstrating the effects of diffusion, as evident by the ascending
line following the arc. The cumulative diffusive behavior of both electrodes
can be better understood via anomalous diffusion owing to their complex
structures (Bisquert et al., 1999). For CF, higher frequencies prompted a
low-angle transition state best described by Warburg diffusion, which is
caused by a short diffusion length compared to the pore length at relatively
high AC signal frequencies (Bastidas, 2007; Lasia, 1995). Thereafter, ions
diffused further through the CF pore matrix against the 3D structure,
while the response further deviated from the Warburg ideal diffusion ac-
cording to the pore capacitance assuming a higher WCA (~78°).
Similarly, PEDOT:PSS exhibited complex behavior as a result of ion dif-
fusion through the CF pores because PSS opposed the charge transfer
through the PEDOT layer. The ionic diffusion through the conductive film
was slower than that through the CF pores as evidenced by the wider War-
burg diffusion response (Hernández-Labrado et al., 2011). The noticeable
diffusion effect proves the deposition and formation of PEDOT:PSS. There-
after, the accumulation of charges and ions in the electrode pores and con-
ductive film EDLC added a capacitive aspect to the diffusion at lower
Fig. 5. (a) CV and (b) EIS results of CF and PEDOT:PSS@CF; (c) magnified view of frequencies by increasing the slope. This behavior is common in ion-
the red box in (b). conductive polymers (Hernández-Labrado et al., 2011; Sharifi-Viand

7
S.-G. Park et al. Science of the Total Environment 856 (2023) 159105

Fig. 6. Performances of CF-MEC and PEDOT:PSS-MEC: (a) CE, (b) H2 production and OHE, and (c) gas composition (CH4 in anodic chamber and H2 in cathodic chamber) of
CF-MEC and PEDOT:PSS-MEC.

et al., 2012). The transition from Warburg to capacitive diffusion occurred (Hernández-Labrado et al., 2011; Sharma et al., 2019b). Meanwhile,
at similar frequencies both inside (0.4 Hz) and outside (0.8 Hz) PEDOT:PSS, CF_outside (~1.6 Hz) exhibited a higher difference in transition frequency
which implies similar film thicknesses, ion conductivities, and pore sizes compared to CF_inside (~0.5 Hz), which indicates differences in the pore

8
S.-G. Park et al. Science of the Total Environment 856 (2023) 159105

Fig. 7. Morphology of bacteria attached to the outside (first layer) and inside (third layer) of (a) CF-MEC and (b) PEDOT:PSS-MEC.

structure and hydrophilicity (Sharma et al., 2019b). The overall diffusion power production can be increased as much as CE increased because the
impedance of CF declined >10 decades upon the introduction of PEDOT: amount of electrons flowing from the anode to the cathode on the external
PSS owing to the enhanced hydrophilicity. wire determines the performance of the MFC (Chae et al., 2010; Kim et al.,
2008). To demonstrate reproducibility, the voltage measurements of all re-
3.3. Performance of the microbial electrolysis cell actors are shown in Fig. S4. Because of the current flowing along the exter-
nal wire, OHE began exceeding 10 % from batch 7 for PEDOT:PSS-MEC and
In Fig. 6, PEDOT:PSS-MEC showed a significantly higher average CE batch 11 for CF-MEC, and the maximum OHE was 73.4 % for PEDOT:PSS-
(41.89 % overall, 56.48 % for the last 10 cycles in the steady state) than MEC and 32.7 % for CF-MEC (Fig. 6b). All batch cycle data of H2 production
CF-MEC (18.25 % overall, 33.38 % for the last 10 cycles in the steady were averaged by using the modified Gompertz model for further compar-
state, p < 0.001; Table S1). In addition, PEDOT:PSS-MEC required 7 batches ison (Fig. S5). The maximum H2 production constant (Pmax) was twice as
to reach 10 % CE as opposed to 12 batches for CF-MEC, and PEDOT:PSS- high with PEDOT:PSS-MEC compared to with CF-MEC, and the H2 produc-
MEC achieved a maximum CE of 73.3 % as opposed to 38.2 % for CF- tion rate constant (Rmax) also increased from 0.68 for CF-MEC to 1.57
MEC (Fig. 6a). From the MFC's perspective, these results indicate that the for PEDOT:PSS-MEC. The lag-phase time constant (λ) was 6.68th batch

9
S.-G. Park et al. Science of the Total Environment 856 (2023) 159105

Fig. 8. (a) NMDS of bacterial communities with centroids and polygons (n = 3) (green: CF-MEC, red: PEDOT:PSS-MEC). (b) Rarefied richness of anode layers with the
annotation indicating statistical differences based on Duncan's test (n = 3) (C: CF-MEC, P: PEDOT:PSS-MEC). (c) 16S rRNA copy number of bacterial phyla present in
each anode layer.

