You are on page 1of 24

Environmental Technology

ISSN: 0959-3330 (Print) 1479-487X (Online) Journal homepage: http://www.tandfonline.com/loi/tent20

Performance evaluation of treating oil-containing


restaurant wastewater in microbial fuel cell using
in-situ graphene/polyaniline modified titanium
oxide anode

ZhiLiang Li, ShengKe Yang, Ya’nan Song, HaiYang Xu, ZongZhou Wang,
WenKe Wang & YaQian Zhao

To cite this article: ZhiLiang Li, ShengKe Yang, Ya’nan Song, HaiYang Xu, ZongZhou Wang,
WenKe Wang & YaQian Zhao (2018): Performance evaluation of treating oil-containing restaurant
wastewater in microbial fuel cell using in-situ graphene/polyaniline modified titanium oxide anode,
Environmental Technology, DOI: 10.1080/09593330.2018.1499814

To link to this article: https://doi.org/10.1080/09593330.2018.1499814

Accepted author version posted online: 17


Jul 2018.

Submit your article to this journal

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tent20
Publisher: Taylor & Francis & Informa UK Limited, trading as Taylor & Francis Group

Journal: Environmental Technology

DOI: 10.1080/09593330.2018.1499814

Performance evaluation of treating oil-containing restaurant


wastewater in microbial fuel cell using in-situ
graphene/polyaniline modified titanium oxide anode
ZhiLiang Lia,b, ShengKe Yanga,*, Ya’nan Songa, HaiYang Xua, ZongZhou Wanga, WenKe Wanga

and YaQian Zhaoc

a.
Key Laboratory of Subsurface Hydrology and Ecology in Arid Areas, Ministry of Education,

Chang’an University, Xi’an 710054, P. R. China.

b.
School of Environment and Energy, South China University of Technology, Guangzhou, 510006,

China.

c.
UCD Dooge Centre for Water Resources Research, School of Civil Engineering, University

College Dublin, Belfield, Dublin 4, Ireland.

E-mail: ysk110@126.com; Tel: +86-29-82339982; Fax: +86-29-85585485.

Abstract Most studies conducted nowadays to boost electrode performance in microbial fuel
cell (MFC) have focused on carbonaceous materials. The titanium suboxides (Ti4O7, TS) is able to

provide a new alternative for achieving better performance in MFC and has been tested and

demonstrated in this study. The Ti4O7 electrode with high electrochemical activity was modified by

graphene/polyaniline via the constant potential method. Electrogenic microorganisms were more

conducive to adhere to the anode electrode due to the presence of graphene/polyaniline. The MFC

reactor with polyaniline /graphene modified TS (TSGP) anode achieves the highest voltage with

980 mV, and produces a peak power density of 2073 mW/m2, which is 2.9 and 12.7 times of those
with the carbon cloth anode at the 1000 Ω external resistance. In addition, this study evaluates the

effects of anolyte conductivity, pH, and COD on the treatment of oil-containing restaurant

wastewater (OCRW) in MFC using TSGP anode. The OCRW amended with 120 mS/cm obtains the

lowest internal resistance (160.3 Ω). The anodic pH gradually from acidic to alkaline conditions (pH

5.5-8.0) resulted in a gradual increase in maximum power density to 576.4 mW/m2 and decrease in

internal cell resistance to 203.7 Ω. The MFC at the COD 1500 mg/L could obtain steady state output

voltage during 103h while removing up to 65.2% of the COD of the OCRW.

Keywords: Microbial fuel cell; Graphene; Polyaniline; Electrode material; Oil-containing

restaurant wastewater

1. Introduction
With the word’s rapid economic growth, an increasing amount of wastewater from restaurants,

cafeterias, and hotels is widely discharged. The characteristics of restaurants wastewater [1, 2] are in

high oil and grease contents with easy spoilage. If the wastewater is not properly handled, it will not

only be an increasing burden on municipal wastewater treatment plants, but can also cause

eutrophication in waterways [3]. Hence, an efficient and cheap oily wastewater treatment

technology is urgently required.

To date, the most common treatment method for the oil-containing wastewater is electroflotation

[1] and electrocoagulation [2]. However, due to the requirements of complex equipment used in the

electro-flotation/coagulation, it is not practical to use this technique. Microbial fuel cell (MFC) is

bioelectrochemical system that can convert chemical energy stored in organic substrate directly into

electrical energy [4]. MFC is the technique for wastewater treatment because of the relatively low

cost and energy consumption compared to conventional methods [5]. It has not previously been

examined whether electricity can be generated using oily wastewater.

However, the low output power density from MFC due to low bacterial loading onto the electrode

and low extracellular electron transfer (EET) efficiency between bacterial and electrodes are the

major bottleneck limiting the practical application of MFC [6, 7]. The anode can directly affect

exoelectrogens attachment and electron transfer, and then further influence the MFC performance
[8]. Development of novel anodic materials that can facilitate bacterial biofilm formation and EET

efficiency is vital to enhance power production of MFC [9].

