You are on page 1of 10

Journal of Hazardous Materials 403 (2021) 123682

Contents lists available at ScienceDirect

Journal of Hazardous Materials


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

The reduction of nitrobenzene by extracellular electron transfer facilitated


by Fe-bearing biochar derived from sewage sludge
Yue Lu a, b, *, Qingqing Xie a, b, Lin Tang a, b, *, Jiangfang Yu a, b, Jingjing Wang a, b,
Zhaohui Yang a, b, Changzheng Fan a, b, Shoujuan Zhang a, b
a
College of Environmental Science and Engineering, Hunan University, Changsha 410082, China
b
Key Laboratory of Environmental Biology and Pollution Control, Hunan University, Ministry of Education, Changsha 410082, China

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

Keywords: In this work, the incorporation of Fe-bearing sludge-derived biochar greatly enhanced both biotic and abiotic
Geobacter reduction of nitrobenzene (NB) to aniline, which was attributed to the concomitant microbial dissimilatory iron
Sewage sludge reduction. Biogenic Fe(II) produced by Geobacter sulfurreducens dominated the anaerobic reduction of NB
Biochar
following the pseudo-first-order kinetic. Besides, the increase of pyrolysis temperature from 600 to 900 ℃ to
Nitrobenzene reduction
generate biochar resulted in an accelerated removal rate of NB in Geobacter-biochar combined system. The
Extracellular electron transfer
morphology and structural characterization of biochar with G. sulfurreducens confirmed the formation of
conductive bacteria-biochar aggregates. Electrochemical measurements suggested the presence of graphitized
domains and quinone-like moieties in biochar as redox-active centers, which might play an important role in
accelerating electron transfer for microbial dissimilatory iron reduction and NB degradation. This study provides
a feasible way of using Fe-bearing sludge as a valuable feedstock for biochar generation and its application with
electrochemically active bacteria for the bioremediation of nitroaromatic compounds-polluted wastewater.

1. Introduction and reduced like catalysts are applied as electron transfer stations be­
tween the initial electron donor and the terminal electron acceptor in
Nitroaromatic compounds (NACs) originated from the industrial multiple redox reactions (Van der Zee and Cervantes, 2009). Among
manufacture of explosives, pharmaceuticals, dyes and pesticides are these, biochar with graphitic structure and/or quinoid-like redox active
ubiquitous in wastewater and groundwater environments as a result of groups was found to be capable of facilitating electron transfer and
massive use and improper disposal (Luan et al., 2010). The mutagenic speeding up the biotic reductive transformation of contaminants (Kappler
and carcinogenic properties of NACs make it urgent to study their et al., 2014; Yu et al., 2015). For instance, Tong et al. (2014) reported that
environmental fate in order to deal with the severe contamination (Wu biochar enhanced the reductive dechlorination of pentachlorophenol by
et al., 2013). The biotic reduction of NACs with biochemical techniques promoting extracellular electron transfer (EET) of Fe(III)-reducing and
exhibits obvious superiority over traditional physical and chemical dechlorinating bacteria in the paddy soil. Beside the biochar, Fe(II)/Fe(III)
methods in terms of avoiding secondary pollution and reducing treat­ redox couple with the extensive reduction potential was found to be
ment costs (Wang et al., 2012). Nitro substituent groups on the aromatic responsible for mediating redox transformation of multiple pollutants,
ring of NACs featuring with strong electron affinity are found to be such as carbon tetrachloride, heavy metal, azo dye (Melton et al., 2014).
resistant to biological oxidation, however, they are vulnerable to elec­ Fe(II)/Fe(III) cycling can be biotically driven by EAB that are capable of
trophilic attack and can be effectively reduced to degradable aniline dissimilatory iron reduction with Fe(III) oxides as terminal electron ac­
under anoxic conditions (Wang et al., 2011). ceptors for their growth (Shi et al., 2016). Highly active species formed by
For the microbial anaerobic conversion of NACs, the efficiency of ferrous iron associated with iron minerals have been demonstrated to
electron transfer between electrochemically active bacteria (EAB, typi­ induce reductive transformation of nitrobenzene (NB) in the presence of
cally Geobacter and Shewanella spp.) and extracellular substances acts as a Shewanella putrefaciens (Luan et al., 2015).
limiting factor (Pereira et al., 2014). To promote the electron transfer, As described above, sole biochar or Fe(II)/Fe(III) redox couple as
redox mediators as a group of chemicals that can be reversibly oxidized intermediate redox reaction might have facilitated the EET process of

* Corresponding authors.
E-mail addresses: yuelu@hnu.edu.cn (Y. Lu), tanglin@hnu.edu.cn (L. Tang).

https://doi.org/10.1016/j.jhazmat.2020.123682
Received 26 May 2020; Received in revised form 4 August 2020; Accepted 6 August 2020
Available online 13 August 2020
0304-3894/© 2020 Published by Elsevier B.V.
Y. Lu et al. Journal of Hazardous Materials 403 (2021) 123682