(≒27 days) for CF-MEC and 3.25th batch (≒13 days) for PEDOT:PSS-MEC, After batch 30 (i.e., the end of MEC operation), the biofilms formed
which indicates that the latter had a faster response to the substrate. More- differently in each anode according to their biocompatibility (Fig. 7). CF
over, Pasupuleti et al. (2015) spent 70 days obtaining a maximum current showed a distinct boundary between the inside and surface, which
density from the enriched bioanode using CF. Similarly, in this study, CF- implies that the anolyte did not sufficiently penetrate the anode. In
MEC showed Pmax at the 16th batch (≒64 days), implying that progressively contrast, a well-developed biofilm was observed in PEDOT:PSS@CF
enriched EAB produce stable performance. PEDOT:PSS-MEC required more from the surface to the most internal area, which implies that
time to get Pmax (18th batch, ≒72 days). However, its Rmax was faster than PEDOT:PSS increased the internal mass transfer within the anode. Fur-
CF-MEC, resulting in a continuously increasing absolute amount of hydro- thermore, a large number of rod-shaped potential EAB were observed
gen production (mL). This implies that more EAB were enriched in the inside PEDOT:PSS@CF (Feng et al., 2014). They formed a rigid connec-
PEDOT:PSS-MEC than in the CF-MEC (Jang et al., 2013). The PEDOT: tion using their pilus to enable active electron transfer (Kumar et al.,
PSS-MEC began surpassing Pmax of CF-MEC from the 10th batch (≒40 2016).
days). These results indicate that the PEDOT:PSS coating reduced the initial
startup time and increased the amount of H2 available for maximum pro- 3.4. Microbial dynamics across the anode layers
duction. The additional cost of coating the basic CF with PEDOT:PSS
is around 14 USD/12.5 cm3CF when calculated based on CONC.8 Fig. 8a presents the bacterial community dynamics in terms of nonmetric
(EDOT: 5.8, NaPSS: 8.1, LiClO4: 0.1, and the power needed to conduct multidimensional scaling (NMDS) ordination. Each polygon and its centroid
electropolymerization). As the long-term effect of the PEDOT:PSS electrode are presented in triplicate. The CF-MEC community shifted greatly toward
has been verified, the initial coating cost can be covered by a continuous hy- the center of the anode (i.e., layer 3), whereas the outer layers (1 and
drogen production. 5) were close to each other, which indicate similar bacterial community
Introducing the PEDOT:PSS coating also improved the maximum purity compositions. In contrast, PEDOT:PSS-MEC showed no significant difference
of H2 in the cathodic chamber from 71.95 % to 82.56 % (Fig. 6c). In partic- among the layers, although the outermost (layer 1) was slightly different
ular, the increase in H2 purity in the cathodic chamber reduced the CH4 from the rest. Fig. 8b presents the microbial diversity within each sample
composition in the anodic chamber because the CH4 and H2 production in terms of the rarefied richness. All layers of PEDOT:PSS-MEC showed no
processes competed for electrons, as demonstrated previously (Park et al., significant difference in rarefied richness, and at least one annotation was
2019). In other words, the amount of EAB that could support H2 production identical to another. In contrast, the innermost layer of CF-MEC (C3) showed
increased. significantly less richness than those of C1, C2, and C5, which indicates poor

