You are on page 1of 14

Chemical Engineering Journal 405 (2021) 126895

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Sugarcane biochar as novel catalyst for highly efficient oxidative removal of T


organic compounds in water
Lingzhi He, Zairan Liu, Jiafeng Hu, Chuanyu Qin, Lan Yao, Yu Zhang, Yunxian Piao

Key Laboratory of Groundwater Resources and Environment (Jilin University), Ministry of Education, Jilin Provincial Key Laboratory of Water Resources and Environment,
Jilin University, Changchun 130021, China

HIGHLIGHTS

• AThenovel PMS based AOPs method was developed with SugBC900 as catalyst.
• HighSugBC900 is easy to prepare and highly cost effective.
• O and 2
sp carbon content, surface area and electron transfer ability are key features.
• It can effectively
1
O •−
are dominant oxidative species instead of SO and OH. •− •


2 2 4
degrade BPA at broad pHs (1–11) and temperatures (10–45 °C).

ARTICLE INFO ABSTRACT

Keywords: Cost effective and environmental friendly biochar catalyst (SugBC900) with high catalytic activity toward
Biochar peroxymonosulfate (PMS) was prepared by simple pyrolysis of sugarcane waste, and utilized for efficient de-
Peroxymonosulfate activation gradation of organic pollutants. Taking advantages of high sp2 hybrid carbon content, large surface area and
Bisphenol A excellent electron transfer ability, the SugBC900 coupled with PMS presented far greater catalytic degradation
Organic pollutant
ability toward bisphenol A (BPA) as model pollutant when compared to the SugBC700, SugBC800 and other
carbonaceous catalysts. The oxidative species of 1O2 and O2•− were demonstrated to be dominant for catalytic
degradation, and the role of SO4•− and •OH was minor. 20 mg/mL of BPA could be completely degraded with less
than 60 min at pHs of 5.0–11.0, and at temperatures of 25–45 °C. And at the extremely acidic (pH of 1.0) or low
temperature (10 °C) conditions, the degradation efficiencies could still reach more than 90%. The SugBC900 was
also successfully utilized for the efficiently catalytic degradation of BPA and other organic pollutants in various
real environmental water. Because of simple preparation by impregnating none of external elements, extremely
low cost property, and robustness characteristic by having broad working conditions, the SugBC900 would have
potential to be a good alternative to the conventional catalysts.

1. Introduction oxidation process (AOPs) and reduction are attractive owing to their
cost-effectiveness and robustness property. AOPs mainly uses catalyst to
Environmental contamination by organic pollutants such as phe- activate oxidant to produce active species with high oxidation ability to
nolic compounds and organic dyes can cause serious threat to human attack pollutant molecules and achieve mineralization [6]. The reduc-
heath and the ecosystem [1]. These organic matters are toxic and dif- tion reaction is the process of promoting the reducing agent to form the
ficult to be degraded by the traditional biological treatment processes in reducing structures and functional groups on the material surface, or
the wastewater treatment plants [2]. So it is of great significance and enhancing the electron transfer between the reducing agent and pol-
desired to explore highly efficient treating method to meet the re- lutant, and enable degradation of the pollutant [7]. Peroxymonosulfate
quirements. (PMS) based AOPs was known to be effective for degradation of organic
Up to now, various methods including adsorption [3], biodegrada- compounds [8]. PMS can be usually activated via UV light, heat, base,
tion [4], and catalytic degradation [5] have been used to treat such metal oxide or carbonaceous materials [9]. The sulfate radicals (SO4•−)
organic pollutants. Among them, catalytic methods including advanced and hydroxyl radicals (•OH) are the main active species in the activation


Corresponding author.
E-mail address: yxpiao@jlu.edu.cn (Y. Piao).

https://doi.org/10.1016/j.cej.2020.126895
Received 20 May 2020; Received in revised form 18 August 2020; Accepted 31 August 2020
Available online 06 September 2020
1385-8947/ © 2020 Elsevier B.V. All rights reserved.
L. He, et al. Chemical Engineering Journal 405 (2021) 126895

process of PMS [10–13], but singlet oxygen and superoxide radicals dried in an oven at 105 °C for 24 h prior to use. Bisphenol A (BPA),
(O2−•) can also be produced when metal and carbonaceous catalysts are Phenol, Humic acid, 5,5-Dimethyl-1-Pyrroline-N-Oxide (DMPO) and
used as catalysts [14–16]. The dominant oxidative species responsible 2,2,6,6-tetramethylpiperidine (TEMP) were purchased from Sigma-
for pollutant degradation is varied depending on the properties of Aldrich (Shanghai, China). Chromatographic purity of methanol and
catalysts and greatly affect the removal performances. Also, the pollu- acetonitrile were obtained from Fisher (USA). Peroxymonosulfate
tant removal performance is greatly affected by PMS activation ability (PMS), nitrotetrazolium blue chloride (NBT, ≥ 98.0 wt%), Rhodamine
of the catalysts [17]. The metallic catalysts (such as Cu, Fe and Co) B dye (RhB) and Methyl orange (MO) were purchased from Aladdin
having excellent PMS activation ability are considered to be highly Chemistry Co., Ltd. Multi-walled carbon nanotubes (MWCNT, 100 m2/
effective for pollutant removal, but the limitation of working conditions g) and graphene oxide (GO, 42 m2/g) were obtained from Nanjing
and the shortcomings of easily causing secondary pollution had hin- Xianfeng Nano Material Technology Co., Ltd. Other chemicals were
dered their application in practice [18–20]. The carbonaceous materials purchased from Sinopharm Chemical Reagent Co., Ltd. Unless other-
including activated carbon, graphene, carbon nanotube and their var- wise stated, chemical reagents are of analytical grade and used without
iously modified forms are promising alternatives to the conventional further purification.
metallic catalysts for their high catalytic activity and eco-friendliness The X-ray diffraction (XRD) patterns were obtained on a D8-
[21]. Advanecd X-ray diffractometer (Brucker, Germany) with Cu Kα radia-
Recently, biochar as newly emerging carbonaceous material might tion (λ = 1.5418 Å) over the 2θ range of 10 − 80°. The surface
be attractive for use as catalyst due to the unique advantages of ex- functional groups of different biochar samples were determined by
tremely low cost and environmental friendly nature. Biochar usually Fourier transform infrared spectroscopy (FTIR) (Thermo Fisher, USA).
can be obtained by simple pyrolysis of waste biomass in anaerobic The size property of biochar samples was investigated by a laser dif-
condition. It has the features of large surface area and abundant func- fraction particle size analyzer (Bettersize 2000, China). Zeta potential is
tional groups, and till now, it was widely applied for soil adaptation and recorded on Zeta potentiometer (Nano Z, Shangguang Microscope
adsorption of pollutants [22–25], and there are also a few reports on its Factory, Shanghai, China) by dispersing samples in ultrapure water
usage as fuel cell and sensor [26–28]. Lately, the applications of bio- with different pH values. The scanning electron microscopy (SEM)
chars as the catalysts for activating sulfate to degrade organic com- images were obtained by a scanning electron microscope (Zeiss EVO
pounds were intensively studied [9,29]. The modification could en- MA10, Germany). Raman spectra were recorded on a dispersive Raman
hance the catalytic activation ability of the biochar compared to the spectrometer (Tokyo Instruments Inc. Nanofinder, Japan) with an
pristine one, but it is complicated and may increase the expense in the argon ion laser (514 nm). Transmission electron microscopy images
preparation process, also the working conditions tend to be limited by (TEM) were recorded on a transmission electron microscope (JEOL Ltd.,
leaching of the modified components, and sometimes causing sec- Tokyo, Japan). A nitrogen adsorption–desorption apparatus (ASAP
ondary pollution. 2020 HD88, USA) was used to determine the surface area and pore size
From the point of cost effectiveness, simplicity and stability, the distribution. The elemental analyser (Elementar, Germany) was used to
development of pristine biochar with high catalytic oxidation ability estimate the elemental composition. X-ray photoelectron spectroscopy
without introducing external reagents was greatly desired, but such (XPS) was conducted using an X-ray photoelectron spectrometer
research was not much accessed in detail. The source of biochar is (Thermo Fisher, USA). The electron paramagnetic resonance (EPR)
ubiquitous and the physicochemical properties of biochar are varied spectra were obtained using an EPR spectroscopy (Bruker Biospin
correspondingly depending on the sources [30–32]. Also the biochar ELEXSYS II E500) in continuous wave X-band mode. The by-products
may have the diverse physicochemical characteristics such as the sur- from the catalytic reaction were identified using the Gas
face area, carbonization, and electron transfer rate via changing of Chromatography and Mass Spectrometry (GC–MS) (Agilent, USA).
preparation methods, as reported previously [3,25,33,34]. In this sense, Total organic carbon (TOC) was analyzed by a TOC Analyzer
the high catalytic oxidation property of biochar could be anticipated by (Shimadzu, TOC-5000).
proper control of the synthesis process. And the application of pristine
biochar directly for catalytic oxidation degradation of pollutant with 2.2. Synthesis of catalysts
comparable or improved quality to the conventional catalysts are
challenging and valuable for robust and low cost treatment of en- Three types of biochars (BCs) were obtained by pyrolysis of su-
vironmental water. garcane waste under different pyrolysis conditions in a tubular furnace
In this research, the biochar catalyst with high catalytic activation (OTF-1200X, Hefeikejing materials technology Co., Ltd. China) with
ability toward PMS was obtained via simple pyrolysis of the sugarcane continuous injection of nitrogen gas. The pyrolysis temperatures were
waste and utilized for efficient degradation of organic pollutant in set to 700, 800, and 900 °C, individually, and the heating rate and
water. By modulation of the pyrolysis temperature, three kinds of the maintenance time were 7 °C/min and 1 h, respectively. The produced
sugarcane derived biochars (SugBC) were prepared, and their catalytic materials were naturally cooled to room temperature under nitrogen
degradation abilities toward bisphenol A (BPA) as model pollutant were protection, and then they were grinded with a grinder for 4 h to pro-
investigated. The SugBC obtained at 900 °C (SugBC900) presented duce fine particle sized SugBCs. The SugBCs obtained at pyrolysis
highly efficient BPA degradation ability. The degradation mechanism, temperature of 700, 800, and 900 °C were denoted as SugBC700,
and the broad working capabilities at various pHs and temperature SugBC800, and SugBC900, respectively. Biochars derived from corn
conditions were pursued. Taking unique advantages of the sugarcane husks (CHBC900) and pine needles (PNBC900) were obtained by pyr-
biochar, the potential usability of SugBC900 for various organic pol- olysis at 900 °C using the same method.
lutants degradation in real environmental water in alternative to the The cobalt oxide (Co3O4) powders were prepared by thermal de-
conventional carbonaceous and metal catalysts, and other biochars was composition of Co(NO3)2 at 500 °C for 2 h [35], and the Fe3O4 magnetic
also demonstrated. particles were obtained by coprecipitation method according to the
previous reports [36]. Magnetic biochar (Mag-SugBC900) was prepared
2. Experimental according to our previous study [25].

