Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Chemical Engineering Journal 434 (2022) 134732

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Efficient degradation of tetracycline by persulfate activation with Fe, Co


and O co− doped g− C3N4: Performance, mechanism and toxicity
Zhibin Wu a, b, *, Zhijun Tong a, b, Yuanyuan Xie c, Haibo Sun a, b, Xiaomin Gong a, b,
Pufeng Qin a, b, *, Yunshan Liang a, b, *, Xingzhong Yuan d, Dongsheng Zou a, b, Longbo Jiang d
a
College of Resources and Environment, Hunan Agricultural University, Changsha 410128, PR China
b
Key Laboratory for Rural Ecosystem Health in the Dongting Lake Area of Hunan Province, Changsha 410128, PR China
c
PowerChina Zhongnan Engineering Corporation Limited, Changsha 410014, PR China
d
College of Environmental Science and Engineering, Hunan University, Changsha 410082, PR China

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

Keywords: In this work, the porously thin layered Fe, Co and O co − doped g − C3N4 (FCOCN) was synthesized as persulfate
Heteroatoms co− doped g− C3N4 (PS) activator for tetracycline (TC) degradation. The characterization demonstrated that Fe, Co and O doping
Persulfate activation could boost the specific surface area and modulate the electronic structure of g − C3N4. With addition of PS, the
Tetracycline
FCOCN2 exhibited the maximum TC removal efficiency of 90.1 % after 120 min reaction, and the kinetic con­
Degradation pathway
Products toxicity
stant was 2.67 times that of pure g–C3N4. Meanwhile, the removal ratio of total organic carbon and reaction
stoichiometric efficiency of PS could reach to 68.6% and 5.2%, respectively. When the addition of catalyst and PS
were 0.6 g/L and 10.5 mM, respectively, the lower pH and TC concentration, as well as higher temperature were
conducive to degradation reaction with activation energy at 23.88 kJ.mol− 1. As indicated by work function, a
lower internal electronic excitation energy barrier enhanced electron transfer, so the active species of •O2− , 1O2,
SO4•− , SO5•− and •OH could be favorably generated on over metal (II/III) redox shuttles, electron − rich O/N
and electron–deficient C active sites. The quantitative structure − toxicity relationship prediction of in­
termediates and the acute toxicity experiment showed that the toxicity of water treated for 60 min to V. fischeri,
E. coli and B. subtilis was significantly reduced. Moreover, after five consecutive degradation cycles, the removal
efficiency of TC still remained at 72.7%. Besides, even in matrices of tap water, river water, municipal waste­
water and medical wastewater, the elimination proportion of TC by FCOCN2/PS were 90.0%, 88.6%, 74.0% and
81.5%, respectively. Hence, the FCOCN2/PS system may be a promising alternative for highly effective treatment
of antibiotic contained wastewater and environmental protection.

1. Introduction and genetic/biological toxicity [5]. Facing this urgent issue, enormous
efforts have been invested in developing efficient physicochemical
As the second–most antibiotic produced and consumed in the world, method, e.g., adsorption [6], flocculation [7], chemical oxidation [8,9],
tetracycline (TC) has been frequently applied in medical and animal etc., for removal of TC from effluent. Inherited from the classical
husbandry for bacterial infections treatment and livestock growth pro­ chemical oxidation, advanced oxidation process (AOPs) with strong
motion [1]. However, due to the poorly metabolizable property of TC in oxidizing free radicals has emerged as a promising alternative for
vivo, vast majority of TC and its metabolites are discharged into water complete degradation of TC to low toxicity substances [10,11].
body through excreta [2]. In hospital and pharmaceutical industry Among several available technologies, sulfate radical (SO4•− ) based
wastewater, the concentration of TC could reach to 1077 mg/L, and AOPs (SR–AOPs) has been regarded as an effective, eco − friendly and
even after biochemical treatment, the level of TC still remains at 0.38 – adaptable technology for organic pollutants mineralization at the wide
2.0 mg/L [3,4]. Moreover, the long − term residues of TC with stable pH range [12]. The SO4•− radical has a comparable or even higher
structure in water environment may destroy the aquatic ecosystem and oxidation potential (SO4•− , 2.6 – 3.1 V vs. •OH, 1.8 – 2.7 V) and longer
threaten the life health by inducing drug − resistant bacteria emergence half–lifespan (SO4•− , 30 – 40 μs vs. •OH, <1 μs) than •OH [13].

