You are on page 1of 10

Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 582–591

Contents lists available at ScienceDirect

Chemical Engineering Research and Design

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

A comparative study of arsenic(V), tetracycline and


nitrate ions adsorption onto magnetic biochars and
activated carbon

Guangcai Tan a,b , Yi Mao a,c , Hongyuan Wang d , Nan Xu a,∗


a Shenzhen Engineering Research Center for Nanoporous Water Treatment Materials, School of Environment and
Energy, Peking University Shenzhen Graduate School, Shenzhen 518055, China
b Department of Civil and Environmental Engineering, Louisiana State University, Baton Rouge, LA 70803, USA
c UCD Dooge Centre for Water Resource Research, School of Civil, Structural and Environmental Engineering,

University College Dublin, Belfield, Dublin 4, Ireland


d Key Laboratory of Nonpoint Source Pollution Control, Ministry of Agriculture, Institute of Agricultural Resources

and Regional Planning, Chinese Academy of Agricultural Sciences, Beijing 100081, China

a r t i c l e i n f o a b s t r a c t

Article history: The removal of contaminants in co-solute systems is a pressing issue. Herein, the adsorp-
Received 19 November 2019 tion capacities of magnetic biochars and activated carbon were investigated and compared
Received in revised form 9 April for various aqueous anions, including As(V), tetracycline (TC) and NO3 − , existed alone or
2020 as a mixture. One-step (pre-treating biomass with FeCl3 before pyrolysis (FB)) and two-
Accepted 9 May 2020 step (pyrolysing biomass before modification with FeCl3 (BF)) magnetization methods were
Available online 23 May 2020 applied to synthesize magnetic biochars. Magnetic activated carbon (AF) was also prepared
with the two-step magnetization method. FB presented the highest adsorption capacities for
Keywords: As(V) and NO3 − respectively at 6.77 and 6.31 mg g−1 . AC (612.38 mg g−1 ) showed the highest
Biochar adsorption capacity for TC, which decreased by 189.76% after magnetization. Magnetiza-
As(V) tion was proved to be a promising method to improve the adsorption of both As(V) and
Tetracycline NO3 − on adsorbents through enhanced electrostatic attraction, while the adsorption of TC
Nitrate depended mainly on high specific surface area of the adsorbent. Results from competitive
Co-adsorption adsorption in the co-solute systems suggested that FB had a great potential for applica-
tion in the co-contaminated environment (heavy metal and inorganic anions), and AF could
be a promising adsorbent to remediate multiple-contaminated environment (i.e. organic
pollution and heavy metal/inorganic anions).
© 2020 Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

1. Introduction et al., 2001; Fujino et al., 2004). Zhang et al. (2002) reported
that the average As content in groundwater was more than
Arsenic (As) contamination in surface water or groundwater is 0.2 mg L−1 in Hetao area (northwest of China) which signifi-
a global health concern because of its various kinds of toxici- cantly exceeded the permissible limits set by the World Health
ties including epidemiological, cytotoxic and genotoxic effects Organization (WHO) (10 ␮g L−1 in drinking water) (Xia et al.,
(Mangwandi et al., 2016; Islam et al., 2018a,b; Song et al., 2019). 2014). As(V) accounted for a greater percentage in the total
In the northwest of China, high levels of As (up to 1.35 mg L−1 ) arsenic, nearly 30% in average (in the range of 9–90%) (Xia
in groundwater have been detected as early as the 1950s (Guo et al., 2014). In the rural areas contaminated by As, the antibi-
otic pollution in groundwater also existed due to the frequent
use of domestic and livestock wastewater containing antibi-

otics for irrigation of farmland (Fytianos and Christophoridis,
Corresponding author.
2004; Wang et al., 2015; Islam et al., 2019). Different kinds of
E-mail address: xunan@pkusz.edu.cn (N. Xu).
https://doi.org/10.1016/j.cherd.2020.05.011
0263-8762/© 2020 Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 582–591 583

