You are on page 1of 14

Science of the Total Environment 749 (2020) 142313

Contents lists available at ScienceDirect

Science of the Total Environment

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

Peroxymonosulfate-enhanced photocatalysis by carbonyl-modified


g-C3N4 for effective degradation of the tetracycline hydrochloride
Yahui Shi a, Jinsong Li a, Dongjin Wan a,⁎, Jinhui Huang b, Yongde Liu a
a
School of Environmental Engineering, Henan University of Technology, Zhengzhou, Henan 450001, China
b
College of Environmental Science and Engineering, Hunan University, Changsha, Hunan 410082, China

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

• The carbonyl-modified g-C3N4 is syn-


thesized via one-step calcination.
• The introduction of carbonyl accelerates
carrier's separation and PMS activation.
• The CO-C3N4 possesses higher photocat-
alytic activity than g-C3N4.
• The performance of CO-C3N4 is largely
promoted with the addition of PMS.
• O2•−, h+ and 1O2 mainly contribute to
TCH degradation relative to •OH and
SO4•−.

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

Article history: In this work, carbonyl-modified g-C3N4 (CO-C3N4) is prepared through one-step calcination of the melamine-
Received 17 June 2020 oxalic acid aggregates. The visible light-assisted photocatalytic degradation efficiency of the tetracycline hydro-
Received in revised form 15 August 2020 chloride (TCH) for CO-C3N4 is significantly enhanced by introducing the peroxymonosulfate (PMS), and the ap-
Accepted 7 September 2020
parent rate constant is greatly increased from 0.01966 min−1 in CO-C3N4/vis system to 0.07688 min−1 in CO-
Available online 12 September 2020
C3N4/PMS/vis system. It is found that carbonyl for CO-C3N4 might offer possible reactive sites for PMS activation
Editor: Yifeng Zhang and collection sites of photo-generated electrons, greatly accelerating carrier's separation for PMS activation. The
favorable conditions, such as the higher catalyst dosage, higher PMS amount and alkaline pH, contribute to TCH
Keywords: degradation. The deleterious effects of co-existing anions on the TCH degradation efficiency are ranked in a de-
Photocatalysis cline: H2PO− 2− − − −
4 > SO4 > HCO3 > NO3 > Cl , and it may be affected by the type and amounts of anions and active
Carbonyl-modified g-C3N4 radicals generated. The radical trapping tests and electron spin resonance (ESR) detection display that the O2•−,
Peroxymonosulfate activation h+, 1O2, •OH and SO4•− all contribute to TCH degradation. Meanwhile, possible degradation mechanism, interme-
TCH degradation diates and degradation pathway of TCH are revealed in CO-C3N4/PMS/vis system. This study will offer a new in-
sight for constructing PMS activation with carbonyl modified g-C3N4 photocatalysis system to achieve effective
treatment of organic wastewater.
© 2018 Elsevier B.V. All rights reserved.

1. Introduction and Ghisi, 2014; Shi et al., 2020). Owing to its overuse, antibiotics
are frequently detected in water environment, including groundwa-
Antibiotics are intensively-applied chemicals for aquaculture ter (Szekeres et al., 2018), river water (Huang et al., 2019), surface
feeding and the treating of animal and human disease (Manzetti water (Danner et al., 2019) and drinking water (Sanganyado and
Gwenzi, 2019). Antibiotics exist for a long time in the environment,
⁎ Corresponding author. resulting in the spreading of antibiotic resistance, thereby
E-mail address: djwan@haut.edu.cn (D. Wan). destructing ecological systems and threatening human health.

https://doi.org/10.1016/j.scitotenv.2020.142313
0048-9697/© 2018 Elsevier B.V. All rights reserved.
Y. Shi, J. Li, D. Wan et al. Science of the Total Environment 749 (2020) 142313

Hence, it is urgent to explore the feasible and effective materials and/ degradation pathways of TCH in CO-C3N4/PMS/vis system were
or methods to eliminate antibiotics in water. discussed. Constructing PMS activation with carbonyl modified g-C3N4
Advanced oxidation processes (AOPs), with the advantage of com- photocatalysts for wastewater treatments will be a promising strategy.
plete decomposition of pollutants by generating active radicals, have
attracted much attention (Guan et al., 2013). AOPs typically include hy- 2. Experimental
droxyl (•OH) and sulfate (SO4•−) radicals. Compared with •OH, SO4•−
based AOPs are being research hot because of such merits: (1) SO4•− ex- 2.1. Materials
hibits the higher redox potential (E0 (SO4•−/SO2− 4 ) = 2.5–3.1VNHE) rel-
ative to •OH (E0 (•OH/H2O) = 1.8–2.7VVHE) (Shao et al., 2017); (2) the Peroxymonosulfate (2KHSO5·KHSO4·K2SO4, PMS) was from Alad-
half-life of 30–40 μs of SO4•− is longer relative to that of •OH (<1 μs) din Chemistry Co., Ltd., China. Melamine, oxalic acid dihydrate, rhoda-
(Lin and Zhang, 2017); (3) it is the independence of pH (Guan et al., mine B (RhB), Congo red (CR), methyl orange (MO), orange G (OG),
2013). The SO4•− expects be obtained via activation of peroxymonosul- tetracycline hydrochloride (TCH), sodium chloride (NaCl), sodium ni-
fate (PMS) via ultraviolet (UV), alkaline, ultrasound, heating, transition trate (NaNO3), sodium dihydrogen phosphate (NaH2PO4), sodium sul-
metal ions, metallic and metal-free catalysts (Wacławek et al., 2017). fate (Na2SO4), sodium bicarbonate (NaHCO3), p-benzoquinone (p-BQ),
Among them, UV, alkaline and thermal activation need the added methanol (CH3OH, MeOH), tert- butyl alcohol (C4H10O, TBA), disodium
chemicals or energy; the use of whether the homogeneous or heteroge- ethylenediaminetetraacetate (EDTA-2Na), hydrogen peroxide (H2O2)
neous metallic catalysts necessarily results in secondary pollution. and L-histidine (C6H9N3O2) were purchased from the Sino pharm
Allowing for economic cost and environmental impact, PMS activation Chemical Reagent Co., Ltd., China. Nitro tetrazolium blue chloride
using non-metal catalysts is promising, and several metal-free catalysts (NBT) and p-Hydroxybenzoic acid (p-HBA) were purchased from the
have been proposed, such as graphene (Duan et al., 2015), graphitic car- Macklin Co., Ltd., China. Benzoic acid (BA) was obtained from the Tianjin
bon nitride (g-C3N4) (Lin and Chen, 2016), carbon nanotubes (CNTs) Kemiou Chemical Reagent Co., Ltd., China. The solution pH was con-
(Lee et al., 2015) and activated carbon (Zhang et al., 2013). Duan et al. trolled by NaOH and H2SO4 with the concentration of 0.1 mol/L in Factor
(2015) reported that sulfur and nitrogen co-doped RGO exhibited re- experiments of pH.
markably improved activity in PMS activation for catalytic degradation
of phenol. Zhang et al. (2013) found that granular activated carbon
2.2. Preparation and characterization of CO-C3N4
(GAC) can activate PMS to degrade Acid Orange (AO7) with much
lower activation energy of 25.13 kJ/mol. Duan et al. (2016) studied
Melamine (6 g) and oxalic acid dihydrate (9 g), according to the
that the nanocarbons with diverse structure and dimensions for PMS ac-
molar ratio (2:3), were respectively dissolved in the de-ionized water.
tivation, and it was revealed that the radical and nonradical oxidations
Two above solutions were mixed for 2 h under electric stirring to obtain
functioned on diverse carbon catalysts, which was dependent on carbon
the white precipitates. Then the precipitate was filtered, washed and
structure.
dried at 60 °C. The precipitate was calcined in a muffle furnace (initial
Considering its characteristics of high specific surface area, good me-
temperature is 120 °C, 2.3 °C/min, 550 °C, 4 h). After being cooled,
chanical property and easy preparation, g-C3N4 is greatly attractive
dark brown products were obtained as CO-C3N4. Meanwhile, bulk g-
(Duan et al., 2020; He et al., 2020; Yi et al., 2020). Recently, g-C3N4-
C3N4 was obtained through a parallel process without addition of oxalic
based photocatalysis coupling with PMS activation has been utilized to
acid dihydrate.
degrade organic pollutants, in which PMS functions as the electron ac-
Characterization of both materials was provided in the Supplemen-
ceptor for formation of SO4•− (HSO− 5 + e