EAB and affected the reductive transformation of contaminants. 2.4. Experiment set up of G. sulfurreducens cultures amended with
Therefore, in order to further study the impact of Fe-bearing biochar on biochar and NB
the rate and extent of NACs reduction by EAB, iron-rich municipal
sewage sludge, an unavoidable by-product in wastewater treatment A series of batch experiments were conducted in sterile 125 mL
process, is selected as a precursor for biochar production. Each year, serum bottles sealed with rubber stopper and aluminum caps. The
massive sewage sludge has been produced, as a consequence, sludge headspace was flashed with N2/CO2 (4:1) gas mixture to ensure an
disposal and management become an intractable environmental issue anaerobic atmosphere. Acetate (20 mM) and NB (initial concentration of
mainly due to its complicated composition (Ho et al., 2017). In this 100 μM) were added to 100 mL of modified DSMZ medium as electron
study, pyrolysis of iron-rich municipal sewage sludge into Fe-bearing donor and electron acceptor, respectively. After autoclave process, two
biochar also provides an environmentally friendly alternative to types of sludge-derived biochars (SDB600 and SDB900) were respec­
reduce the volume of the solid residue and to eliminate inherent path­ tively added to the culture medium to realize a final concentration of 1
ogens (Zhang et al., 2018; Yu et al., 2019). g/L biochar. An 5% (v/v) inoculum of G. sulfurreducens strain PCA was
Therefore, this study is a worthwhile attempt to combine Fe-bearing spiked to reach initial OD600 of 0.035. Biotic controls without adding
biochar with G. sulfurreducens PCA (as a model EAB) for NACs reduction. biochar (only PCA cells + NB) and abiotic controls without adding PCA
Iron-rich municipal sewage sludge was calcined into biochar at different cells (only SDB900/SDB600 + NB) were prepared in parallel. Another
pyrolysis temperatures, and the impact of Fe-bearing biochar on the control treatment to investigate microbial iron reduction was prepared
transformation of NB (as a model NAC) coupled with electron transfer with PCA cells and SDB900/SDB600 but without NB addition. All serum
process of G. sulfurreducens was investigated. Herein, we showed the Fe- bottles were set up in triplicate and cultivated at 30 ℃ with 150 rpm
bearing biochar stimulated EET of G. sulfurreducens PCA and the shaking. During the cultivation, 1 mL of sample was sacrificed for the
dissimilatory iron reduction enhanced the reduction of NB to aniline. analysis of HCl-extractable Fe(II) and total Fe concentration following
the standard protocol (Kappler et al., 2014).
2. Materials and methods
2.5. Analytical methods of NB
2.1. Materials
Another 1 mL of sample was periodically withdrawn from each bottle
NB and aniline were purchased from Shanghai Chemical Corp. All and immediately filtered through a 0.22 μm filter to remove biochar and
chemicals utilized in this study were of analytical grade and used bacteria cells for NB measurement. The analyses of NB and other
without further purification. reducing products were performed by high-performance liquid chro­
matography (HPLC, 1260 Infinity, Agilent Co.) equipped with a C18
column (4 μm, 150 × 4.6 mm, Agilent Co.) at detection wavelength of
2.2. Synthesis of sludge-derived biochar 265 nm (NB) and 231 nm (aniline), respectively. The acetonitrile/water
mixture (1:1) was employed as a mobile phase with flow rate of 1.0 mL/
Municipal sewage sludge cake formed by filter-press dewatering in min at 30 ℃. The filtered liquid sample was extracted with 0.5 mL
Xingsha Wastewater Treatment Plant (Changsha, China, 28 ◦ N, 113 ◦ E) hexane by overnight shaking. The upper organic phase (20 μL) was
was collected with airtight plastic bags and directly transported to the harvested after centrifugation (5000 rpm, 10 min) and injected into
Key Laboratory of Environmental Biology and Pollution Control, Hunan HPLC.
University, China. The sewage sludge was stored at room temperature
without any pretreatment. The method used to prepare biochar was 2.6. Characterization of G. sulfurreducens and biochar
slightly modified from the previous report (Tang et al., 2018). The
sewage sludge was dried under ambient temperature, crushed and Electron paramagnetic resonance (EPR, JES FA200) spectrometer
screened through a 100-mesh sieve (0.15 mm). Approximately 27 g of was utilized for the detection of persistent free radicals in as-prepared
sieved sludge was transferred to a quartz boat and put into a tubular sludge-derived biochar. Scanning electron microscope (SEM, Hitachi,
reactor (SK-G08123 K, China) with a constant heating rate of 5 ℃/min S4800) equipped with energy dispersive spectrometer (EDS) was
under continuous flow of nitrogen (5 mL/min). Samples were pyrolyzed applied to observe the morphology and correlation between
for 2 h after reaching 600 ℃ (SDB600) or 900 ℃ (SDB900), respectively. G. sulfurreducens and biochar. For SEM analysis, the samples with cells
After natural cooling to ambient temperature, the produced biochar was were collected by centrifugation (5000 rpm, 5 min) and fixed with 2.5%
washed with deionized water and ethanol till reaching neutral pH, and glutaraldehyde overnight at 4 ℃. After being washed twice with 0.1 M
then dried at 80 ℃ for 12 h in a vacuum oven. The as-prepared biochar phosphate buffer solution, the samples were serially dehydrated with
was stored in a desiccator for use without further modification. 30%, 50%, 70%, 85%, 95% and 100% ethanol, and dried in a vacuum
freeze-dryer. The BET specific surface area and pore volumes were
calculated by N2 sorption isotherms using a Quantachrome NovaWin
2.3. Cultivation of G. sulfurreducens (NOVA 2000e) at 77 K. The crystallographic structure of biochar was
confirmed by X-ray diffraction (XRD, D8 Advance) with the 2θ scanning
G. sulfurreducens strain PCA was provided by Prof. Hanqing Yu from range from 10 to 80◦ . The element composition was determined by X-ray
the University of Science and Technology of China. The strain was photoelectron spectroscopy (XPS, ESCALAB 250Xi, Thermo Fisher)
routinely cultured in modified DSMZ medium with 10 mL/L of Wolfe’s under an Al-Kα X-ray radiation of 1486.6 eV. XPS spectra were decon­
vitamin solution, and with 20 mM acetate and 50 mM fumarate as volved using Avantage 5.52 software. Fourier transform infrared spec­
electron donor and electron acceptor, respectively (Li et al., 2014). trometer (FTIR, Nicolet 5700) characterized the surface functional
Resazurin was excluded from the medium. For further experiment, groups ranging from 400 to 4000 cm− 1 wave number. The Raman
G. sulfurreducens in the exponential growth phase were harvested by spectrum was obtained at the excitation wavelength of 532 nm using a
centrifugation (5000 rpm, 5 min) in an artificial atmospheric chamber LabRAM HR800 spectrometer (Horiba Jobin Yvon, France).
(520 L, NO. 690321, Spilfyter) filled with nitrogen. Subsequently, the
cell pellet was resuspended and then washed twice by repeated centri­ 2.7. Electrochemical measurements
fugation and resuspension in the DSMZ basal medium without fumarate.
After being washed, cells were then resuspended in fumarate-free me­ Electrochemical measurements were applied to investigate the
dium for further experiment. electrochemical activity of G. sulfurreducens and sludge-derived biochar.

2
Y. Lu et al. Journal of Hazardous Materials 403 (2021) 123682

Solid samples were gathered via centrifugation at 8500 rpm for 5 min 3. Results
and resuspended in phosphate buffer solution (pH 6.8, 100 mM), and the
supernatant was collected separately. Prior to the electrochemical tests, 3.1. Characterizations of sludge-derived biochar
the electrolyte (supernatant or solid samples in phosphate buffer solu­
tion) was deoxygenated by sparging pure N2 gas for 20 min and glassy Two different types of biochar were produced from municipal sewage
carbon electrode (GCE) was pretreated as reported previously (Peng sludge with pyrolysis temperature of 600 ℃ (SDB600) and 900 ℃
et al., 2010). Cyclic voltammetry (CV) and Differential pulse voltam­ (SDB900), respectively. FTIR spectra of sludge-derived biochar and raw
metry (DPV) were tested in an electrochemical workstation (CHI 760E, sludge were measured to investigate the surface functional groups, as
China) equipped with the three-electrode system, namely saturated presented in Fig. 1a. The common bands ranging from 1500 to 1720 cm− 1
calomel electrode (SCE) as the reference electrode, the GCE as the reflected the vibration of aromatic C=C and carbonyl C=O in quinones,
working electrode and Pt electrode as the counter electrode. The CV ketones and carboxylic acids, respectively (Amezquita-Garcia et al.,
measurement was executed from -0.7 V to 0.8 V at the scan rate of 0.01 2013). Also, the appearance of a small peak at 1389 cm− 1 in SDB900
V/s. And the DPV measurement was performed from -0.6 V to 0.4 V with reflected the C–
– C stretching in aromatic ring carbons (Peng et al., 2016).
the following parameters: pulse height of 50 mV, pulse width of 0.2 s, The strong peak located at 1056 cm− 1 was assigned to C–O stretching
pulse period of 0.5 s. All measurements were conducted at ambient vibration and showed a slight change, which was attributed to the direct
temperature and the electrochemical data of each sample were collected combination of various forms of oxygen with adjacent carbon elements
in triplicate. during pyrolysis process (Ho et al., 2017; Chen et al., 2014; Pourhosseini
et al., 2018). A broad peak located at 3428 cm− 1 was assigned to the
2.8. Statistical analysis stretching vibration of − OH groups (hydroxyl or carboxyl), and its
decrease indicated that − OH groups were gradually decomposed at
OriginPro 8 was utilized to conduct the statistical analyses. One-way elevated temperature (Zhu et al., 2017). The two bands assigned to C–H
analysis of variance (ANOVA) was performed using SPSS 20.0 (SPSS on aromatic ring (785 cm− 1) and − CH2– (682 cm− 1) group weakened
Inc., Chicago, USA). with increasing temperature, suggesting that the dehydrogenation reac­
tion occurred to enhance aromaticity (Tang et al., 2018; Chen et al.,
2014). Meanwhile, the two peaks with low intensity at 2855 and 2923
cm− 1 were attributed to symmetric and asymmetric stretching of