However, most of the studies conducted nowadays to boost electrode performance in MFC

focuse on carbonaceous materials. Recently, studies have demonstrated that ceramics can provide

stability, improve power output and treatment efficiencies, create a better environment for the

electro-active bacterial and contribute towards resource recovery [10]. In addition, featuring low

cost, biocompatibility and chemical stability, titanium oxides have gained wide interest. Titanium

substoichiometric oxides of general formula Ti nO2n-1 (with n between 4 and 10) [11], have already

been used to prepare electrically conducting macro-porous supports, making them attractive for

building ceramic electrodes [12]. Massazza et al. demonstrated the ice-templated titanium-based

ceramics (ITTC) and combined the chemistry of titanium suboxides (Ti 4O7) with ice-templating

technique (ISSA). This new electrode material has been demonstrated to outperform graphite when

used as an anode in bioelectrochemical reactors, obtaining volumetric current densities of 9500

A/m3 [13].

Polyaniline (PANI) has been widely employed as electrode modifier due to its low cost, safe

environment, easy of synthesis, good biocompatibility and unique acid-base doping/dedoping

property [14, 15]. Graphene (G) has attracted tremendous interests due to its unique properties

including high electronic conductivity, high modulus, and high specific surface area (up to ⁓2600

m2 g-1) [8]. However, PANI presents smaller specific surface and reluctant conductivity but

graphene possesses poor biocompatibily [16], which distinctly influence the bacterial loading

capacity and EET efficiency. The combined modification by PANI and graphene would make best

use of their advantages, which is suitable to serve as the MFC anode. Sun et al. fabricated a novel

MFC anode by electro-depositing PANI onto graphene coated graphite paper (GO/PANIOS), and

achieved power density of 381 mW/m2 [17]. Li et al. fabricated the graphene/PANI carbon cloth

(PANI+G+CC) as anode of MFC, generating the maximum power density of 884±89 mW/m2 [16].

Although attention has been paid to MFC with several unique advantages over the conventional

bioenergy technologies [18], previous study suggested that the low bioelectricity output because of

the high internal resistance (Rint) [19]. One of the factors that strongly influence the Rint of MFC is

ionic conductivity of electrolyte [20]. Generally, power densities of MFC could be facilitated by

high conductivities of up to 20 mS/cm [21]. Lefebvre et al [22]found although NaCl addition (up to
20 g/L) could increase power density, the highest coulombic efficiency (CE) was obtained at a much

lower NaCl (5g/L). On the other hand, anolyte pH is another critical factor that affects the R int of a

MFC [20]. However, the impact of anolyte pH on MFC performance is somewhat contradictory. For

instance, Yuan et al [23] reported a contrasting outcome with a two-chamber MFC that the electron

transfer resistance of their anodic biofilm was much lower under alkaline condition (pH 9.0)

compared to acidic and neutral conditions (pH 5 and 7). But the anolyte pH was actively maintained

at neutral assuming the electrical output would be most favorable under neutral condition [24]. In

addition, anolyte COD also has an influence the performance of MFC equipped with an oxygen

cathode [25]. Clearly, understanding the impact of anolyte conductivity, pH, and COD towards

using MFC for the treatment of oily wastewater is necessary.

The objective of this study was to evaluate the MFC in terms of electrochemical properties

(maximum power density, open circuit potential (OCP), and Rint) and the maximum current density

(Im). Special focuses are placed on: (i) MFC with the PANI/G composite modified Ti4O7 electrode,

which is firstly prepared via constant potential method; and (ⅱ) the effect of anolyte conductivity

(COND), pH, and COD on the bioelectricity generation performance using OCRW as the MFC

feedstock.

2. Materials and methods

2.1 Fabrication of titanium suboxides with G/PANI

The monolithic porous Ti4O7, produced from high-temperature reduction TiO2 by H2 was supplied

by Ti-Dynamics Co. Ltd. (China). Titanium suboxides (Ti4O7) (TS) and carbon cloth were

immersed in propanone and 0.1 mol/L ethanedioic acid solution for 6 h to remove impurities. They

were then rinsed with deionized water to neutral, and dried at 60 ℃ for 12 h. The pretreated Ti4O7

and carbon cloth were then cut into small pieces with geometry size of 2.5 cm × 3.0 cm and

connected to copper wire. The prepared TS electrode was either directly used as MFC anode or

further electrodeposited via constant potential method to fabricate the composite electrode.

The fabrication of TS electrode with graphene/polyaniline (synthesis method of graphene and

PANI meets supplementary information) requires two steps [26]. The potential was controlled to be

0.8 V. The electrodeposition time was 15min. The Pt wire was used as counter electrode and the

calomel electrode was employed as reference. In the first step, TS electrode was fixed in an
electrolytic bath, which contained aqueous reduced graphene oxide suspension of 5g/L. The

titanium suboxides electrode with graphene was marked as TSG. In the second step, TSG electrode

was fixed in an electrolytic bath, which contained aqueous PANI suspension. The TS electrode with

graphene, and TS electrode with graphene /PANI were named as TSG and TSGP, respectively.

2.2 MFC setup and construction

All experiments were carried out using a single-chamber air cathode MFC as illustrated in Fig.1.

The effective empty volume for anode chamber was 288 mL (6.0 cm × 6.0 cm × 8.0 cm). The

dimensions of the anode were 2.5 cm × 3.0 cm, and the dimensions of the cathode were 4.0 cm × 4.0

cm. The cathode electrode was made of carbon cloth, without any catalyst. Proton exchange

membranes (DongYue Federation, DMR100, China) were placed to separate the anode chamber

and the cathode electrode. Copper wire was used to connect the circuit, and the fuel cell was placed

under a constant load by connecting the anode and cathode to an external resistance (Rex) of 1000 Ω,

unless otherwise stated. The MFC was operated in fed-batch mode by conducting the batch mode

test until the voltage decreased to a low value of 100 ± 10 mV from the maximum. The anode

medium contained 0.13 g/L KCl, 0.31 g/L NH4Cl, 3.32 g/L NaH2PO4•2H2O, 10.36 g/L

NaH2PO4•2H2O, carbon source and 1ml/L trace elements [27].