→ SO4•− + OH−) (Tao
tary Material.
et al., 2015). However, g-C3N4 suffers from poor photocatalytic activity
due to small specific surface area, poor separation of photogenerated
2.3. TCH degradation by CO-C3N4 activating PMS under visible light
carriers and weak light absorption (Wang et al., 2018, 2019a, 2019b).
In practice, several ways are used to modify g-C3N4 to improve PMS ac-
Briefly, the catalyst (20 mg) was dispersed into TCH solution
tivation for organic pollutant removal under visible light irradiation, in-
(10 mg/L, 100 mL) and stirred for 30 min in the dark to reach an
cluding nanostructure designing (g-C3N4 nanosheet) (Jiang et al., 2017),
adsorption-desorption equilibrium. The reaction was further conducted
constructing the heterojunctions with other semiconductors (PI/g-
via simultaneously adding PMS and turning on Xenon lamp (300 W,
C3N4, Co3O4/g-C3N4) (Gao et al., 2018a; Shao et al., 2017; Zhang et al.,
λ > 420 nm). Herein, unless otherwise specified, all experiments do
2018b), noble metal loading (Ag/g-C3N4) (Wang et al., 2017b), and cou-
not adjust initial pH. The aliquot of 2 mL was sampled at given times,
pling with activated carbon (g-C3N4/AC) (Dangwang Dikdim et al.,
and mixed with methanol of 2 mL at once. The sample was filtered
2019). As a matter of fact, the nonmetal doping (e.g., sulfur, nitrogen
using the microfiltration membrane (0.22 um) and determined at
and born co-doping) has been demonstrated to achieve high photocat-
357 nm by UV–vis spectrum analysis. The possible structure of TCH in-
alytic activity and PMS activation (Chen et al., 2019; Lin and Zhang,
termediates was analyzed through the high-performance liquid
2017). Such a carbon and oxygen co- doping g-C3N4 consists of earth-
chromatography-mass spectrometry (HPLC-MS) system in m/z range
abundant nonmetal elements, and can be simply synthesized by one-
of 50–500. The degradation of RhB (10 mg/L, 100 mL), CR (50 mg/L,
step preparation, making it a promising material with high photocata-
100 mL), OG (50 mg/L, 100 mL) and MO (50 mg/L, 100 mL) was mea-
lytic performance and PMS activation based on its sustainability. How-
sured respectively at 554 nm, 497 nm, 475 nm and 463 nm.
ever, for all we know, there are no studies on PMS activation by
All the degradation experiments were conducted in triplicate.
carbon and oxygen co-doping g-C3N4 under visible light irradiation so
far to remove TCH in water.
In this study, we have developed the non-metallic carbonyl modified 2.4. Photoelectrochemical experiments
g-C3N4 (CO-C3N4, i.e. C, O co-doping g-C3N4) with PMS effective activa-
tion for TCH degradation under visible light irradiation. CO-C3N4 was Details experiment process is presented in the Supplementary
characterized via various technologies, and its photocatalytic perfor- Material.
mance was investigated. The possible roles of carbonyl in CO-C3N4 for
PMS activation were revealed. Meanwhile, the effects of different factors 2.5. The quantification of the common free radicals (O2•−, •OH, SO4•−)
on TCH degradation by CO-C3N4 under visible light irradiation were
studied in details. Based on the radical trapping test, ESR results and Details quantification methods are given in the Supplementary
TCH intermediates analysis, the possible degradation mechanism and Material.

2
Y. Shi, J. Li, D. Wan et al. Science of the Total Environment 749 (2020) 142313

3. Results and discussion 13.1° is marked as the (100) phase of inplane structural packing motif
of tri-s-triazine unit (Shi et al., 2018; Yan, 2012). Obviously, the peaks
3.1. Structure and morphology characterization of CO-C3N4 are far weaker relative to those of g-C3N4, suggesting a
lower crystallinity degree for CO-C3N4. The overall stacked layer mor-
The crystal structure of the catalysts is characterized by XRD phologies of g-C3N4 and CO-C3N4 are present in Fig. 2 based on SEM
(Fig. 1a). The g-C3N4 and CO-C3N4 exhibit similar XRD spectra, resulting and TEM techniques, whereas the layer structure of CO-C3N4 with
from graphite-like structure. A stronger peak at 27.4° is marked as the some irregular holes is more clear and distinct compared with that of
(002) phase of aromatic segment interlayer stacking with interlayer dis- g-C3N4, likely because of release of more gas (NH3 and CO) during calci-
tance of 0.326 nm (Liao et al., 2014; Shi et al., 2018). A weaker peak at nation of melamine-oxalic acid supramolecular aggregates.

Fig. 1. (a) XRD patterns, (b) FTIR spectra, (c) Survey spectra, (d) C1s, (e) N1s and (f) O1s of XPS spectra of g-C3N4 and CO-C3N4, and (g) 13C solid state NMR spectrum of CO-C3N4.

3
Y. Shi, J. Li, D. Wan et al. Science of the Total Environment 749 (2020) 142313

Fig. 2. SEM and TEM images of (a, b) g-C3N4 and (c, d) CO-C3N4.

Both g-C3N4 and CO-C3N4 have similar FTIR spectra (Fig. 1b), sug- from Fig. 1d, C1s peaks at 288.40 eV and 284.80 eV of both materials fol-
gesting the retention of tri-s-triazine-based structure (Gao et al., low the sp2-hybridized C atoms (N-C=N) (Li et al., 2018b) and the C-
2018b). The broad peaks in range of 3500–3000 cm−1 are because of C/C=C bond (Gao et al., 2018b), respectively. Fig. 1e provides the fitting
the existence of NH2 or NH at the structure defect sites by the incom- of N1s into the four nitrogen chemical states at 398.70 eV, 400.42 eV,
plete condensation (Guo et al., 2018; Shi et al., 2018). The strong ab- 401.02 eV and 404.59 eV, corresponding to the sp2-hybridized aromatic
sorption between 1800 cm−1 and 1100 cm−1 is owing to the N bonding to C atom (C=N-C), the tertiary N bonding to C atom (N-(C)
stretching vibration of the heptazine heterocyclic ring (C6N7), specifi- 3), amino functions with hydrogen (C-N-H) related to structural defects,
cally for C\\N (1412, 1324 and 1242 cm−1) and C_N stretching and π excitation, respectively (Shi et al., 2018; Wei et al., 2016). As seen
modes (1636 and 1574 cm−1) (Guo et al., 2018). A sharp peak at in Fig. 1f, both g-C3N4 and CO-C3N4 possess O1s peak at 532 eV, owing to
814 cm−1 is marked as s-triazine ring mode. However, the characteristic H2O adsorbing on the surface or hydroxyl groups (Shi et al., 2019; Zhang
peaks of C_O bending vibration can be not found in the FTIR spectrum et al., 2018a). In addition, CO-C3N4 exhibits a new peak at 530.62 eV that
of CO-C3N4, possibly because of similarity of C_O and C_N bending vi- corresponds to the carbonyl groups (C=O) (Zhao et al., 2014). The solid
bration as well as the intrinsic limits from analytical instrument. state 13C NMR spectrum of CO-C3N4 is performed to further confirm the
The elemental composition of materials is determined via the ele- existence of carbonyl carbon. As presented in Fig. 1g, the characteristic
mental analysis (Table 1). The C/N value and O content of g-C3N4 are peak at δ 161.36 ppm corresponds to amide carbonyl, and it may be
0.577 and 9.342, respectively, while increase to 0.583 and 10.696 in formed by dehydration of amino in melamine and carbonyl in the bro-
CO-C3N4, indicating the existence of characteristic carbon and oxygen ken oxalic acid during the calcination. Besides, it is worth noting that
doping. The XPS is used to further analyze g-C3N4 and CO-C3N4. Three characteristic peak at δ 121.40 ppm correspond to -N=C=O (isocya-
elements including carbon, nitrogen and oxygen are found in g-C3N4 nate). Meanwhile, the surface element composition of g-C3N4 and CO-
and CO-C3N4 materials (Fig. 1c), and there is no obvious shift of binding C3N4 based on XPS is given in Table S1, in which the percentage of car-
energy of C1s and N1s core electrons, suggesting that g-C3N4 and CO- bon increases from 39.31% (g-C3N4) to 41.27% (CO-C3N4), and the oxy-
C3N4 possess the similar chemical states of both C and N. As present gen percentage increases from 2.41% (g-C3N4) to 2.50% (CO-C3N4). Thus,
combined XPS results with EA analysis and 13C NMR spectrum, CO-C3N4
presents the characteristic carbon and oxygen doping by the existence
Table 1 of carbonyl.
Elemental analysis (EA) of materials. The N2 adsorption-desorption isotherms and pore size distributions
are given in Table S2 and Fig. 3. The N2 adsorption-desorption isotherms
N (wt%) C (wt%) H (wt%) C/N (at.) O (wt%)
of both materials can be marked as the type I with a hysteresis loop of
g-C3N4 61.39 35.45 1.905 0.577 9.342 H3 type, suggesting that their mesoporous structures are from accumu-
CO-C3N4 59.98 34.98 2.267 0.583 10.696
lation of sheet-like particles, which accords with SEM and TEM results.