Fig. 1. The FTIR (a) and XRD (c) spectra of the sludge-derived biochar at pyrolysis temperature of 600 ℃ (SDB600) and 900 ℃ (SDB900). SEM pattern of
G. sulfurreducens attached to biochar (b) during NB bioreduction. SEM elemental mapping of SDB900 (red stands for Fe, blue stands for O, green stands for N) (d-g).

3
Y. Lu et al. Journal of Hazardous Materials 403 (2021) 123682

aliphatic − CH3, respectively, which disappeared once pyrolysis con­ predominant compositions in sludge-derived biochar.
ducting at higher temperature to generate biochar (Gao et al., 2019). XPS analysis was executed to further characterize the elemental
Therefore, presented results indicated that higher temperature led to the compositions and surficial structure of SDB600 and SDB900. The XPS
decrease of surface functional groups due to the transformation of − OH full survey presented that C, N, O, Fe, Si and Al were the main elements
and − CH3 groups into aromatic structures by gas emission and thermal of sludge-derived biochar (Fig. S1 and Table S2), consisting with the
decomposition (Tang et al., 2019). The bands around 464 cm− 1 and 557 results of XRD. Deconvolution of the C 1s spectra of SDB600 and SDB900
cm− 1 corresponded to Fe2O3/Si− O− Si and Fe− OH, respectively, indi­ was fitted into four peaks, namely, aromatic C–C at 284.28-284.47 eV,
cating the formation of Fe oxides or Fe-O complexes as a consequence of C–O at 285.41-286.08 eV and C=O at 288.25-289.31 eV (Fig. 2a and b).
the interaction of iron and oxygen on the surface of biochar during py­ Moreover, C–C bond accounted for 62.2% (SDB600) and 71.0%
rolysis process. In addition, the low-intensity peak at 2366 cm− 1 was (SDB900) of the total carbon content, which verified the formation of
assigned to the presence of the goethite (Fe2O3⋅H2O) (Yang et al., 2018). highly carbonized and aromatic structures in sludge-derived biochar
The XRD pattern showed that the form of Fe oxides in sludge-derived (Zhao et al., 2018). These results were in good agreement with that of
biochar was mainly γ-Fe2O3 (JCPDS No. 39-0238) with a minor FTIR spectra. As shown in Fig. 2d, the Fe 2p3/2 spectra of SDB900 was
amount of Fe3O4 (JCPDS No. 65-3107) (Fig. 1c). Apart from Fe oxides, deconvoluted into two peaks at 711.17 and 714.73 eV, corresponding to
carbon (JCPDS No. 26-1080), quartz (JCPDS No. 65-0466) and γ-Fe2O3 and Fe(III), respectively (Han et al., 2007). While for SDB600,
aluminum oxide (JCPDS No. 10-0425) were also identified as the the Fe 2p spectra shifted slightly to 711.30 and 714.98 eV (Fig. 2c).

Fig. 2. C1s XPS spectra of the sludge-derived biochar at pyrolysis temperature of 600 ℃ (SDB600) (a) and 900 ℃ (SDB900) (b), Fe2p XPS spectra of SDB600 (c) and
SDB900 (d). Curve-fitting of Raman spectra of SDB600 (e) and SDB900 (f).