10
S.-G. Park et al. Science of the Total Environment 856 (2023) 159105

microbial growth and distribution at the center of the anode. These results CRediT authorship contribution statement
further imply that the PEDOT:PSS treatment made the entire 3D structure
of CF a suitable environment for microbial growth. Fig. 8c presents the pop- Sung-Gwan Park: Conceptualization, Methodology, Visualization,
ulation sizes of the bacterial phyla. Overall, PEDOT:PSS-MEC had a larger Writing – original draft. Chaeyoung Rhee: Data curation, Software, Visual-
bacterial population than CF-MEC. For CF-MEC, the population size signifi- ization, Writing – original draft. Dipak A. Jadhav: Writing – original draft,
cantly decreased toward the center of the anode. In contrast, PEDOT:PSS- Writing – review & editing. Tasnim Eisa: Writing – original draft. Riyam
MEC showed no significant differences in population size among layers B. Al-Mayyahi: Writing – review & editing. Seung Gu Shin: Resources.
except for layer 1, which had a significantly larger population owing to Mohammad Ali Abdelkareem: Supervision. Kyu-Jung Chae: Supervi-
the enhanced biocompatibility resulting from the positive effects of the sion, Writing – review & editing.
overlying PEDOT:PSS. In particular, CF-MEC showed a 98 % reduction
in the population size for layer 3 compared with layer 1, whereas Data availability
PEDOT:PSS-MEC showed an average reduction of only 63 %.
Proteobacteria was the most dominant bacterial phylum in all samples. Data will be made available on request.
The largest contributor to the phylum Proteobacteria was the genus
Dechloromonas, which was assigned to OTU 7 (Table S2). Dechloromonas,
with Geobacter, is a known EAB with both nitrifying and electrogenic func- Declaration of competing interest
tions, and it is commonly observed in the anodes of bioelectrochemical
systems such as MECs and MFCs (Hao et al., 2020; Li et al., 2021). Thus, the The authors declare the following financial interests/personal relation-
increase in the Proteobacteria population, including Dechloromonas, most likely ships which may be considered as potential competing interests: Kyu-Jung
enhanced the electron transfer and consequently improved H2 production. Chae reports financial support was provided by National Research Founda-
tion of Korea (NRF). Sung-Gwan Park reports financial support was pro-
3.5. Effect of bifunctional properties on the anode in bioelectrochemical systems vided by National Research Foundation of Korea (NRF). Kyu-Jung Chae
reports financial support was provided by Rural Development Administra-
Compared with the control anode, the PEDOT:PSS-CF anode clearly tion in Republic of Korea.
showed a heavier microbial load on both the surface and internal layers
(Fig. 7). Additionally, the total copy number of the bacterial population, es- Acknowledgments
pecially that of EAB including Dechloromonas, was significantly higher for
the PEDOT:PSS-CF anode (Fig. 8c). According to these results, surface char- This work was supported by the National Research Foundation of Korea
acteristics, such as roughness and hydrophilicity, affect the formation of (NRF) funded by the Korea government (No. 2019R1A2C1006356), Basic
biofilms in previous studies. Rough surfaces provide suitable conditions Science Research Program through the NRF funded by the Ministry of
for biofilm formation compared to smooth surfaces (Zhang et al., 2012). Education (2022R1I1A1A01072864) and “Cooperative Research Program
As confirmed in Section 3.1, PEDOT:PSS@CF has a rougher surface than for Agriculture Science and Technology Development (Project No.
CF. Moreover, as discussed in Section 3.2.1, the superhydrophilicity of PJ016259022021)” funded by Rural Development Administration in
PEDOT:PSS boosted the internal mass transfer of the substrate and inocu- Republic of Korea.
lum, resulting in a well-established microbial community throughout the
3D structure of the anode. The increased microbial population further con- References
tributed to H2 production by transferring more electrons. Microbial diver-
sity was more stable and greater across the layers with PEDOT:PSS, while Barbieri, O., Hahn, M., Herzog, A., Kötz, R., 2005. Capacitance limits of high surface area ac-
tivated carbons for double layer capacitors. Carbon 43, 1303–1310.