4
Y. Shi, J. Li, D. Wan et al. Science of the Total Environment 749 (2020) 142313

3.2. Optical property and photoluminescence

The diffused reflectance spectra (DRS) of both materials are pro-


vided in Fig. 4a. g-C3N4 has an adsorption edge of ~475 nm, whereas
CO-C3N4 possesses a wider and stronger light absorption across the
overall spectrum ranging from 300 to 800 nm. Meanwhile, absorption
edge of CO-C3N4 also shows red-shift likely because of the existence of
carbonyl groups. The optical absorption of semiconductor and its elec-
tronic structure are closely related (Yang et al., 2016), and the band
gap energy of semiconductors can be estimated based on Kubelka-
Munk transformation (αhv = A(hv-Eg)n/2). The n value for g-C3N4 (indi-
rect transition) is 4. According to the (αhv)1/2 versus hv curves in Fig. 4b,
the Eg value of CO-C3N4 is 1.91 eV, smaller than that of g-C3N4 (2.28 eV),
revealing that CO-C3N4 has improved light absorption performance.
The photoluminescence (PL) is adopted to investigate charge
carrier's recombination in semiconductors. As showed in Fig. 4c, both
g-C3N4 and CO-C3N4 exhibit the wide emission peaks in range of 400
to 700 nm. Obviously, CO-C3N4 shows diminished PL intensity relative
Fig. 3. N2 adsorption-desorption isotherms and pore size distribution (inset) of materials. to g-C3N4 likely owing to the existence of carbonyl, and it has been re-
ported that the carbonyl can function as collection site of
photogenerated electrons, contributing to carriers separation (Kong
BET surface area and pore volume of CO-C3N4 are 47.97 m2/g and et al., 2019). In addition, the clearer layer structure for CO-C3N4 also
0.17 cm3/g, respectively, and those of g-C3N4 are 27.93 m2/g and might contribute to separation and transfer of the carriers. Meanwhile,
0.12 cm3/g, respectively. Introducing the oxalic acid dihydrate has the fluorescence decay time curves of both materials are provided
only slight influence on BET surface area, suggesting that the surface from Fig. 4d and Table 2. The lifetime of the photogenerated charges di-
area might not be a key factor to cause different photocatalytic perfor- rectly determines its reaction with H2O to generate reactive oxygen rad-
mance of g-C3N4 and CO-C3N4. icals, or its mineralization efficiency by direct reaction with target

Fig. 4. (a) DRS, (b) (αhv)1/2 versus hv curves, (c) PL spectra and (d) the fluorescence decay curves of materials.

5
Y. Shi, J. Li, D. Wan et al. Science of the Total Environment 749 (2020) 142313

Table 2 semiconductor, respectively (Dong et al., 2020b, 2020c). As seen in


The lifespan of photogenerated charge of materials. Fig. 5c, both g-C3N4 and CO-C3N4 show positive slopes in the linear re-
Samples τ1 τ2 mean (ns) gion, confirming that they are n-type semiconductors. The VFB values
ns Rel (%) ns Rel (%)
of g-C3N4 and CO-C3N4 are respectively −0.60 and −0.58 V (vs. SCE),
corresponding to −0.36 and −0.34 V (vs. NHE). Meanwhile, the valence
g-C3N4 1.0667 32.69 5.8883 67.31 4.3121
band XPS (VB-XPS) is performed (Fig. 5d), and the valence band (EVB) of
CO-C3N4 0.6671 37.56 6.8000 62.44 4.4965
g-C3N4 and CO-C3N4 are 1.97 and 1.55 eV, respectively. According to the
Mott-Schottky and VB-XPS, energy band gap of g-C3N4 and CO-C3N4 can
pollutants. Specifically, the longer the photogenerated charge lifetime be calculated to 2.33 and 1.89 eV, which is basically consistent with their
is, more conductive its migration to reaction active sites is; the shorter estimated band gap from the Tauc plots (Fig. 4b).
the photogenerated charges lifetime is, the higher its recombination
rate is (Zhao et al., 2014). The short life (τ1) and long life (τ2) in CO- 3.4. Photocatalytic performance with PMS and stability
C3N4 are decreased and increased, respectively, along with increase of
τ1 percentage, causing an increase of its mean life. All in all, CO-C3N4 The photocatalytic performance of materials is evaluated by
possesses higher separation ability of photogenerated charges, thus degrading TCH in presence of PMS (Fig. 6a). The blank experiments in-
contributing to its stronger photocatalytic performance. dicate that visible light irradiation alone barely degrade TCH, whereas
PMS alone is capable of degrading TCH (58.12%). For the control exper-
3.3. Photoelectrochemical properties iments, before introducing the PMS and irradiation, it achieves an
adsorption-desorption balance in the dark for 30 min. Both g-C3N4
In order to further evaluate the separation and transfer efficiencies of and CO-C3N4 exhibit no great adsorption toward TCH. After adding
the photogenerated charges of CO-C3N4, the photocurrent response and PMS without the irradiation, TCH degradation is clearly improved
the Nyquist plot are measured. In the Fig. 5a, the photocurrent density of from 71.34% (CN + PMS) to 85.71% (CO-CN + PMS), suggesting car-
CO-C3N4 is obviously higher than that of g-C3N4, indicating the excellent bonyl great reactivity on PMS activation without irradiation. It was
separation efficiency of the photogenerated electron-hole pairs of CO- found that C_O groups act as principal reactive sites to activate PMS
C3N4. The Nyquist plot in the Fig. 5b presents that CO-C3N4 has a (Guo et al., 2018; Wei et al., 2016). Specifically, Guo et al. (2018) studied
lower charge transfer resistance with a smaller semi-circle than those the catalytic properties of the graphitic carbon nitride modified mag-
of g-C3N4, suggesting that CO-C3N4 possesses a faster charge transfer netic carbon nanocomposites (Fe3O4@C/g-C3N4) in activating PMS for
by effectively separating electron and holes (Dong et al., 2020a; Li degradation of Acid Orange 7, and proposed that the C_O groups on
et al., 2018a). the carbon surface of Fe3O4@C functioned as the principal active sites
The Mott-Schottky (M-S) plots of g-C3N4 and CO-C3N4 materials are for PMS activation. Wei et al. (2016) found that the delocalized elec-
tested at 1000 Hz in the dark to investigate the semiconductor nature. trons from the triazine units of g-C3N4, especially the C_O groups had
The flat band potential (VFB) from the x-intercept of the M-S plot can a high chemical activity and could be transferred to activate the PMS
be used to approximately estimate the conduction band potential to produce radicals in g-C3N4/AC + PMS system. Further, the photocat-
(VCB) or the valence band potential (VVB) in the n-type or p-type alytic activity of g-C3N4 and CO-C3N4 is compared via TCH degradation

Fig. 5. (a) Transient photocurrent response, (b) EIS Nyquist in the dark, (c) the M-S plots at a frequency of 1000 Hz in the dark and (d) valence band XPS spectra of materials.