The carbon source was made of raw OCRW, which was obtained from Chongqing Hot pot

shop (Xi’an SAGA, China). The OCRW was moved to the lab and then left for 24 h. After 24h, the

supernatant was collected and then stored in a deep-freezer (Forma Scientic, Marietta, OH, USA) at

-18℃ to avoid biodegration and changes in the wastewater properties. The ORCW was removed

from the freezer and used as a substrate as needed. The raw ORCW of COND, pH, COD, and total

dissolved solids is 13.4 mS/cm, 7.2, 14514 mg/L, and 12.1g/L, respectively. In order to prepare the

anolyte with different anolyte conductivity (10, 15, 40, 60, 80, 120 mS/cm) and COD (500, 800,

1000, 1200, 1500 mg/L), different dosage of Na2SO4 and raw OCRW were added to the individual

anodic chamber. The desired pH (5.0, 5.5, 6.0, 6.5, 7.0, 7.5, 8.0) of the anolyte was obtained by

using requisite amounts of 0.5 mol/L NaOH and 0.5 mol/L HCL solutions.

The MFC was inoculated with a mixed bacterial culture that was originally enriched from

domestic wastewater (Qu Jiang Wastewater Treatment Plant, Xi’an, China), and operated in batch
mode for six months, sequencing analysis of the microbial consortia using 16S rDNA

pyrosequencing (Supplementary information as shown). The MFC was conducted in batch mode

and the temperature was maintained at 30 ± 2 ℃.

2.3 Analysis

The morphology of prepared electrode and modified electrode were characterized with an

environmental scanning electron microscope (ESEM, Quanta 200, FEI). The high resolution

transmission electrode microscope (HRTEM, G2F30300kV, FEI, USA) was used to test the treated

electrode. FTIR was conducted in Nexus 470 FTIR (Thermo Fisher Nicolet, USA) from 400 to 4000

cm-1. Raman spectra were surveyed from 500 to 3500 cm-1 via DXR Raman Microscope

(Thermo-Fisher; USA) with 623 nm laser. The cell voltage (U) and current were automatically

recorded using a data acquisition unit (Agilent Technologies, Malaysia).

The specific surface area measurements were carried out via the gas adsorption technique (BET)

on a surface area analyzer (Beckman Coulter, SA3100). The crystal phase of Ti4O7 was assessed by

X-ray diffraction (XRD), and the XRD data were collected with a Bruker D8 Advance

diffractometer (Bruker, Cu Kα1 source, λ=1.5406 Å). The diffraction data were collected in the

range of 10°< 2θ <90° at a rate of 0.01°/s.

Electrochemical analysis was performed using an electrochemical workstation (CHI660C,

Shanghai Chenhua Instruments Co., Ltd, China) with a three-electrode model. Specific measuring

instruments and measuring methods are in the supplementary material.

COD was measured using fast digestion-spectrophotometric method with COD digester and

photometer (Lianhua 5B-3C, China). The anolyte pH was measured using a pH meter (LP115,

Mettler Toledo ). The anolyte conductivity was measured using a conductivity meter

(Shanghaisanxin SX713, China).

The polarization and power density curves were calculated through adjusting external

resistance from 99 kΩ to 10.2 Ω [28]. The internal resistance (Rint) was obtained in the region of

interest from the slope of the polarization curve [29].

In this study, all potentials quoted were referred to SHE, unless otherwise stated. All tests were

conducted in triplicate and the averages were presented.


3. Results and discussion

3.1 Structural and morphological characterization

To confirm the identity of the TS material, high-resolution XRD patterns were used. The results are

presented in Fig. 2, which shows that the XRD patterns, including the fingerprint peaks and their

relative intensities, are almost identical to standard values (PDF2-2004), indicating that the main

phase of the TS material is Ti4O7 with excellent conductivity. BET measurement was also

conducted on the TS, and gave a surface area of 1.4 m2/g.

To examine the morphology of the TS, TSG, and TSGP, high-resolution SEM images were

collected, as shown in Fig. 2 B, C, and D. The pristine TS surface is mesoporous with a smooth and

clean surface (Fig. 2B). The TSG electrode was mostly covered with graphene and a wrinkled

graphene layer was observed on the surface of TS (Fig. 2C). As expected, for the TSGP electrode, a

layer of PANI matrix was densely packed on the TSG (Fig. 2D). The morphology of the G/PANI

modified TS surface was significantly differently from the G modified TS electrode, which was

attributed to the incorporation of graphene sheets into the PANi nanofibers during the

electrodeposition process. TEM further verifies that a few layers graphene and PANI formed on the

surface layers of TSGP (Fig. S3, Supplementary information).