* Corresponding authors at: College of Resources and Environment, Hunan Agricultural University, Changsha 410128, PR China.
E-mail addresses: wzbaaa11@hunau.edu.cn, wzbaaa11@163.com (Z. Wu), qinpufeng@126.com (P. Qin), lyss3399@126.com (Y. Liang).

https://doi.org/10.1016/j.cej.2022.134732
Received 30 October 2021; Received in revised form 2 January 2022; Accepted 12 January 2022
Available online 17 January 2022
1385-8947/© 2022 Elsevier B.V. All rights reserved.
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

Generally, it could be generated by activation of persulfates (PS), e.g., 2. Materials and methods
peroxymonosulfate (E0(HSO5–/SO42–) = 1.75 VNHE) or peroxydisulfate
(E0(S2O82–/SO42–) = 1.96 VNHE) through various approaches, including 2.1. Materials
light irradiation [14,15], thermal treatment [16], ultrasonic irradiation
[17], electro–chemistry [8], semiconductors [15], magnetic and non­ The Fe(NO3)3⋅9H2O (>98.5%), Co(NO3)2⋅6H2O (>99%), urea
–magnetic carbon/transition metal catalysts [9,12,18]. Taking advan­ (CH4N2O, > 99%) Na2S2O8 (PS, >90%) and tetracycline (TC, 99.5%)
tage of no extra energy input, carbon (e.g., biochar, graphene, carbon were purchased from Macklin Reagent Network (Shanghai, China).
nanotubes, etc.) and transition metal (e.g., Fe2+, Ni2+, Co2+, Mn2+, etc.) River water, municipal wastewater and medical wastewater were ob­
catalysts have been intensively investigated for PS activation, in which tained from Liuyang River, a Municipal Sewage Treatment Plants and a
free radicals (e.g., SO4•− , •OH, •O2− , etc.) and/or nonradical species (e. Community Hospital, respectively, in Changsha, China. All other
g., 1O2, surface − bound complex, etc.) may be simultaneously generated analytical grade chemicals and the ultrapure water resistivity at 18.25
for organic pollutants degradation [19–22]. Nevertheless, a large num­ MΩ.cm were used throughout the experiments.
ber of previous studies have proved that the secondary pollution caused
by metal ions leaching and the limited catalytic capacity of pristine 2.2. Synthesis of Fe, Co and O co − doped g − C3N4
carbon material are the intrinsic drawbacks [20,23]. Therefore, the
strategy of coupling transition metal species into carbon skeleton has The Fe, Co and O co − doped g − C3N4 were synthesized by a simple
been attempted to tailor the carbon/transition metal catalysts properties calcination route with the mixture of urea, Fe3+ and Co2+. Briefly, 1
[24–26]. For instance, after Co embedded into N, S co − doped carbon, mmol metal sourcewith different mole ratio of Fe3+/Co2+ was mixed
the Co and doped N,S element have synergistic effect to improve the with 18 g of urea and 10 ml ultrapure water in a 100 ml ceramic crucible
Fermi energy and activity of C atoms, which makes C atoms could act as by ultrasonication for 30 min. Then, the suspensions were thoroughly
catalytic active center with more positive charge [27]. Likewise, the dried at 333 K, and subsequently heated at 823 K for 4 h under the
coordination of iron with N contained species (e. g., –NH2, –NH–, N–) self–sustained atmosphere with a heating rate of 10 K/min in a muffle
and carbonyl group (C = O) could provide more electron − rich sp2 furnace. The products with mole ratio Fe3+/Co2+ at 7:1, 9:1 and 11:1 are
carbon and variable chemical valenced iron as active sites to achieve denoted as FCOCN1, FCOCN2 and FCOCN3, respectively. For compari­
favorable nonradical and radical reaction in Fe@N − doped graphene − son, the Fe doped g − C3N4, Co doped g − C3N4 and pure g − C3N4 were
like carbon/PS system [28]. prepared in absent of Co2+, Fe3+ and metal (Fe3+ and Co2+) source,
Last several years, much attention have focused on the modification respectively, which is referred to as FCN, CCN and CN, respectively.
of emerging graphitic carbon nitride (g − C3N4) as PS activator for Also, by setting Fe3+/Co2+ ration at 9:1, the CoFe sample was prepared
pollutants elimination [29]. g − C3N4, an earth − abundant carbon and without urea and the Fe and Co co − doped g − C3N4 (FCCN2) was
nitrogen elements conjugated polymer, has superior advantages of synthesized under N2 atmosphere for investigation of catalytic degra­
unique two − dimensional structure, high stability, easy preparation and dation discrepancy.