antibiotics (e.g. tetracycline (TC), lincomycin and sul- 2. Materials and methods
famethoxazole) and their metabolites have been detected in
groundwater in the northern China (Hu et al., 2010). The spread 2.1. Preparation of the adsorbents
of antibiotics residues can potentially induce adverse effects
on ecosystem such as acute and chronic toxicity and prop- Air-dried corn straw (Zea mays L.) was washed three times
agation of antibiotic resistance in microbes (Li et al., 2016). using ultrapure water (18.2 M cm) firstly, and then oven-dried
Besides, with the unbalanced and/or excessive application of (80 ◦ C) for 12 h. Biochar (BC) was produced using slow pyrolysis
fertilizers, nitrate (NO3 − ) was widely detected in groundwater at 600 ◦ C in a furnace (Olympic 1823HE, USA) under N2 envi-
due to its high solubility (Kilpimaa et al., 2014; Xu et al., 2016). ronment for 1.5 h. For synthesis of activated carbon (AC), dried
Chen et al. reported that the average NO3 − content in ground- corn straw was firstly impregnated in phosphoric acid (85%)
water was as high as 12.15 ± 12.92 mg L−1 in Ningxia (Hetao for 24 h, then dried for 4 h at 105 ◦ C in an oven, and pyrolyzed
area, northwest of China) (Chen et al., 2016), indicating the under the same condition as for producing BC.
potential health risk (Divband Hafshejani et al., 2016). Besides, Two synthesis methods were applied to prepare two dif-
heavy metals, organic chemicals and inorganic anions were ferent kinds of magnetic biochars (Zhang et al., 2013; Baig
often detected simultaneously in wastewater and surface et al., 2014; Tan et al., 2016b). In the first method (two-step
water (Lian et al., 2013). Therefore, it is a pressing issue to procedure), the magnetic biochar was produced with BC as the
investigate the removal of contaminants in co-solute systems. precursor. Briefly, 5 g of BC was suspended in FeCl3 solutions
As an organic carbon-rich material, biochar is synthe- (dissolving 40 g of FeCl3 ·6H2 O in 500 mL of ultrapure water) for
sized through pyrolyzing biomass under an oxygen-limited 24 h under continuous agitation using a magnetic stirrer, and
environment (Lehmann and Joseph, 2015). Because of its abun- then oven-dried (105 ◦ C) for 12 h under air. Therefore, the mag-
dant oxygen functional groups and high surface-to-volume netic biochar synthesized with this method was termed as BF.
ratio, biochar has been successfully used for treatment of In the second method, one-step procedure was applied, dur-
organically and inorganically contaminated water, which is ing which dried corn straw was immersed into the same FeCl3
beneficial to both wastewater treatment and greenhouse gas solution for 24 h, and then oven-dried (105 ◦ C) for 12 h under
mitigation (Xu et al., 2015; Wang et al., 2017; Wu et al., 2019; air. The pre-treated corn straw was pyrolyzed in a furnace at
Tan et al., 2019b). However, the raw biochar was reported to be the same condition as for producing BC. Therefore, the mag-
less efficient for treatment of anionic contaminants including netic biochar synthesized with the one-step procedure was
As(V), TC (at high pH) and NO3 − due to its negatively charged termed as FB. As the corn straw needs to be firstly activated
surface and limited functional groups (Tan et al., 2016b). In by phosphoric acid to produce AC, only the two-step procedure
addition, the powdered biochar is difficult to be separated was employed to produce magnetic activated carbon with AC
after application in aqueous solution. In order to overcome as the precursor, referred as AF. All the five kinds of adsorbents
these drawbacks, magnetic modification was employed to pro- were thoroughly washed with ultrapure water and sieved to
duce Fe2 O3 /biochar composites (Zhang et al., 2013; Zhu et al., get <1 mm sized particle before being used for the following
2014a). Zhang et al. (2013) reported that magnetic biochar can experiments.
be prepared directly from FeCl3 treated biomass through pyrol-
ysis, which showed excellent adsorption capacity for arsenic 2.2. Characterization of adsorbents
in water compared to the raw one. However, little information
was documented about the performance of magnetic biochar The CHNOS analyzer (Thermo Flash 2000, USA) was used to
in the adsorption of organic pollutant (e.g. TC) and inorganic measure the contents of different elements (C, H, O, N) in
anions (e.g. NO3 − ), and the competitive adsorption among the adsorbents. Surface structural characteristics of the adsor-
them is even less studied. Activated carbon is also a kind of bents were examined by a NOVA 2000 sorptiometer (Quan-
adsorbent which has been widely used in wastewater reme- tachrome Instruments, USA). The Brunauer–Emmett–Teller
diation owing to its large specific surface area. Therefore, it (BET) method was used to measure the pore volume and
is worthwhile to compare the adsorption ability of magnetic total surface area. The Barrett–Joiner–Halenda (BJH) method
biochars with that of magnetic activated carbon. was used to determine the pore size distributions. A scan-
Herein, for the first time, we investigated and compared ning electron microscope (SEM, Hitachi S4800, Japan) coupled
the adsorption capacities of magnetic biochars and activated with an energy dispersive X-ray spectroscopy (EDS, Oxford
carbon for various aqueous anions, As(V), TC and NO3 − , in Instruments, USA) was used to examine the morphology and
single/binary-mixture systems. Both biochar and activated elemental composition of adsorbents. The spectral properties
carbon were produced from corn-straw. Magnetic biochars of adsorbents were examined by a Fourier transform infrared
and magnetic activated carbon were produced using fer- (FT-IR) spectrometer (Nicolet Magna-IR 750, USA). The crystal
ric chloride. These adsorbents were used to evaluate the structure of the adsorbents was evaluated through Powder X-
adsorption of individual As(V), TC and NO3 − , as well as their ray diffraction (XRD) (X’Pert Pro, PANalytical, Netherland). The
binary mixtures. The main objectives were to: (1) investi- X-ray photoelectron spectroscopy (XPS) of the adsorbents was
gate and compare the adsorption capacities of the adsorbents obtained using a PHI 5100 series ESCA spectrometer (Perkin
for As(V), TC and NO3 − , and reveal the effects of magne- Elmer, USA). Magnetic properties were obtained on a vibrating
tization on the adsorption performance; (2) investigate the sample magnetometer (VSM, Lake Shore 7410, USA). The ther-
competitive adsorption in binary-mixture systems among mal stability of adsorbents was examined by a Mettler Toledo’s
As(V), TC and NO3 − ; and (3) investigate the mechanisms for TGA/DSC analyzer (TGA/SDTA851, USA). Zeta potential (ZP)
the competitive adsorption of As(V), TC and NO3 − onto these of the adsorbents was determined following the procedures
adsorbents. described in our previous work (Tan et al., 2016a).
584 Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 582–591

2.3. Adsorption experiments 2.3.5. Analytical methods


The supernatant of each sample at a given time interval
2.3.1. Chemicals was firstly centrifuged and then filtered. The concentra-
Sodium arsenate dibasic heptahydrate (Na2 HAsO4 ·7H2 O) and tions of As(V) were analyzed using an inductively coupled
potassium nitrate (KNO3 ) were purchased from Fisher Sci- plasma-atomic emission spectrometry (ICP-AES, Perkin–Elmer
entific (USA). TC hydrochloride (>98%) was purchased from Plasma 3200, USA). The high-performance liquid chromato-
Aldrich Chemical Co. (USA). The chemicals used in the graph (Agilent 1100, USA) coupled with a UV detector at 355 nm
present study were of analytical grade and dissolved in ultra- absorbing wavelength was used to determine the concentra-
pure water. The stock solutions of As(V) (1000 mg L−1 ), TC tions of TC. The targeted compound was separated on an
(1000 mg L−1 ) as well as NO3 − (1000 mg L−1 ) were prepared and Agilent Eclipse Plus C18 column (4.6 × 250 mm, 5 ␮m) using a
stored at 4 ◦ C. binary mobile phase of CH3 CN and 0.01 mol L−1 NaH2 PO4 of
20:80 (v/v). The flow rate was 1.0 mL min−1 with the injection
2.3.2. Effect of pH volume of 10 ␮L, and the retention time for TC was 3.9 min. The
All the adsorption experiments were performed through a ion chromatograph (ICS90, Dionex Inc., USA) equipped with a
batch equilibrium method. The effects of pH on the adsorp- conductivity detector was used to analyze the concentrations
tion of As(V) (5 mg L−1 ), TC (100 mg L−1 ) and NO3 − (10 mg L−1 ) of NO3 − .
onto these adsorbents were examined at five different pH val-
ues (3, 5, 7, 9, and 11). The pH was adjusted with HCl and NaOH
solutions. 3. Results and discussion