Raman spectra and FTIR analyses (Fig. S4 & S5,Supplementary information) were further

carried out to characterize the chemical bonding and composition of the prepared electrodes. Fig.S4

compares Raman spectra of the TSG electrode, pure PANI, and TSGP electrode. The spectrum of

the pure PANI displayed some typical bans of the C-H bending at 1164 cm-1, C-N+ stretching at

1333 cm-1, C=N stretching vibration at 1486 cm-1 and C=C stretching at 1591 cm-1 [16, 30]. To the

TSG electrode, there was a couple of remarkable peaks around 1333 cm-1 and 1596 cm-1, matching

the well recorded D and G band [31], respectively. Similarly, the D band at 1324 cm-1 and G band at

1621 cm-1 were detected in the Raman spectrum of the TSGP electrode. For the spectrum of the

TSGP, some new peaks were visibly observed different from TSG, such as C-H bending at 1159

cm-1, C=N stretching vibration at 1472 cm-1 [32]. As can be seen from Fig.S5, the FTIR spectrum of

the TSGP presented the PANI here is the conductive emeraldine salt form [33, 34] because of
oxidized and reduced repeat units (Fig.S5, presence of benzenoid at 1316 cm-1 and 1436 cm-1

(reduced stated), and quinoid ring at 1057 cm-1 and 1612 cm-1 (oxidized state)), which is beneficial

to improve EET efficiency.

The results demonstrate the successful preparation of G/PANI composite on the TS surface.

3.2 Electrochemical and MFC performance improved with the modified electrode

The electrochemical performance of electrodes of TS, TSGP, and C was illustrated in Fig.3. As

revealed in Fig. 3A, the cyclic voltammetry curve of the TSGP showed lager currents than that of

the TS and carbon cloth anode electrodes. A peak current (Ipa) of 7.5 mA at -0.31 V in the oxidation

scan was observed at the TSGP. Ipa for TS and C are 4.4 mA and 4.0 mA, respectively. Higher Ipa

indicates larger electrochemical active surface area [17]. The peak current of the TSGP was much

larger than that of TS and C, which indicated the active surface area was greatly increased when G

and PANI structure formed and could be contributed to the enhanced electron-transfer efficiency

resulted from the inherent properties of G/PANI.

Electrochemical impedance spectroscopy (EIS) analysis was then carried out to investigate the

charge-transfer resistance of the different anodes transfer efficiency resulted from the inherent

properties. The results are plotted as Nyquist curves and further fitted with an equivalent circuit (Fig.

3B and the inset). The EIS Nyquist plot is the superimposition of a preceding frequency-dependent

semicircle (high frequency region) and a subsequent straight line (low frequency region), the

diameter of the former represents the charge-transfer resistance [35]. By fitting the data of the

Nyquist plots using the ZSimpWin software, the value of each parameter was obtained. The solution

resistances (Rs) were all about 12 Ω for different electrodes. Rct at the electrode/electrolyte interface,

corresponding to the diameter of the semicircles, follows the order of C (~ 6.27 Ω) >TS (~ 5.02 Ω) >

TSGP (~ 1.92 Ω). Because all three electrodes share the same electrolyte, the decrease in the Rct can

only be attributed to the presence of graphene and PANI. This result highlights the effectiveness of

G/PANI in enhancing charge transfer ability of TS. It confirms that the electron transfer efficiency

of the TSGP is much higher than that of the C. In addition, the straight line region over low

frequency of TSGP electrode was obviously smaller than that of TS and C electrode. This indicates
that G and PANI could accelerate the diffusion of electrolyte toward the electrode surface.

Looking at Fig. 4, it was apparent that different electrodes (TS, TSGP, and C) were applied in

MFC and used for performance comparison (The anode medium contained: 0.13 g/L KCl, 0.31 g/L

NH4Cl, 3.32 g/L NaH2PO4•2H2O, 10.36 g/L NaH2PO4•2H2O, 0.5 g/L C6H12O6 and 1 ml/L trace

elements [27].). MFC with TS outperformed C. The MFC reactor with PANI/graphene modified TS

(TSGP) anode achieves the highest voltage with 980 m V, and produces a peak power density of

2073 mW/m2 at the 1000 Ω external resistance , which is 2.9 and 12.7 times of those with the

carbon cloth.

3.3 Microbial community analyses in the MFC

Using a complex substrate like OCRW that contains high strength organic matter could help

establish a multiplex and exoelectrogens in the MFC system. The rarefaction curve (Fig.S1,

Supplementary information as shown) indicated that the number of sequences from sample reached

plateau implying sufficient sampling of genetic diversity in the sample. Tolumonas were the most

dominant genus (18.0 %) in the inoculation of the anolyte, followed by Bacteroidales (11.3%),

Porphyromonadaceae (10.8%) and Rhodobacteraceae (4.6%) (Fig.S2, Supplementary information

as shown). This composition reflected the exoelectrogenic families in the anolyte where anaerobic

bacteria were observed [36, 37, 38, 39]. The TSGP electrode was generally of sandwich structure

with TS or PANI on both sides and G in the center. As reported previously, the π-π stacking

force between the phenyl of aniline and basal planes of graphene was beneficial to PANI

polymerization on the surface of graphene layers [40]. Several reports had shown that PANI

nanowire structure was suitable for bioelectrochemical systems application since they increased the

chance for the interaction between the material and nanostructured electrochemical active bacteria

[17, 41]. Electrogenic microorganisms were more conducive to adhere to the anode electrode due to

the presence of G/PANI, which greatly improved the power production and extracellular electron

transfer efficiency.