2.1. Materials and apparatus 2.3. Catalytic removal of organic pollutants

The feedstock of sugarcane waste was obtained from the local For degradation of BPA, 50 mg of SugBCs were suspended in a
market, washed by deionized (DI) water to remove the impurities, then 100 mL of BPA (20 mg/L in ultrapure water) solution, and then 100 mg

2
L. He, et al. Chemical Engineering Journal 405 (2021) 126895

Fig. 1. (A) SEM, (B) Size distribution and (C) TEM images of (a) SugBC700, (b) SugBC800 and (c) SugBC900, respectively.

of PMS was added to initiate the reaction. The reacting solution was 280 nm, and the mobile phase was consisted of methanol and water
incubated in a full-temperature air bath shaking incubator with shaking (65:35, v/v) at a flow rate of 1 mL/min. For phenol, the detection
speed of 200 rpm. wavelength was set at 270 nm, and the mobile phase was composed of
At specific time intervals, 0.5 mL of the reaction solution was acetonitrile and water (50:50, v/v) at a flow rate of 1.0 mL/min.
sampled and the reaction was stopped by immediately addition of The RhB and MO concentrations were determined using an UV–vis
0.5 mL of methanol as the quenching agent. And then it was filtered spectrophotometer (UV-2450, Shimadzu, Japan), the detection wave-
using a 0.45 μm cut-off sized filter membrane for the following quan- lengths were set to 464 nm and 554 nm, respectively.
titative analysis. The PMS concentration during the catalytic reaction was de-
To reuse the catalyst, the catalyst was collected by centrifugation of termined by an ABTS method according to previous study [37]. The
the reacting solution at 10000 rpm for 3 min, followed by thorough DMPO was used as spin-trapping agents for detecting SO4•−, •OH, and,
washing with deionized water and drying in a vacuum oven. O2•−, and TEMP for detecting 1O2.
For real environmental water treatment, the tap water (TW) from The reaction rate was fitted with a pseudo first-order-kinetic model
our laboratory, lake water (LW) from the vicinity of the South Lake (Eq. (1)).
(Changchun, China) and municipal wastewater (MW) from the sewage
ln(C/C0) = -kt (1)
treatment plant (Songyuan, China) without any pretreatment were
spiked with 20 mg/L of BPA, 20 mg/L of phenol, 100 mg/L of MO and where C0, C and k correspond to the initial concentration of BPA, the
10 mg/L of RhB, individually. And they were treated by the SugBC900 concentration of BPA at certain times, and reaction constant, respec-
and PMS. All experiments were performed in triplicate to ensure the tively.
accuracy.
2.5. Electrochemical measurements
2.4. Analytical methods
The electrochemical property of the SugBCs was investigated with
The BPA and phenol concentrations were determined using the high an electrochemical workstation (Shanghai Chenhua Instrument Co.,
performance liquid chromatography system equipped with UV–vis de- Ltd, China) from CH Instruments. For the preparation of working
tector and a Zorbax SB C18 column (150 × 4.6 mm, particle size of electrodes, 2 mg of the SugBCs was dispersed in 1 mL of ultrapure water
5 μm, Agilent, USA). For BPA, the detection wavelength was set at containing 0.5 μL of HCl (1 M), and then ultrasonically dispersed for

3
L. He, et al. Chemical Engineering Journal 405 (2021) 126895

1 min to form a suspension, then 6 μL of it was deposited on a well TEM analysis. As illustrated in Fig. 2B and Table 1, the average pore
polished glassy carbon electrode with a diameter of 3 mm, and in- size of SugBC900 (1.764 nm) was smallest compared to the SugBC700
cubated for 40 min under an ambient condition [33]. Ag/AgCl elec- (7.485 nm) and SugBC800, which was due to a relatively higher pyr-
trode and a platinum wire were used as reference and counter elec- olysis temperature could produce more narrow mouthed pores [39,40].
trodes, respectively. The linear sweep voltammetry (LSV) was However, the surface area and pore volume of SugBC900 (602.52 m2
conducted at the potentials of −1.0–1.0 V with the scan rate of 50 mV/ g−1, 0.274 cm3 g−1) were significantly larger than that of SugBC700
s, and the chronoamperometry was carried out at an applied potential (26.53 m2 g−1, 0.050 cm3 g−1) and SugBC800 (34.38 m2 g−1,
of 0.0 V by using 50 mM Na2SO4 as the electrolyte. The cyclic vol- 0.056 cm3 g−1). This is related to the volatilization of more gas mole-
tammetry (CV) and electrochemical impedance spectroscopy (EIS) cules at higher carbonization temperature [41] and smaller size of
measurements were conducted in a 5 mM K3Fe(CN)6 solution con- biochar could produce larger surface area. which could facilitate the
taining 0.1 M KCl. mass transfer and expose more catalytic sites to activate PMS for pol-
lutants removal. It was also found the surface area of SugBC850
(172.27 m2 g−1) was larger than SugBC800 and smaller than
3. Results and discussion SugBC900, and the surface area of SugBC700 could be greatly enlarged
to 572.36 m2 g−1 by further pyrolysis of the SugBC700 at 900 °C (Fig.
3.1. Physicochemical properties of SugBC catalyst S1), suggested that the surface area of biochar was greatly affected by
the pyrolysis temperature.
By changing the pyrolysis temperatures, three types of SugBC cat- The crystal structural property of the SugBCs prepared at different
alysts (SugBC700, SugBC800 and SugBC900) were obtained. The cat- pyrolysis temperatures was evaluated by the XRD analysis. As presented
alytic properties were mostly related to the characteristics of the cata- in Fig. 2C, the three types of SugBCs (BC700, SugBC800 and SugBC900)
lysts, so various physicochemical properties of the SugBCs were all had the broad peaks at approximately 2θ = 25° that attributed to the
investigated. (0 0 2) crystal planes of parallel stacked graphene sheet [42], suggested
The morphology of SugBCs was analyzed by SEM. As shown in that the SugBCs contain graphitized structures. There were also dif-
Fig. 1A, all SugBCs were irregularly shaped, and the particle size be- fraction peaks at 2θ = 43.2° that are corresponded to the (1 0 0) crystal
came smaller as pyrolysis temperature was increased from 700 °C to plane of sp2 carbon with honeycomb structure [39]. It was noticed the
900 °C. This should be due to that SugBC obtained at 900 °C contains a peak of SugBC900 was a little sharper than that of SugBC700 and
larger amount of carbon with a graphite structure compared to the SugBC800, meaning higher graphitization degree and more sp2 carbon
other two samples, which could be more easily pulverized during the content of SugBC900 than others, by which high efficiency PMS acti-
same grinding process. From the size distribution obtained by the laser vation for pollutant degradation could be enabled.
diffraction particle size analyzer (Fig. 1B). it was found that the particle The graphitization property of the SugBCs was also accessed by the
size became smaller as the increment of the pyrolysis temperature. The Raman spectrometry. As presented in Fig. 2D, the D band at 1320 cm−1
particle sizes in the range of 0.02–2.0 μm for SugBC700, SugBC800 and represented the defects and heteroatoms of disordered sp2 hybrid
SugBC900 were 4.12%, 50.56% and 84.35%, and average particle sizes carbon in the graphite lattice, and the G band at 1594 cm−1 referred to
were 29.74, 1.96 and 0.81 μm, respectively. Smaller particle size the crystalline graphite carbon in the SugBCs [35]. The ID /IG value of
property would contribute to larger surface area and exposure of more SugBC900 (0.96) was higher than that of SugBC700 (0.84) and
catalytic sites for efficient catalytic oxidation reaction [38]. SugBC800 (0.88), indicating more defects in carbon lattice produced
The change of structural property was analyzed by TEM. As pre- with pyrolysis temperature of 900 °C, which was consistent with the
sented in Fig. 1C, the biochar pyrolyzed at 700 °C (SugBC700) showed XRD analysis. For carbon materials, the key factor determining the
disordered carbon, while SugBC900 presented more ordered carbon heterogeneous electron transfer rate is the state of electronic density ,
structure, indicating that with the increase of pyrolysis temperature, the which comes from the defect density of carbon materials [43]. The high
biochar structure would change from disordered channel to highly or- defect density property of SugBC900 would lead to higher electron
dered one. This also suggested that at the higher pyrolysis temperature, transfer rate, thus an enhanced catalytic performance would be an-
the more reversible pore structure in the SugBC could be produced due ticipated.
to the presence of graphene sheets with parallel arrangement [39]. The functional groups of SugBCs were characterized with the FTIR.
The element composition, pore size, and surface area of the SugBCs As shown in Fig. 2E. the three types of SugBCs all presented peaks at
were also analyzed. As indicated in Table 1, the C contents of three 1574 and 1110 cm−1 corresponding to aromatic C = C (sp2 carbon)
types of SugBCs were all around 80%, indicating a high degree of stretching, and 3424 cm−1 to the stretching of –OH [34,39]. It was
carbonization of SugBCs, and the O contents were gradually decreased noted that the peak intensities of aromatic C = C were increased with
as the pyrolysis temperatures became higher. From the nitrogen ad- the elevation of pyrolysis temperatures, suggesting higher temperature
sorption–desorption isotherm of SugBC900 (Fig. 2A), it was found the would result in increment of the sp2 content in the final structure of
type IV isotherms with H2 hysteresis loops and capillary condensation biochar, which was consistent with XRD analysis.
occurred at the relative pressure (P/P0) of 0.4–0.6, suggested the XPS analysis was also performed to analyze the carbon forms in
SugBC900 has the mesoporous structure, which was in accordance with different SugBC samples. As displayed in Fig. S2, as pyrolysis tem-
peratures higher, the peak of 284.4 eV responsible for the sp2 hybrid
Table 1 carbon was increased from 26.49% (SugBC700) to 47.83% (SugBC900),
Physicochemical properties of the obtained SugBCs. and the peak of 285.0 eV responsible for sp3 hybrid carbon decreased
BC types from 50.99% (SugBC700) to 31.03% (SugBC900), demonstrated that
pyrolysis at high temperature would improve the sp2 hybrid carbon
SugBC700 SugBC800 SugBC900 content, which was consistent with the former characterization results
C (%) 80.63 81.66 80.16 (Fig. 2C and E). The surface charges of the SugBCs were also analyzed.
O (%) 11.12 9.45 8.64 As shown in Fig. S3, surfaces of both SugBC700 and SugBC800 were
H (%) 2.47 1.75 1.26 positively charged, and the potential change amplitude was small under
O/C 0.14 0.12 0.11 different pHs. Instead, the surface of BC900 was negatively charged at
BET Surface Area (m2 g−1) 26.53 34.38 602.52
pHs of 1.0–11.0.
Pore volume (cm3 g−1) 0.050 0.056 0.274
Average pore size (nm) 6.460 7.485 1.764