4
Y. Lu et al. Journal of Hazardous Materials 403 (2021) 123682

FeOOH was also demonstrated to exist in both SDB900 (724.41 eV, Fe


2p1/2) and SDB600 (724.81 eV, Fe 2p1/2).
The SEM graph of SDB900 presented irregular surface with many
impurity particles and the pore distribution was not obvious, as well as
those for SDB600 (Fig. S2). Sludge-derived biochar provided a carrier for
G. sulfurreducens PCA so that the cells could closely contact the material
surface (Fig. 1b). The EDS elemental mapping analyses of SDB600
(Fig. S1b-e) and SDB900 (Fig. 1d–g) further supported the relatively
uniform distribution of O, N and Fe on sludge-derived biochar. All of the
aforementioned results suggested that iron oxides and abundant func­
tional groups (mainly − OH, C=O, C=C) formed within pyrolysis process
established the basis for imparting sludge-derived biochar with good
catalytic properties.
The carbon structures of two sludge-derived biochars were further
characterized by Raman spectroscopy. The deconvolution of Raman
spectra was carried out to further understand the structural changes of
biochars during pyrolysis process (Li et al., 2006), as shown in Fig. 2e
and f. D and G bands were typically observed in the curve-fitted Raman
spectrum of both biochars, corresponding to the disordered carbon/­
structural defects and graphitic carbon, respectively (Jiang et al., 2013).
A weak and broad 2D band centered at 2800 cm− 1 was observed for the
multilayer graphene phase structure (Li et al., 2017). As a significant
parameter, the intensity ratio of the D and G bands (ID/IG) based on the
peak area values was calculated to evaluated the degree of graph­
itization/defectiveness in carbon materials (Shao et al., 2017). The Fig. 3. The concentrations of HCl-extractable Fe(II) from SDB600 and SDB900
resulting ID/IG values showed a decrease from 1.41 for SDB600 to 1.17 with or without the presence of G. sulfurreducens PCA at each time interval.
for SDB900. The change implied that higher graphitized degree was Error bar represents the standard deviation of triplicate.
generated from the conversion of amorphous carbon to ordered sp2
carbon crystallites with increasing pyrolysis temperature (Rhim et al., phenylhydroxylamine) were observed in HPLC chromatograms. In the
2010). The increase of the pyrolysis temperature decreased the full initial 2 h, the removal of NB by two sludge-derived biochars was
width half maximum (FWHM) of the G band in SDB900 (73.83 cm− 1) different and the generation of aniline was not observed (Fig. S8). Also,
compared to that of SDB600 (88.07 cm− 1), suggesting an enhancement no NB degradation products in the cell-free biochar medium were
of the crystallinity in biochar (Sousa et al., 2020). Meanwhile, the peak observed in HPLC chromatograms, indicating that the lag period of
position of the G band was shifted to higher wavenumbers with the aniline production was mainly attributed to the NB adsorption property
pyrolysis temperature increasing from 600 to 900 ℃ (Table S1). These of biochar. And the equilibrium concentrations of NB were approxi­
results demonstrated the formation of more graphitized structure in mately 24 μM and 57 μM in the presence of SDB600 or SDB900 after 2 h,
SDB900 than that in SDB600. All results supported the increase of respectively. Subsequently, only negligible aniline production occurred
graphitic content during pyrolysis treatments, which were in line with in the treatments amended with SDB600 or SDB900 while without
the conclusion from previous literature (Rhim et al., 2010). G. sulfurreducens, suggesting that sole sludge-derived biochar mainly
achieved NB removal by adsorption rather than reduction process under
3.2. Reduction of Fe(III) with sludge-derived biochar anaerobic conditions. For the treatment with pure G. sulfurreducens, the
NB concentration decreased from 100 μM to 75 μM with an aniline
Compared to bacteria-free controls, the reduction extent of Fe(III) in production of 25 μM after incubating for 15 days, illustrating the bio­
those two sludge-derived biochars (SDB600 and SDB900) was signifi­ logical reduction of NB by G. sulfurreducens occurred. Compared with
cantly enhanced by the stimulation of G. sulfurreducens PCA within 10 the biotic or abiotic controls, the removal rate of NB was increased to
days (p < 0.05) (Fig. 3). Both SDB600 and SDB900 inherently contained 99% for G. sulfurreducens-SDB900 system and 89% for G. sulfurreducens-
100− 200 μM of HCl-extractable Fe(II), which was possibly originated SDB600 system in the end of anaerobic incubation, respectively. As re­
from iron oxides found on the surface of sludge-derived biochar. With ported in previous study, other detected components of SDB600 and
the addition of G. sulfurreducens PCA, the concentrations of HCl- SDB900, such as quartz and Al2O3, had no effect on the degradation of
extractable Fe(II) increased from 160.9 ± 23.6 μM to 427.7 ± 0.6 μM NB (Luan et al., 2015). Therefore, in the co-existence system of
(SDB600) and 116.6 ± 6.1 μM to 487.6 ± 53.8 μM (SDB900) after G. sulfurreducens and sludge-derived biochar, the fate of NB could be
incubating for 10 days. Consistent with the previous literatures, divided into three routes, namely, i) absorption by biochar, ii) degra­
G. sulfurreducens PCA was capable of reducing crystalline iron oxides or dation of NB to aniline and iii) the residual part. Interestingly, all re­
poorly crystalline iron oxides (Notini et al., 2019). Our results further action systems exhibited mass balance during the 15-days removal of NB
confirmed that G. sulfurreducens PCA could utilize Fe-bearing complexes (Fig. 4c and S3). It could be clearly seen that the total amounts of NB
on biochars as electron acceptors to enhance the transformation of Fe reduced by G. sulfurreducens-SDB600 or G. sulfurreducens-SDB900 sys­
(III) to Fe(II) by EET process. The generation of biogenic Fe(II) by tems were 56 μM and 66 μM after 15 Day, respectively, which were
G. sulfurreducens PCA was beneficial to reduce the redox potential of the higher than that of pure cells. Thus, in addition to the biological
reaction system and thus facilitated the anaerobic transformation of NB. reduction of G. sulfurreducens itself, biogenic Fe(II) may also be involved
in the abiotic reduction of NB to aniline.
3.3. Biochar-stimulated reduction of NB The reduction of NB in these reaction systems was fitted with the
pseudo-first-order kinetic model. As shown in Fig. 4d, the first-order rate
The reduction of NB by G. sulfurreducens with the presence of sludge- constants for NB removal were 0.135 ± 0.011 d− 1 and 0.230 ± 0.020
derived biochar was investigated within a time span of 15 days. Fig. 4a d− 1 in G. sulfurreducens with SDB600 and SDB900, respectively. These
and b illustrated the reduction of NB to aniline as the corresponding values were 7.9- to 13.5-times higher than that of pure G. sulfurreducens
product and no other reaction intermediates (e.g., nitrosobenzene or PCA (0.017 ± 0.005 d− 1). These results demonstrated that the anaerobic

5
Y. Lu et al. Journal of Hazardous Materials 403 (2021) 123682

Fig. 4. The adsorption and reduction of nitrobenzene (NB) (a) with the production of aniline (b) in different controls and treatments. The mass balance of NB
concentrations during the cultivation of G. sulfurreducens with the presence of SDB900 (c). NB concentrations transformed to ln([Ct]/[C0]) to calculate first-order
reduction rates (d). The initial concentration of NB was 100 μM, and the biochar concentration was 1 g/L. Error bar represents the standard deviation of triplicate.