environment friendly in cost − effectively catalytic application [30,31].
Moreover, the heptazine rings with N based ligands of six nitrogen lone 2.3. Characterization methods
pair electrons in g − C3N4 could serve as electron donors for high­
–affinity coordination with metal ions, which can overcome poor elec­ The morphologies of the samples were observed using the environ­
tronic mobility and relatively low surface area (10 – 20 m2/g) of bulk g mental scanning electron microscope (SEM, FEI QuANTA 200, USA),
− C3N4 [32,33]. Previous reports have proved that the metal doping in g and transmission electron micro − scopy (TEM, JEM − 3010, Japan).
− C3N4 could not only modulate the electron density on g − C3N4 surface The crystal structure was recorded by using X–ray diffraction (XRD)
through formation of Me − Nx (Me = Fe, Co, Mn, Cu, etc.) bond, but also patterns with Bruker AXS D8 Advance diffractometer with Cu–Ka beam
significantly enlarge the specific surface area and active sites for PS source (λ = 1.541 Å). The surface elemental composition were analyzed
activation and various pollutants degradation [32–36]. What particu­ by the X − ray photoelectron spectroscopy (XPS, Thermo Fisher Scien­
larly striking is that when metal ions and non − metallic elements are co tific–K–Alpha 1063, UK) with a resolution of 0.5 eV. The ultraviolet
− doped into g − C3N4, a synergistic effect of electronic reconstruction photoemission spectroscopy (UPS) were detected by a Thermo spec­
could be obtained [37,38]. After formation of C − O − Cu − N bonds in trometer (ESCALAB 250 XI). The Brunauere–Emmette–Teller (BET)
Cu, O co − doped g − C3N4, the electron − rich Cu, O sites and electron specific surface area and pore size were measured by using automatic
− deficient C site could react with PS to generate metastable in­ surface analyzer (Quantachrome, USA). The zeta potentials analysis of
termediates and 1O2 for synergistic decomposition of pollutant [39]. FCOCN2 in water solution at pH 3.0–11.0 (adjusted by 0.1 M NaOH or
Therefore, inspired by such previous excellent works, Fe, Co and O are HCl) was determined with a zeta potential meter (Zetasizer Nano–ZS90,
expected to hybridize into g − C3N4 for PS activation, and to best our Malvern). The electron spin resonance (ESR) signals were conducted on
knowledge, rare attention have been concerned on degradation pathway a Bruker model ESR JES − FA200 spectrometer with trapping agent of
and toxicity change of TC degradation over Fe, Co and O co − doped g − 100 mM, in which •SO4− , •OH and •O2− radicals were spin − trapped by
C3N4. 5, 5 − dimethyl − l − pyrroline N − oxide (DMPO), and 1O2 was spin −
Herein, Fe, Co and O co − doped g − C3N4 was synthesized via trapped by 2,2,6,6 − tetramethylpiperidine (TEMP). Electrochemical
one–step calcination method and the catalytic activity for PS activation properties measurements were performed on a CHI 660D electro­
was measured by TC degradation. The physicochemical properties chemical workstation (Chenhua, China) using a three–electrode cell
including morphology structure, elemental composition, electro­ with the nanostructure materials on FTO as the working electrode,
chemical properties, etc. of Fe, Co and O co − doped g − C3N4 are sys­ saturated Ag/AgCl and platinum electrode as the counter electrode and
tematically characterized. The influence factors of PS dosage, solid the reference electrode, respectively.
liquid ratio, initial TC concentration, reaction pH and temperature on TC
degradation were examined. The production of radicals and non − 2.4. Catalytic performance evaluation
radical, and degradation intermediates were detected to reflect the
degradation mechanism and pathways. In addition, the toxicity of All the catalytic degradation experiments were carried out in the
degradation intermediates, the treatment of real wastewater and the dark. Typically, 60 mg catalysts was magnetically stirred with 100 ml TC
reusability of Fe, Co and O co − doped g − C3N4 were measured to (15 mg/L) until it reaches the adsorption − desorption equilibrium.
evaluate its economic and environmental friendliness. After that, 250 mg PS was rapidly added into the solution to initiate TC
degradation process. In order to monitor the concentration change of TC