4
L. He, et al. Chemical Engineering Journal 405 (2021) 126895

Fig. 2. (A) Nitrogen adsorption–desorption isotherms, (B) Pore diameter distribution, (C) XRD, (D) Raman, and (E) FTIR spectra of SugBC700, SugBC800 and
SugBC900, respectively.

3.2. Catalytic degradation of BPA by SugBC together with the SugBCs, the removal ability was greatly improved,
suggested that PMS alone could not effectively degrade BPA, but it may
The ability of the three types of SugBCs for activating PMS for de- catalytically be activated by the SugBCs and benefit to the fast and
gradation of BPA was investigated by treating BPA (20 mg/L) with the efficient BPA degradation. It was interestingly found that, with the
SugBCs (0.5 g/L) and PMS (1.0 g/L) in ultrapure water (pH 7.0) at SugBC700 and SugBC800 the BPA degradation efficiencies in 180 min
25 °C. As indicated in Fig. 3A, the BPA removal efficiency by only PMS of reaction were 18.5% and 28.7%, respectively, however, with the
was very low (5.4% in 180 min). When the BPA was treated using PMS SugBC900, the degradation efficiency reached to 93.7% in 30 min, and

5
L. He, et al. Chemical Engineering Journal 405 (2021) 126895

Fig. 3. (A) BPA removal and (B) PMS decomposition by SugBC700, SugBC800 and SugBC900. Reaction conditions: [SugBCs] = 0.5 g/L, [PMS] = 1.0 g/L,
[BPA]0 = 20 mg/L, [pH] = 7.0, [Temperature] = 25 °C.

100% in 60 min. These results demonstrated that the three kinds of 3.3.2. Influence of PMS concentration
SugBCs coupled with PMS all can degrade BPA, and among them As presented in Fig. S5C and D, the performance of BPA (20 mg/L)
SugBC900 have the highest activation ability. degradation was also affected by PMS concentrations. When the PMS
To realize whether the adsorption can contribute to the removal of concentrations were changed from 0.4 to 0.7, 1.0 and 1.3 g/L, the re-
BPA, the BPA was treated with the same amounts of only SugBCs, and action rate constants became higher from 0.096 to 0.106, 0.126 and
found that the removal efficiencies by SugBC700, SugBC800, and 0.137 min−1, respectively, indicating the SugBC900 possesses abun-
SugBC900 were 9.52%, 3.92%, and 3.55% in 60 min, respectively (Fig. dant catalytic active sites. Also suggested that the higher concentration
S4), suggested very low or neglectable adsorption removal abilities of of PMS will lead to production of more amounts of oxidized active
the SugBCs toward BPA. species resulting in the faster degradation of target pollutant. When
The activation ability of SugBCs toward PMS was also investigated 1.0 g/L of PMS was used, the BPA removal efficiency could reach 100%
by treating PMS with SugBCs for 180 min at 25 °C, and measuring the in 60 min (Fig. S5C), so this concentration was used in the following
residual amounts of PMS. As presented in Fig. 3B, the PMS could be experiments.
decomposed by the SugBC900 with the decomposition efficiency of
58.7%, which was much higher than by SugBC700 (8.1%) and 3.3.3. Influence of SugBC900 concentration
SugBC800 (17.5%). The abilities of three kinds of SugBCs for PMS de- The profile of BPA degradation depending on the change of
composition were consistent with that of BPA degradation. This result SugBC900 dosage was investigated. As shown in Fig. S5E and F, the
suggested the SugBC900 presenting far greater BPA degradation ability BPA (20 mg/L) removal efficiency reached 46.26%, 85.01%, and 100%
is due to its higher catalytic activation effect toward the PMS by su- in 60 min by using 0.1, 0.3, and 0.5 g/L of SugBC900, respectively, and
perior characteristics of the SugBC900 such as the high leveled gra- efficiency of 100% was achieved in 15 min with 0.7 g/L of SugBC900,
phitization degree, more sp2 carbon content, and significantly large suggesting BPA removal ability was improved as increment of the
surface area (Fig. 2C, Table 1). SugBC900 dosage. The reaction rate constants were also elevated from
0.022 to 0.345 min−1 as the increment of the SugBC900 concentrations
from 0.1 to 0.7 g/L, which is because more amounts of the SugBC900 in
3.3. Factors affecting catalytic reaction of SugBC900 the solution could provide more catalytic sites for PMS activation and
enable more fast and efficient degradation of BPA. Although 0.7 g/L of
The catalytic activity of SugBC900 toward PMS for BPA degradation SugBC900 presented faster reaction rate than that of using 0.5 g/L,
can be affected by various factors. So BPA degradation profiles at var- since both two could completely degrade BPA in short time, the con-
ious conditions including concentrations of BPA, PMS, and SugBC900, centration of 0.5 g/L was used in the following study.
reaction temperature and pH, were investigated.
3.3.4. Temperature effect
The BPA degradation profiles at various temperature conditions
3.3.1. Influence of BPA concentration were accessed by treating 20 mg/L of BPA with 0.5 g/L of SugBC900
As shown in Fig. S5A and B, the BPA degradation efficiencies were and 1.0 mg/L of PMS at pH of 7.0. As shown in Fig. 4A, the BPA could
varied depending on the concentrations of BPA. When initial con- be efficiently degraded at various temperature conditions. When the
centration of BPA was 20 mg/L, 100% of BPA could be removed within temperatures were 25 °C, 35 °C and 45 °C, the BPA removal efficiencies
60 min, while the removal efficiencies reached 68.2% and 39.1% for could all reach 100% in 60 min of reaction. And even at the low tem-
the BPA concentrations of 40 mg/L and 60 mg/L, respectively. And the perature of 10 °C, the removal efficiency could reach 94.7%. The re-
reaction rate constants were 0.126, 0.056, and 0.026 min−1 for BPA action rate constants were found to be increased from 0.060 to
concentrations of 20, 40, and 60 mg/L, respectively.The reduced de- 0.150 min−1 as the temperatures became higher from 10 to 45 °C
gradation rate should be due to that high concentration of BPA solution (Fig. 4B) that is related to the activation capacity of heat treatment
may occupy the catalytic active sites of the SugBC900, and block the [44]. The activation energy (Ea) of BPA degradation by SugBC900 and
contact with the sulfate oxidant, resulting in reduction of the oxidized PMS was 38.49 kJ/mol as calculated by the Arrhenius equation with the
species in the reaction solution. linear regression coefficient of 0.995 between ln(K) and 1/T. This was

6
L. He, et al. Chemical Engineering Journal 405 (2021) 126895

Fig. 4. Effects of (A, B) reaction temperature and (C, D) initial pH values on BPA removal.

lower than that of other catalysts coupled with PMS for BPA degrada- the negative charged SugBC900 catalyst, occupying the active sites and
tion (Table S1), suggesting that SugBC900 has an excellent ability to reducing the degradation rate.
activate PMS for pollutants degradation. Overall, the ability of BPA
degradation at such broad temperature conditions, especially even at
the low temperature, is beneficial for the in-situ water remediation in 3.4. Catalytic mechanism
field.
3.4.1. Characterization of oxidizing species
In PMS catalyzed reaction, the sulfate (SO4•−) and hydroxyl radical