3.1. Characterization of adsorbents


2.3.3. Adsorption kinetics for As(V), TC and NO3 −
According to the tentative experiments and previous studies
Table 1 presents the physico-chemical properties of the five
(Tofighy et al., 2012; Peng et al., 2014; He et al., 2018), 40 mg L−1
adsorbents. After magnetization, both the ratios of (O + N)/C
As(V), 400 mg L−1 TC and 30 mg L−1 NO3 − were used in kinetic
and O/C obviously increased for FB, BF and AF as compared
experiments. In detail, 0.05 g adsorbent samples were added
to BC and AC, respectively, suggesting the increase of O-
into 60 mL As(V) solution (40 mg L−1 ) in a conical flask (500 mL)
containing groups. The BET surface area of FB (115.43 m2 g−1 )
with cover. The HCl and NaOH solutions were used to adjust
was more than twice that of BC (56.18 m2 g−1 ), which could
the initial solution pH to 6.0 ± 0.1. All the conical flasks were
be attributed to the reaction between iron particle and car-
then shaken at 110 rpm and 25 ± 1 ◦ C on a water bath shaker
bonaceous compound during the pyrolysis in the one-step
in the dark. Similarly, the adsorption kinetics for TC and NO3 −
procedure (Zhang et al., 2013). However, the surface area of
were conducted with 60 mL corresponding solution (at 400 and
BF (48.43 m2 g−1 ) and AF (217.87 m2 g−1 ) decreased in compari-
30 mg L−1 , respectively) and 0.05 g adsorbent samples under
son with BC (85.06 m2 g−1 ) and AC (563.93 m2 g−1 ), respectively.
the same experimental conditions. At appropriate time inter-
This could be due to the blockage of internal pores by the sur-
vals, one milliliter solution was sampled for analysis of As(V),
face complex of iron in the two-step procedure (Baig et al.,
TC and NO3 − concentration. Kinetic data were fitted using
2014). Meanwhile, the total pore volume and mesopore volume
different kinetic models (the details were given in Supplemen-
increased for FB and decreased for BF compared to BC, and
tary Materials).
also decreased for AF in comparison with AC. In addition, the
average pore size of FB, BF and AF increased compared to BC
2.3.4. Adsorption isotherms for As(V), TC and NO3 − and AC, respectively. The N2 adsorption–desorption isotherms
Adsorption isotherms experiments for As(V) were conducted and pore size distribution are exhibited in Figs. S1 and S2,
by adding 0.05 g adsorbent into 60 mL As(V) solutions (the con- respectively.
centrations were ranging from 2 to 60 mg L−1 ) in the conical Fig. S3 shows the XRD patterns of the adsorbents. For the
flasks. Then 50 mg L−1 of TC or 30 mg L−1 of NO3 − was addi- magnetic adsorbents (FB, BF and AF), the sharp and strong
tionally spiked, respectively, so as to investigate the effect of reflection peaks indicated the as-prepared iron oxides were
TC or NO3 − on As(V) adsorption. The initial solution pH was well crystallized within the biochar or activated carbon matrix.
adjusted to 6.0 ± 0.1 using HCl and NaOH solutions. The con- Maghemite (␥-Fe2 O3 ) was found as the major crystalline phase
ical flasks were shaken at 110 rpm and 25 ± 1 ◦ C on a water for the magnetic adsorbents (Zhang et al., 2013; Wang et al.,
bath shaker for 24 h in the dark. Similar to As(V), adsorp- 2014). Fig. S4 shows the SEM images and the corresponding
tion isotherms of TC and NO3 − were conducted by adding EDS spectra of the adsorbents. Element analysis from EDS sug-
0.05 g adsorbent samples into 60 mL corresponding working gested that FB (22.79%) had the highest iron weight, followed
solutions in the conical flasks at different concentrations rang- by BF (14.35%) and AF (9.91%). As a result, FB possessed the
ing from 20 to 400 and 0.5 to 50 mg L−1 , respectively. Then highest residual mass (>35%) during TGA, indicating its higher
30 mg L−1 of As(V) or NO3 − was additionally spiked, respec- stability than BF and AF (Fig. S5). High-resolution XPS spectra
tively, so as to investigate the effect of As(V) or NO3 − on TC of magnetic adsorbents in the Fe2p region are shown in Fig. S6.
adsorption. Similarly, 30 mg L−1 of As(V) or 50 mg L−1 TC was The two bands at 710.43 and 723.76 eV were always present in
additionally spiked, respectively, so as to investigate the effect the spectra for FB, BF and AF, which were attributed to the
of As(V) or TC on NO3 − adsorption. All the adsorption exper- peaks of Fe3+ including Fe 2p3/2 and Fe 2p1/2 (Zhang et al.,
iments were conducted in triplicate, and the average values 2013). The magnetic properties of FB, BF and AF are shown in
were reported. Control experiments with no adsorbents were Fig. S7. The saturation magnetization of FB was 17.5 emu g−1 ,
carried out simultaneously under the same conditions, indi- the highest among the magnetic adsorbents, suggesting the
cating that the pollutants removal was negligible. Isotherm easiest separation of FB from the aqueous solution with exter-
data were fitted with different isotherm models (the details nal magnetic field. These results further demonstrated that
were described in Supplementary Materials). biochar and activated carbon were successfully magnetized.
Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 582–591 585

Table 1 – Elemental compositions, atomic ratios and porous characterization of the biochar (BC), magnetic biochars (FB,
BF), activated carbon (AC) and magnetic activated carbon (AF).
BC FB BF AC AF

C (%) 85.06 56.27 64.13 75.81 53.24


H (%) 0.84 0.76 1.11 1.01 1.32
O (%) 4.14 10.47 8.29 10.89 14.95
N (%) 0.61 0.45 0.67 0.75 0.52
H/C 0.0099 0.0135 0.0173 0.0133 0.0248
O/C 0.0487 0.1861 0.1293 0.1436 0.2808
(O + N)/C 0.0558 0.1941 0.1397 0.1535 0.2906
BET surface area (m2 g−1 ) 56.18 115.43 48.41 563.93 217.87
Total pore volume (cm3 g−1 ) 0.0313 0.0693 0.0159 1.0872 0.4287
Mesopore volume (2–10 nm) (cm3 g−1 ) 0.0128 0.0160 0.0032 0.0728 0.0428
Average pore size (nm) 3.33 5.42 5.18 2.26 3.47

Fig. 1 – The effects of pH values on As(V), tetracycline (TC) and NO3 − adsorption by the biochar (BC), magnetic biochars (FB,
BF), activated carbon (AC) and magnetic activated carbon (AF) (T = 25 ◦ C, adsorbent dosage = 0.05 g).