reduction rate of NB was appreciably accelerated by sludge-derived peak around − 16 mV vs. SCE appeared in the forward scan, while its
biochars prepared with the elevated pyrolysis temperature from 600 reduction counterpart was not discernible in the reverse scan. The slight
to 900 ℃. intensity of this peak indicated that its contribution to electron transfer
was insignificant. For the DPV of pure SDB600, a minor reduction peak
at +14 mV vs. SCE appeared in the reverse scan, but its counterpart was
3.4. Impacts of biochar on the microbial electron flow route
too weak to be detected in the forward scan. Nevertheless, the redox
peaks in the combined system of SDB600 and G. sulfurreducens nega­
Differential pulse voltammetry (DPV) and Cyclic voltammetry (CV)
tively shifted to − 216 mV (reduction peak) and − 36 mV vs. SCE
were performed to characterize the electrochemical properties of
(oxidation peak) compared to pure SDB600 (Fig. S5b). The comparison
SDB600 and SDB900 before the reduction of NB. DPV with high selec­
of SDB600 and SDB900 suggested that SDB900 possessed more redox
tivity/sensitivity was applied to show the redox peaks masked by the
substances and higher electrochemical activity, which were essential to
catalytic wave in CV. DPV on pure G. sulfurreducens PCA showed a major
support EET process.
pair of redox peaks at around − 422/− 387 mV vs. saturated calomel
DPV analysis was also conducted to detect redox properties of su­
electrode (SCE) (Fig. S5a). The value of major redox peaks gave prox­
pernatant of culture medium with pure G. sulfurreducens, SDB900 or
imity to the potential reported for c-type cytochrome OmcB (− 430 mV
SDB600, respectively (Fig. 5c). No obvious redox peaks were detected in
vs. SCE) located at the outer membrane of G. sulfurreducens (Richter
fresh growth medium. An apparent reduction peak at +225 mV vs. SCE
et al., 2009). Besides, the reduction peak at − 387 mV vs. SCE had a
was detected in the supernatant of pure PCA cells, which was likely
shoulder peak at approximately − 309 mV vs. SCE and another minor
attributed to the redox protein secreted by strain PCA itself (Li et al.,
oxidation peak appeared at − 141 mV vs. SCE, which were presumably
2016). The supernatant of SDB900 and SDB600 without strain PCA
assigned to c-type cytochromes of G. sulfurreducens (Magnuson et al.,
exhibited two pairs of DPV peaks, suggesting that some redox substances
2001).
would be leached from sludge-derived biochar into the medium.
SDB900 displayed strong DPV signals near − 136 mV and − 296 mV
Additional evidences to support sludge-derived biochar as electron
vs. SCE, indicating the presence of redox structures that might act as
shuttle resulted from CV analysis (Fig. 5d). The redox couples at -273
electron shuttle for electron transfer (Fig. 5a). The reduction peak at
and − 370 mV vs. SCE were attributed to the existence of c-type cyto­
− 296 mV was broaden by the adjacent peak at − 172 mV. SDB900
chromes in G. sulfurreducens, which was confirmed to be directly
combined with G. sulfurreducens showed three distinct redox couples
involved in EET process as previously reported (Katuri et al., 2010). Both
(Fig. 5b). The pair at − 481 mV and − 448 mV vs. SCE were potentially
SDB900 and SDB600 with the presence of G. sulfurreducens displayed a
attributed to c-type cytochrome OmcZ (− 465 mV vs. SCE) or OmcS
single redox couple in CV with potentials ranging from +300 mV to
(− 456 mV vs. SCE) (Inoue et al., 2010; Jain et al., 2011). Besides, the
+500 mV vs. SCE. The range of the redox potentials resembled the
pair at − 265 mV and − 199 mV vs. SCE might represent the redox sub­
electrochemical properties of quinone/hydroquinone redox couple as
stances in the composite system, which were postulated to be c-type
previously reported (Quan et al., 2007). Compared to the pure bacterial
cytochrome or quinone-like substances (Leng et al., 2014). An additional

6
Y. Lu et al. Journal of Hazardous Materials 403 (2021) 123682

Fig. 5. DPV analyses of pure SDB900 (a), SDB900-G. sulfurreducens PCA combined system (b) and supernatant of different controls (c). Cyclic voltammetric analyses
of glassy carbon electrode (GCE), pure G. sulfurreducens PCA and G. sulfurreducens PCA with SDB900/ SDB600 (d). The arrows represented the scan direction of DPV.

cells, the combined system of SDB900/SDB600 and G. sulfurreducens of electron exchange (Kappler et al., 2014).
together had larger peak currents. The increase of peak current During the pyrolysis process, inherent iron of sludge was converted
demonstrated that sludge-derived biochar contributed in promoting into iron oxides (mainly γ-Fe2O3 and Fe3O4), and the Fe-bearing biochar
electron transfer from G. sulfurreducens to glassy carbon electrode (GCE). showed magnetic property (Fig. S7). The generation of magnetism was
Notably, SDB900 generated higher peak currents than SDB600, further beneficial to the growth of EAB like Geobacter and could stimulate EET
supporting that SDB900 exhibited higher electrical conductivity for EET. process (Zhou et al., 2019). Besides, the standard redox potential of Fe
Further EPR analysis was conducted on SDB600 and SDB900 (II)/Fe(III) couple was known to be +770 mV (Pereira et al., 2016).
(Fig. S6). The emergence of signals for both biochars showed that Accordingly, both SDB900 (+263 mV vs. SCE) and SDB600 (+129 mV
persistent free radicals were formed in the process of sludge pyrolysis. vs. SCE) could theoretically act as feasible redox mediators to accept
The g value was calculated from EPR spectrum to enable the determi­ electrons from G. sulfurreducens (− 415 mV vs. SCE) and further donate
nation of free radical type (Roden et al., 2010). SDB600 and SDB900 those electrons to Fe(III).
exhibited the same g value of 2.0031, which could be attributed to the Besides, the addition of sludge-derived biochar triggered microbial
mixed carbon- and oxygen-centered radicals (Xu et al., 2016). However, dissimilatory iron reduction process by G. sulfurreducens, in which Fe
the EPR signal intensity markedly decreased with the increase of the (III) on sludge-derived biochar could be firstly reduced to Fe(II) by
pyrolysis temperature. Similar scenarios were found in previous studies G. sulfurreducens, and those generated biogenic Fe(II) provided reactive
(Zhu et al., 2018; Duan et al., 2018). These results were ascribed to the sites to facilitate electrons transfer to NB occurring at the cell/biochar
elimination of oxygen functionalities and condensation/graphitization interface. According to previous reports, adsorbed Fe(II) species
of carbon clusters (Zhu et al., 2018). Herein, we deduced that electron possessed higher reductive activity than aqueous Fe(II) for NB reduction
transfer was accelerated by highly graphitic biochars to improve both (Luan et al., 2015; Lu et al., 2019). Therefore, biogenic Fe(II) in Geo­
biotic and abiotic reduction of NB. bacter-biochar combined system may be the major driving factor for the
NB reduction. Besides, some biogenic Fe(II) might be quickly adsorbed
4. Discussion on the surface of sludge-derived biochar to decrease its redox potential,
which may explain the negative shift of redox pair in the DPV diagram of
In this study, Fe-bearing biochar was generated from municipal Geobacter-biochar combined systems. Additionally, the Fe(II)/total iron
sewage sludge, and with G. sulfurreducens together they were jointly ratio remained almost constant during 15 days of cultivation (Fig. S4),
responsible for the enhanced extracellular reduction of NB to produce which suggested biogenic Fe(II) could be re-oxidized to Fe(III) and a
aniline as end product. During the cultivation, bacterial cells formed dynamic cycling of Fe(II)/Fe(III) occurred in the Geobacter-biochar
compact aggregates with Fe-bearing biochar during the NB reduction combined system during long-term reduction of NB (Lu et al., 2019).
process, which was accordance with the results found with the SEM In our study, the increase of pyrolysis temperature (from 600 ℃ to
micrographs. The dense aggregate was in favor of shortening the dis­ 900 ℃) resulted in the generation of more graphitized structure, which
tance of EET between G. sulfurreducens and biochar to increase the rate was endowed with higher mediation activity and favored the EET of