3.3.5. pH effect ( OH) are usually the major oxidizing species that degrade pollutants,
The profile of BPA degradation depending on the various pH con- and the singlet oxygen (1O2) and the superoxide radical (O2−•) may also
ditions was accessed by treating 20 mg/L of BPA with the SugBC900 play important roles in the activation of PMS by non-metallic carbo-
(0.5 mg/L) and PMS (1.0 g/L) at room temperature. As illustrated in naceous catalysts. To confirm the main oxidizing species generated by
Fig. 4C, after 60 min reaction, the BPA degradation efficiency could the SugBC900 and PMS during BPA degradation, the EPR analysis was
reach 95.2% when pH was 3.0, and when the pHs were 5.0, 7.0, and performed by using the DMPO and TEMP as the chemical scavengers to
9.0, 100% of BPA could be degraded within 60, 50, and 40 min. Even at trap the oxidizing species.
the strong acidic (pH = 1.0) and alkali (pH = 11.0) conditions, the As presented in Fig. 5A, the presence of DMPO adducts at weak
BPA could be effectively degraded with efficiencies of 91.7% in 60 min strength, suggested that SO4•− and •OH were generated. When the
and 100% in 15 min respectively. This suggested SugBC900 together DMPO capture experiment was performed in a methanol solution, the
with PMS could effectively degrade BPA at all range of pHs from 1.0 to DMPO-O2−• adduct signal also appeared in the EPR spectrum. To fur-
11.0, such robustness property is significantly advantageous in treat- ther prove the generation of O2−•, NBT was used as the qualitative
ment of complicated wastewater. indicator as it could be reduced by O2−• (k = 5.88 × 104 M−1 s−1),
From Fig. 4D, it was also noted that the BPA degradation rates be- and the existence of O2−• could be verified by the absorbance change of
came faster (0.086 to 0.238 min−1) as the pH values became higher. NBT at 259 nm [46]. As shown in Fig. S6, the typical absorption peak of
This is because higher leveled hydroxide ions in the alkali solution can NBT in the catalytic reaction system was decreased in 60 min, which
activate PMS for BPA degradation [45], and the higher concentration of demonstrated the existence of O2−•. The presence of singlet oxygen in
hydrogen ions in the acidic solution can remove oxidative species in- the catalytic reaction was evaluated with TEMP as the spin capture
cluding sulfate and hydroxyl radicals [17]. In addition, because the reagent. As shown in Fig. 5B, there generated strong signals with the
surface of SugBC900 is negatively charged according to zeta potential relative intensity ratio of 1: 1: 1 in the EPR spectra, which corresponded
analysis (Fig. S3), and the pHzpc of BPA is 9.73 [29], so the hydrogen to the TEMP-1O2 adduct. These results suggested that 1O2 and O2−•
ions and the plus charged BPA might competitively adsorb to surface of were produced during the process of catalytic reaction.

7
L. He, et al. Chemical Engineering Journal 405 (2021) 126895

Fig. 5. EPR spectra obtained using (A) DMPO and (B) TEMP as spin-trapping agent, (C) BPA removal and (D) reaction rate by using SugBC900/PMS system in the
presence of different scavengers. Reaction conditions: [SugBC900] = 0.5 g/L, [PMS] = 1.0 g/L, [BPA] = 20 mg/L, [Temperature] = 25 °C, [EtOH] = 100 mM, [p-
BQ] = [L-his] = 24 mM, [DMPO] = [TEMP] = 80 mM, [pH] = 7.0. The control experiment was conducted without the addition of scavenger.

To realize main contributions of the different kinds of oxidative property of the SugBC was also accessed by EIS (Fig. 6B), in which, the
species in the catalytic reaction for BPA degradation, the quenching obtained Nyquist diagram showed a semicircle in the high frequency
assay was performed by using ethanol (EtOH), L-histidine (L-his), and region and a linearity in the low frequency region, which corresponded
p-benzoquinone (p-BQ) as the quenchers for SO4•− and •OH, 1O2, and to the electron transfer resistance (RCT) and the diffusion limit process
O2−•, respectively [46]. As shown in Fig. 5C and D, after treating with The RCT for SugBC900 was 22 Ω that was 13.5 and 5.6 times lower than
quenchers the BPA degradation abilities were all lowered than the that of SugBC700 (297 Ω) and SugBC800 (124 Ω), respectively. These
control. In response to the EtOH, L-his, and p-BQ, the degradation ef- results suggested that SugBC900 has superior conductivity and electron
ficiencies in 60 min were dropped to 92.7%, 51.9%, and 49.2%, and the transferring ability than the others, which would contribute to sig-
reaction rate constants also decreased to 0.065, 0.036, and 0.032 min-1, nificantly promoting the performance of PMS activation for efficient
respectively. This indicated that L-his and p-BQ had relatively stronger pollutant degradation.
inhibitory effect on the degradation of BPA (Fig. S6), while the influ- The charge transfer property between PMS and SugBCs was in-
ence of EtOH was weaker, this suggested SO4•−, •OH, 1O2, and O2−• are vestigated by monitoring chronoamperometric responses of the su-
all responsible for catalytic degradation, and among them, contribu- garcane SugBC electrode after sequential addition of PMS (1.0 g/L) and
tions by 1O2 and O2−• are dominant. BPA (20 mg/L) to the Na2SO4 solution. As shown in Fig. 6C, after ad-
dition of PMS, the currents were abruptly changed, suggesting that the
surfaces of PMS and SugBCs are likely to undergo electron transfer
3.4.2. Electrochemical property analysis through the formation of charge transfer complexes while SugBC may
Generally, the ability of catalyst for degradation of pollutants by serve as the mediator [1]. The current shift for SugBC900 (60.5 μA) was
PMS activation are usually related to the electrochemical property of much larger than that of SugBC700 (3.6 μA) and SugBC800 (32.54 μA),
the catalyst [47], this is because the catalytic reaction involves the that is due to the superior electron transfer ability of the SugBC900
electron transfer between catalyst, PMS and target pollutants. There- enabled more efficient activation reaction between SugBC and PMS.
fore, the electrocatalytic properties of the sugarcane derived biochars Subsequently, after addition of BPA, another current shift was observed
were evaluated by measuring CV, EIS in 5 mM K3Fe(CN)6 and 0.1 M KCl by SugBC900 electrode that should be due to the electron transfer from
solution. BPA to the complex of SugBC and PMS. The current shift in response to
As shown in the CV (Fig. 6A), the redox peak at around 0.27 V with BPA were none or lower with SugBC700 and SugBC800 electrodes,
the SugBC900 (81.0 μA) modified electrode was much higher than that respectively, that is because of the lower electron transfer ability.
of SugBC700 (46.83 μA) and SugBC800 (30.7 μA). The resistance

8
L. He, et al. Chemical Engineering Journal 405 (2021) 126895

3.4.3. The main active structure for PMS activation


The catalytic performance of the biochar should be greatly related
to the structural properties of the biochar. To investigate relation be-
tween the structural properties and the catalytic degradation perfor-
mance of the SugBC900, the change of various physical and chemical
properties of SugBC900 after first time use for BPA degradation was
pursued.
From the FTIR analysis shown in Fig. S8A, after the first use, the
characteristic peak intensity of the aromatic C = C (sp2 carbon) func-
tional groups at 1574 and 1110 cm−1 were significantly weakened
compared to that of the fresh SugBC900, indicating that it may act as an
effective activation structure for PMS activation. As illustrated in the
C1s spectra of XPS (Fig. 7A and B), after the first time use, the peak of
284.4 eV responsible for the sp2 hybrid carbon was decreased from
47.83% (fresh SugBC900) to 40.62%, and the peak of 285.0 eV re-
sponsible for sp3 hybrid carbon increased from 31.03% (fresh
SugBC900) to 39.79%. This suggested that the portion of sp2 hybrid
carbon structure in the biochar would be responsible for high catalytic
activity of the SugBC900 catalyst, which can effectively activate PMS
and produce oxidation active species to degrade BPA. It is known that
the contents of sp2 hybrid carbon and sp3 hybrid carbon in the biochar
correspond to contents of the graphite carbon and the disordered
carbon, respectively [47]. The decrease of sp2 hybrid carbon content
represents the decrease of graphitization degree of biochar, which
makes it difficult to transfer electrons, and the ability of activating PMS
to produce oxidized species for BPA degradation would be also reduced.
The decrease of graphitization degree was also analyzed by Raman
spectroscopy. As illustrated in Fig. S8B, the relative intensity ratios of D
and G bands (ID/IG) were 0.96 and 0.91 for fresh SugBC900 and the
used SugBC900, respectively, suggested that the graphitization degree
of biochar decreased after the catalytic reaction, which was consistent
with XPS analysis.
From the BET analysis presented in Fig. S8C, the surface area of the
used Sug900 has decreased from 602.52 m2 g−1 to 133.25 m2 g−1. This
is possibility due to the adsorption of the intermediate products pro-
duced in the system to the surface of the biochar, that would result in
decrease of the available catalytic sites.
According to the above results, the catalytic mechanism of BPA
degradation using SugBC900 together with PMS can be realized
(Scheme 1). There should be two pathways coexist in the SugBC900
based catalytic reaction. One is the radical reaction that was mediated
by SO4•−, •OH and O2−•, and the other one is the non-radical catalytic
pathway mediated by 1O2. The oxidation reaction of 1O2 and O2−•
generation is the prevailing pathway, while the formation of SO4•− and