2
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

during the degradation process, 3 ml of sample solution was withdrawn of 2θ = 13.0◦ and 27.5◦ , which correspond to the (1 0 0) and (0 0 2) facets
by pipette at a certain interval. After solid − liquid separation, the re­ with d − spacing of 0.68 nm and 0.32 nm, respectively, agreeing with g
sidual concentration of TC was measured at wavelength of 357 nm by − C3N4 reported elsewhere [40]. These two characteristic peaks are
UV–vis spectrophotometer (UV − 2250, SHIMADZU Corporation, arisen from the structural packing of tri − s − triazine units in the lattice
Japan). The concentration before and after adsorption equilibrium was plane and the stacking of sp2 hybrid orbitals for conjugated aromatic
defined as the initial concentration (C0) and post–adsorption concen­ system, respectively [32]. Compared with CN patterns, the main peak of
tration (Ca). The total organic carbon (TOC) level of samples were (0 0 2) plane for doped CN is still preserved, suggesting that the doping
determined by using a TOC analyzer (TOC–2000, Shanghai Metash In­ has not essential effect on graphitic stacking of g − C3N4. But it is worth
struments Corporation. China). The PS concentration before and after noting that the position of (0 0 2) diffraction peak for FCOCN has been
reaction was determined by UV–VIS spectrophotometer at 352 nm, in slightly blue − shifted from to 2θ = 27.3◦ (Fig. S1, Supporting Infor­
which 0.1 ml sample was pretreated for 15 min by 2 g KI and 0.1 g mation), implying the expansion of interlayer distances after interstitial
NaHCO3 in diluent with a total volume of 20 ml. All experiments were sites doping of heteroatom into in − planar g − C3N4 [32,36].
run in triplicates and the results were reproducible within 5 % of relative Furthermore, after Fe, Co and O co − doped into the CN, the peak in­
standard deviation. The generation and role of radicals SO4•− , •OH, tensity of (1 0 0) plane is greatly decreased and even disappeared. During
SO5•− , •O2− and 1O2 was investigated by radical capture experiment. reaction process, the heteroatoms of Fe, Co and O may exhibit the
The FCOCN2 after reaction was collected centrifugation and washed inhibitory effect and host–guest interaction with carbon − nitrogen
with water to remove residual impurities, and then dried at 60 ◦C for framework, in which the in − plane structure of CN could be partially
recycling. fragmented into smaller crystallites, leading to the formation of edge
defects and in − plane cavities as new active sites for redox reaction
[34]. Interestingly, there are no impurity peaks belonging to Fe or Co
2.5. Analysis of degradation intermediates and toxicity measurement species, e.g., oxides, which suggests that the Fe and Co are appropriately
coordinated into the host of g − C3N4 in the form of Fe − Nx and Co − Nx
The degradation intermediates of TC during catalytic process were bonds [38].
detected by high performance liquid chromatography mass spectrom­ The morphological features of CN, FCN, CCN and FCOCN2 were
etry (HPLC − MS, Thermo Scientific Q Exactive, USA). The theoretical observed by scanning electron microscopy (SEM). As illustrated in
toxicity of degradation intermediates was predicted by Toxicity Esti­ Fig. 2a, the pure CN presents loosely thin flake network with abundant
mation Software Tool program (T.E.S.T., version 5.1.1) depended on porous resulted from gas release in self − polymerization of urea. After
Quantitative Structure Activity Relationships (QSAR) method. The real doping of heteroatoms into CN, the FCN (Fig. 2b), CCN (Fig. 2c) and