7
Y. Lu et al. Journal of Hazardous Materials 403 (2021) 123682

G. sulfurreducens as well as the NB reduction. The results of FTIR, Raman mediate electron transfer. On the other hand, quinone/hydroquinone
and EPR suggested the elevated pyrolysis temperature accelerated the redox couple in biochar, i.e. combining quinone groups as electron
decomposition of sludge to generate the condensed aromatic structure acceptor and hydroquinones as electron donor, could form an electron
and increased graphitized degree. Higher graphitized degree favored transfer conduit by redox reaction for EET. And according to our results,
biochar with higher conductivity to promote electrons transfer from the conductor mechanism by graphitized structure of biochar, instead of
G. sulfurreducens to contaminants. These results were in good agreement the mediation of quinone moieties, played a dominant role in mediating
with a previous study, where Sun et al. (2017) reported the higher py­ electron transfer during NB reduction. NB was transformed to the end
rolysis temperature led to an increase in the formation of graphitic product aniline through Geobacter-biochar combined system under
structures of biochar for rapid electron transfer. In addition, quinone anaerobic conditions, which was consistent with the results of previous
and hydroquinone couple has previously been proved to act as a surface studies (Wang et al., 2011; Luan et al., 2015). Besides, aniline was
redox-active center, allowing biochar to accept and contribute electrons relatively less toxic and easier to be degraded than the parent com­
reversibly (Yu et al., 2015; Zhang et al., 2018). Therefore, quinoid-like pounds. In order to completely deal with the NB contaminant in prac­
groups might distribute on the surface of biochar to mediate electron tical applications, subsequent treatments should be conducted for
transfer in NB reduction and microbial iron reduction, which was in further biodegradation and mineralization of aniline.
good agreement with the characterization results of FTIR and XPS as
described above. Similar to our study, Saquing et al. (2016) showed that 5. Conclusion
wood biochar accelerated electron transfer from G. metallireducens to
nitrate and promoted the microbial denitrification mediated by the Our results demonstrated the addition of Fe-bearing biochar
charging and discharging functions of quinone/hydroquinone redox (generated from sludge by one-step pyrolysis) significantly enhanced the
couple. reduction of NB to aniline by G. sulfurreducens. The enhancement effect
In the end, a systematical mechanism was proposed for the biotic and was primarily attributed to the important role of Fe-bearing biochar in
abiotic reduction of NB (Fig. 6). A relatively small portion of NB was both mediating EET of G. sulfurreducens and assisting the formation of
directly reduced by pure G. sulfurreducens, while the majority of NB was biogenic Fe(II) as natural iron source. Graphitized structure and
indirectly reduced by surface-adsorbed biogenic Fe(II) produced by quinone-like moieties in generated biochar might act as electron shut­
G. sulfurreducens from Fe-bearing biochar. Besides, as reflected by the tles, and thus accelerated EET process and enhanced the rate of NB
electrochemical measurements, most probably c-type cytochromes of reduction. Besides, adding Fe-bearing biochar triggered microbial
G. sulfurreducens were involved in EET process, in which intracellular dissimilatory iron reduction by G. sulfurreducens and the biogenic Fe(II)
electrons were released to external environment for dissimilatory iron in Geobacter-biochar combined system was the major driving factor for
reduction or the biotic reduction of NB (Inoue et al., 2010). Fe-bearing the NB reduction, which might act as an important but overlooked
biochar played the role of electron transfer conductor and natural iron electron transfer route in the anaerobic environments. The great mag­
source in the Geobacter-biochar combined reduction system. Besides, to netic property of sludge-derived biochar makes it easy to separate from
explain the mechanisms of electron transfer mediated by biochar, two reaction system, therefore, further investigation will conduct on the
major routes were proposed. On one hand, the graphitized domains of regeneration property of deactivated sludge-derived biochar to improve
biochar as direct electron transfer conduit had the potential to directly the reusability. Future applications of the combination of

Fig. 6. Schematic of the proposed mechanism of integrated biotic/abiotic reduction of nitrobenzene (NB) by the combination of G. sulfurreducens PCA and Fe-
bearing biochar.