6
Y. Shi, J. Li, D. Wan et al. Science of the Total Environment 749 (2020) 142313

Fig. 6. (a) TCH degradation efficiency by different catalysts under different systems, (b) the corresponding kinetic constants, (c) the cycling tests and (d) the influence of different water
sources on TCH degradation efficiency in CO-C3N4/PMS/vis system. Conditions: [Cat.] = 0.2 g/L, [TCH]0 = 10 mg/L, [PMS] = 200 mg/L, λ > 420 nm, ambient temperature.

only under visible light irradiation. 71.60% TCH degradation within activate PMS; (2) it is the collection site of the photogenerated electrons
60 min can be achieved in CO-CN + vis system, which is greatly higher and contributes to PMS activation into the active radicals. TCH degrada-
than that of CN + vis system (23.71%). It is speculated that CO-C3N4 tion accords well with the pseudo-first-order kinetics in different sys-
possesses the improved photocatalytic activity likely owing to existence tems (Zhang et al., 2019a), and corresponding rate constants are given
of carbonyl (Kong et al., 2019; Zhang et al., 2016). Kong et al. (2019) in Fig. 6b, which are calculated to be 0.01466 (R2 = 0.99551, PMS),
prepared one new carbonyl linked g-C3N4 through CO2-assisted thermal 0.01872 (R2 = 0.99251, PMS + vis), 0.0045 (R2 = 0.99958, CN + vis),
polymerization of urea, and it was found that the = C=O groups func- 0.02096 (R2 = 0.98777, CN + PMS), 0.02952 (R2 = 0.98934,
tioned as the collection sites of photogenerated electrons, thus contrib- CN + PMS + vis), 0.01966 (R2 = 0.95738, CO-CN + vis), 0.03146
uting to the carriers separation. In addition, the higher surface area of (R2 = 0.94269, CO-CN + PMS) and 0.07668 min−1 (R2 = 0.9644, CO-
CO-C3N4 can provide more surface reaction sites; meanwhile the more CN + PMS + vis), respectively. Those results indicate that PMS largely
distinct layered porous structure of CO-C3N4 also facilitates TCH diffu- improves TCH photocatalytic degradation efficiency of CO-C3N4. A com-
sion and improves the reaction kinetics. With PMS adding, TCH degra- parison of TCH degradation efficiency in CO-C3N4/PMS/vis system with
dation is greatly increased up to 97.77% within 40 min. Such results recently reported studies are provided in Table 3. Meanwhile, the re-
indicate that CO-C3N4 is more capable of activating PMS, and meanwhile moval efficiencies of TOC and COD in the TCH solution are also mea-
visible light irradiation also contributes to PMS activation by CO-C3N4, sured. As given in Fig. S1, the removal efficiencies of TOC and COD in
which is ascribed to carbonyl: (1) it may be the main reactive sites to CO-C3N4/PMS/vis system are higher than those in g-C3N4/PMS/vis

Table 3
The comparison with other reported studies for TCH degradation.

Photocatalyst Pollutant Concentration Dosage Time Degradation efficiency Light source Reference
(mg/L) (g/L) (min) (%)

ZnS-SnS2 TC 10 0.3 120 93.46 300 W XL (λ ≥ 420 nm) (Xia et al., 2020)
MnO2/CNK-OH-Mn TC 10 0.5 120 96.7 (H2O2) 300 W XL (λ ≥ 400 nm) (Zhang et al., 2020)
CuInS2/Bi2WO6 TCH 10 0.3 120 92.4 300 W XL (λ ≥ 420 nm) (Lu et al., 2019)
Ag/AgIn5S8 TCH 10 0.3 120 95.3 300 W XL (λ ≥ 400 nm) (Deng et al., 2018)
SnFe2O4/ZnFe2O4 TC 10 0.3 120 93.2 300 W XL (λ ≥ 400 nm) (Wang et al., 2020)
α-ZnTcPc/g-C3N4 TC 30 1.5 120 91.43 XL (λ ≥ 400 nm) (He et al., 2019)
ZnTCPP/g-C3N4 TC 30 1.5 120 80.3 XL (λ ≥ 400 nm) (Ma et al., 2019)
CO-C3N4 TCH 10 0.2 60 97.77 (PMS) 300 W XL (λ ≥ 420 nm) This work

XL: Xenon lamp; MnO2/CNK-OH-Mn: MnO2/Mn-modified alkalinized g-C3N4.

7
Y. Shi, J. Li, D. Wan et al. Science of the Total Environment 749 (2020) 142313

system, about 38.26% of TOC and 25% of COD can be removed for 5 h vis- 3.5. The influence of different water sources
ible light irradiation. Those results indicate that CO-C3N4/PMS/vis sys-
tem shows excellent TCH degradation performance with favorable Considering the practical application, three kinds of water sources
mineralization ability and potential application prospects. are used as reaction solution for TCH degradation in CO-C3N4/PMS/vis
Cycling tests are conducted to evaluate the stability of CO-C3N4/ system, including tap water (Zhengzhou Running-water Company),
PMS/vis system (Fig. 6c). TCH degradation efficiency by CO-C3N4 pre- lake water (Henan University of Technology) and reclaimed water
sents the gradually decreasing trend, which is because of possible (Wulongkou Sewage Treatment Plant, Zhengzhou). Deionized water is
blocking of active site by TCH degradation intermediates. Gratifyingly, used for comparison. Three kinds of water sources are treated as the sol-
TCH degradation efficiency still can achieve up to 85.63% after five cy- vents and then mixed with TCH to study the photocatalytic activity of
cling, suggesting the satisfactory reusability of CO-C3N4. Meanwhile, CO-C3N4/PMS/vis system and evaluate whether different water sources
XRD and FTIR comparison of the fresh and reused CO-C3N4 are provided will affect the photocatalytic performance. As presented in Fig. 6d, TCH
in Fig. S2, and there are no obvious change in the composition and struc- degradation efficiencies in the deionized water, tap water, lake water
ture for the reused sample, further indicating high stability and reusabil- and reclaimed water are respectively 98.69%, 90.69%, 84.24% and
ity of catalyst (Deng et al., 2017a, 2017b). 80.84%. Notably, TCH degradation efficiencies in real samples are

Fig. 7. Factors on the TCH degradation efficiency in CO-C3N4/PMS/vis system: (a) the initial TCH concentration, (b) catalyst dosage, (c) PMS concentration, (d) initial pH and (e) anions
(5 mM); (f) the degradation of other pollutants in the CO-C3N4/PMS/Vis system. Conditions: [RhB]0 = 10 mg/L, [CR]0 = 50 mg/L, [MO]0 = 50 mg/L, [OG]0 = 50 mg/L, [PMS] =
200 mg/L, [Cat.] = 0.2 g/L, λ > 420 nm, ambient temperature.

8
Y. Shi, J. Li, D. Wan et al. Science of the Total Environment 749 (2020) 142313

lower than that in deionized water because of the existence of compet- The influence of anions on TCH degradation is given in Fig. 7e. It is
itive or quenching substances (e.g., ions and natural organic substances) known that adding HCO− −
3 /H2PO4 can cause the higher pH in the solu-
(Qin et al., 2020), whereas they are still satisfactory. Thus, CO-C3N4/ tion compared to the addition of Cl−, NO− 2−
3 and SO4 , while such
PMS/vis system exhibits great potential for realistic wastewater above anions cause inhibition effects on TCH degradation efficiency in
treatment. an increase: Cl− < NO− −
3 < HCO3 < SO4
2−
< H2PO− 4 , which indicates
that pH change caused by anions in the solution is not main reason
3.6. Factors on the photocatalytic degradation with PMS and it may be ascribed to anions competition with TCH for the active
sites and their quenching effects on the radicals. Specifically, Cl− can
The influence of various conditions on TCH degradation by CO-C3N4 react with SO4•− to form chloride radicals (Cl•) with weaker oxidizing
are investigated in Fig. 7a–e. As observed in Fig. 7a, TCH degradation ef- (Wang et al., 2014); NO− 2− −
3 , phosphate ions (HPO4 /H2PO4 ) and HCO3