3.4 Effect of anolyte conductivity on the performance of MFC

MFC requires inorganic salts to maintain ionic conductivity in the electrolyte to facilitate charge

transfer. Low ionic conductivity limits the transfer of cations from anode to cathode and would
retard the balance of electro-neutrality of the system [42]. Fig. 5 and Table 1 present the changes in

ionic conductivity under different operating cycles. In general, the ionic conductivity of the three

cycles exhibited the expected order throughout the period (i.e. 60> 40> 15> 80> 10> 120 mS/cm).

This implies that the dissolved ions contribute the charge carrier and reduce the solution resistance

under low ionic strength values. Hence the MFC at the anolyte conductivity 60 mS/cm should

exhibit a stable and controlled utilization of organic matter for electricity generation and facilitate a

higher power density, which makes known that the dissolved ions and the activity of bacteria would

deliver simple and soluble substrate for the enhancement of the current generation in MFCs.

However, with the anolyte conductivity continues to increase (i.e. 80 & 120 mS/cm), the

electrogenic organisms activity is inhibited due to higher osmotic pressure [43], which leads to the

performance of MFC being suppressed.

3.5 Effect of pH on the performance of MFC

The experimental results clearly demonstrated the influence of the anodic feed pH on the

performance of the MFC. This relationship was confirmed by the analysis of the voltage-time curve,

polarization curve (Fig. 6), and by the calculations derived from these curves (see Table 2). The

peak biovoltage and peak power density increased with the increase in the pH of the feed solution up

to pH 8.0. The peak biovoltage and peak power density at the pH 6.5 were 398.9 mV and 338.2

mW/m2, which was increased substantially when pH was raised to 8.0 and were showing

corresponding values of 489.7 mV and 576.4 mW/m2, respectively.

By comparing the results with some previous studies dealing with the effect of pH on MFC, the

alike result was found. In this study enhanced performance was observed at anodic pH 8.0. Kaushik

and Chetal [44] found that enhanced performance was observed at anodic pH 9 whereas at pH 5

more severe inhibitions on the current generation were observed. Raghavulu et al. [45], however,

reported that the optimal pH for their MFC was between 6 and 8 while pH 6 was more favorable for

high MFC output. The variable results on effect of anodic pH for different MFCs indicate that it is

important to optimize the pH for efficient operation of different MFCs. It has been reported that

increased anolyte pH leads to more negative anode potential, thus resulting in enhanced

performance of the MFC [46]. Moreover, under alkaline conditions, the exoelectrogens at the anode
have a better chance of outcompeting methanogens for the degradation of organic matter in the

wastewater [44].

3.6 Effect of COD on the performance of MFC

Different anodic COD conditions were applied in MFC and used for performance comparison (Fig.

7). Fig. 6A shows the voltage output of MFC operated under batch mode (V-t curve). This

relationship was confirmed by the analysis of the voltage-time curve, polarization curve (Fig. 7),

and by the calculations derived from these curves (see Table 3). The peak biovoltage and peak

power density increased with the increase in the COD of the feed solution up to COD 1500 mg/L. As

demonstrated in Fig. 7A, the peak biovoltage at the COD 500 mg/L had the worst performance, only

398.9 mV steady state voltage was reached. Further improvement in maximum steady state output

voltage was achieved at the COD 1500 mg/L, which was 631.2 mV. The above result is consistent

with previous works [47, 48].

During polarization, the maximum power densities of different COD were obtained. The

maximum power densities was 176% higher at the COD 1500 mg/L, compared to the COD

500mg/L. With the increase of COD, the slower slope in the V-j curve (voltage vs. current density)

was decreasing, which indicated smaller R int. As with COD 1500 mg/L, the slowest slope was

achieved in MFC (Fig. 7B). This phenomenon could be related to the enhancement of the

electrogenic organisms population. With this enhancement organisms population was manifested

when the influent COD concentration was increased. In this case, the higher activity reduced the

resistance to the electrical current, which resulted in the best electricity production performance

obtained at the COD 1500 mg/L. The MFC (the COD of 1500 mg/L) could obtain steady state

output voltage during 103 h while removing up to 65.2% of the COD of the OCRW.

4. Conclusions
In this study TS electrode of ceramic materials, being a promising electrode in the MFC, has been

well demonstrated. G and PANI were stably modified on the surface of macro porous TS by

electrochemical techniques. The MFC with PANI/graphene modified TS (TSGP) anode achieved

the highest voltage with 980 mV, and produced a peak power density of 2073 mW/m2, which was
2.9 and 12.7 times of those with the carbon cloth control at the 1000 Ω external resistance. In

addition, this study evaluates the effects of ionic conductivity, pH, and COD on the treatment of

OCRW in MFC using TSGP anode. The OCRW amended with 120 mS/cm enabled the lowest

internal resistance (160.3 Ω). Increasing the anodic pH gradually from acidic to alkaline conditions

(pH 5.5-8.0) resulted in a gradual increase in maximum power density to 576.4 mW/m2 and

decrease in internal cell resistance to 203.7 Ω. The open circuit voltage is 760.5 mV , and the

maximum power density (Pm) at the COD 1500 mg/L is 934.4 mW/m2, respectively. The MFC at

the COD 1500 mg/L could obtain steady state output voltage during 103 h while removing up to

65.2% of the COD of the OCRW. The results suggest that maintaining an appropriate operation is

essential for efficient bioelectricity production using MFC for treatment of oily wastewater.
Acknowledgements
The work was financially supported by National Natural Science Foundation of China (NO.