Fig. S8 shows the FTIR patterns of the adsorbents. For all the there was electrostatic repulsion between the As(V) species
five adsorbents, visible bands appeared around 3400, 1580, and the negatively charged adsorbents at high pH. Although
1380 and 1080 cm−1 , corresponding to the vibration peaks of BC was generally negatively charged at pH 3–11, some of the
OH, C C, -COOH and C-O, respectively (Baig et al., 2014; He functional groups could be protonated and thus positively
et al., 2018; Tan et al., 2019a). The intensity of OH and C O charged. Thereby, BC had a relatively higher As(V) adsorption
peaks slightly increased for FB, BF and AF in comparison with at low pH. Similarly, the NO3 − adsorption also decreased with
BC and AC, respectively. In addition, the band around 580 cm−1 increasing solution pH for all the adsorbents (Fig. 1c), which
was assigned to Fe O vibration in the magnetic adsorbents was similar with the result of Usman et al. (2016). At low pH,
(Lim et al., 2009; Baig et al., 2014). These results indicated there were more positively charged sites on the surface of
that the functional groups on the surface of adsorbents were adsorbents due to the reduction of negative charges by the
altered after magnetization and the oxygen-containing groups excessive protons in solution (Usman et al., 2016).
were increased (i.e. OH and C O). Fig. S9 depicts the zeta As an amphoteric molecule, TC has three pKa values (3.3,
potentials of adsorbents. The zeta potentials of all magnetic 7.7 and 9.7). The predominant species of TC will be H4 TC+
adsorbents obviously shifted toward higher pH in comparison at pH < 3.3, H3 TC at 3.3 < pH < 7.7, H2 TC− at 7.7 < pH < 9.7, and
with BC and AC, indicating the increase of positive charges at HTC2− at pH > 9.0, respectively (Liu et al., 2012; Zhu et al.,
the surface of FB, BF and AF. Therefore, these magnetic adsor- 2014a). Fig. 1b presents the effect of pH on TC adsorption.
bents could have higher adsorption capacity for the anionic Similar to the findings of Zhao et al. (2011) and Lian et al.
pollutants due to the electrostatic interaction. (2013), the adsorption of TC on all the adsorbents gradu-
ally increased in this study with the increase of pH, and
3.2. Effects of solution pH on As(V), TC and NO3 − reached the maximum level at pH 7, except for BC. For FB, BF,
adsorption AC and AF, electrostatic repulsion would occur between the
adsorbents and cationic TC at acidic pH. With the increase
The amounts of As(V) retained by the adsorbents at differ- of pH, TC molecules gradually became neutral or negatively
ent pH values are shown in Fig. 1a. The As(V) adsorption charged through deprotonation, resulting in higher adsorption
decreased with increasing solution pH for all the adsorbents affinity with the adsorbents through electrostatic attractions.
especially at pH > 7, which was consistent with the results of However, when pH further increased, the adsorbents became
Baig et al. (2014) and Zhou et al. (2017). According to He et al. negatively charged, resulting in electrostatic repulsion with
(2018), the predominant AS(V) species were H2 AsO4 − at pH < 7 anionic TC and therefore the decline in adsorption. Regarding
and HAsO4 2− at pH > 7. The zeta potentials of adsorbents BC, the adsorption of TC slightly decreased with the increase
gradually changed from positive at low pH to negative at in solution pH from 3 to 11, attributed to the negatively charged
high pH, except for BC. As a result, these adsorbents could surface of BC that exhibited higher adsorption affinity with
adsorb negatively charged As(V) species at low pH, while cationic TC at low pH.
586 Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 582–591

Fig. 2 – Adsorption kinetics and isotherms of As(V) with/without 50 mg L−1 TC or 30 mg L−1 NO3 − onto the biochar (BC),
magnetic biochars (FB, BF), activated carbon (AC) and magnetic activated carbon (AF) (solid lines represent Langmuir model
and dashed lines represent Freundlich model in adsorption isotherm, T = 25 ◦ C, adsorbent dosage = 0.05 g).

Table 2 – Kinetic parameters for As (V), TC and NO3 − adsorption onto the biochar (BC), magnetic biochars (FB, BF),
activated carbon (AC) and magnetic activated carbon (AF).
Model Parameters 30 mg L−1 As (V) 400 mg L−1 TC 30 mg L−1 NO3 −

BC FB BF AC AF BC FB BF AC AF BC FB BF AC AF

Pseudo first qe (mg g−1 ) 0.797 5.790 3.225 1.054 1.974 12.038 42.519 37.054 535.82 192.26 0.567 5.285 2.642 0.845 1.131
order
k1 (h−1 ) 0.022 0.256 0.156 0.105 0.132 0.011 0.015 0.011 0.00238 0.0031 0.040 0.092 0.016 0.043 0.070
R2 0.623 0.877 0.888 0.624 0.924 0.952 0.901 0.982 0.984 0.981 0.753 0.968 0.749 0.776 0.729
Pseudo qe (mg g−1 ) 0.822 5.901 4.371 1.332 3.586 12.024 41.165 35.745 546.913 189.815 0.59 5.469 2.795 0.897 1.189
second order
k2 (g mg−1 h−1 ) 0.030 0.066 0.040 0.070 0.034 0.0014 5 × 10−4 6 × 10−4 6 × 10−6 2 × 10−5 0.12 0.031 0.011 0.079 0.095
R2 0.803 0.990 0.983 0.855 0.991 0.908 0.885 0.853 0.926 0.902 0.915 0.968 0.839 0.953 0.939

Table 3 – Adsorption isotherm parameters for As (V) in single solute or co-solute system with tetracycline (50 mg L−1 ) or
NO3 − (30 mg L−1 ) onto the biochar (BC), magnetic biochars (FB, BF), activated carbon (AC) and magnetic activated carbon
(AF).
Model Parameters Single-solute Co-solute (with 50 mg L−1 TC) Co-solute (with 30 mg L−1 NO3 − )

BC FB BF AC AF BC FB BF AC AF BC FB BF AC AF

Langmuir Qm (mg g−1 ) 1.14 6.77 4.20 1.59 2.70 0.94 5.76 3.44 1.54 2.57 0.97 6.43 4.01 1.33 2.52
KL (L mg−1 ) 0.057 0.120 0.075 0.045 0.067 0.045 0.114 0.052 0.040 0.052 0.035 0.098 0.056 0.048 0.045
R2 0.987 0.986 0.963 0.988 0.975 0.965 0.989 0.958 0.990 0.985 0.969 0.986 0.959 0.978 0.990
Freundlich KF (mg g−1 )(mg L−1 )−n 0.276 1.805 1.226 0.281 1.094 0.159 1.422 0.732 0.244 0.788 0.117 1.474 0.931 0.223 0.659
1/n 0.497 0.409 0.441 0.529 0.460 0.542 0.409 0.497 0.546 0.519 0.599 0.436 0.479 0.528 0.537
R2 0.952 0.981 0.957 0.953 0.972 0.936 0.939 0.972 0.964 0.946 0.946 0.967 0.972 0.916 0.966

3.3. Adsorption kinetics and capacity As(V) was adsorbed through monolayer adsorption onto the
surface of the adsorbent. According to the Langmuir model,
3.3.1. As(V) removal in a monocomponent solution the maximum adsorption capacities of FB (6.77 mg g−1 ) and BF
Fig. 2a presents the adsorption kinetics of As(V), and the (4.20 mg g−1 ) increased by 4.94 and 2.68 times in comparison
parameters obtained from pseudo-first-order and pseudo- with BC (1.14 mg g−1 ), respectively. Similarly, the adsorption
second-order models are listed in Table 2. Rapid adsorption capacity of AF (2.70 mg g−1 ) increased by 70% in compari-
of As(V) was observed within the first 3 h, especially on son with AC (1.59 mg g−1 ). These results indicated that ferric
magnetic adsorbents, and then quick approaching to the chloride magnetization was an effective way to improve the
equilibrium. For all the adsorbents, the R2 values of the removal of aqueous As(V). Besides, the maximum adsorption
pseudo-first-order model (0.623–0.924) were lower than those capacity of FB was much higher than that of many previously
of the pseudo-second-order model (0.803–0.991). The results reported Fe-impregnated adsorbents (Table S1), suggesting
suggested that the number of active sites on the surface of that the produced magnetic biochar can be an efficient adsor-
adsorbents played a key role in the reaction rate and chemi- bent to remove aqueous As(V).
cal adsorption was the rate-limiting step (Zhou et al., 2017).
Fig. 2b shows the adsorption isotherms of As(V). Table 3 3.3.2. TC removal in a monocomponent solution
gives the correlation coefficients and adsorption constants Fig. 3a presents the effect of contact time on TC adsorption by
determined from Langmuir and Freundlich isotherms. The various adsorbents. Table 2 presents the parameters obtained
Langmuir model described the adsorption of As(V) on the from pseudo-first-order and pseudo-second-order models.
adsorbents better than the Freundlich model, suggesting that Unlike As(V), it took nearly 7 d for the TC adsorption to reach
Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 582–591 587