ficiency is decreased from 97.77% to 79.77% for 60 min with TCH initial exhibit the quenching effects on SO4• and/or •OH (Wang et al.,
concentration increasing from 10 mg/L to 40 mg/L. It may be that a 2017b; Zhang et al., 2018a). Totally, the order of the inhibition efficien-
higher TCH concentration will need a longer time to achieve same deg- cies by various anions in this study differs from the other studies (Gao
radation efficiency because of same reactive radicals generated by the et al., 2018a; Wang et al., 2017b; Zhang et al., 2018a), and it is closely re-
same amounts of CO-C3N4 and PMS. lated to the concentration of anions and types and amounts of reactive
Fig. 7b gives the effect of catalyst dosage on TCH degradation effi- radicals generated in those studies (Dangwang Dikdim et al., 2019;
ciency. Specifically, TCH degradation is greatly improved with catalyst Gao et al., 2018a; Wang et al., 2017b). Specifically, Wang et al.
dosage increasing from 0.1 g/L to 0.2 g/L, which is because of the in- (2017b) reported that Cl− had the inhibitory effect at low concentra-
creased availability of active sites in the catalysts to react with PMS, gen- tions and presented the acceleration effect over a critical concentration
erating more reactive radicals (Dangwang Dikdim et al., 2019). for BPA degradation in 10% Ag/mpg-C3N4/PMS/vis system.
Nevertheless, increasing catalyst dosage to 0.4 g/L has no improvement Different organic dyes are also adopted to study the universal appli-
on TCH degradation efficiency, possibly due to the obstruction of light cability of CO-C3N4 (Fig. 7f). RhB, CR, MO and OG are effectively de-
penetration at the higher catalyst dosage, reducing the surface area of graded up to 99.85%, 98.20%, 97.33% and 98.03% in 60 min,
catalysts exposed to irradiation (Dangwang Dikdim et al., 2019). respectively, indicating that CO-C3N4/PMs/vis system possesses excel-
Different PMS amounts (0.05 g/L–0.4 g/L) show the similar effect on lent versatility toward multiple pollutants degradation.
TCH degradation efficiency, especially for 0.2 g/L and 0.4 g/L (Fig. 7c). Al-
though higher PMS concentration is expected to form more reactive 3.7. Active species detection and mechanism insights
radicals, the amount of photogenerated electrons produced by certain
catalyst dosage might be not enough for PMS activation (Wang et al., The reactive oxygen species (ROS), the photogenerated electron
2017a; Wang et al., 2017b). Besides, Dangwang Dikdim et al. (2019) re- (e−) and hole (h+) may exhibit roles for TCH degradation in such CO-
ported that the excessive PMS could quench SO4•− to form SO5•− with C3N4/PMs/vis system. Active species trapping tests are carried out by
lower reactivity (E0 = 1.1 V). using various scavengers, in which p-BQ, EDTA-Na, L-histidine and
Fig. 7d presents TCH degradation efficiencies by CO-C3N4/PMS/vis TBA are used to trap O2•−, h+, 1O2 and •OH, and MeOH is adopted to
system at different initial pH. With initial pH increasing from 3.0 to trap both SO4•− and •OH. As showed in Fig. 8 and Table 4, the inhibition
11.3, there is no obvious influence on TCH adsorption efficiency. Mean- rates of p-BQ, EDTA-Na, L-histidine, MeOH and TBA toward TCH degra-
while, CO-C3N4/PMS/vis system under different initial pH shows satis- dation are respectively 90.61%, 92.99%, 82.14%, 44.75% and 23.41%, and
factory TCH degradation efficiencies; while it is worth noting that TCH thus the role of SO4•− for TCH degradation is 21.34%. The results indicate
degradation rates with the initial pH of 3.0–9.0 are similar, which that O2•−, h+, 1O2, •OH and SO4•− all contribute to TCH degradation in
might be owing to PMS acidification. The pH change after reaction in CO-C3N4/PMS/vis system, of which O2•−, h+ and 1O2 are the dominant
CO-C3N4/PMS/vis system is provided in Fig. S3, and the final pH is oxidized species.
2.88, 3.47, 4.52, 5.36 and 10.48, respectively. The pKa of PMS is 9.4 To further verify the reactive species produced in CO-C3N4/PMS/vis
(HSO− 5 → SO5
2−
+ H+, pKa = 9.4), and when the pH of solutions is system, ESR tests using DMPO, BMPO and TEMP as spin-trapping agents
below 9.4, PMS mainly occurs by the H2SO5 type, and PMS activation is conducted under dark and visible light irradiation. There occur no sig-
is inhibited (Ren et al., 2014). As a result, TCH degradation efficiency is nals of active species in PMS/vis system without catalysts (Fig. S4). For
highest at initial pH 11.3, which is owing to that PMS is better the CO-C3N4/PMS/vis system (Fig. 9a), no ESR signals occur without
decomposed to more reactive radicals under alkaline conditions for adding PMS under dark condition, and the SO4•− and •OH signals
the enhanced TCH degradation efficiency. occur under visible light irradiation and PMS. The h+ in CO-C3N4

Fig. 8. (a) TCH degradation in the presence of different scavengers and (b) the corresponding kinetic constants in CO-C3N4/PMS/vis system. Conditions: [Cat.] = 0.2 g/L, [TCH]0 = 10 mg/L,
[PMS] = 200 mg/L, λ > 420 nm, ambient temperature.

9
Y. Shi, J. Li, D. Wan et al. Science of the Total Environment 749 (2020) 142313

Table 4 that of •OH (0.0199 mmol/L) and SO4•− (0.0112 mmol/L) at 60 min in
The trapping agents, reactive species trapped, and the kinetic constants by the trapped re- CO-C3N4/PMS/vis system. The amount of SO4•− is lower during the reac-
active species.
tion, varying over the range of 0.0016–0.0112 mmol/L. This is likely
Trapping agents Reactive species k (min−1) Inhibition rate (%) owing to the transformation from SO4•− to •OH during the reaction.
blank – 0.04994 (R2 = 0.97083) 0 Herein, the proposed degradation mechanism for TCH in CO-C3N4/
p-BQ O2•− 0.00469 (R2 = 0.95985) 90.61 PMS/vis system is schematized in Fig. 10. Firstly, in the non-radical pro-
EDTA-2Na h+ 0.00350 (R2 = 0.99619) 92.99 cess, the oxidant PMS is able to degrade part of TCH (Eq. (1)). Secondly,
1
L-histidine O2 0.00892 (R2 = 0.90109) 82.14
in the radical process without irradiation, g-C3N4 possesses electron-
MeOH SO•− + •OH 0.02759 (R2 = 0.91261) 44.75
TBA •OH 0.03825 (R2 = 0.96785) 23.41 rich properties (Wei et al., 2016), the delocalized electrons from the tri-
azine units of g-C3N4, especially the C_O groups have a high chemical
Note: TCH degradation in the presence of different scavengers follows pseudo-first-order
kinetics, and the fitting correlation coefficients (R2) are given.
activity and could be transferred to activate PMS to produce •OH and
SO4•− radicals, and degrade TCH to various intermediates (Eqs. (2)–
(3)). Thirdly, after visible light irradiation, CO-C3N4 is excited to produce
(EVB = 1.55 eV) cannot react with H2O molecule or OH− to generate h+ and e− in its VB and CB, respectively (Eq. (4)). The e− can reduce the
•OH (E(•OH/OH−) = 1.99 eV, E(•OH/H2O) = 2.37 eV) (Chang et al., adsorbed O2 to O2•− (Eq. (5)), and also can be trapped by the PMS to
2015), whereas reaction between H2O/OH− and SO4•− (+2.5–3.1 produce SO4•− (Eq. (2)). The formed O2•− and SO4•− further are trans-
VNHE) can produce •OH (Tang et al., 2015). The •OH signals show clearly formed via a series of reactions to form •OH and 1O2 (Eqs. (6)–(9)).
stronger intensity than SO4•− signals. Meanwhile, O2•− and 1O2 signals Meanwhile, the h+ can either directly degrade TCH or react with PMS
also are found in CO-C3N4/PMS/vis system (Fig. 9b–c), which is owing to to form SO4•− (Eqs. (10)–(11)).
that e− in CO-C3N4 (ECB = −0.34 eV) is able to trap O2 to form O2•− (E HSO5 − þ TCH ! Products ð1Þ
(O2/O2•−) = −0.33 eV), and O2•− further reacts with •OH to form 1O2
− − − −
(Wang et al., 2014; Zhang et al., 2018a). Obviously, both O2•− and 1O2 HSO5 þe ! SO4 ˙ þ OH ð2Þ
signals are greatly stronger than •OH and SO4•− signals, indicating that
O2•− and 1O2 should be the dominant oxidized radicals in CO-C3N4/ HSO5 − þ e− ! SO4 2− þ ˙OH ð3Þ
PMS/vis system.
þ
To further quantify the concentration of common free radicals (O2•−, CO−C3 N4 þ hv ! h þ e− ð4Þ
•OH and SO4•−) in CO-C3N4/PMS/vis system, NBT, BA and p-HBA are
used as the indicators of O2•− (Li et al., 2019), •OH (Joo et al., 2005) O2 þ e− ! O2 ˙− ð5Þ
and SO4•− (Oh et al., 2017), respectively. As showed in Fig. S5, the
amount of the formed O2•− (0.0497 mmol/L) is greatly stronger than O2 ˙− þ e− þ 2Hþ ! H2 O2 ð6Þ

Fig. 9. The ESR spectra of CO-C3N4 in the presence of PMS under visible light irradiation by using (a) DMPO, (b) BMPO and (c) TEMP as the trapping agents. Conditions: [Cat.] = 0.2 g/L,
[PMS] = 200 mg/L, [DMPO] = [BMPO] = [TEMP] = 100 mM, λ > 420 nm, ambient temperature.