significantly lower diversity and more drastic microbial shifts were ob- Bastidas, D.M., 2007. Interpretation of impedance data for porous electrodes and diffusion
served in the control anode (Fig. 8a and b). Here, the higher EAB population processes. Corrosion 63, 515–521.
and stable microbial diversity across the PEDOT:PSS-CF anode layers likely Benoudjit, A., Bader, M.M., Wan Salim, W.W.A., 2018. Study of electropolymerized PEDOT:
PSS transducers for application as electrochemical sensors in aqueous media. Sens. Bio-
contributed to the reduced startup period in PEDOT:PSS-MEC. The electro-
Sens. Res. 17, 18–24.
chemical experiments (Section 3.2.2) also demonstrated that enhanced Biessmann, L., Kreuzer, L.P., Widmann, T., Hohn, N., Moulin, J.F., Muller-Buschbaum, P.,
electrical conductivity and reduced internal resistance were important fac- 2018. Monitoring the swelling behavior of PEDOT:PSS electrodes under high humidity
conditions. ACS Appl. Mater. Interfaces 10, 9865–9872.
tors in improving CE. These properties enabled the rapid transfer of elec-
Bisquert, J., Garcia-Belmonte, G., Fabregat-Santiago, F., Compte, A., 1999. Anomalous trans-
trons generated by the densely formed EAB on the PEDOT:PSS@CF anode port effects in the impedance of porous film electrodes. Electrochem. Commun. 1,
to the cathode with minimal energy loss. 429–435.
Chae, K.-J., Choi, M.-J., Kim, K.-Y., Ajayi, F.F., Chang, I.-S., Kim, I.S., 2010. Selective inhibi-
tion of methanogens for the improvement of biohydrogen production in microbial elec-
4. Conclusions trolysis cells. Int. J. Hydrog. Energy 35, 13379–13386.
Chang, S.H., Chiang, C., Kao, F., Tien, C., Wu, C., 2014. Unraveling the enhanced electrical
PEDOT:PSS successfully improved the CE of the anode; therefore, H2 conductivity of PEDOT:PSS thin films for ITO-free organic photovoltaics. IEEE Photonics
J. 6, 1–7.
yield was approximately two times higher than that of the original CF. Chen, C.-h., Kine, A., Nelson, R.D., LaRue, J.C., 2015. Impedance spectroscopy study of con-
The major contributing factors were increased EAB population size ducting polymer blends of PEDOT:PSS and PVA. Synth. Met. 206, 106–114.
and density at the anode and enhanced electrical conductivity. The Chmiola, J., Yushin, G., Dash, R., Gogotsi, Y., 2006. Effect of pore size and surface
area of carbide derived carbons on specific capacitance. J. Power Sources 158,
superhydrophilicity of PEDOT:PSS allowed for the rapid exchange of the so- 765–772.
lution across the carbon fibers, inducing a rich microbial distribution in the Choi, J.-H., 2010. Fabrication of a carbon electrode using activated carbon powder and appli-
anode. The FE-SEM results and NGS analysis confirmed the maximized oc- cation to the capacitive deionization process. Sep. Purif. Technol. 70, 362–366.
Cui, H.-F., Du, L., Guo, P.-B., Zhu, B., Luong, J.H.T., 2015. Controlled modification of carbon
cupancy of the EAB population in the 3D structural anode after the PEDOT: nanotubes and polyaniline on macroporous graphite felt for high-performance microbial
PSS treatment. This first attempt to apply the bioelectrocatalytic material fuel cell anode. J. Power Sources 283, 46–53.
(PEDOT:PSS) to modify the MEC anodes proved its potential applicability Feng, C., Ma, L., Li, F., Mai, H., Lang, X., Fan, S., 2010. A polypyrrole/anthraquinone-2,6-
disulphonic disodium salt (PPy/AQDS)-modified anode to improve performance of mi-
in bioelectrochemical systems, including MFCs, microbial solar cells,
crobial fuel cells. Biosens.Bioelectron. 25, 1516–1520.
MEC-coupled anaerobic digestion, and dark fermentation. Additionally, Feng, C., Li, J., Qin, D., Chen, L., Zhao, F., Chen, S., et al., 2014. Characterization of
the importance of fully utilizing the anode for the superior performance exoelectrogenic bacteria Enterobacter strains isolated from a microbial fuel cell exposed
of MECs has been highlighted. to copper shock load. PLOS ONE 9, e113379.
Funda, S., Ohki, T., Liu, Q., Hossain, J., Ishimaru, Y., Ueno, K., et al., 2016. Correlation be-
Supplementary data to this article can be found online at https://doi. tween the fine structure of spin-coated PEDOT:PSS and the photovoltaic performance
org/10.1016/j.scitotenv.2022.159105. of organic/crystalline-silicon heterojunction solar cells. 120.