41672224 and No. 41372259) and the National Key Research and Development Program of China

(No.2016YFC0400701).

References
1. Ji M, Jiang X, Wang F. A mechanistic approach and response surface optimization
of the removal of oil and grease from restaurant wastewater by electrocoagulation and
electroflotation. Desalination & Water Treatment. 2015:1-9.
2. Qin X, Yang B, Gao F, et al. Treatment of Restaurant Wastewater by Pilot-Scale
Electrocoagulation-Electroflotation: Optimization of Operating Conditions. Journal of
Environmental Engineering. 2013;139(7):1004-1016.
3. Xu X, Zhu X. Treatment of refectory oily wastewater by electro-coagulation
process. Chemosphere. 2004;56(10):889.
4. Logan BE, Hamelers B, Rozendal R, et al. Microbial fuel cells: methodology and
technology. Environmental Science & Technology. 2006;40(17):5181-5192.
5. Silva AJ, Varesche MB, Foresti E, et al. Sulphate removal from industrial
wastewater using a packed-bed anaerobic reactor. Process Biochemistry. 2002;37(9):927-935.
6. Lai B, Tang X, Li H, et al. Power production enhancement with a polyaniline
modified anode in microbial fuel cells. Biosens Bioelectron. 2011;28(1):373-377.
7. Qiao Y, Bao SJ, Li CM, et al. Nanostructured Polyaniline/Titanium Dioxide
Composite Anode for Microbial Fuel Cells. Acs Nano. 2008;2(1):113-9.
8. Batzill M. ChemInform Abstract: The Surface Science of Graphene: Metal
Interfaces, CVD Synthesis, Nanoribbons, Chemical Modifications, and Defects. Cheminform.
2013;44(27):83-115.
9. Yong YC, Dong XC, Chanpark MB, et al. Macroporous and Monolithic Anode
Based on Polyaniline Hybridized Three-Dimensional Graphene for High-Performance
Microbial Fuel Cells. Acs Nano. 2012;6(3):2394.
10. Winfield J, Gajda I, Greenman J, et al. A review into the use of ceramics in
microbial fuel cells. Bioresour Technol. 2016;215:296-303.
11. Smith JR, Walsh FC, Clarke RL. Electrodes based on Magnéli phase titanium
oxides: the properties and applications of Ebonex® materials. Journal of Applied
Electrochemistry. 1998;28(10):1021-1033.
12. Kitada A, Hasegawa G, Kobayashi Y, et al. Selective Preparation of Macroporous
Monoliths of Conductive Titanium Oxides TinO2n–1 (n = 2, 3, 4, 6). Journal of the American
Chemical Society. 2012;134(26):10894.
13. Massazza D, Parra R, Busalmen JP, et al. New ceramic electrodes allow reaching
the target current density in bioelectrochemical systems. Energy & Environmental Science.
2015;8(9):2707-2712.
14. Ginder JM, Richter AF, MacDiarmid AG, et al. Insulator-to-metal transition in
polyaniline. Solid State Communications. 1987 July;63(2):97-101. doi:
10.1016/0038-1098(87)91173-2. PubMed PMID: INSPEC:2965428.
15. Wang H, Hao Q, Yang X, et al. A nanostructured graphene/polyaniline hybrid
material for supercapacitors. Nanoscale. 2010;2(10):2164-2170.
16. Huang L, Li X, Ren Y, et al. In-situ modified carbon cloth with
polyaniline/graphene as anode to enhance performance of microbial fuel cell. International
Journal of Hydrogen Energy. 2016 2016/07/13/;41(26):11369-11379. doi:
https://doi.org/10.1016/j.ijhydene.2016.05.048.
17. Sun DZ, Yu YY, Xie RR, et al. In-situ growth of graphene/polyaniline for
synergistic improvement of extracellular electron transfer in bioelectrochemical systems.
Biosens Bioelectron. 2017;87:195-202.
18. Li WW, Yu HQ. From wastewater to bioenergy and biochemicals via two-stage
bioconversion processes: a future paradigm. Biotechnology Advances. 2011;29(6):972-982.
19. Li XM, Cheng KY, Selvam A, et al. Bioelectricity production from acidic food
waste leachate using microbial fuel cells: Effect of microbial inocula. Process Biochemistry.
2013;48(2):283-288.
20. Li XM, Cheng KY, Wong JW. Bioelectricity production from food waste leachate
using microbial fuel cells: effect of NaCl and pH. Bioresour Technol. 2013;149(12):452.
21. Logan BE, Rabaey K. Conversion of wastes into bioelectricity and chemicals by
using microbial electrochemical technologies. Science. 2012;337(6095):686-90.
22. Lefebvre O, Tan Z, Kharkwal S, et al. Effect of increasing anodic NaCl
concentration on microbial fuel cell performance. Bioresour Technol. 2012;112(3):336.
23. Yuan Y, Zhao B, Zhou S, et al. Electrocatalytic activity of anodic biofilm responses
to pH changes in microbial fuel cells. Bioresour Technol. 2011;102(13):6887-91.
24. Choi JD, Chang HN, Han JI. Performance of microbial fuel cell with volatile fatty
acids from food wastes. Biotechnol Lett. 2011;33(4):705-714.
25. Li JT, Zhang SH, Hua YM. Performance of denitrifying microbial fuel cell
subjected to variation in pH, COD concentration and external resistance. Water Science &
Technology. 2013;68(1):250-256.
26. Zhang K, Li J, Li Q, et al. Improvement on electrochemical performance by
electrodeposition of polyaniline nanowires at the top end of sulfur electrode. Applied Surface