10
Y. Shi, J. Li, D. Wan et al. Science of the Total Environment 749 (2020) 142313

Fig. 10. The possible degradation mechanism for TCH in CO-C3N4/PMS/vis system.

H2 O2 þ e− ! OH− þ ˙OH ð7Þ 4. Conclusions

In our work, CO-C3N4 is obtained through one-step calcination


SO4 ˙− þ OH− ! SO4 2− þ ˙OH ð8Þ
using melamine-oxalic acid aggregates as precursors, and has
greater photocatalytic performance than g-C3N4. With PMS adding,
O2 ˙− þ ˙OH ! 1 O2 þ OH− ð9Þ TCH photocatalytic degradation efficiency by CO-C3N4 is greatly im-
proved and the corresponding apparent rate constant is improved
þ
h þ HSO5 − ! SO5 ˙− þ Hþ ð10Þ from 0.01966 min−1 to 0.07688 min−1. The enhanced activity may
be attributed to the stronger absorption of visible light and faster
separation of photogenerated carriers, which greatly contributes to
2SO5 ˙− ! 2SO4 ˙− þ O2 ð11Þ PMS activation into active radicals. Radical trapping tests and ESR
detection indicate that the O 2 •− , h+ and 1O 2 play the dominant
Overall, the above reactive species, including h+, O2•−,1O2, •OH and roles rather than •OH and SO4 •− for TCH degradation in CO-C3N 4/
SO4•−, contribute together to the TCH degradation. PMS/vis system. Therefore, PMS activation with carbonyl-modified
g-C 3N 4 seems to be a greatly promising approach for wastewater
3.8. The possible pathway of TCH degradation in CO-C3N4/PMS/vis system purification.

The analysis of intermediate products of TCH degradation in CO- CRediT authorship contribution statement
C3N4/PMS/vis system based on the HPLC-MS is conducted, and the
results are given from Fig. S6 and Table S3. Several intermediates Yahui Shi: Validation, Formal analysis, Visualization, Writing - Orig-
with the proposed m/z values of 476, 433, 429, 386, 362, 361, 320, inal Draft.
318, 274, 258, 242, 238, 124 and 61 are found in mass spectrometry. Jingsong Li: Conceptualization, Methodology, Resources, Data
It is known that TCH mainly contains the dimethyl amine structure Curation.
(A), ketone-amide-enol system (B) and phenol-acetone system (C), Dongjin Wan: Conceptualization, Methodology, Resources, Data
and those structures are prone to group removal, dehydration, dehy- Curation, Writing - Review & Editing, Funding acquisition.
drogenation and heterogeneous rearrangement (Dalmázio et al., Jinhui Huang: Resources, Supervision.
2007; Jiao et al., 2008; Wang et al., 2011; Zhang et al., 2019b). Herein, Yongde Liu: Resources, Supervision.
based on the analyzed intermediates, three potential routes of TCH
degradation are presented (Fig. 11). Both the I and II routes present
TCH decomposition process via a gradual removal of amine groups, Declaration of competing interest
fracture of C\\C double bonds and attack of α-position of carbonyl,
which are similar to the previous reports (Gao et al., 2019). For the The authors declare that they have no known competing financial
III routes, •OH firstly replaces the hydrogen of -N(CH 3 ) 2 (Tang interests or personal relationships that could have appeared to influ-
et al., 2017), then removes -CH3 and -CONH2, and further is oxidized ence the work reported in this paper.
to open loop and decarburized. It is worth noting that there are other
ion peaks in MS, which is due to the non-selective oxidation charac- Acknowledgments
teristic of photocatalysis. Meanwhile, based on TOC results (Fig. S1),
it is supposed that the majority of organic carbon compounds are This study was supported by the National Natural Science Founda-
converted to inorganic carbon in CO-C3N4/PMS/vis system. Hence, tion of China (51878251, 51578222) and the Doctoral Scientific Re-
at last, all the intermediates will be decomposed to small molecules, search Start-up Foundation from Henan University of Technology
like H2O, CO2, NH+ 4 , etc. (2020BS005).

11
Y. Shi, J. Li, D. Wan et al. Science of the Total Environment 749 (2020) 142313

Fig. 11. The possible TCH degradation pathways in CO-C3N4/PMS/vis system.

Appendix A. Supplementary data Dalmázio, I., Almeida, M.O., Augusti, R., Alves, T.M.A., 2007. Monitoring the degradation of
tetracycline by ozone in aqueous medium via atmospheric pressure ionization mass
spectrometry. J. Am. Soc. Mass Spectrom. 18, 679–687.
Supplementary data to this article can be found online at https://doi. Dangwang Dikdim, J.M., Gong, Y., Noumi, G.B., Sieliechi, J.M., Zhao, X., Ma, N., Yang, M.,
org/10.1016/j.scitotenv.2020.142313. Tchatchueng, J.B., 2019. Peroxymonosulfate improved photocatalytic degradation of
atrazine by activated carbon/graphitic carbon nitride composite under visible light ir-
radiation. Chemosphere 217, 833–842.
Danner, M.-C., Robertson, A., Behrends, V., Reiss, J., 2019. Antibiotic pollution in surface
References fresh waters: occurrence and effects. Sci. Total Environ. 664, 793–804.
Deng, F., Zhong, F., Lin, D., Zhao, L., Liu, Y., Huang, J., Luo, X., Luo, S., Dionysiou, D.D., 2017a.
Chang, F., Li, C., Luo, J., Xie, Y., Deng, B., Hu, X., 2015. Enhanced visible-light-driven photo- One-step hydrothermal fabrication of visible-light-responsive AgInS2/SnIn4S8
catalytic performance of porous graphitic carbon nitride. Appl. Surf. Sci. 358, heterojunction for highly-efficient photocatalytic treatment of organic pollutants
270–277. and real pharmaceutical industry wastewater. Appl. Catal. B Environ. 219, 163–172.
Chen, X., Duan, X., Oh, W.-D., Zhang, P.-H., Guan, C.-T., Zhu, Y.-A., Lim, T.-T., 2019. Insights Deng, F., Zhong, F., Zhao, L., Luo, X., Luo, S., Dionysiou, D.D., 2017b. One-step in situ hydro-
into nitrogen and boron-co-doped graphene toward high-performance peroxymono- thermal fabrication of octahedral CdS/SnIn4S8 nano-heterojunction for highly effi-
sulfate activation: maneuverable N-B bonding configurations and oxidation path- cient photocatalytic treatment of nitrophenol and real pharmaceutical wastewater.
ways. Appl. Catal. B Environ. 253, 419–432. J. Hazard. Mater. 340, 85–95.