11
S.-G. Park et al. Science of the Total Environment 856 (2023) 159105

Guselnikova, O.A., Postnikov, P.S., Fitl, P., Tomecek, D., Sajdl, P., Elashnikov, R., et al., 2017. electrobiofuels in microbial electrolysis cells through an on-demand control strategy
Tuning of PEDOT:PSS properties through covalent surface modification. J. Polym. Sci. B using the coenzyme M and 2-bromoethanesulfonate. Environ. Int. 131, 105006.
Polym. Phys. 55, 378–387. Park, S.-G., Rajesh, P.P., Hwang, M.-H., Chu, K.H., Cho, S., Chae, K.-J., 2021. Long-term ef-
Han, Y.-K., Chang, M.-Y., Li, H.-Y., Ho, K.-S., Hsieh, T.-H., Huang, P.-C., 2014. In situ synthesis fects of anti-biofouling proton exchange membrane using silver nanoparticles and
and deposition of conducting PEDOT:PSS nanospheres on ITO-free flexible substrates. polydopamine on the performance of microbial electrolysis cells. Int. J. Hydrog. Energy
Mater. Lett. 117, 146–149. 46, 11345–11356.
Hao, D.C., Li, X.J., Xiao, P.G., Wang, L.F., 2020. The utility of electrochemical Systems in Mi- Park, S.-G., Rajesh, P.P., Sim, Y.-U., Jadhav, D.A., Noori, M.T., Kim, D.-H., et al., 2022. Ad-
crobial Degradation of polycyclic aromatic hydrocarbons: discourse,diversity and design. dressing scale-up challenges and enhancement in performance of hydrogen-producing
Front. Microbiol. 11, 557400. microbial electrolysis cell through electrode modifications. Energy Rep. 8, 2726–2746.
Hernández-Labrado, G.R., Contreras-Donayre, R.E., Collazos-Castro, J.E., Polo, J.L., 2011. Pasupuleti, S.B., Srikanth, S., Venkata Mohan, S., Pant, D., 2015. Development of
Subdiffusion behavior in poly(3,4-ethylenedioxythiophene): polystyrene sulfonate exoelectrogenic bioanode and study on feasibility of hydrogen production using abiotic
(PEDOT:PSS) evidenced by electrochemical impedance spectroscopy. J. Electroanal. VITO-CoRE and VITO-CASE electrodes in a single chamber microbial electrolysis cell
Chem. 659, 201–204. (MEC) at low current densities. Bioresour. Technol. 195, 131–138.
Hryniewicz, B.M., Winnischofer, H., Vidotti, M., 2018. Interfacial characterization and Pawar, A.A., Karthic, A., Lee, S., Pandit, S., Jung, S.P., 2022. Microbial electrolysis cells for
supercapacitive behavior of PEDOT nanotubes modified electrodes. J. Electroanal. electromethanogenesis: materials, configurations and operations. Environ.Eng.Res. 27,
Chem. 823, 573–579. 200484.
Jadhav, D.A., Das, I., Ghangrekar, M.M., Pant, D., 2020. Moving towards practical applica- Rognes, T., Flouri, T., Nichols, B., Quince, C., Mahé, F., 2016. VSEARCH: a versatile open
tions of microbial fuel cells for sanitation and resource recovery. J. Water Process Eng. source tool for metagenomics. PeerJ 4 e2584-e2584.
38, 101566. Santoro, C., Bollella, P., Erable, B., Atanassov, P., Pant, D., 2022. Oxygen reduction reaction
Jadhav, D.A., Park, S.-G., Pandit, S., Yang, E., Ali Abdelkareem, M., Jang, J.-K., et al., 2022. electrocatalysis in neutral media for bioelectrochemical systems. Nat.Catal. 5, 473–484.
Scalability of microbial electrochemical technologies: applications and challenges. Sharifi-Viand, A., Mahjani, M.G., Jafarian, M., 2012. Investigation of anomalous diffusion and
Bioresour. Technol. 345, 126498. multifractal dimensions in polypyrrole film. J. Electroanal. Chem. 671, 51–57.
Jang, J.K., Hong, S.H., Ryou, Y.S., Lee, E.Y., Chang, I.S., Kang, Y.K., et al., 2013. Microbial Sharma, M., Alvarez-Gallego, Y., Achouak, W., Pant, D., Sarma, P.M., Dominguez-Benetton,