Science. 2013;285(11):900-906.
27. Catal T, Cysneiros D, O'Flaherty V, et al. Electricity generation in single-chamber
microbial fuel cells using a carbon source sampled from anaerobic reactors utilizing grass silage.
Bioresour Technol. 2011;102(1):404-410.
28. Zhang L, Zhu X, Li J, et al. Biofilm formation and electricity generation of a
microbial fuel cell started up under different external resistances. Journal of Power Sources.
2011;196(15):6029-6035.
29. Kim H, Kim B, Kim J, et al. Electricity generation and microbial community in
microbial fuel cell using low-pH distillery wastewater at different external resistances. Journal
of Biotechnology. 2014;186:175-180.
30. Hou J, Liu Z, Zhang P. A new method for fabrication of graphene/polyaniline
nanocomplex modified microbial fuel cell anodes. Journal of Power Sources. 2013 Feb
15;224:139-144. doi: 10.1016/j.jpowsour.2012.09.091. PubMed PMID:
WOS:000313390400020.
31. Najafabadi AT, Ng N, Gyenge E. Electrochemically exfoliated graphene anodes
with enhanced biocurrent production in single-chamber air-breathing microbial fuel cells.
Biosens Bioelectron. 2016 Jul 15;81:103-110. doi: 10.1016/j.bios.2016.02.054. PubMed PMID:
WOS:000374811800016.
32. Yan J, Wei T, Shao B, et al. Preparation of a graphene nanosheet/polyaniline
composite with high specific capacitance. Carbon. 2010 Feb;48(2):487-493. doi:
10.1016/j.carbon.2009.09.066. PubMed PMID: WOS:000272764300023.
33. Sun D-Z, Yu Y-Y, Xie R-R, et al. In-situ growth of graphene/polyaniline for
synergistic improvement of extracellular electron transfer in bioelectrochemical systems.
Biosensors and Bioelectronics. 2017 2017/01/15/;87:195-202. doi:
https://doi.org/10.1016/j.bios.2016.08.037.
34. Qiao Y, Li CM, Bao S-J, et al. Carbon nanotube/polyaniline composite as anode
material for microbial fuel cells. Journal of Power Sources. 2007 Jun 30;170(1):79-84. doi:
10.1016/j.jpowsour.2007.03.048. PubMed PMID: WOS:000247793500012.
35. Hou J, Liu Z, Zhang P. A new method for fabrication of graphene/polyaniline
nanocomplex modified microbial fuel cell anodes. Journal of Power Sources.
2013;224(4):139-144.
36. Quan XC, Quan YP, Tao K. Effect of anode aeration on the performance and
microbial community of an air-cathode microbial fuel cell. Chemical Engineering Journal. 2012
Nov;210:150-156. doi: 10.1016/j.cej.2012.09.009. PubMed PMID: WOS:000312617100018.
37. Luo JM, Yang J, He HH, et al. A new electrochemically active bacterium
phylogenetically related to Tolumonas osonensis and power performance in MFCs. Bioresour
Technol. 2013 Jul;139:141-148. doi: 10.1016/j.biortech.2013.04.031. PubMed PMID:
WOS:000321163100021.
38. Kiely PD, Rader G, Regan JM, et al. Long-term cathode performance and the
microbial communities that develop in microbial fuel cells fed different fermentation
endproducts. Bioresour Technol. 2011 Jan;102(1):361-366. doi:
10.1016/j.biortech.2010.05.017. PubMed PMID: WOS:000285658300049.
39. Daghio M, Gandolfi I, Bestetti G, et al. Anodic and cathodic microbial communities
in single chamber microbial fuel cells. New Biotechnology. 2015 2015/01/25/;32(1):79-84. doi:
https://doi.org/10.1016/j.nbt.2014.09.005.
40. Xu J, Wang K, Zu SZ, et al. Hierarchical Nanocomposites of Polyaniline Nanowire
Arrays on Graphene Oxide Sheets with Synergistic Effect for Energy Storage. Acs Nano.
2010;4(9):5019.
41. Ding C, Liu H, Lv M, et al. Hybrid bio-organic interfaces with matchable nanoscale
topography for durable high extracellular electron transfer activity. Nanoscale.
2014;6(14):7866-71.
42. Karthikeyan R, Selvam A, Cheng KY, et al. Influence of ionic conductivity in
bioelectricity production from saline domestic sewage sludge in microbial fuel cells. Bioresour
Technol. 2016;200:845-852.
43. Pilizota T, Shaevitz JW. Plasmolysis and Cell Shape Depend on Solute
Outer-Membrane Permeability during Hyperosmotic Shock in E. coli. Biophysical Journal.
2013;104(12):2733.
44. Kaushik A, Chetal A. Power generation in microbial fuel cell fed with post
methanation distillery effluent as a function of pH microenvironment. Bioresour Technol.
2013;147(7):77-83.
45. Raghavulu SV, Mohan SV, Goud RK, et al. Effect of anodic pH microenvironment
on microbial fuel cell (MFC) performance in concurrence with aerated and ferricyanide
catholytes. Electrochemistry Communications. 2009;11(2):371-375.
46. Puig S, Serra M, Coma M, et al. Effect of pH on nutrient dynamics and electricity
production using microbial fuel cells. Bioresour Technol. 2010;101(24):9594-9599.
47. Tamakloe RY. Effect of COD and H2O2 concentration on DC-MFC. Renewable
Energy. 2015;83:1299-1304.
48. Kim H, Kim B, Yu J. Power generation response to readily biodegradable COD in
single-chamber microbial fuel cells. Bioresour Technol. 2015;186:136-140.
Tables