8
Y. Lu et al. Journal of Hazardous Materials 403 (2021) 123682

microorganism and biochar should evaluate the contaminant removal in Jiang, J.H., Zhang, L., Wang, X.Y., Holm, N., Rajagopalan, K., Chen, F.L., Ma, S.G., 2013.
Highly ordered macroporous woody biochar with ultra-high carbon content as
a long-term period with the change of biochar surface structure. Such
supercapacitor electrodes. Electrochim. Acta 113, 481–489.
assessments can contribute to advance the understanding on the fate of Kappler, A., Wuestner, M.L., Ruecker, A., Harter, J., Halama, M., Behrens, S., 2014.
nitroaromatic compounds in anoxic environment and help to predict the Biochar as an electron shuttle between bacteria and Fe(III) minerals. Environ. Sci.
natural attenuation at in situ fields amended with biochar. Technol. Lett. 1, 339–344.
Katuri, K.P., Kavanagh, P., Rengaraj, S., Leech, D., 2010. Geobacter sulfurreducens
biofilms developed under different growth conditions on glassy carbon electrodes:
Declaration of competing interest insights using cyclic voltammetry. Chem. Commun. (Camb.) 46, 4758–4760.
Leng, Y.Q., Guo, W.L., Shi, X., Li, Y.Y., Wang, A.Q., Hao, F.F., Xing, L.T., 2014.
Degradation of Rhodamine B by persulfate activated with Fe3O4: effect of
The authors declare that they have no known competing financial polyhydroquinone serving as an electron shuttle. Chem. Eng. J. 240, 338–343.
interests or personal relationships that could have appeared to influence Li, X.J., Hayashi, J., Li, C.Z., 2006. FT-Raman spectroscopic study of the evolution of char
the work reported in this paper. structure during the pyrolysis of a Victorian brown coal. Fuel 85, 1700–1707.
Li, D.B., Cheng, Y.Y., Li, L.L., Li, W.W., Huang, Y.X., Pei, D.N., Tong, Z.H., Mu, Y., Yu, H.
Q., 2014. Light-driven microbial dissimilatory electron transfer to hematite. Phys.
CRediT authorship contribution statement Chem. Chem. Phys. 16, 23003–23011.
Li, S.W., Sheng, G.P., Cheng, Y.Y., Yu, H.Q., 2016. Redox properties of extracellular
polymeric substances (EPS) from electroactive bacteria. Sci. Rep. 6, 39098.
Yue Lu: Conceptualization, Methodology, Investigation, Writing - Li, H.Q., Hu, J.T., Meng, Y., Su, J.H., Wang, X.J., 2017. An investigation into the rapid
review & editing, Funding acquisition. Qingqing Xie: Conceptualiza­ removal of tetracycline using multilayered graphene-phase biochar derived from
tion, Investigation, Writing - original draft, Data curation, Formal waste chicken feather. Sci. Total Environ. 603-604, 39–48.
Lu, Y., Li, J.F., Li, Y.M., Liang, L.P., Dong, H.P., Chen, K., Yao, C.X., Li, Z.F., Li, J.X.,
analysis. Lin Tang: Supervision, Project administration, Funding
Guan, X.H., 2019. The roles of pyrite for enhancing reductive removal of
acquisition, Investigation. Jiangfang Yu: Conceptualization, Formal nitrobenzene by zero-valent iron. Appl. Catal. B-Environ. 242, 9–18.
analysis, Methodology, Writing - review & editing. Jingjing Wang: Luan, F.B., Burgos, W.D., Xie, L., Zhou, Q., 2010. Bioreduction of nitrobenzene, natural
Investigation, Data curation, Writing - review & editing. Zhaohui Yang: organic matter, and hematite by Shewanella putrefaciens CN32. Environ. Sci. Technol.
44, 184–190.
Resources, Writing - review & editing. Changzheng Fan: Resources, Luan, F.B., Liu, Y., Griffin, A.M., Gorski, C.A., Burgos, W.D., 2015. Iron(III)-bearing clay
Writing - review & editing. Shoujuan Zhang: Methodology, Software, minerals enhance bioreduction of nitrobenzene by Shewanella putrefaciens CN32.
Writing - review & editing. Environ. Sci. Technol. 49, 1418–1426.
Magnuson, T.S., Isoyama, N., Hodges-Myerson, A.L., Davidson, G., Maroney, M.J.,
Geesey, G.G., Lovley, D.R., 2001. Isolation, characterization and gene sequence
Acknowledgements analysis of a membrane-associated 89 kDa Fe(III) reducing cytochrome c from
Geobacter sulfurreducens. Biochem. J. 359, 147–152.
Melton, E.D., Swanner, E.D., Behrens, S., Schmidt, C., Kappler, A., 2014. The interplay of
We thank the research group of Prof. Hanqing Yu (University of microbially mediated and abiotic reactions in the biogeochemical Fe cycle. Nat. Rev.
Science & Technology of China) to provide Geobacter sulfurreducens PCA. Microbiol. 12, 797–808.
This work was supported by the National Natural Science Foundation of Notini, L., Byrne, J.M., Tomaszewski, E.J., Latta, D.E., Zhou, Z., Scherer, M.M.,
Kappler, A., 2019. Mineral defects enhance bioavailability of goethite toward
China (Grant No. 51709100, 51679084, 51579096), the Fundamental microbial Fe(III) reduction. Environ. Sci. Technol. 53, 8883–8891.
Research Funds for the Central Universities (531107050936), the Key Peng, L., You, S.J., Wang, J.Y., 2010. Carbon nanotubes as electrode modifier promoting
Research and Development Project of Hunan Province of China direct electron transfer from Shewanella oneidensis. Biosens. Bioelectron. 25,
1248–1251.
(2017SK2241), the National Innovative Talent Promotion Program of
Peng, C., Zhai, Y.B., Zhu, Y., Xu, B.B., Wang, T.F., Li, C.T., Zeng, G.M., 2016. Production
China (2017RA2088), and the Funds for Innovative Province Con­ of char from sewage sludge employing hydrothermal carbonization: char properties,
struction of Hunan Province of China (2019RS3012). combustion behavior and thermal characteristics. Fuel 176, 110–118.
Pereira, R.A., Pereira, M.F.R., Alves, M.M., Pereira, L., 2014. Carbon based materials as
novel redox mediators for dye wastewater biodegradation. Appl. Catal. B-Environ.
Appendix A. Supplementary data 144, 713–720.
Pereira, L., Pereira, R., Pereira, M.F.R., Alves, M.M., 2016. Effect of different carbon
Supplementary material related to this article can be found, in the materials as electron shuttles in the anaerobic biotransformation of nitroanilines.
Biotechnol. Bioeng. 113, 1194–1202.
online version, at doi:https://doi.org/10.1016/j.jhazmat.2020.123682. Pourhosseini, S.E.M., Norouzi, O., Salimi, P., Naderi, H.R., 2018. Synthesis of a novel
interconnected 3D pore network algal biochar constituting iron nanoparticles
References derived from a harmful marine biomass as high-performance asymmetric
supercapacitor electrodes. ACS Sustain. Chem. Eng. 6, 4746–4758.
Quan, M., Sanchez, D., Wasylkiw, M.F., Smith, D.K., 2007. Voltammetry of quinones in
Amezquita-Garcia, H.J., Razo-Flores, E., Cervantes, F.J., Rangel-Mendez, J.R., 2013.
unbuffered aqueous solution: reassessing the roles of proton transfer and hydrogen
Activated carbon fibers as redox mediators for the increased reduction of
bonding in the aqueous electrochemistry of quinones. J. Am. Chem. Soc. 129,
nitroaromatics. Carbon 55, 276–284.
12847–12856.
Chen, T., Zhang, Y.X., Wang, H.T., Lu, W.J., Zhou, Z.Y., Zhang, Y.C., Ren, L.L., 2014.
Rhim, Y.R., Zhang, D.J., Fairbrother, D.H., Wepasnick, K.A., Livi, K.J., Bodnar, R.J.,
Influence of pyrolysis temperature on characteristics and heavy metal adsorptive
Nagle, D.C., 2010. Changes in electrical and microstructural properties of
performance of biochar derived from municipal sewage sludge. Bioresour. Technol.
microcrystalline cellulose as function of carbonization temperature. Carbon 48,
164, 47–54.
1012–1024.
Duan, X.G., Ao, Z.M., Zhang, H.Y., Saunders, M., Sun, H.Q., Shao, Z.P., Wang, S.B., 2018.
Richter, H., Nevin, K.P., Jia, H.F., Lowy, D.A., Lovley, D.R., Tender, L.M., 2009. Cyclic
Nanodiamonds in sp2/sp3 configuration for radical to nonradical oxidation: core-
voltammetry of biofilms of wild type and mutant Geobacter sulfurreducens on fuel cell
shell layer dependence. Appl. Catal. B-Environ. 222, 176–181.
anodes indicates possible roles of OmcB, OmcZ, type IV pili, and protons in
Gao, X.T., Tan, W.B., Zhao, Y., Wu, J.Q., Sun, Q.H., Qi, H.S., Xie, X.Y., Wei, Z.M., 2019.
extracellular electron transfer. Energy Environ. Sci. 2, 506–516.
Diversity in the mechanisms of humin formation during composting with different
Roden, E.E., Kappler, A., Bauer, I., Jiang, J., Paul, A., Stoesser, R., Konishi, H., Xu, H.F.,
materials. Environ. Sci. Technol. 53, 3653–3662.
2010. Extracellular electron transfer through microbial reduction of solid-phase
Han, Q., Liu, Z.H., Xu, Y.Y., Chen, Z.Y., Wang, T.M., Zhang, H., 2007. Growth and
humic substances. Nat. Geosci. 3, 417–421.
properties of single-crystalline gamma-Fe2O3 nanowires. J. Phys. Chem. C 111,
Saquing, J.M., Yu, Y.H., Chiu, P.C., 2016. Wood-derived black carbon (biochar) as a
5034–5038.
microbial electron donor and acceptor. Environ. Sci. Technol. Lett. 3, 62–66.
Ho, S.H., Chen, Y.D., Yang, Z.K., Nagarajan, D., Chang, J.S., Ren, N.Q., 2017. High-
Shao, J.Q., Ma, F.W., Wu, G., Dai, C.C., Geng, W.D., Song, S.J., Wan, J.F., 2017. In-situ
efficiency removal of lead from wastewater by biochar derived from anaerobic
MgO (CaCO3) templating coupled with KOH activation strategy for high yield
digestion sludge. Bioresour. Technol. 246, 142–149.
preparation of various porous carbons as supercapacitor electrode materials. Chem.
Inoue, K., Qian, X.L., Morgado, L., Kim, B.C., Mester, T., Izallalen, M., Salgueiro, C.A.,
Eng. J. 321, 301–313.
Lovley, D.R., 2010. Purification and characterization of OmcZ, an outer-surface,
Shi, L., Dong, H.L., Reguera, G., Beyenal, H., Lu, A.H., Liu, J., Yu, H.Q., Fredrickson, J.K.,
octaheme c-type cytochrome essential for optimal current production by Geobacter
2016. Extracellular electron transfer mechanisms between microorganisms and
sulfurreducens. Appl. Environ. Microbiol. 76, 3999–4007.
minerals. Nat. Rev. Microbiol. 14, 651–662.
Jain, A., Gazzola, G., Panzera, A., Zanoni, M., Marsili, E., 2011. Visible
Sousa, D.V., Guimaraes, L.M., Felix, J.F., Ker, J.C., Schaefer, C., Rodet, M.J., 2020.
spectroelectrochemical characterization of Geobacter sulfurreducens biofilms on
Dynamic of the structural alteration of biochar in ancient Anthrosol over a long
optically transparent indium tin oxide electrode. Electrochim. Acta 56,
timescale by Raman spectroscopy. PLoS One 15, e0229447.
10776–10785.