communities of the microbial fuel cell using swine wastewater in the enrichment step X., 2019a. Electrode material properties for designing effective microbial electrosynthesis
with the lapse of time. J.Korean Soc.Environ.Eng. 35, 973–977. systems. J. Mater. Chem. A 7, 24420–24436.
Jones, P.D., Moskalyuk, A., Barthold, C., Gutöhrlein, K., Heusel, G., Schröppel, B., et al., 2020. Sharma, T., Holm, T., Diaz-Real, J.A., Mérida, W., 2019b. Experimental verification of pore
Low-impedance 3D PEDOT:PSS ultramicroelectrodes. 14. impedance theory: drilled graphite electrodes with gradually more complex pore size dis-
Kim, I.S., Chae, K.-J., Choi, M.-J., Verstraete, W., 2008. Microbial fuel cells: recent advances, tribution. Electrochim. Acta 317, 528–541.
bacterial communities and application beyond electricity generation. Environ. Eng. Res. Shin, S.G., Han, G., Lim, J., Lee, C., Hwang, S., 2010. A comprehensive microbial insight into
13, 51–65. two-stage anaerobic digestion of food waste-recycling wastewater. Water Res. 44,
Kumar, R., Singh, L., Zularisam, A.W., 2016. Exoelectrogens: recent advances in molecular 4838–4849.
drivers involved in extracellular electron transfer and strategies used to improve it for mi- Stöcker, T., Köhler, A., Moos, R., 2012. Why does the electrical conductivity in PEDOT:PSS de-
crobial fuel cell applications. Renew. Sustain. Energy Rev. 56, 1322–1336. crease with PSS content?A study combining thermoelectric measurements with imped-
Lane, D.J., 1991. 16S/23S rRNA sequencing. Nucleic Acid Techniques in Bacterial Systematic. ance spectroscopy. 50, pp. 976–983.
John Wiley and Sons, New York, pp. 115–175. Sunghoon, S., Youngjin, K., Myeong Woon, K., Sokhee, P.J., 2021. Recent trends and pros-
Lasia, A., 1995. Impedance of porous electrodes. J. Electroanal. Chem. 397, 27–33. pects of microbial fuel cell technology for energy positive wastewater treatment plants
Lee, J.-H., Ahn, H.-J., Cho, D., Youn, J.-I., Kim, Y.-J., Oh, H.-J., 2015. Effect of surface modi- treating organic waste resources. J.Korean Soc.Environ.Eng. 43, 623–653.
fication of carbon felts on capacitive deionization for desalination. Carbon Lett. 16, Susanti, E., Wulandari, P., Herman, 2018. Effect of localized surface plasmon resonance from
93–100. incorporated gold nanoparticles in PEDOT:PSS hole transport layer for hybrid solar cell
Li, J., Liu, L., Zhang, D., Yang, D., Guo, J., Wei, J., 2014. Fabrication of polyaniline/silver applications. J. Phys. Conf. Ser. 1080, 012010.
nanoparticles/multi-walled carbon nanotubes composites for flexible microelectronic cir- Volkov, A.V., Wijeratne, K., Mitraka, E., Ail, U., Zhao, D., Tybrandt, K., et al., 2017. Under-
cuits. Synth. Met. 192, 15–22. standing the capacitance of PEDOT:PSS. 27.
Li, F., Wang, D., Liu, Q.Z., Wang, B., Zhong, W.B., Li, M.F., et al., 2019. The construction of Ware, A., Power, N., 2017. Modelling methane production kinetics of complex poultry slaugh-
rod-like polypyrrole network on hard magnetic porous textile anodes for microbial fuel terhouse wastes using sigmoidal growth functions. Renew. Energy 104, 50–59.
cells with ultra-high output power density. J. Power Sources 412, 514–519. Wu, Z., Pittman, C.U., Gardner, S.D., 1995. Nitric acid oxidation of carbon fibers and the ef-