OH is the recessive pathway. It has been reported that the catalytic
performance of carbon materials is closely related to the surface func-
tional groups such as hydroxyl and carboxyl groups [48]. BC and sur-
face oxygen-containing functional groups can act as electron transfer
media and activation sites [49], react with PMS to produce sulfate ra-
dical, which then reacts with water molecules to form hydroxyl radical
[50]. In this study, the oxygen content of SugBC900 was only 8.64%
Fig. 6. (A) Cyclic voltammetric and (B) electrochemical impedance spectro- (Table 1), indicating that most of the oxygen containing functional
metric profiles, and (C) i-t profiles of SugBC700, SugBC800 and SugBC900 groups were removed at high temperature. At the same time, FTIR also
modified GCE. confirmed that there were only a small amount of hydroxyl functional
groups. Therefore, less sulfate radical and hydroxyl radical would be
According to the LSV analysis in the Na2SO4 solution (Fig. S7), with produced.
the elevation of PMS concentrations from 1.0 to 10.0 g/L, the current Combined with quenching experiment and EPR analysis, superoxide
responses by the SugBCs were all increased, and current response by radical and singlet oxygen play a leading role in BPA degradation,
SugBC900 was greatly higher than by SugBC700 and SugBC800, which which indicates that SugBC900 may react with dissolved oxygen in
was consistent with the previous results (Fig. 6C), and demonstrated solution and transform to superoxide radical. As the reaction solution is
excellent catalytic activity of SugBC900 to PMS. It was concluded that acidic (pH nearly 4.0) after adding PMS, superoxide radical will react
SugBC900 with higher electron transfer ability and catalytic activity with hydrogen ion to form singlet oxygen (Eqs. (2)–(4)). Also, the an-
can be used as a super effective PMS activator, thus significantly im- ionic radicals of PMS (SO5•−) may be generated by electron transfer
proving the catalytic performance in degradation of organic pollutants. from PMS to SugBC900, and then SO5•− react with water molecules to
form singlet oxygen (Eqs. (5)–(6)). Moreover, sp2 hybrid carbon
structure was likely to be the main active site participating in the

9
L. He, et al. Chemical Engineering Journal 405 (2021) 126895

Fig. 7. C1s XPS of (A) fresh SugBC900 and (B) 1st used SugBC900.

activation process of PMS according to the XPS and FTIR analysis, including p-Isopropenylphenol, 2,4-ditert-butylphenol, benzoquinone,
combined with large surface area and high electron transfer property of hydroquinone and phenol were identified. On this basis, two pathways
SugBC900, enhanced reaction rate and high performance of BPA de- for BPA degradation were proposed (Fig. S9). On the one hand, BPA is
gradation could be endowed. attacked by reactive species via breaking C–C bond, and leading to
formation of p-Isopropenylphenol and phenol, then the p-Iso-
O2 + SugBC900 → SugBC900+ + O2−• (2)
propenylphenol is further oxidized to p-hydroquinone through the hy-
2H +
+ O2−• + O2−• 1
→H2O2 + O2 (3) droxylation process. Subsequently, benzoquinone is generated through
−• • + 1 −
the oxidation of phenol and p-hydroquinone. On the other hand, the
O2 + OH + 2H → O2 + OH (4) BPA is firstly oxidized by the reactive oxygen species, generating iso-
HSO5− + SugBC900 → SO5•− +H +
(5) propenylphenol radicals, which would couple with BPA and further
decompose into 2,4-ditert-butylphenol and phenol, and phenol would
2SO5•− + H2O → 2HSO4− + 1.5 O21
(6) be transformed into benzoquinone. In addition, ring-opened inter-
mediates, such as oxalic acid, formic acid, acetic acid and alkane may
also produce and could be mineralized into CO2 and H2O.
3.5. Intermediates and pathway of BPA degradation

To realize the process of BPA degradation by SugBC900 and PMS, 3.6. Reusability of SugBC900 for catalytic degradation of BPA
the intermediates produced during degradation reaction were analyzed
by the GC–MS. After 60 min of catalytic reaction, the byproducts It is meaningful for the catalyst to have good reusability in terms of

Scheme 1. Schematic illustration of the synthesis of SugBC900 and proposed mechanism of BPA degradation by SugBC900 and PMS.

10
L. He, et al. Chemical Engineering Journal 405 (2021) 126895

Fig. 8. (A) Degradation kinetic (inset: reaction rate profile) and (B) removal efficiency of phenol (20 mg/L), MO (100 mg/L) and RhB (10 mg/L) using SugBC900/
PMS system in different types of water samples. Reaction conditions: [SugBC900] = 0.5 g/L, [PMS] = 1.0 g/L, [BPA]0 = 20 mg/L, [Temperature] = 25 °C. (C)
Degradation kinetic and (D) reaction rate of SugBC900-catalyzed oxidation of BPA in comparison to MWCNT, GO, CHBC900 and PNBC900. Reaction conditions:
[cata] = 0.5 g/L, [PMS] = 1.0 g/L, [BPA]0 = 20 mg/L, [pH] = 7.0, [Time] = 60 min, [Temperature] = 25 °C.

reducing cost. The reusability of the SugBC900 catalyst was in- humic acid are the typical inhibitors that may possess the ability of
vestigated by repeated treating of BPA (20 mg/L) with SugBC900 radical trapping and competitive adsorption [52], so influences of such
(0.5 g/L) and PMS (1.0 g/L). As shown in Fig. S10, at the first time use reagents were investigated.
of SugBC900, the BPA could be completely degraded within 60 min, As illustrated in Fig. S11, when 20 mM of SO42−, H2PO4−, HCO3−,
and at the second and third time usages, the removal efficiencies were Cl− and 20 mg/L of humic acid were coexisted with 20 mg/L of BPA,
81.58%, 57.63%, respectively. At the fourth and fifth times usages, the the BPA degradation efficiencies were 100%, 97.23%, 87.64%, 99.39%,
removal efficiencies (38.53% and 10.11%) were less than 50%. The and 93.96%, respectively. This indicated that the influences by SO42−,
decrease in BPA removal efficiency may be due to decrement of the sp2 H2PO4−, and Cl− were negligible. Although the Cl− was reported to
hybrid carbon, and increment of the sp3 hybrid carbon that can not react with SO4•− and •OH (Eq. (7) and (8)) and inhibit relating catalytic
produce enough oxidizing species and it may also be due to the ad- reaction [53], but in the SugBC900 based catalytic reaction, the role of
sorption of by products on the catalyst and the reduction of effective SO4•− and •OH are minor, instead 1O2 and O2−• are the major oxidizing
active centers for PMS activation [51]. species (Fig. 5), so the BPA degradation was rarely effected by the Cl−.
The regeneration ability of the biochar catalyst was evaluated by However, with the presence of HCO3− the BPA degradation efficiency
anaerobic heat treatment (900 °C) of the catalyst after fifth reuse. And dropped 12.36% than the control (100%) that is due to the reaction of
found the BPA degradation efficiency by the regenerated SugBC900 HCO3− with the SO4•− and •OH and producing H+, which may gen-
could recovered to more than 89%, which would be due to the partial erate radical scavenging effect and make the pH of solution lowered
recovery of carbonization and removal of possible impurities attached (Eq. (9) and (10)). The lower pH would result in reduced BPA de-
on the catalyst surface at high temperature. gradation efficiency (Fig. 4C and D). The humic acid as a typical natural
organic materials resulted in about 6% drop of BPA degradation effi-
3.7. The effect of anions and natural organic matters on catalytic ciency that is because of the competitive reaction with the reactive
degradation of BPA species, this is similar with the results reported previously [54].

Cl− + SO4•− ↔ Cl• + SO42− (7)


Considering the complexity of environmental water, the effect of
− • •−
some typical anions and natural organic matters on catalytic degrada- Cl + OH ↔ ClOH (8)
tion of BPA should be concerned. SO42−, H2PO4−, HCO3−, Cl− and

11
L. He, et al. Chemical Engineering Journal 405 (2021) 126895

HCO3− + SO4•− → SO42− + CO3•− + H+ (9) Table 2


Structural and electrochemical properties of various catalyst.
HCO3− •
+ OH → H2O + CO3•− (10)
Samples Ipa (μA) Rctb (Ω) ID/IGc Surface aread (m2 g−1) sp2 contente

SugBC900 81.0 22 0.96 602.52 47.83%


3.8. BPA degradation in real environmental water PNBC900 49.9 89 0.92 182.19 40.26%
CHBC900 31.8 102 0.90 121.68 34.64%
The performance of BPA degradation in real environmental water MWCNT 52.1 64 1.01 100 38.08%
was conducted by spiking tap water (TW), lake water (LW) and muni- GO 30.4 135 0.82 42 28.88%