toxic effects of solution samples at different degradation time on FCOCN2 (Fig. 2d) tend to display small sheets with irregular wrinkles,
bioluminescence inhibition of V. fischeri (NRRL B − 11177) was per­ which may be ascribed to the disruption and inhibition poly­
formed according to standard biotest (ISO11348 − 3, 2007). The effect condensation reaction for the formation of smaller microparticles
of TC and degradation intermediates on E. coli and B. subtilis growth was [41,42]. This change in morphological structure will significantly pro­
evaluated by colony method. 100 μL of bacteria solution in logarithmic mote the exposure of redox–active centers and the accessibility of per­
growth period that had been diluted 103 times was evenly spread on the sulfate and pollutant, so as to improve catalytic activity for pollutants
agar plate containing 10% TC degradation liquid for 12 h culture at degradation [43]. Especially, FCOCN2 (Fig. 2d and 2e) exhibits relative
37 ◦ C. Before that, all solution samples were treated by adding ascorbic laminar stacked phase with porous composited of thin wrinkles, indi­
acid for termination of degradation reaction. cating the successful co − doping of Fe, Co and O into g − C3N4. It can
also be confirmed by energy dispersive spectrum (EDS) of FCOCN2
3. Results and discussions (Fig. 2f). The signals of C, N, Fe, Co and O elements are clearly detected,
and all of them are uniformly distributed on FCOCN2 surface supported
3.1. Characterization by elemental mapping analysis. The doping of O element is mainly
ascribed to the oxidation of C and N, in which the interactions of Fe, Co
The X − ray diffraction patterns (XRD) of pristine CN, FCN, CCN and with urea molecules could capture oxygen into g − C3N4 structure[34].
FCOCN samples were presented in Fig. 1. As can be seen, two typical Further detailed structure of CN, FCN, CCN and FCOCN2 were
polymeric layered structure peaks of g − C3N4 locate at diffraction angle characterized by transmission electron microscopy (TEM). The crinkly
CN (Fig. 3a) exhibits very thin silk − veil morphology with numerous
irregular pore. As revealed in Fig. 3b, the bulk FCN presents thickly
stacked block characteristics with remarkable decrease in particle size,
agreeing with SEM observations. After implantation of Co into CN
(Fig. 3c), the framework of CCN turns to become more loose with more
pores, which could provide more edge defects and in − plane cavities as
redox active site for PS activation and pollutant degradation [41]. In
Fig. 3d, the above − mentioned porous morphology of aggregated
irregular block − based flakiness with wrinkles and curls could be
clearly observed in FCOCN2. Furthermore, as illustrated in Fig. S2
(Supporting Information), high resolution TEM images (HRTEM) and
selected–area electron diffraction (SAED) patterns exhibit that all CN,
FCN, CCN and FCOCN2 have the similar chaotic diffraction ripple and
amorphous diffraction ring, and no other diffraction owing to iron and/
or cobalt oxide nanoparticles could be observed in FCOCN2, which
verifies that Fe and Co are substantially doped into the g − C3N4.
The surface chemical composition and elements valence states of
FCOCN2 sample were analyzed by X − ray photoelectron spectroscopy
(XPS) technique. As presented in survey scan spectra (Fig. 4a), the ele­
ments of C 1 s, N 1 s, O 1 s, Fe 2p and Co 2p are observed at binding
Fig. 1. XRD patterns of various catalysts. energies of 288.1 eV, 398.7 eV, 532.0 eV, 710.0 eV and 785.1 eV,