Li, Z.-L., Zhu, Z.-L., Lin, X.-Q., Chen, F., Li, X., Liang, B., et al., 2021. Microbial fuel cell-upflow fects of subsequent treatment in refluxing aqueous NaOH. Carbon 33, 597–605.
biofilter coupling system for deep denitrification and power recovery: efficiencies, bacte- Xiong, S., Zhang, L., Lu, X., 2013. Conductivities enhancement of poly(3,4-
rial succession and interactions. Environ. Res. 196, 110331. ethylenedioxythiophene)/poly(styrene sulfonate) transparent electrodes with diol addi-
Lindfors, T., Boeva, Z.A., Latonen, R.-M., 2014. Electrochemical synthesis of poly(3,4- tives. Polym. Bull. 70, 237–247.
ethylenedioxythiophene) in aqueous dispersion of high porosity reduced graphene Xu, S., Liu, C., Xiao, Z., Zhong, W., Luo, Y., Ou, H., et al., 2017. Cooperative effect of carbon
oxide. RSC Adv. 4, 25279–25286. black and dimethyl sulfoxide on PEDOT:PSS hole transport layer for inverted planar pe-
Lu, B., Yuk, H., Lin, S., Jian, N., Qu, K., Xu, J., et al., 2019. Pure PEDOT:PSS hydrogels. Nat. rovskite solar cells. Sol. Energy 157, 125–132.
Commun. 10, 1043. Yang, E., Omar Mohamed, H., Park, S.-G., Obaid, M., Al-Qaradawi, S.Y., Castaño, P., et al.,
Mei, B.-A., Munteshari, O., Lau, J., Dunn, B., Pilon, L., 2018. Physical interpretations of 2021. A review on self-sustainable microbial electrolysis cells for electro-biohydrogen
Nyquist plots for EDLC electrodes and devices. J. Phys. Chem. C 122, 194–206. production via coupling with carbon-neutral renewable energy technologies. Bioresour.
Mengistie, D.A., Wang, P.-C., Chu, C.-W., 2013. Effect of molecular weight of additives on the Technol. 320, 124363.
conductivity of PEDOT:PSS and efficiency for ITO-free organic solar cells. J. Mater. Chem. Yu, Y., Lee, C., Kim, J., Hwang, S., 2005. Group-specific primer and probe sets to detect me-
A 1, 9907–9915. thanogenic communities using quantitative real-time polymerase chain reaction. 89,
Nuramdhani, I., Gokceoren, A.T., Odhiambo, S.A., De Mey, G., Hertleer, C., Van Langenhove, pp. 670–679.
L., 2018. Electrochemical impedance analysis of a PEDOT:PSS-based textile energy stor- Zajdel, T.J., Baruch, M., Mehes, G., Stavrinidou, E., Berggren, M., Maharbiz, M.M., et al.,
age device. 11, p. 48. 2018a. PEDOT:PSS-based multilayer bacterial-composite films for bioelectronics. Sci.
Olivares, A.J., Cosme, I., Sanchez-Vergara, M.E., Mansurova, S., Carrillo, J.C., Martinez, H.E., Rep. 8, 15293.
et al., 2019. Nanostructural modification of PEDOT:PSS for high charge carrier collection Zhang, X., Pei, X., Jia, Q., Wang, Q., 2009. Effects of carbon fiber surface treatment on the tri-
in hybrid frontal Interface of solar cells. Polymers 11, 1034. bological properties of 2D woven carbon fabric/polyimide composites. Appl. Phys. A 95,
Park, S.-G., Chae, K.-J., Lee, M., 2017. A sulfonated poly(arylene ether sulfone)/polyimide 793–799.
nanofiber composite proton exchange membrane for microbial electrolysis cell applica- Zhang, Y., Sun, J., Hu, Y., Li, S., Xu, Q., 2012. Bio-cathode materials evaluation in microbial
tion under the coexistence of diverse competitive cations and protons. J. Membr. Sci. fuel cells: a comparison of graphite felt, carbon paper and stainless steel mesh materials.
540, 165–173. Int. J. Hydrog. Energy 37, 16935–16942.
Park, S.G., Rhee, C., Shin, S.G., Shin, J., Mohamed, H.O., Choi, Y.J., et al., 2019.
Methanogenesis stimulation and inhibition for the production of different target

12

You might also like