cipal water (MW) samples with 20 mg/L of BPA, and treating with a
Determined by CV.
SugBC900 (0.5 g/L) and PMS (1.0 g/L). As illustrated in Fig. S12, after b
Determined by EIS.
60 min reaction in water samples of TW, LW and MW, 86.72%, 82.91% c
Calculated by Raman spectra.
and 65.22% of BPA removals were achieved, respectively. During the d
Determined by BET analysis.
experiment, TOC can be used to characterize the mineralization of BPA. e
Estimated by XPS analysis.
Fig. S11B shows the TOC removal efficiency in different water en-
vironments, As the time pass, the removal efficiency of TOC has in- sugarcane has the superior characteristics for the high performance of
creased. After 3 h, the removal efficiency increased from 60.44% to catalytic reaction.
90.33% in DW, 54.13% to 82.52% in TW, 47.12% to 74.41% in LW, and The catalytic activity of SugBC900 was also compared to the con-
37.75% to 69.55% in MW, respectively. The decrease in removal effi- ventional catalysts including multi-walled carbon nanotube (MWCNT)
ciency in the real water may be due to the presence of various anions and graphene oxide (GO). From the CV, EIS, Raman, and XPS analysis
and organics that occupy the active sites of catalysts and react with free (Fig. S14, Table 2), it is noticed that the SugBC900 has better electron
radicals [55]. However, the occurrence of BPA mineralization indicates transfer ability, higher sp2 carbon content and lower resistance than
that part of the BPA was oxidized to carbon dioxide during the ex- MWCNT and GO, but the ID/IG ratio of SugBC900 was lower than that
periment. At the same time, the relatively high degree of mineralization of MWCNT indicating higher defect degree of MWCNT. Under the same
in various water environments means that the SugBC900 based PMS conditions with the same amounts of catalyst and PMS, the BPA de-
activation method has a good application prospect in the restoration of gradation efficiencies and reaction rates with the GO and MWCNT as
real organic polluted water bodies. the catalysts reached 14.3% and 0.007 min−1, and 64.9% and
0.070 min−1, respectively, which were much lower or slower than that
3.9. Catalytic degradation of other pollutants with SugBC900 and PMS with the SugBC900 (Fig. 8C and D) that should be related to their
electron transfer abilities and the carbonic structures.
The catalytic degradation abilities of SugBC900 and PMS toward Given the high efficiency of metal catalysts based PMS activating
other organic pollutants including phenol, methyl orange (MO) and system, the catalytic performance of SugBC900 was compared to the
Rhodamine B (RhB) were accessed. As presented in Fig. 8A, phenol metal catalyst. As depicted in Fig. S15. with the same amount of Co3O4
(20 mg/L) and MO (100 mg/L) could be completely removed within and PMS, 100% of BPA removal could be achieved within 30 min, and
30 min and 60 min, the removal efficiency for RhB (10 mg/L) could the reaction rate was 0.214 min−1, this enables the maximum activa-
reach 92.6% in 60 min. The degradation reaction rates of phenol, MO tion of PMS due to the cobalt catalysts [6].
and RhB were 0.296, 0.235 and 0.107 min−1, accordingly. In real en- The BPA degradation efficiencies using the Fe3O4 and Fe3O4 mod-
vironmental water, satisfactory degradation performance in these or- ified SugBC900 (Mag-SugBC900) were 38.2% and 61.4%, and the re-
ganics polluted water was also achieved (Fig. 8B). It could be concluded action rates were calculated to be 0.020 and 0.054 min−1, respectively
that the sugarcane SugBC900 and PMS based catalytic oxidation (Fig. S15). The incorporation of magnetic nanoparticles to the
method showed wide adaptability in the complex environments and SugBC900 is good to separate the catalyst from the solution, but due to
offered a promising prospect in the water treatment contaminated by the poor catalytic property of the Fe3O4, the control of the modified
organic pollutants. amount of the Fe3O4 is needed to favor the practical use.
Considering the economic feasibility of the catalyst, the cost of the
various catalysts were evaluated. As shown in Table S3, the unit price of
3.10. Comparison of the SugBC900 with other catalysts
SugBC900 (0.052 $/g) was less than Mag-SugBC900, Fe3O4, Co3O4,
MWCNT and GO suggesting that SugBC900 has better economic bene-
To investigate the catalytic performances of biochar from other
fits than the conventional carbonaceous and metallic catalysts. Al-
types of biomass, biochars from the corn husks and pine needles were
though its BPA degradation efficiency is slightly lower than Co3O4, but
prepared at 900 °C, and depicted as CHBC900 and PNBC900, respec-
SugBC900 should be significantly perspective for broad application in
tively. Their catalytic performances were compared with the SugBC900,
pollutant degradation in terms of environmental friendliness and cost
and relationship between the structural property and catalytic activity
effectiveness.
was evaluated by several characterization methods including CV, EIS,
In addition, the performance of catalytic oxidation reaction of
Raman, BET, and XPS analysis. It was found that the BPA degradation
SugBC900 with PMS for BPA degradation was compared with the other
efficiency and reaction rate using CHBC900 and PNBC900 were 29.4%
reported advanced oxidation process using modified or non-modified
and 0.016 min−1, and 39.7% and 0.019 min−1 (Fig. 8C and D), re-
biochar (Table 3). It was found the proposed SugBC900 catalyst pre-
spectively, which were lower than that of using SugBC900 as the cat-
sented wider range of working pHs (1.0–11.0) and temperatures
alyst (100% and 0.1259 min−1). From the results by the CV, EIS,
(10–45 °C) than other modified or non-modified biochars, suggested the
Raman, BET, and XPS analysis (Fig. S13, Table 2), it was noted that the
superior robustness property of the SugBC900 in catalytic degradation
oxidation current, graphitization degree (ID/IG), surface area, and sp2
of BPA. These results strongly revealed that the method of pollutant
hybridized carbon content of SugBC900 (81.0 μA, 0.96, 602.52 m2 g−1,
degradation using the SugBC900 as the catalyst for PMS activation
47.83%) were all higher than that of CHBC900 (31.8 μA, 0.90,
could be a powerful approach for environmental remediation in terms
121.68 m2 g−1, 34.64%) and PNBC900 (49.9 μA, 0.92, 182.19 m2 g−1,
of comparable efficiency, robustness and highly cost effectiveness.
40.26%). And the resistance (22 Ω) was also lower than that of
CHBC900 (102 Ω) and PNBC900 (89 Ω). These results suggested that
the catalytic performance of the biochar was greatly related to the
structural properties of the biochar, and the SugBC900 obtained from

12
L. He, et al. Chemical Engineering Journal 405 (2021) 126895

Table 3
Comparison of BPA degradation performances with various biochars as catalysts.
Resource BPA concentration Catalyst and Oxidant Modification Conditions Removal efficiency References

Pine pollen 20 mg/L biochar@CoFe2O4-Ag3PO4 0.5 g/L (MBA-3) CoFe2O4-Ag3PO4 UV 300 W 91.0%, 60 min [56]
Green tea 20 mg/L BC-iron (2 g/L) Iron US 60 W, pH 3.0 97.8%, 80 min [57]
PS (3 mM)
Sawdust 20 mg/L Fe-BC-700 (0.15 g/L) Iron pH 3.0, 5.0, 7.0, 9.0 100%, 15 min [29]
PMS (0.2 g/L)
Banana peels 20 mg/L Fe2O3@BC 0.3 g/L Fe2O3 pH 6.28 100%, 20 min [58]
PS 5 mM
Sewage sludge 10 mg/L SBC-600 0.2 g/L None pH 4.0–11.0 100%, 30 min [59]
PMS 0.1 g/L
Spent coffee 5 mg/L NBC-1000 0.2 g/L Nitrogen pH 4.0 99.0%, 60 min [51]
PMS 0.3 g/L
Rice husk 22.9 mg/L MnO2/BCs 0.5 g/L MnO2 US 130 W pH 7.0 100%, 120 min [60]
H2O2 10 mM
Sugarcane waste 20 mg/L SugBC900 0.5 g/L None pH 1.0, 3.0 90% − 95%, 60 min This work
PMS 0.4–1.0 g/L pH 5.0, 7.0, 9.0, 11.0 100%, 15–60 min