Table 1 Effect of anolyte conductivity over the MFC performance. Number of replicates =3,

standard deviation (SD) < 5%.

anolyte conductivity COD removal OCP (mV) Um (mV) Rint (Ω)

(mS/cm) (%)

10 27.5±0.8 498.3±1.8 382.4±1.5 309±0.7

15 34.0±1.4 526.2±2.4 398.9±2.7 264.9±1.7

40 43.0±1.2 538.6±1.4 446.8±1.1 203±1.9

60 47.9±0.4 665.8±3.7 568.5±4.1 169.4±2.2

80 41.9±1.1 413.6±2.9 349.7±3.4 164.5±1.3

120 36.8±0.7 180.6±0.7 148.6±2.3 160.3±0.4

Table 2 Effect of pH over the MFC performance. Number of replicates =3, standard deviation (SD)

< 5%.

pH COD removal OCP (mV) Um (mV) Rint (Ω)

(%)

5.5 34.6±1.7 450.9±1.3 301.1±2.7 450.6±1.7

6.0 38.0±1.3 483.3±2.8 348.6±4.1 382.6±2.8

6.5 43.0±1.1 526.2±0.8 398.9±5.1 264.9±3.3

7.0 47.5±1.6 530.1±3.1 420.7±1.9 242.7±4.2

7.5 50.7±0.7 560.4±2.4 451.2±3.4 213.7±3.7

8.0 53.1±0.6 610.2±4.6 489.7±1.8 203.7±0.9


Table 3 Effect of COD over the MFC performance. Number of replicates =3, standard deviation

(SD) < 5%.

COD (mg/L) COD removal OCP (mV) Um (mV) Rint(Ω)

(%)

500 43.0±0.7 526.2±2.3 398.9±1.3 264.9±1.8

800 59.2±1.1 603.4±1.2 479.9±0.9 230.2±2.3

1000 63.7±0.3 721.3±2.9 582.9±1.4 212.5±1.7

1200 64.3±1.4 740.2±3.4 610.6±2.7 203.8±1.3

1500 65.2±0.9 760.5±5.7 631.2±4.1 195.7±0.5


Figures

Fig. 1 Schematic of the air cathode MFC. PEM- Proton exchange membrane; CA – cathode; AN –

anode; Pt – platinum wire; RE – reference electrode; R – external resistance; PC – computer; DA –

data acquisition; and CW - electrochemical workstation.

A B

C D

Fig. 2 XRD image of the TS (A); SEM images of the TS (B), TSG (C), and TSGP (D).
A B

Fig. 3 Cyclic voltammetry curve (A) of TS, TSGP, and C; Electrochemical impedance

spectroscopy (B) curve of TS, TSGP, and C.

A
B

Fig. 4 Voltage-time curve (A) of MFC equipped with C, TS, and TSGP; Polarization curve (B) of

MFC equipped with C, TS, and TSGP.


A
B

Fig.5 Voltage-time curve (A) of MFC at different anodic anolyte conductivity; Polarization curve

(B) of MFC at different anolyte conductivity. Number of replicates =3, standard deviation (SD) <

5%.

A B

Fig. 6 Voltage-time curve (A) of MFC at different anodic pH conditions; Polarization curve (B) of

MFC at different anodic pH conditions. Number of replicates =3, standard deviation (SD) < 5%.
A B

Fig. 7 Voltage-time curve (A) of MFC at different anodic COD conditions; Polarization curve (B) of

MFC at different anodic COD conditions. Number of replicates =3, standard deviation (SD) < 5%.
Figures captions
Fig. 1 Schematic of the air cathode MFC. PEM- Proton exchange membrane; CA – cathode; AN –

anode; Pt – platinum wire; RE – reference electrode; R – external resistance; PC – computer; DA –

data acquisition; and CW - electrochemical workstation.

Fig. 2 XRD image of the TS (A); SEM images of the TS (B), TSG (C), and TSGP (D).

Fig. 3 Cyclic voltammetry curve (A) of TS, TSGP, and C; Electrochemical impedance spectroscopy

(B) curve of TS, TSGP, and C.

Fig. 4 Voltage-time curve (A) of MFC equipped with C, TS, and TSGP; Polarization curve (B) of

MFC equipped with C, TS, and TSGP.

Fig.5 Voltage-time curve (A) of MFC at different anodic anolyte conductivity; Polarization curve

(B) of MFC at different anolyte conductivity. Number of replicates =3, standard deviation (SD) <

5%.

Fig. 6 Voltage-time curve (A) of MFC at different anodic pH conditions; Polarization curve (B) of

MFC at different anodic pH conditions. Number of replicates =3, standard deviation (SD) < 5%.

Fig. 7 Voltage-time curve (A) of MFC at different anodic COD conditions; Polarization curve (B) of
MFC at different anodic COD conditions. Number of replicates =3, standard deviation (SD) < 5%.

You might also like