Fig. 3 – Adsorption kinetics and isotherms of tetracycline (TC) with/without 30 mg L−1 As(V) or 30 mg L−1 NO3 − onto the
biochar (BC), magnetic biochars (FB, BF), activated carbon (AC) and magnetic activated carbon (AF) (solid lines represent
Langmuir model and dashed lines represent Freundlich model in adsorption isotherm, T = 25 ◦ C, adsorbent dosage = 0.05 g).

Table 4 – Adsorption isotherm parameters for TC in single solute or co-solute system with As (V) (30 mg L−1 ) or NO3 −
(30 mg L−1 ) onto the biochar (BC), magnetic biochars (FB, BF), activated carbon (AC) and magnetic activated carbon (AF).
Model Parameters Single-solute Co-solute (with 30 mg L−1 As (V)) Co-solute (with 30 mg L−1 NO3 − )

BC FB BF AC AF BC FB BF AC AF BC FB BF AC AF

Langmuir Qm (mg g−1 ) 12.24 49.57 44.07 612.38 211.34 8.77 42.59 34.15 562.46 191.32 9.64 45.55 38.39 578.49 196.48
KL (L mg −1 ) 0.066 0.037 0.019 0.076 0.047 0.024 0.026 0.017 0.066 0.037 0.044 0.025 0.017 0.066 0.039
R2 0.976 0.963 0.968 0.935 0.935 0.926 0.958 0.978 0.978 0.953 0.952 0.959 0.951 0.977 0.958
Freundlich KF (mg g−1 )(mg L−1 )−n 4.276 8.674 5.032 131.39 46.63 1.122 5.882 3.231 95.01 35.846 2.007 6.364 3.689 95.04 37.15
1/n 0.180 0.269 0.359 0.311 0.265 0.303 0.326 0.381 0.355 0.291 0.216 0.305 0.379 0.358 0.285
R2 0.963 0.987 0.967 0.938 0.942 0.968 0.978 0.930 0.932 0.909 0.966 0.959 0.952 0.939 0.916

equilibrium. Fig. 3b presents the adsorption isotherms of TC, magnetization was an efficient way to improve NO3 − adsorp-
and the Langmuir and Freundlich model fittings. The best-fit tion as well. Additionally, the maximum adsorption capacity
model parameters can be found in Table 4. Both isotherms of FB was much higher than that of many other adsorbents
can well describe the equilibrium data, indicating that the (Table S1). Taking into account the high NO3 − adsorption
TC adsorption on the adsorbents included not only mono- capacity and facile preparation method, the magnetic biochar
layer physical adsorption process but also some chemical can be an excellent alternative to remove aqueous NO3 − .
interactions (Liu et al., 2012). According to the Langmuir
model (Table 4), the maximum adsorption capacities of FB 3.4. Competitive adsorption
(49.56 mg g−1 ) and BF (44.07 mg g−1 ) increased by 3.05 and
2.60 times in comparison with BC (12.24 mg g−1 ). However, As shown in Figs. 2–4, the adsorption isotherms of As(V),
the adsorption capacity of AF (211.34 mg g−1 ) decreased by TC and NO3 − in the co-solute systems were shifted down-
1.90 times in comparison with AC. The adsorption capacity ward compared to their corresponding isotherms in the single
of AC (612.38 mg g−1 ) was comparable and even higher than solute systems, indicating the presence of competitive adsorp-
that of many other similar adsorbents (Table S1). Neverthe- tion between different anionic pollutants. Under the condition
less, the use of phosphoric acid (85%) should be taken into of 50 mg L−1 TC, the As(V) adsorption capacities decreased
consideration when employing the produced AC. by 17.54%, 14.92%, 18.09%, 3.14% and 4.81%, while the NO3 −
adsorption capacities decreased by 17.5%, 16.32%, 20.98%,
3.3.3. NO3 − removal in a monocomponent solution 7.81% and 10.19% for BC, FB, BF, AC and AF, respectively, com-
Fig. 4a presents the adsorption kinetics of NO3 − on the five pared to the single solute system. As shown in the graphical
adsorbents, and parameters fitted with the kinetic models are illustration of postulated adsorption mechanisms (Fig. 5), in
presented in Table 2. The apparent adsorption equilibrium was a co-solute system, the adsorbed TC on the surface of adsor-
reached in approximately 4 h, the pseudo-second-order model bents may block the access of As(V) and NO3 − to the surface
gave the best fitting. Similar to As(V), the result indicated because of its large molecular size, thus suppress the adsorp-
that chemical adsorption could be the rate-limiting step. The tion of As(V) and NO3 − . Besides, as a kind of bulky molecule, TC
adsorption isotherms data were evaluated using the Langmuir can induce a marked pore filling effect when being adsorbed
and Freundlich models (Fig. 4b), and the calculated parameters on carbonaceous material (Zhu et al., 2014b), thereby decrease
are listed in Table 5. The adsorption of NO3 − tended to increase the adsorption of As(V) and NO3 − .
with ascending NO3 − concentration until equilibrium was Under the condition of 30 mg L−1 NO3 − , the As(V) adsorp-
achieved. The Langmuir maximum NO3 − adsorption capacity tion capacities decreased by 14.91%, 5.02%, 4.52%, 16.35% and
on FB amounted to 6.31 mg g−1 , followed by BF (3.48 mg g−1 ), 6.67% for BC, FB, BF, AC and AF, respectively, compared to the
AF (2.06 mg g−1 ), AC (1.28 mg g−1 ) and BC (0.82 mg g−1 ). The single solute system. On the other hand, the NO3 − adsorption
order was similar to that for the As(V) adsorption, suggesting capacities were inhibited by 24.39%, 18.07%, 26.44%, 35.94%
that similar mechanism was involved and ferric chloride and 27.67% in the presence of 30 mg L−1 As(V) for BC, FB, BF,
588 Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 582–591

Fig. 4 – Adsorption kinetics and isotherms of NO3 − with/without 30 mg L−1 As(V) or 50 mg L−1 tetracycline (TC) onto the
biochar (BC), magnetic biochars (FB, BF), activated carbon (AC) and magnetic activated carbon (AF) (solid lines represent
Langmuir model and dashed lines represent Freundlich model in adsorption isotherm, T = 25 ◦ C, adsorbent dosage = 0.05 g).