4. Conclusion [3] X. Dong, L. He, H. Hu, N. Liu, S. Gao, Y. Piao, Removal of 17Β-estradiol by using
highly adsorptive magnetic biochar nanoparticles from aqueous solution, Chem.
Eng. J. 352 (2018) 371–379, https://doi.org/10.1016/j.cej.2018.07.025.
A highly efficient organic pollutant degradation method was de- [4] S. Wang, Y. Yin, J. Wang, Enhanced biodegradation of triclosan by means of gamma
veloped by using the sugarcane derived pristine biochar (SugBC900) as irradiation, Chemosphere 167 (2017) 406–414, https://doi.org/10.1016/j.
the novel catalyst coupled with PMS. The SugBC900 presented excellent chemosphere.2016.10.028.
[5] R. Yin, W. Guo, H. Wang, J. Du, Q. Wu, J.S. Chang, N. Ren, Singlet oxygen-domi-
features benefit to catalytic activation such as higher surface area nated peroxydisulfate activation by sludge-derived biochar for sulfamethoxazole
(602.52 m2 g−1) and sp2 hybrid carbon contents (47.83%), and greater degradation through a nonradical oxidation pathway: Performance and mechanism,
electron transfer ability than that of the SugBC700, SugBC800 and Chem. Eng. J. 357 (2019) 589–599, https://doi.org/10.1016/j.cej.2018.09.184.
[6] Y. Du, W. Ma, P. Liu, B. Zou, J. Ma, Magnetic CoFe2O4 nanoparticles supported on
other biomass derived CHBC900 and PNBC900. The quenching and EPR titanate nanotubes (CoFe2O4/TNTs) as a novel heterogeneous catalyst for perox-
test revealed that the oxidative species of 1O2 and O2•− were dominant ymonosulfate activation and degradation of organic pollutants, J. Hazard. Mater.
for the catalytic reaction, and the role of the SO4•− and •OH was minor. 308 (2016) 58–66, https://doi.org/10.1016/j.jhazmat.2016.01.035.
[7] L. Ding, P. Zhang, H. Luo, Y. Hu, M. Norouzi Banis, X. Yuan, N. Liu, Nitrogen-Doped
The decrease of the sp2 carbon content after catalytic degradation of
Carbon Materials as Metal-Free Catalyst for the Dechlorination of Trichloroethylene
BPA was also confirmed. It was likely the sp2 carbon structure of the by Sulfide, Environ. Sci. Technol. 52 (2018) 14286–14293, https://doi.org/10.
SugBC900 be the active site for catalyzing PMS decomposition to de- 1021/acs.est.8b03565.
grade pollutants. The most significant achievement is fast and highly [8] K.Y.A. Lin, B.J. Chen, Prussian blue analogue derived magnetic carbon/cobalt/iron
nanocomposite as an efficient and recyclable catalyst for activation of perox-
efficient BPA degradation ability and the broad pH (1.0–11.0) and ymonosulfate, Chemosphere 166 (2017) 146–156, https://doi.org/10.1016/j.
temperature (10–45 °C) conditions for catalytic reaction. The SugBC900 chemosphere.2016.09.072.
with advantages properties including extremely low cost, highly effi- [9] C. Liu, L. Chen, D. Ding, T. Cai, From rice straw to magnetically recoverable ni-
trogen doped biochar: Efficient activation of peroxymonosulfate for the degradation
cient catalytic activation ability and broad working conditions, would of metolachlor, Appl. Catal. B 254 (2019) 312–320, https://doi.org/10.1016/j.
be significantly perspective for degradation of various organic pollutant apcatb.2019.05.014.
in the complicated practical environment water. [10] Z. Huang, P. Wu, J. Liu, S. Yang, M. Chen, Y. Li, W. Niu, Q. Ye, Defect-rich carbon
based bimetallic oxides with abundant oxygen vacancies as highly active catalysts
for enhanced 4-aminobenzoic acid ethyl ester (ABEE) degradation toward perox-
Declaration of Competing Interest ymonosulfate activation, Chem. Eng. J. 395 (2020) 124936, , https://doi.org/10.
1016/j.cej.2020.124936.
[11] S. Yang, X. Duan, J. Liu, P. Wu, C. Li, X. Dong, N. Zhu, D.D. Dionysiou, Efficient
The authors declare that they have no known competing financial peroxymonosulfate activation and bisphenol A degradation derived from mineral-
interests or personal relationships that could have appeared to influ- carbon materials: Key role of double mineral-templates, Appl. Catal. B 267 (2020)
118701, , https://doi.org/10.1016/j.apcatb.2020.118701.
ence the work reported in this paper.
[12] M. Chen, P. Wu, N. Zhu, Z. Dang, Y. Bi, F. Pei, Re-utilization of spent Cu 2 +
-immobilized MgMn-layered double hydroxide for efficient sulfamethoxazole de-
Acknowledgments gradation : Performance and metals synergy, Chem. Eng. J. 392 (2020) 123709, ,
https://doi.org/10.1016/j.cej.2019.123709.
[13] S. Yang, P. Wu, J. Liu, M. Chen, Z. Ahmed, N. Zhu, Efficient removal of bisphenol A
This work was funded by the National Natural Science Foundation by superoxide radical and singlet oxygen generated from peroxymonosulfate acti-
of China (51809111), and the open fund from the Key Lab of Eco-re- vated with Fe0-montmorillonite, Chem. Eng. J. 350 (2018) 484–495, https://doi.
storation of Regional Contaminated Environment (Shenyang org/10.1016/j.cej.2018.04.175.
[14] S. Wang, H. Liu, J. Gu, H. Sun, M. Zhang, Y. Liu, Technology feasibility and eco-
University), Ministry of Education of China (2019_ERRCE_K2). nomic viability of an innovative integrated ceramic membrane bioreactor and re-
verse osmosis process for producing ultrapure water from municipal wastewater,
Appendix A. Supplementary data Chem. Eng. J. 375 (2019) 122078, , https://doi.org/10.1016/j.cej.2019.122078.
[15] S. Yang, Z. Huang, P. Wu, Y. Li, X. Dong, C. Li, N. Zhu, X. Duan, D.D. Dionysiou,
Rapid removal of tetrabromobisphenol A by Α-Fe2O3-x@Graphene@
Supplementary data to this article can be found online at https:// Montmorillonite catalyst with oxygen vacancies through peroxymonosulfate acti-
doi.org/10.1016/j.cej.2020.126895. vation: Role of halogen and Α-hydroxyalkyl radicals, Appl. Catal. B 260 (2020)
118129, , https://doi.org/10.1016/j.apcatb.2019.118129.
[16] J.T. Gray, F. Che, J.S. McEwen, S. Ha, Field-assisted suppression of coke in the
References methane steam reforming reaction, Appl. Catal. B 260 (2020), https://doi.org/10.
1016/j.apcatb.2019.118132.
[17] F. Ghanbari, M. Moradi, Application of peroxymonosulfate and its activation
[1] H. Lee, H. Il Kim, S. Weon, W. Choi, Y.S. Hwang, J. Seo, C. Lee, J.H. Kim, Activation
methods for degradation of environmental organic pollutants: Review, Chem. Eng.
of Persulfates by Graphitized Nanodiamonds for Removal of Organic Compounds,
J. 310 (2017) 41–62, https://doi.org/10.1016/j.cej.2016.10.064.
Environ. Sci. Technol. 50 (2016) 10134–10142, https://doi.org/10.1021/acs.est.
[18] Z. Li, M. Wang, C. Jin, J. Kang, J. Liu, H. Yang, Y. Zhang, Q. Pu, Y. Zhao, M. You,
6b02079.
Z. Wu, Synthesis of novel Co3O4 hierarchical porous nanosheets via corn stem and
[2] J. Wang, S. Wang, Activation of persulfate (PS) and peroxymonosulfate (PMS) and
MOF-Co templates for efficient oxytetracycline degradation by peroxymonosulfate
application for the degradation of emerging contaminants, Chem. Eng. J. 334
activation, Chem. Eng. J. 392 (2020) 123789, , https://doi.org/10.1016/j.cej.2019.
(2018) 1502–1517, https://doi.org/10.1016/j.cej.2017.11.059.