12
Y. Shi, J. Li, D. Wan et al. Science of the Total Environment 749 (2020) 142313

Deng, F., Zhao, L., Luo, X., Luo, S., Dionysiou, D.D., 2018. Highly efficient visible-light pho- Ma, Z., Zeng, C., Hu, L., Zhao, Q., Yang, Q., Niu, J., Yao, B., He, Y., 2019. A high-performance
tocatalytic performance of Ag/AgIn5S8 for degradation of tetracycline hydrochloride photocatalyst of ZnTCPP sensitized porous graphitic carbon nitride for antibiotic deg-
and treatment of real pharmaceutical industry wastewater. Chem. Eng. J. 333, radation under visible light irradiation. Appl. Surf. Sci. 484, 489–500.
423–433. Manzetti, S., Ghisi, R., 2014. The environmental release and fate of antibiotics. Mar. Pollut.
Dong, H., Hong, S., Zhang, P., Yu, S., Wang, Y., Yuan, S., Li, H., Sun, J., Chen, G., Li, C., 2020a. Bull. 79, 7–15.
Metal-free Z-scheme 2D/2D VdW heterojunction for high-efficiency and durable Oh, W.-D., Dong, Z., Ronn, G., Lim, T.-T., 2017. Surface-active bismuth ferrite as superior
photocatalytic H2 production. Chem. Eng. J. 395, 125150. peroxymonosulfate activator for aqueous sulfamethoxazole removal: performance,
Dong, H., Xiao, M., Yu, S., Wu, H., Wang, Y., Sun, J., Chen, G., Li, C., 2020b. Insight into the mechanism and quantification of sulfate radical. J. Hazard. Mater. 325, 71–81.
activity and stability of RhxP Nano-species supported on g-C3N4 for photocatalytic Qin, F., Peng, Y., Song, G., Fang, Q., Wang, R., Zhang, C., Zeng, G., Huang, D., Lai, C., Zhou, Y.,
H2 production. ACS Catal. 10, 458–462. Tan, X., Cheng, M., Liu, S., 2020. Degradation of sulfamethazine by biochar-supported
Dong, H., Zhang, X., Li, J., Zhou, P., Yu, S., Song, N., Liu, C., Che, G., Li, C., 2020c. Construction bimetallic oxide/persulfate system in natural water: performance and reaction mech-
of morphology-controlled nonmetal 2D/3D homojunction towards enhancing photo- anism. J. Hazard. Mater. 398, 122816.
catalytic activity and mechanism insight. Appl. Catal. B Environ. 263, 118270. Ren, H.T., Jia, S.Y., Wu, Y., Wu, S.H., Han, X., 2014. Improved photochemical reactivities of
Duan, X., O’Donnell, K., Sun, H., Wang, Y., Wang, S., 2015. Sulfur and nitrogen co-doped Ag2O/g-C3N4 in phenol degradation under UV and visible light. Ind. Eng. Chem. Res.
graphene for metal-free catalytic oxidation reactions. Small 11, 3036–3044. 53, 17645–17653.
Duan, X., Ao, Z., Zhou, L., Sun, H., Wang, G., Wang, S., 2016. Occurrence of radical and Sanganyado, E., Gwenzi, W., 2019. Antibiotic resistance in drinking water systems: occur-
nonradical pathways from carbocatalysts for aqueous and nonaqueous catalytic oxi- rence, removal, and human health risks. Sci. Total Environ. 669, 785–797.
dation. Appl. Catal. B Environ. 188, 98–105. Shao, H., Zhao, X., Wang, Y., Mao, R., Wang, Y., Qiao, M., Zhao, S., Zhu, Y., 2017. Synergetic
Duan, Y., Deng, L., Shi, Z., Liu, X., Zeng, H., Zhang, H., Crittenden, J., 2020. Efficient sulfadi- activation of peroxymonosulfate by Co3O4 modified g-C3N4 for enhanced degradation
azine degradation via in-situ epitaxial grow of Graphitic Carbon Nitride (g-C3N4) on of diclofenac sodium under visible light irradiation. Appl. Catal. B Environ. 218,
carbon dots heterostructures under visible light irradiation: synthesis, mechanisms 810–818.
and toxicity evaluation. J. Colloid Interface Sci. 561, 696–707. Shi, Y., Huang, J., Zeng, G., Cheng, W., Yu, H., Gu, Y., Shi, L., Yi, K., 2018. Stable, metal-free,
Gao, H., Yang, H., Xu, J., Zhang, S., Li, J., 2018a. Strongly coupled g-C3N4 nanosheets- Co3O4 visible-light-driven photocatalyst for efficient removal of pollutants: mechanism of
quantum dots as 2D/0D Heterostructure composite for peroxymonosulfate activa- action. J. Colloid Interface Sci. 531, 433–443.
tion. Small 14, 1801353. Shi, Y., Huang, J., Zeng, G., Cheng, W., Hu, J., Shi, L., Yi, K., 2019. Evaluation of self-cleaning
Gao, Y., Zhu, Y., Lyu, L., Zeng, Q., Hu, C., 2018b. Electronic structure modulation of graphitic performance of the modified g-C3N4 and GO based PVDF membrane toward oil-in-
carbon nitride by oxygen doping for enhanced catalytic degradation of organic pol- water separation under visible-light. Chemosphere 230, 40–50.
lutants through peroxymonosulfate activation. Environ. Sci. Technol. 52. Shi, Y., Wan, D., Huang, J., Liu, Y., Li, J., 2020. Stable LBL self-assembly coating porous mem-
Gao, X., Ma, C., Liu, Y., Xing, L., Yan, Y., 2019. Self-induced Fenton reaction constructed by brane with 3D heterostructure for enhanced water treatment under visible light irra-
Fe(III) grafted BiVO4 nanosheets with improved photocatalytic performance and diation. Chemosphere 252, 126581.
mechanism insight. Appl. Surf. Sci. 467-468, 673–683.
Szekeres, E., Chiriac, C.M., Baricz, A., Szőke-Nagy, T., Lung, I., Soran, M.-L., Rudi, K., Dragos,
Guan, Y.-H., Ma, J., Ren, Y.-M., Liu, Y.-L., Xiao, J.-Y., Lin, L.-q., Zhang, C., 2013. Efficient deg- N., Coman, C., 2018. Investigating antibiotics, antibiotic resistance genes, and micro-
radation of atrazine by magnetic porous copper ferrite catalyzed peroxymonosulfate bial contaminants in groundwater in relation to the proximity of urban areas. Envi-
oxidation via the formation of hydroxyl and sulfate radicals. Water Res. 47, ron. Pollut. 236, 734–744.
5431–5438.
Tang, D., Zhang, G., Guo, S., 2015. Efficient activation of peroxymonosulfate by manganese
Guo, F., Lu, J., Liu, Q., Zhang, P., Zhang, A., Cai, Y., Wang, Q., 2018. Degradation of Acid Or-
oxide for the degradation of azo dye at ambient condition. J. Colloid Interface Sci. 454,
ange 7 by peroxymonosulfate activated with the recyclable nanocomposites of g-
44–51.
C3N4 modified magnetic carbon. Chemosphere 205, 297–307.
Tang, X., Ni, L., Han, J., Wang, Y., 2017. Preparation and characterization of ternary mag-
He, Y., Huang, Z., Ma, Z., Yao, B., Liu, H., Hu, L., Zhao, Q., Yang, Q., Liu, D., Du, D., 2019.
netic g-C3N4 composite photocatalysts for removal of tetracycline under visible
Highly efficient photocatalytic performance and mechanism of α-ZnTcPc/g-C3N4
light. Chin. J. Catal. 38, 447–457.
composites for methylene blue and tetracycline degradation under visible light irra-
Tao, Y., Ni, Q., Wei, M., Xia, D., Li, X., Xu, A., 2015. Metal-free activation of peroxymonosul-
diation. Appl. Surf. Sci. 498, 143834.
fate by g-C3N4 under visible light irradiation for the degradation of organic dyes. RSC
He, Y.Q., Zhang, F., Ma, B., Xu, N., Binnah Junior, L., Yao, B., Yang, Q., Liu, D., Ma, Z., 2020.
Adv. 5, 44128–44136.
Remarkably enhanced visible-light photocatalytic hydrogen evolution and antibiotic
Wacławek, S., Lutze, H.V., Grübel, K., Padil, V.V.T., Černík, M., Dionysiou, D.D., 2017. Chem-
degradation over g-C3N4 nanosheets decorated by using nickel phosphide and gold
istry of persulfates in water and wastewater treatment: a review. Chem. Eng. J. 330,
nanoparticles as cocatalysts. Appl. Surf. Sci. 517, 146187.
44–62.
Huang, Y.-H., Liu, Y., Du, P.-P., Zeng, L.-J., Mo, C.-H., Li, Y.-W., Lü, H., Cai, Q.-Y., 2019. Occur-
rence and distribution of antibiotics and antibiotic resistant genes in water and sed- Wang, Y., Zhang, H., Zhang, J., Lu, C., Huang, Q., Wu, J., Liu, F., 2011. Degradation of tetra-
cycline in aqueous media by ozonation in an internal loop-lift reactor. J. Hazard.
iments of urban rivers with black-odor water in Guangzhou, South China. Sci. Total
Environ. 670, 170–180. Mater. 192, 35–43.
Jiang, X., Li, J., Fang, J., Gao, L., Cai, W., Li, X., Xu, A., Ruan, X., 2017. The photocatalytic per- Wang, S., Li, D., Cheng, S., Yang, S., Yuan, G., He, H., 2014. Synthesis and characterization of
formance of g-C3N4 from melamine hydrochloride for dyes degradation with peroxy- g-C3N4/Ag3VO4 composites with significantly enhanced visible-light photocatalytic
monosulfate. J. Photochem. Photobiol. A Chem. 336, 54–62. activity for triphenylmethane dye degradation. Appl. Catal. B Environ. 144, 885–892.
Jiao, S., Zheng, S., Yin, D., Wang, L., Chen, L., 2008. Aqueous photolysis of tetracycline and Wang, G., Chen, S., Quan, X., Yu, H., Zhang, Y., 2017a. Enhanced activation of peroxymono-
toxicity of photolytic products to luminescent bacteria. Chemosphere 73, 377–382. sulfate by nitrogen doped porous carbon for effective removal of organic pollutants.
Joo, S.H., Feitz, A.J., Sedlak, D.L., Waite, T.D., 2005. Quantification of the oxidizing capacity Carbon 115, 730–739.
of nanoparticulate zero-valent iron. Environ. Sci. Technol. 39, 1263–1268. Wang, Y., Zhao, X., Cao, D., Wang, Y., Zhu, Y., 2017b. Peroxymonosulfate enhanced visible
Kong, L., Yan, J., Liu, S.F., 2019. Carbonyl linked carbon nitride loading few layered MoS2 light photocatalytic degradation bisphenol a by single-atom dispersed Ag mesopo-
for boosting photocatalytic hydrogen generation. ACS Sustain. Chem. Eng. 7, rous g-C3N4 hybrid. Appl. Catal. B Environ. 211, 79–88.
1389–1398. Wang, M., Guo, P., Zhang, Y., Lv, C., Liu, T., Chai, T., Xie, Y., Wang, Y., Zhu, T., 2018. Synthesis
Lee, H., Lee, H.-J., Jeong, J., Lee, J., Park, N.-B., Lee, C., 2015. Activation of persulfates by car- of hollow lantern-like Eu(III)-doped g-C3N4 with enhanced visible light photocata-
bon nanotubes: oxidation of organic compounds by nonradical mechanism. Chem. lytic perfomance for organic degradation. J. Hazard. Mater. 349, 224–233.
Eng. J. 266, 28–33. Wang, M., Jin, C., Li, Z., You, M., Zhang, Y., Zhu, T., 2019a. The effects of bismuth (III) dop-
Li, C., Yu, S., Dong, H., Liu, C., Wu, H., Che, H., Chen, G., 2018a. Z-scheme mesoporous ing and ultrathin nanosheets construction on the photocatalytic performance of gra-
photocatalyst constructed by modification of Sn3O4 nanoclusters on g-C3N4 nano- phitic carbon nitride for antibiotic degradation. J. Colloid Interface Sci. 533, 513–525.
sheets with improved photocatalytic performance and mechanism insight. Appl. Wang, M., Li, Z., Tian, L., Xie, Y., Han, J., Liu, T., Jin, C., Wu, Z., 2019b. A facile synthesis of
Catal. B Environ. 238, 284–293. nano-layer structured g-C3N4 with efficient organic degradation and hydrogen evolu-
Li, Y., Ho, W., Lv, K., Zhu, B., Lee, S.C., 2018b. Carbon vacancy-induced enhancement of the tion using a MDN energetic material as the starting precursor. Int. J. Hydrog. Energy
visible light-driven photocatalytic oxidation of NO over g-C3N4 nanosheets. Appl. 44, 4102–4113.
Surf. Sci. 430, 380–389. Wang, J., Zhang, Q., Deng, F., Luo, X., Dionysiou, D.D., 2020. Rapid toxicity elimination of
Li, H., Deng, F., Zheng, Y., Hua, L., Qu, C., Luo, X., 2019. Visible-light-driven Z-scheme rGO/ organic pollutants by the photocatalysis of environment-friendly and magnetically
Bi2S3–BiOBr heterojunctions with tunable exposed BiOBr (102) facets for efficient recoverable step-scheme SnFe2O4/ZnFe2O4 nano-heterojunctions. Chem. Eng. J. 379,
synchronous photocatalytic degradation of 2-nitrophenol and Cr(VI) reduction. Envi- 122264.
ron. Sci.: Nano 6, 3670–3683. Wei, M., Gao, L., Li, J., Fang, J., Cai, W., Li, X., Xu, A., 2016. Activation of peroxymonosulfate
Liao, Y., Zhu, S., Ma, J., Sun, Z., Di, Z., 2014. Tailoring the morphology of g-C3N4 by self- by graphitic carbon nitride loaded on activated carbon for organic pollutants degra-
assembly towards high photocatalytic performance. Chemcatchem 6, 3419–3425. dation. J. Hazard. Mater. 316, 60–68.
Lin, K.-Y.A., Chen, Y.-C., 2016. Accelerated decomposition of Oxone using graphene-like Xia, B., Deng, F., Zhang, S., Hua, L., Luo, X., Ao, M., 2020. Design and synthesis of robust Z-
carbon nitride with visible light irradiation for enhanced decolorization in water. scheme ZnS-SnS2 n-n heterojunctions for highly efficient degradation of pharmaceu-
J. Taiwan Inst. Chem. Eng. 60, 423–429. tical pollutants: performance, valence/conduction band offset photocatalytic mecha-
Lin, K.-Y.A., Zhang, Z.-Y., 2017. Degradation of bisphenol A using peroxymonosulfate acti- nisms and toxicity evaluation. J. Hazard. Mater. 392, 122345.
vated by one-step prepared sulfur-doped carbon nitride as a metal-free heteroge- Yan, Hongjian, 2012. Soft-templating synthesis of mesoporous graphitic carbon nitride
neous catalyst. Chem. Eng. J. 313, 1320–1327. with enhanced photocatalytic H2 evolution under visible light. Chem. Commun. 48,
Lu, X., Che, W., Hu, X., Wang, Y., Zhang, A., Deng, F., Luo, S., Dionysiou, D.D., 2019. The fac- 3430–3432.
ile fabrication of novel visible-light-driven Z-scheme CuInS2/Bi2WO6 heterojunction Yang, X., Qian, F., Zou, G., Li, M., Lu, J., Li, Y., Bao, M., 2016. Facile fabrication of acidified g-
with intimate interface contact by in situ hydrothermal growth strategy for extraor- C3N4/g-C3N4 hybrids with enhanced photocatalysis performance under visible light
dinary photocatalytic performance. Chem. Eng. J. 356, 819–829. irradiation. Appl. Catal. B Environ. 193, 22–35.