Table 5 – Adsorption isotherm parameters for NO3 − in single solute or co-solute system with As (V) (30 mg L−1 ) or TC
(50 mg L−1 ) onto the biochar (BC), magnetic biochars (FB, BF), activated carbon (AC) and magnetic activated carbon (AF).
Model Parameters Single-solute Co-solute (with 30 mg L−1 As (V)) Co-solute (with 50 mg L−1 TC)

BC FB BF AC AF BC FB BF AC AF BC FB BF AC AF
−1
Langmuir Qm (mg g ) 0.82 6.31 3.48 1.28 2.06 0.62 5.17 2.56 0.82 1.49 0.66 5.28 2.75 1.18 1.85
KL (L mg−1 ) 0.125 0.502 0.224 0.098 0.102 0.070 0.283 0.163 0.102 0.074 0.073 0.338 0.198 0.097 0.066
R2 0.997 0.946 0.970 0.994 0.973 0.975 0.916 0.973 0.972 0.991 0.930 0.954 0.986 0.977 0.963
Freundlich KF (mg g−1 )(mg L−1 )−n 0.156 2.272 0.927 0.208 0.354 0.075 1.537 0.576 0.138 0.195 0.082 1.816 0.933 0.184 0.238
1/n 0.398 0.291 0.338 0.433 0.421 0.487 0.336 0.377 0.423 0.471 0.480 0.312 0.280 0.434 0.488
R2 0.946 0.962 0.970 0.936 0.976 0.912 0.975 0.963 0.911 0.984 0.842 0.944 0.929 0.890 0.932

Fig. 5 – Graphical illustration of postulated mechanisms for the competitive adsorption of As(V), tetracycline (TC) and NO3 −
on magnetic adsorbent (e.g. FB) surface.

AC and AF, respectively. Compared to others, FB showed a Under the condition of 30 mg L−1 As(V), the TC adsorp-
much smaller decrease in the adsorption capacities for As(V) tion capacities decreased by 28.43%, 14.06%, 22.51%, 8.15% and
and NO3 − in the co-solute system, indicating its great poten- 9.47% for BC, FB, BF, AC and AF, respectively, compared to the
tial for application in the co-contaminated environment with single solute system. Similarly, the TC adsorption capacities
heavy metals and inorganic anions. Besides, NO3 − imposed decreased by 21.24%, 8.09%, 12.89%, 5.53% and 7.03% for BC, FB,
less impact on the As(V) removal, especially for FB, BF and AF. BF, AC and AF, respectively, when 30 mg L−1 NO3 − was spiked.
This could be attributed to the stronger electrostatic attrac- The presence of either As(V) or NO3 − inhibited the adsorption
tion of As(V) with Fe-adsorbent than that of NO3 − (An et al., of TC in a co-solute system compared to the single system. In
2011; He et al., 2018). For the same reason, TC exerted stronger addition, As(V) exerted stronger negative effect on TC adsorp-
inhibitory effect on NO3 − adsorption than on As(V) adsorption. tion than NO3 − . This could be firstly due to the formation of
Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 582–591 589

strong inner-sphere complexes between As(V) and the func- TC into the inner pore network (Table 1). Some other factors
tional groups on the adsorbent surface, which results in the can also influence TC adsorption such as electrostatic attrac-
blockage to the pores. Secondly, the electrostatic interaction tion and functional groups. Compared to BC, there were more
of As(V) with the adsorption sites of adsorbent was much O-containing functional groups on the surface of FB, BF and
stronger than that of NO3 − , because multivalent anion with AF after magnetization (Fig. S8). The functional groups can
higher charge density was more readily adsorbed in compari- facilitate the formation of hydrogen bonds with TC molecules,
son with monovalent anion (Divband Hafshejani et al., 2016). therefore improve the adsorption of TC (Jing et al., 2014). The
Thirdly, As(V) of larger aqueous ionic radius induced stronger surficial Fe may also play a part in TC adsorption through
steric repulsion than NO3 − , and thus stronger inhibitory effect surface reactions and ion exchange (e.g. cation bridging and
on the adsorption of TC (He et al., 2018). It is noteworthy that complexation) (MacKay and Vasudevan, 2012; Lian et al., 2013).
AC and AF showed a much smaller decrease in the adsorption Additionally, both biochar and activated carbon possessed
capacity for TC in the co-solute system. Due to its relatively short stacks of small graphite-like sheets arranged in disor-
higher adsorption capacities for As(V) and NO3 − , AF could be dered form (Liu et al., 2012). Due to the electron-withdrawing
a promising adsorbent to remediate the aquatic environment ability of ketone groups, the TC molecules could act as
with multiple contaminants. effective ␲-electron-acceptors. While the adsorbents could be
␲-electron-donors because of the polarized ␲-electron-rich
3.5. Adsorption mechanisms graphite surface. Therefore, the ␲–␲ electron donor-acceptor
(EDA) interaction could be formed between the TC molecules
It has been reported that the adsorption of As(V) was primarily and the adsorbents (Fig. 5) (Zhu et al., 2014a). Van der Waals
influenced by two factors: the As(V) speciation and the surface forces can also exist between the graphite surface of adsor-
charge of the adsorbent (Zhang et al., 2013; Wang et al., 2014). bents and TC (Zhao et al., 2011; Liu et al., 2012). On the other
As the adsorption kinetics and isotherms were conducted at hand, according to the FTIR spectra (Fig. S8), several func-
pH 6, As(V) mainly existed as HAsO4 2− under this circum- tional groups (i.e. C O and Fe O) showed variations among
stance, and generally FB, BF, AC and AF were all positively post-adsorption samples, suggesting that As(V), TC and NO3 −
charged. Besides, some functional groups on the surface of would compete for finite adsorption sites in a co-solute system
biochar were positively charged because they could be pro- (Fig. 5). In addition, they will repulse each other in a co-solute
tonated at pH 6. As a result, electrostatic attractions were system as anions. Therefore, their adsorption affinity to the
present between all the five adsorbents and As(V). In partic- adsorbent would be significantly inhibited.
ular, the positive iron oxide particles on FB, BF and AF could
act as efficient adsorption sites for As(V) in aqueous solution. 4. Conclusions
The adsorption capacity for As(V) increased with ascending
iron ratio among FB, BF and AF, suggesting that the electro- Magnetic biochar synthesized from one-step method showed
static anionic attraction on the H2 + –O–Fe-adsorbent played a the highest As(V) and NO3 − adsorption capacities, which were
key role in As(V) adsorption (Usman et al., 2016). Neither the 5.94 and 7.69 times that of the pristine biochar, respectively.
XRD nor the XPS analyses of Fe2p showed obvious change for Competitive adsorption among As(V), TC and NO3 − occurred
the post-adsorption adsorbents compared to the original ones on all adsorbents as they could all be anions. One-step mag-
(Figs. S3 and S6). These results demonstrated that As(V) in netization of biochar was an efficient way to improve the
the initial solution was only loaded onto the surface of the adsorption capacities of As(V) and NO3 − , showing a great
adsorbents, and electrostatic attractions could be the main potential for application in the co-contaminated environment
mechanism for the As(V) removal. Similar to As(V), FB showed with heavy metals and inorganic anions. Activated carbon
the highest adsorption capacity for NO3 − and followed by revealed the highest adsorption capacity for TC which was
BF and AF, indicating that the elevated amount of iron can decreased after magnetization. However, magnetic activated
adsorb more negatively charged NO3 − through electrostatic carbon could still be a promising adsorbent to remediate the
attraction. In addition, as shown in Fig. S9 (pHpzc ), there were aquatic environment with multiple contaminants.
more surface positive charges on FB, followed by BF and AF,
which could provide more surface functional groups and con-
Conflict of interest
sequently a higher affinity for NO3 − . AC exhibited higher NO3 −
adsorption, 1.39 times that of BC. This could be due to the
None declared.
surface positive charges as well as high specific surface area.
Unlike As(V) and NO3 − , TC was primarily present as
Acknowledgements
zwitterion at pH 6, therefore hydrophobic effect could play
an important role in TC removal (Lian et al., 2013). Gen-
The study was supported by the National Natural Science
erally, as the specific surface area decreased in the order
Foundation of China (41301311, 51579003), and the Shenzhen
of AC > AF > FB > BC > BF (Table 1), the maximum adsorption
Municipal Development and Reform Commission (Shenzhen
capacities of these adsorbents for TC also decreased except
Engineering Research Center for Nanoporous Water Treat-
for BF and BC (Fig. 3). Larger surface areas can provide a
ment Materials, and Discipline construction of watershed
larger number of active adsorption sites for TC molecules
ecological engineering).
(Zhu et al., 2014b). Besides, the more porous and aromatic
structure of AC was more beneficial for TC removal, which
favored both specific interaction and hydrophobic effect (Zhu
Appendix A. Supplementary data
et al., 2014b). Although the specific surface area of BF was
slightly lower than that of BC, the adsorption capacity of BF Supplementary material related to this article can be
for TC was much higher. One possible reason could be the found, in the online version, at https://doi.org/10.1016/
large average pore size in BF which favored the diffusion of j.cherd.2020.05.011.
590 Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 582–591