13
L. He, et al. Chemical Engineering Journal 405 (2021) 126895

123789. [40] K. Qian, A. Kumar, H. Zhang, D. Bellmer, R. Huhnke, Recent advances in utilization
[19] H. Lu, M. Sui, B. Yuan, J. Wang, Y. Lv, Efficient degradation of nitrobenzene by Cu- of biochar, Renew. Sustain. Energy Rev. 42 (2015) 1055–1064, https://doi.org/10.
Co-Fe-LDH catalyzed peroxymonosulfate to produce hydroxyl radicals, Chem. Eng. 1016/j.rser.2014.10.074.
J. 357 (2019) 140–149, https://doi.org/10.1016/j.cej.2018.09.111. [41] F. Zheng, D. Liu, G. Xia, Y. Yang, T. Liu, M. Wu, Biomass waste inspired nitrogen-
[20] Y. Pang, Y. Ruan, Y. Feng, Z. Diao, K. Shih, L. Hou, D. Chen, L. Kong, Ultrasound doped porous carbon materials as high-performance anode for lithium-ion batteries,
assisted zero valent iron corrosion for peroxymonosulfate activation for J. Alloy. Compd. 693 (2017) 1197–1204, https://doi.org/10.1016/j.jallcom.2016.
Rhodamine-B degradation, Chemosphere 228 (2019) 412–417, https://doi.org/10. 10.118.
1016/j.chemosphere.2019.04.164. [42] G.T.K. Fey, D.C. Lee, Y.Y. Lin, T. Prem Kumar, High-capacity disordered carbons
[21] X. Duan, H. Sun, Z. Shao, S. Wang, Nonradical reactions in environmental re- derived from peanut shells as lithium-intercalating anode materials, Synth. Met.
mediation processes: Uncertainty and challenges, Appl. Catal. B 224 (2018) 139 (2003) 71–80, https://doi.org/10.1016/S0379-6779(03)00082-1.
973–982, https://doi.org/10.1016/j.apcatb.2017.11.051. [43] A. Nsabimana, X. Bo, Y. Zhang, M. Li, C. Han, L. Guo, Electrochemical properties of
[22] N.M. Noor, R. Othman, N.M. Mubarak, E.C. Abdullah, Agricultural biomass-derived boron-doped ordered mesoporous carbon as electrocatalyst and Pt catalyst support,
magnetic adsorbents: Preparation and application for heavy metals removal, J. J. Colloid Interface Sci. 428 (2014) 133–140, https://doi.org/10.1016/j.jcis.2014.
Taiwan Inst. Chem. Eng. 78 (2017) 168–177, https://doi.org/10.1016/j.jtice.2017. 04.044.
05.023. [44] Y. Guo, Z. Zeng, Y. Zhu, Z. Huang, Y. Cui, J. Yang, Catalytic oxidation of aqueous
[23] M. Ruthiraan, E.C. Abdullah, N.M. Mubarak, M.N. Noraini, A promising route of organic contaminants by persulfate activated with sulfur-doped hierarchically
magnetic based materials for removal of cadmium and methylene blue from waste porous carbon derived from thiophene, Appl. Catal. B 220 (2018) 635–644, https://
water, J. Environ. Chem. Eng. 5 (2017) 1447–1455, https://doi.org/10.1016/j.jece. doi.org/10.1016/j.apcatb.2017.08.073.
2017.02.038. [45] M. Nie, Y. Deng, S. Nie, C. Yan, M. Ding, W. Dong, Y. Dai, Y. Zhang, Simultaneous
[24] C. Huang, Z. Tan, P. Ai, Y. Wang, A. Kasiulienė, Cadmium removal potential by rice removal of bisphenol A and phosphate from water by peroxymonosulfate combined
straw-derived magnetic biochar, Clean Technol. Environ. Policy 19 (2016) with calcium hydroxide, Chem. Eng. J. 369 (2019) 35–45, https://doi.org/10.
761–774, https://doi.org/10.1007/s10098-016-1264-2. 1016/j.cej.2019.03.046.
[25] L. He, Y. Yang, J. Kim, L. Yao, X. Dong, T. Li, Y. Piao, Multi-layered enzyme coating [46] Z. Wei, D. Liu, W. Wei, X. Chen, Q. Han, W. Yao, X. Ma, Y. Zhu, Ultrathin TiO2(B)
on highly conductive magnetic biochar nanoparticles for bisphenol A sensing in Nanosheets as the Inductive Agent for Transfrering H2O2 into Superoxide Radicals,
water, Chem. Eng. J. 384 (2019) 123276, , https://doi.org/10.1016/j.cej.2019. ACS Appl. Mater. Interfaces 9 (2017) 15533–15540, https://doi.org/10.1021/
123276. acsami.7b03073.
[26] T. Huggins, H. Wang, J. Kearns, P. Jenkins, Z.J. Ren, Biochar as a sustainable [47] J. He, Y. Xiao, J. Tang, H. Chen, H. Sun, Science of the Total Environment Persulfate
electrode material for electricity production in microbial fuel cells, Bioresour. activation with sawdust biochar in aqueous solution by enhanced electron donor-
Technol. 157 (2014) 114–119, https://doi.org/10.1016/j.biortech.2014.01.058. transfer effect, Sci. Total Environ. 690 (2019) 768–777, https://doi.org/10.1016/j.
[27] C. Kalinke, V. Wosgrau, P.R. Oliveira, G.A. Oliveira, G. Martins, A.S. Mangrich, scitotenv.2019.07.043.
M.F. Bergamini, L.H. Marcolino-Junior, Green method for glucose determination [48] X. Duan, H. Sun, Z. Ao, L. Zhou, G. Wang, S. Wang, Unveiling the active sites of
using microfluidic device with a non-enzymatic sensor based on nickel oxyhydr- graphene-catalyzed peroxymonosulfate activation, Carbon 107 (2016) 371–378,
oxide supported at activated biochar, Talanta 200 (2019) 518–525, https://doi.org/ https://doi.org/10.1016/j.carbon.2016.06.016.
10.1016/j.talanta.2019.03.079. [49] Y. Wu, J. Guo, Y. Han, J. Zhu, L. Zhou, Y. Lan, Insights into the mechanism of
[28] Y. Xiang, H. Liu, J. Yang, Z. Shi, Y. Tan, J. Jin, R. Wang, S. Zhang, J. Wang, Biochar persulfate activated by rice straw biochar for the degradation of aniline,
decorated with gold nanoparticles for electrochemical sensing application, Chemosphere 200 (2018) 373–379, https://doi.org/10.1016/j.chemosphere.2018.
Electrochim. Acta 261 (2018) 464–473, https://doi.org/10.1016/j.electacta.2017. 02.110.
12.162. [50] S. Yang, X. Yang, X. Shao, R. Niu, L. Wang, Activated carbon catalyzed persulfate
[29] S.F. Jiang, L.L. Ling, W.J. Chen, W.J. Liu, D.C. Li, H. Jiang, High efficient removal of oxidation of Azo dye acid orange 7 at ambient temperature, J. Hazard. Mater. 186
bisphenol A in a peroxymonosulfate/iron functionalized biochar system: (2011) 659–666, https://doi.org/10.1016/j.jhazmat.2010.11.057.
Mechanistic elucidation and quantification of the contributors, Chem. Eng. J. 359 [51] A. Giannis, A. Veksha, J.-W. Lim, R.D. Webster, G. Lisak, T.-T. Lim, J.G.S. Moo, Y.-
(2019) 572–583, https://doi.org/10.1016/j.cej.2018.11.124. N. Liang, W.-D. Oh, Insights into the thermolytic transformation of lignocellulosic
[30] A. Aghababaei, M.C. Ncibi, M. Sillanpää, Optimized removal of oxytetracycline and biomass waste to redox-active carbocatalyst: Durability of surface active sites, Appl.
cadmium from contaminated waters using chemically-activated and pyrolyzed Catal. B 233 (2018) 120–129, https://doi.org/10.1016/j.apcatb.2018.03.106.
biochars from forest and wood-processing residues, Bioresour. Technol. 239 (2017) [52] R. Yin, W. Guo, H. Wang, J. Du, X. Zhou, Q. Wu, H. Zheng, J. Chang, N. Ren,
28–36, https://doi.org/10.1016/j.biortech.2017.04.119. Selective degradation of sulfonamide antibiotics by peroxymonosulfate alone:
[31] M.L. Crowley, L.B. Ricchetti, O.A. Todd, T.E. Mlsna, R. Anderson, C.U. Pittman, Direct oxidation and nonradical mechanisms, Chem. Eng. J. 334 (2018) 2539–2546,
A.G. Karunanayake, Rapid removal of salicylic acid, 4-nitroaniline, benzoic acid https://doi.org/10.1016/j.cej.2017.11.174.
and phthalic acid from wastewater using magnetized fast pyrolysis biochar from [53] X. Zhang, M. Feng, L. Wang, R. Qu, Z. Wang, Catalytic degradation of 2-phe-
waste Douglas fir, Chem. Eng. J. 319 (2017) 75–88, https://doi.org/10.1016/j.cej. nylbenzimidazole-5-sulfonic acid by peroxymonosulfate activated with nitrogen
2017.02.116. and sulfur co-doped CNTs-COOH loaded CuFe2O4, Chem. Eng. J. 307 (2017)
[32] Y. Zhang, X. Jiao, N. Liu, J. Lv, Y. Yang, Enhanced removal of aqueous Cr(VI) by a 95–104, https://doi.org/10.1016/j.cej.2016.08.078.
green synthesized nanoscale zero-valent iron supported on oak wood biochar, [54] H. Song, L. Yan, J. Ma, J. Jiang, G. Cai, W. Zhang, Z. Zhang, J. Zhang, T. Yang,
Chemosphere 245 (2020) 125542, , https://doi.org/10.1016/j.chemosphere.2019. Nonradical oxidation from electrochemical activation of peroxydisulfate at Ti/Pt
125542. anode: Efficiency, mechanism and influencing factors, Water Res. 116 (2017)
[33] X. Dong, L. He, Y. Liu, Y. Piao, Preparation of highly conductive biochar nano- 182–193, https://doi.org/10.1016/j.watres.2017.03.035.
particles for rapid and sensitive detection of 17β-estradiol in water, Electrochim. [55] D. Dai, Z. Yang, Y. Yao, L. Chen, G. Jia, L. Luo, Highly efficient removal of organic
Acta 292 (2018) 55–62, https://doi.org/10.1016/j.electacta.2018.09.129. contaminants based on peroxymonosulfate activation by iron phthalocyanine:
[34] K. Kameyama, T. Miyamoto, Y. Iwata, T. Shiono, Effects of biochar produced from Mechanism and the bicarbonate ion enhancement effect, Catal. Sci. Technol. 7
sugarcane bagasse at different pyrolysis temperatures on water retention of a cal- (2017) 934–942, https://doi.org/10.1039/c6cy02317g.
caric dark red soil, Soil Sci. 181 (2016) 20–28, https://doi.org/10.1097/SS. [56] Y. Zhai, Y. Dai, J. Guo, L. Zhou, M. Chen, H. Yang, L. Peng, Novel
0000000000000123. Biochar@CoFe2O4/Ag3PO4 Photocatalysts for Highly Efficient Degradation of
[35] H. Sun, S. Liu, G. Zhou, H.M. Ang, M.O. Tadé, S. Wang, Reduced graphene oxide for Bisphenol A under Visible-light Irradiation, J. Colloid Interface Sci. 560 (2019)
catalytic oxidation of aqueous organic pollutants, ACS Appl. Mater. Interfaces 4 111–121, https://doi.org/10.1016/j.jcis.2019.08.065.
(2012) 5466–5471, https://doi.org/10.1021/am301372d. [57] Z.-H. Diao, F.-X. Dong, L. Yan, Z.-L. Chen, W. Qian, L.-J. Kong, Z.-W. Zhang,
[36] S. Zhang, H. Niu, Y. Cai, Z. Meng, X. Zhang, D. Zhang, Humic acid coated Fe3O4 T. Zhang, X.-Q. Tao, J.-J. Du, D. Jiang, W. Chu, Synergistic oxidation of Bisphenol A
magnetic nanoparticles as highly efficient Fenton-like catalyst for complete mi- in a heterogeneous ultrasound-enhanced sludge biochar catalyst/persulfate process:
neralization of sulfathiazole, J. Hazard. Mater. 190 (2011) 559–565, https://doi. Reactivity and mechanism, J. Hazard. Mater. (2019) 121385, , https://doi.org/10.
org/10.1016/j.jhazmat.2011.03.086. 1016/j.jhazmat.2019.121385.
[37] Y. Wang, M. Liu, X. Zhao, D. Cao, T. Guo, B. Yang, Insights into heterogeneous [58] X. Rong, M. Xie, L. Kong, V. Natarajan, L. Ma, J. Zhan, The magnetic biochar de-
catalysis of peroxymonosulfate activation by boron-doped ordered mesoporous rived from banana peels as a persulfate activator for organic contaminants de-
carbon, Carbon 135 (2018) 238–247, https://doi.org/10.1016/j.carbon.2018.01. gradation, Chem. Eng. J. 372 (2019) 294–303, https://doi.org/10.1016/j.cej.2019.
106. 04.135.
[38] M. Ahmadi, B. Kakavandi, S. Jorfi, M. Azizi, Oxidative degradation of aniline and [59] B.C. Huang, J. Jiang, G.X. Huang, H.Q. Yu, Sludge biochar-based catalysts for im-
benzotriazole over PAC@FeIIFe2IIIO4: A recyclable catalyst in a heterogeneous proved pollutant degradation by activating peroxymonosulfate, J. Mater. Chem. A 6
photo-Fenton-like system, J. Photochem. Photobiol., A 336 (2017) 42–53, https:// (2018) 8978–8985, https://doi.org/10.1039/c8ta02282h.
doi.org/10.1016/j.jphotochem.2016.12.014. [60] K.W. Jung, S.Y. Lee, Y.J. Lee, J.W. Choi, Ultrasound-assisted heterogeneous Fenton-
[39] S.W. Han, D.W. Jung, J.H. Jeong, E.S. Oh, Effect of pyrolysis temperature on carbon like process for bisphenol A removal at neutral pH using hierarchically structured
obtained from green tea biomass for superior lithium ion battery anodes, Chem. manganese dioxide/biochar nanocomposites as catalysts, Ultrason. Sonochem. 57
Eng. J. 254 (2014) 597–604, https://doi.org/10.1016/j.cej.2014.06.021. (2019) 22–28, https://doi.org/10.1016/j.ultsonch.2019.04.039.

14

You might also like