13
Y. Shi, J. Li, D. Wan et al. Science of the Total Environment 749 (2020) 142313

Yi, X., Yuan, J., Tang, H., Du, Y., Hassan, B., Yin, K., Chen, Y., Liu, X., 2020. Embedding few- Zhang, C., Wang, W., Duan, A., Zeng, G., Huang, D., Lai, C., Tan, X., Cheng, M., Wang, R.,
layer Ti3C2Tx into alkalized g-C3N4 nanosheets for efficient photocatalytic degrada- Zhou, C., Xiong, W., Yang, Y., 2019a. Adsorption behavior of engineered carbons and
tion. J. Colloid Interface Sci. 571, 297–306. carbon nanomaterials for metal endocrine disruptors: experiments and theoretical
Zhang, J., Shao, X., Shi, C., Yang, S., 2013. Decolorization of Acid Orange 7 with peroxymo- calculation. Chemosphere 222, 184–194.
nosulfate oxidation catalyzed by granular activated carbon. Chem. Eng. J. 232, Zhang, C., Zeng, G., Huang, D., Lai, C., Chen, M., Cheng, M., Tang, W., Tang, L., Dong, H.,
259–265. Huang, B., Tan, X., Wang, R., 2019b. Biochar for environmental management: mitigat-
Zhang, J.-J., Ge, J.-M., Wang, H.-H., Wei, X., Li, X.-H., Chen, J.-S., 2016. Activating oxygen ing greenhouse gas emissions, contaminant treatment, and potential negative im-
molecules over carbonyl-modified graphitic carbon nitride: merging supramolecular pacts. Chem. Eng. J. 373, 902–922.
oxidation with photocatalysis in a metal-free catalyst for oxidative coupling of Zhang, Q., Peng, Y., Deng, F., Wang, M., Chen, D., 2020. Porous Z-scheme MnO2/Mn- mod-
amines into imines. Chemcatchem 8, 3441–3445. ified alkalinized g-C3N4 heterojunction with excellent Fenton-like photocatalytic ac-
Zhang, J., Zhao, X., Wang, Y., Gong, Y., Cao, D., Qiao, M., 2018a. Peroxymonosulfate- tivity for efficient degradation of pharmaceutical pollutants. Sep. Purif. Technol.
enhanced visible light photocatalytic degradation of bisphenol A by perylene 246, 116890.
imide-modified g-C3N4. Appl. Catal. B Environ. 237, 976–985. Zhao, H., Yu, H., Quan, X., Chen, S., Zhang, Y., Zhao, H., Wang, H., 2014. Fabrication of
atomic single layer graphitic-C3N4 and its high performance of photocatalytic disin-
Zhang, J., Zhao, X., Wang, Y., Gong, Y., Cao, D., Qiao, M., 2018b. Peroxymonosulfate-
fection under visible light irradiation. Appl. Catal. B Environ. 152-153, 46–50.
enhanced visible light photocatalytic degradation of bisphenol A by perylene
imide-modified g-C3N4. Appl. Catal. B Environ. 237, 976–985.

14

You might also like