References biomass and its application in removal of tetracycline from


aqueous solution. Bioresour. Technol. 121, 235–240.
MacKay, A.A., Vasudevan, D., 2012. Polyfunctional ionogenic
An, B., Liang, Q., Zhao, D., 2011. Removal of arsenic(V) from spent
compound sorption: challenges and new approaches to
ion exchange brine using a new class of starch-bridged
advance predictive models. Environ. Sci. Technol. 46,
magnetite nanoparticles. Water Res. 45, 1961–1972.
9209–9223.
Baig, S.A., Zhu, J., Muhammad, N., Sheng, T.T., Xu, X.H., 2014.
Mangwandi, C., Suhaimi, S.N.A., Liu, J.T., Dhenge, R.M., Albadarin,
Effect of synthesis methods on magnetic Kans grass biochar
A.B., 2016. Design, production and characterisation of
for enhanced As(III, V) adsorption from aqueous solutions.
granular adsorbent material for arsenic removal from
Biomass Bioenergy 71, 299–310.
contaminated wastewater. Chem. Eng. Res. Des. 110, 70–81.
Chen, J., Wu, H., Qian, H., 2016. Groundwater nitrate
Peng, L., Ren, Y., Gu, J., Qin, P., Zeng, Q., Shao, J., Lei, M., Chai,
contamination and associated health risk for the rural
L.J.E.S., Research, P., 2014. Iron improving bio-char derived
communities in an agricultural area of Ningxia, Northwest
from microalgae on removal of tetracycline from aqueous
China. Exposure Health 8, 349–359.
system. Environ. Sci. Pollut. Res. 21, 7631–7640.
Divband Hafshejani, L., Hooshmand, A., Naseri, A.A.,
Song, X., Zhang, Y., Luo, X., Chen, P., Liu, J., 2019. 2D magnetic
Mohammadi, A.S., Abbasi, F., Bhatnagar, A., 2016. Removal of
scallion sheathing-based biochar composites design and
nitrate from aqueous solution by modified sugarcane bagasse
application for effective removal of arsenite in aqueous
biochar. Ecol. Eng. 95, 101–111.
solutions. Chem. Eng. Res. Des. 152, 384–392.
Fujino, Y., Guo, X., Liu, J., You, L., Miyatake, M., Yoshimura, T.,
Tan, G., Mao, Y., Wang, H., Junaid, M., Xu, N., 2019a. Comparison
Japan Inner Mongolia Arsenic Pollution Study G, 2004. Mental
of biochar-and activated carbon-supported zerovalent iron for
health burden amongst inhabitants of an arsenic-affected
the removal of Se(IV) and Se(VI): influence of pH, ionic
area in Inner Mongolia, China. Soc. Sci. Med. 59, 1969–1973.
strength, and natural organic matter. Environ. Sci. Pollut. Res.
Fytianos, K., Christophoridis, C., 2004. Nitrate, arsenic and
26 (21), 21609–21618.
chloride pollution of drinking water in Northern Greece.
Tan, G., Sun, W., Xu, Y., Wang, H., Xu, N., 2016a. Sorption of
Elaboration by applying GIS. Environ. Monit. Assess. 93, 55–67.
mercury (II) and atrazine by biochar, modified biochars and
Guo, X., Fujino, Y., Kaneko, S., Wu, K., Xia, Y., Yoshimura, T., 2001.
biochar based activated carbon in aqueous solution.
Arsenic Contamination of Groundwater and Prevalence of
Bioresour. Technol. 211, 727–735.
Arsenical Dermatosis in the Hetao Plain Area, Inner Mongolia,
Tan, G., Wang, H., Xu, N., Junaid, M., Liu, H., Zhai, L., 2019b.
China, Molecular Mechanisms of Metal Toxicity and
Effects of biochar application with fertilizer on soil microbial
Carcinogenesis. Springer, pp. 137–140.
biomass and greenhouse gas emissions in a peanut cropping
He, R., Peng, Z., Lyu, H., Huang, H., Nan, Q., Tang, J., 2018.
system. Environ. Technol., 1–11.
Synthesis and characterization of an iron-impregnated
Tan, X.F., Liu, Y.G., Gu, Y.L., Xu, Y., Zeng, G.M., Hu, X.J., Liu, S.B.,
biochar for aqueous arsenic removal. Sci. Total Environ. 612,
Wang, X., Liu, S.M., Li, J., 2016b. Biochar-based
1177–1186.
nano-composites for the decontamination of wastewater: a
Hu, X., Zhou, Q., Luo, Y., 2010. Occurrence and source analysis of
review. Bioresour. Technol. 212, 318–333.
typical veterinary antibiotics in manure, soil, vegetables and
Tofighy, M.A., Mohammadi, T.J.C.E.R., Design, 2012. Nitrate
groundwater from organic vegetable bases, northern China.
removal from water using functionalized carbon nanotube
Environ. Pollut. 158, 2992–2998.
sheets. Chem. Eng. Res. Des. 90, 1815–1822.
Islam, M.A., Ali, I., Karim, S.M.A., Hossain Firoz, M.S., Chowdhury,
Usman, A.R., Ahmad, M., El-Mahrouky, M., Al-Omran, A., Ok, Y.S.,
A.-N., Morton, D.W., Angove, M.J., 2019. Removal of dye from
Sallam, A., El-Naggar, A.H., Al-Wabel, M.I., 2016. Chemically
polluted water using novel nano manganese oxide-based
modified biochar produced from conocarpus waste increases