9
Y. Lu et al. Journal of Hazardous Materials 403 (2021) 123682

Sun, T., Levin, B.D.A., Guzman, J.J.L., Enders, A., Muller, D.A., Angenent, L.T., Yang, Q., Wang, X.L., Luo, W., Sun, J., Xu, Q.X., Chen, F., Zhao, J.W., Wang, S.N., Yao, F.
Lehmann, J., 2017. Rapid electron transfer by the carbon matrix in natural pyrogenic B., Wang, D.B., Li, X.M., Zeng, G.M., 2018. Effectiveness and mechanisms of
carbon. Nat. Commun. 8, 14873. phosphate adsorption on iron-modified biochars derived from waste activated
Tang, L., Yu, J.F., Pang, Y., Zeng, G.M., Deng, Y.C., Wang, J.J., Ren, X.Y., Ye, S.J., sludge. Bioresour. Technol. 247, 537–544.
Peng, B., Feng, H.P., 2018. Sustainable efficient adsorbent: alkali-acid modified Yu, L.P., Yuan, Y., Tang, J., Wang, Y.Q., Zhou, S.G., 2015. Biochar as an electron shuttle
magnetic biochar derived from sewage sludge for aqueous organic contaminant for reductive dechlorination of pentachlorophenol by Geobacter sulfurreducens. Sci.
removal. Chem. Eng. J. 336, 160–169. Rep. 5, 10.
Tang, Y., Alam, M.S., Konhauser, K.O., Alessi, D.S., Xu, S.N., Tian, W.J., Liu, Y., 2019. Yu, J.F., Tang, L., Pang, Y., Zeng, G.M., Wang, J.J., Deng, Y.C., Liu, Y.N., Feng, H.P.,
Influence of pyrolysis temperature on production of digested sludge biochar and its Chen, S., Ren, X.Y., 2019. Magnetic nitrogen-doped sludge-derived biochar catalysts
application for ammonium removal from municipal wastewater. J. Clean. Prod. 209, for persulfate activation: internal electron transfer mechanism. Chem. Eng. J. 364,
927–936. 146–159.
Tong, H., Hu, M., Li, F.B., Liu, C.S., Chen, M.J., 2014. Biochar enhances the microbial Zhang, P., Zheng, S.L., Liu, J., Wang, B.C., Liu, F.H., Feng, Y.J., 2018. Surface properties
and chemical transformation of pentachlorophenol in paddy soil. Soil Biol. Biochem. of activated sludge-derived biochar determine the facilitating effects on Geobacter
70, 142–150. co-cultures. Water Res. 142, 441–451.
Van der Zee, F.P., Cervantes, F.J., 2009. Impact and application of electron shuttles on Zhao, H.Q., Liu, Q., Wang, Y.X., Han, Z.Y., Chen, Z.G., Mu, Y., 2018. Biochar enhanced
the redox (bio)transformation of contaminants: a review. Biotechnol. Adv. 27, biological nitrobenzene reduction with a mixed culture in anaerobic systems: short-
256–277. term and long-term assessments. Chem. Eng. J. 351, 912–921.
Wang, A.J., Cheng, H.Y., Liang, B., Ren, N.Q., Cui, D., Lin, N., Kim, B.H., Rabaey, K., Zhou, H.H., Zhang, D.W., Zhang, Y.F., Yang, Y., Liu, B.F., Ren, N.Q., Xing, D.F., 2019.
2011. Efficient reduction of nitrobenzene to aniline with a biocatalyzed cathode. Magnetic cathode stimulates extracellular electron transfer in bioelectrochemical
Environ. Sci. Technol. 45, 10186–10193. systems. ACS Sustain. Chem. Eng. 7, 15012–15018.
Wang, D.Z., Zheng, G.Y., Zhou, L.X., 2012. Isolation and characterization of a Zhu, S.S., Ho, S.H., Huang, X.C., Wang, D.W., Yang, F., Wang, L., Wang, C.Y., Cao, X.D.,
nitrobenzene-degrading bacterium Klebsiella ornithinolytica NB1 from aerobic Ma, F., 2017. Magnetic nanoscale zerovalent iron assisted biochar: interfacial
granular sludge. Bioresour. Technol. 110, 91–96. chemical behaviors and heavy metals remediation performance. ACS Sustain. Chem.
Wu, J.H., Yin, W.Z., Gu, J.J., Li, P., Wang, X.D., Yang, B., 2013. A biotic Fe0-H2O system Eng. 5, 9673–9682.
for nitrobenzene removal from groundwater. Chem. Eng. J. 226, 14–21. Zhu, S.S., Huang, X.C., Ma, F., Wang, L., Duan, X.G., Wang, S.B., 2018. Catalytic removal
Xu, S.N., Adhikari, D., Huang, R.X., Zhang, H., Tang, Y.Z., Roden, E., Yang, Y., 2016. of aqueous contaminants on N-doped graphitic biochars: inherent roles of adsorption
Biochar-facilitated microbial reduction of hematite. Environ. Sci. Technol. 50, and nonradical mechanisms. Environ. Sci. Technol. 52, 8649–8658.
2389–2395.

10

You might also like