3
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

Fig. 2. The SEM images of CN (a), FCN (b), CCN (c), FCOCN2 (d, e) and the EDS spectra of FCOCN2 (f) and corresponding element mapping.

respectively. From Fig. 4b, the de–convoluted peaks of C 1 s are found at (782.4 eV) and Fe − Nx (709.7 eV) bonds by hybridization of empty orbit
binding energies of 284.7 eV, 286.5 eV and 288.8 eV, which are of dopants and lone pair electrons of N, which is consistent with pre­
attributed to the C − C, N − C − O and N − C = N moieties respectively vious researches [36,48]. Therefore, the Fe, Co accompanied O atoms
[39]. The high–resolution spectra of N 1 s (Fig. 4c) can be split into four have been successfully inserted into the interstitial position of CN.
convex peaks ascribing to sp2 hybridized aromatic nitrogen of C = N–C The specific surface area and porosity of CN, FCN, CCN and FCOCN2
(398.7 eV), tertiary nitrogen of N–(C)3 linking structural motif C6N7 were measured by conducting N2 adsorption–desorption isotherms, and
(400.2 eV), amino groups of C–NH2 (401.1 eV) and the π–excitations the results are demonstrated in Fig. 5a. Evidently, all tested samples
(404.6 eV) [44,45]. Fig. 4d shows the O 1 s spectrum that contains three exhibit the type IV isotherms with hysteresis loops of H3 in the range of
kinds of oxygen species at binding energies of 530.8 eV (C − O), 531.9 relative pressure (P/P0) from 0.8 to 1.0, hinting the formation of mes­
eV (–OH) and 533.3 eV (O − M) [39]. The presence of oxygenated in­ opore in CN, FCN, CCN and FCOCN2 [49]. Moreover, the doped samples
gredients suggests that the O atoms have directly inserted in lattice of g show stronger adsorption capacity than pure g − C3N4 towards N2,
− C3N4 by connecting with C atoms, which is resulted from the which suggests that all doped specimens have significantly increased
incomplete polymerization and disruption of aromatic layers stacking specific surface area. According to the model of Brunauere–Emmette–­
degree after metal doping [46]. Further analysis of Fe 2p and Co 2p Teller (BET), the specific surface area of CN, FCN, CCN and FCOCN2 are
spectrum with shakeup satellite peaks (Fe 2p: 715.7 eV and 718.8 eV, Co calculated to be 28.59 m2/g, 36.70 m2/g, 51.56 m2/g and 55.66 m2/g,
2p: 787.6 eV and 804.0 eV) are depicted in Fig. 4e and Fig. 4f, respec­ respectively. The improvement in specific surface area may be associ­
tively. The binding energies of Fe 2p3/2 and Fe 2p1/2 with Gaussian ated to the promotion of secondary particles formation caused by doping
curves fitting at 709.7 eV and 723.7 eV are ascribed to Fe2+, and two inhibitory effect on crystal growth of g − C3N4, which could benefit the
other peaks at 712.0 eV and 730.8 eV are attributed to Fe3+ [47]. adsorption and diffusion of PS and pollutant molecules with more
Analogously, the electronic configuration of Co 2p3/2 and Co 2p1/2 at exposed active sites [5]. The related porosity distribution calculated by
binding energy of 779.5 eV and 793.8 eV belong to the Co3+ state, while original density functional theory are depicted in Fig. 5b. It can be
that at 782.4 eV and 799.5 eV correspond to the Co2+ state [32]. These observed that doped samples show broader pore size distribution be­
results indicate that Fe and Co in doped g − C3N4 exist in the form of tween 0 and 180 nm, and the FCOCN2 has the largest pore volume
mixed valence states due to the reduction of Fe3+/Co3+ by reducing among them. Concretely, the average pore volume of CN, FCN, CCN and
agents (i.e., CO and CNH2). The presence of Fe3+/Fe2+ and Co3+/Co2+ FCOCN2 is 0.142 cm3/g, 0.159 cm3/g, 0.287 cm3/g and 0.325 cm3/g,
redox shuttles may act as the active sites in PS activation via inter­ respectively, and the corresponding pore size is 19.76 nm, 17.36 nm,
–valence electron transfer. Furthermore, since the greater electronega­ 22.18 nm 23.35 nm, respectively. Besides, the average particle size of
tivity of N (3.04) than that of Co (1.88), Fe (1.83) and C (2.55), the Co CN, FCN, CCN and FCOCN2 is estimated to be 209.88 nm, 163.49 nm,
and Fe atoms could partially substitute for C atom to form Co − Nx 116.38 nm and 107.79 nm, respectively, which confirms the