materials. J. Water Process. Eng. 32, 100911.
NO3 − removal from aqueous solutions. Environ. Geochem.
Islam, M.A., Morton, D.W., Johnson, B.B., Mainali, B., Angove, M.J.,
Health 38, 511–521.
2018a. Manganese oxides and their application to metal ion
Wang, B., Wang, Z., Jiang, Y., Tan, G., Xu, N., Xu, Y., 2017.
and contaminant removal from wastewater. J. Water Process.
Enhanced power generation and wastewater treatment in
Eng 26, 264–280.
sustainable biochar electrodes based bioelectrochemical
Islam, M.A., Morton, D.W., Johnson, B.B., Pramanik, B.K., Mainali,
system. Bioresour. Technol. 241, 841–848.
B., Angove, M.J., 2018b. Metal ion and contaminant sorption
Wang, H., Zhang, D., Mou, S., Song, W., Al-Misned, F.A., Mortuza,
onto aluminium oxide-based materials: a review and future
M.G., Pan, X., 2015. Simultaneous removal of tetracycline
research. J. Environ. Chem. Eng. 6, 6853–6869.
hydrochloride and As(III) using poorly-crystalline manganese
Jing, X.-R., Wang, Y.-Y., Liu, W.-J., Wang, Y.-K., Jiang, H., 2014.
dioxide. Chemosphere 136, 102–110.
Enhanced adsorption performance of tetracycline in aqueous
Wang, S., Gao, B., Zimmerman, A.R., Li, Y., Ma, L., Harris, W.G.,
solutions by methanol-modified biochar. Chem. Eng. J. 248,
Migliaccio, K.W., 2014. Removal of arsenic by magnetic biochar
168–174.
prepared from pinewood and natural hematite. Bioresour.
Kilpimaa, S., Runtti, H., Kangas, T., Lassi, U., Kuokkanen, T., 2014.
Technol. 175C, 391–395.
Removal of phosphate and nitrate over a modified carbon
Wu, P., Ata-Ul-Karim, S.T., Singh, B.P., Wang, H., Wu, T., Liu, C.,
residue from biomass gasification. Chem. Eng. Res. Des. 92,
Fang, G., Zhou, D., Wang, Y., Chen, W., 2019. A scientometric
1923–1933.
review of biochar research in the past 20 years (1998–2018).
Lehmann, J., Joseph, S., 2015. Biochar for Environmental
Biochar 1, 23–43.
Management: Science, Technology and Implementation.
Xia, S., Shen, S., Xu, X., Liang, J., Zhou, L., 2014. Arsenic removal
Routledge.
from groundwater by acclimated sludge under
Li, A., Pi, S., Wei, W., Chen, T., Yang, J., Ma, F., 2016. Adsorption
autohydrogenotrophic conditions. J. Environ. Sci. 26, 248–255.
behavior of tetracycline by extracellular polymeric substrates
Xu, N., Tan, G., Wang, H., Gai, X., 2016. Effect of biochar additions
extracted from Klebsiella sp. J1. Environ. Sci. Pollut. Res. 23,
to soil on nitrogen leaching, microbial biomass and bacterial
25084–25092.
community structure. Eur. J. Soil Biol. 74, 1–8.
Lian, F., Song, Z., Liu, Z., Zhu, L., Xing, B., 2013. Mechanistic
Xu, N., Zhang, B., Tan, G., Li, J., Wang, H., 2015. Influence of
understanding of tetracycline sorption on waste tire powder
biochar on sorption, leaching and dissipation of bisphenol A
and its chars as affected by Cu(2+) and pH. Environ. Pollut.
and 17␣-ethynylestradiol in soil. Environ. Sci. Process. Impact
178, 264–270.
17, 1722–1730.
Lim, S.-F., Zheng, Y.-M., Chen, J.P., 2009. Organic arsenic
Zhang, H., Ma, D.S., Hu, X.X., 2002. Arsenic pollution in
adsorption onto a magnetic sorbent. Langmuir 25, 4973–4978.
groundwater from Hetao Area, China. Environ. Geol. 41,
Liu, P., Liu, W.J., Jiang, H., Chen, J.J., Li, W.W., Yu, H.Q., 2012.
638–643.
Modification of bio-char derived from fast pyrolysis of
Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 582–591 591

Zhang, M., Gao, B., Varnoosfaderani, S., Hebard, A., Yao, Y., Inyang, magnetic gelatin-modified biochar. Chem. Eng. J. 314, 223–231.
M., 2013. Preparation and characterization of a novel magnetic Zhu, X., Liu, Y., Qian, F., Zhou, C., Zhang, S., Chen, J., 2014a.
biochar for arsenic removal. Bioresour. Technol. 130, 457–462. Preparation of magnetic porous carbon from waste hydrochar
Zhao, Y., Geng, J., Wang, X., Gu, X., Gao, S., 2011. Tetracycline by simultaneous activation and magnetization for
adsorption on kaolinite: pH, metal cations and humic acid tetracycline removal. Bioresour. Technol. 154, 209–214.
effects. Ecotoxicology 20, 1141–1147. Zhu, X., Liu, Y., Zhou, C., Luo, G., Zhang, S., Chen, J., 2014b. A
Zhou, Z., Liu, Y.-g., Liu, S.-b., Liu, H.-y., Zeng, G.-m., Tan, X.-f., novel porous carbon derived from hydrothermal carbon for
Yang, C.-p., Ding, Y., Yan, Z.-l., Cai, X.-x., 2017. Sorption efficient adsorption of tetracycline. Carbon 77, 627–636.
performance and mechanisms of arsenic(V) removal by

You might also like