4
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

Fig. 3. TEM image of CN (a), FCN (b), CCN (c) and FCOCN2 (d).

fragmentation of g − C3N4 into secondary particles by doping inhibitory However, the excessive Fe (FCOCN3) or Co (FCOCN1) will occupy a part
effect. of active sites of Co (or Fe), leading to the weakened PS activation re­
action for TC decomposition [38,52]. Therefore, the optimum doping
3.2. Catalytic performance evaluation ratio of Fe/CO is selected at 9:1 in the follow–up study. As presented in
Fig. S3 and Fig.S4 (Supporting Information), the FCOCN2 with best
The effects of several factors including doping ratio, PS dosage, catalytic ability among doped g − C3N4 samples also has the higher TC
catalysts dosage, initial TC concentration, pH and temperature on TC removal efficiency than CoFe (80.9%) prepared without urea and
degradation efficiency were investigated. As demonstrated in Fig. 6a, FCCN2 (84.0%) prepared under N2 (Fig.S4), indicating that the co −
the removal efficiency of TC is only 30.3% in presence of PS without doping of Fe, Co and O has the synergistic effect on PS activation for TC
catalyst, which implies that the oxidation ability of the inactive PS is too removal over g − C3N4.
weak to decompose TC. As the addition of catalysts, the degradation The catalytic degradation rates of TC were obtained by fitting tested
efficiency of TC is significantly enhanced, in which the TC removal ef­ data with the Langmuir–Hinshelwood (L–H) first–order model as
ficiency of CN, FCN, CCN, FCOCN1, FCOCN2 and FCOCN3 is 65.2%, equation of − ln(C/Ca) = k1t, where Ca and C was TC concentration at
79.3%, 68.8%, 83.9%, 90.1% and 87.4%, respectively, after catalytic time of t = 0 and intervals (min), respectively, k1 was the pseudo − first
reaction for 120 min, implying that catalysts and PS may have a syn­ − order kinetic constant (min− 1). As presented in Fig. 7a, the degrada­
ergistic effect on TC removal. The synergistic factor (SF) can be calcu­ tion rates of all samples are in the following order: CN < CCN < FCN <
lated by the formula of SF = TC removed by (catalyst + PS) / (TC removed FCOCN1 < FCOCN3 < FCOCN2. It is apparently that FCOCN catalysts
by catalyst adsorption + TC removed by PS) [43]. Accordingly, the SF exhibit greatly higher kinetic constant than the CN, FCN and CCN.
value for CN, FCN, CCN, FCOCN1, FCOCN2 and FCOCN3 is 1.21, 1.62, Among them, the k1 value of FCOCN2 towards TC degradation is 0.021
1.15, 1.66, 1.68 and 1.64, respectively. All SF values are large than 1, min− 1, which is about 2.67 times, 1.60 times and 2.18 times that of pure
verifying the superior synergistic effect the between catalysts and PS CN (0.009 min− 1), FCN (0.015 min− 1) and CCN (0.011 min− 1),
[50]. As previously reported, the conjugated graphite structure of pure respectively. Furthermore, it is also much higher than both value of
g − C3N4 could promote TC degradation by PS via electron transfer 0.016 min− 1 for CoFe (Fig. S3, Supporting Information) and negative
pathway [36]. But, the catalytic efficiency is still limited due to the lack value for FCCN2 that may be caused by the continuous desorption of
of free − flowing π electrons in carbon and nitrogen conjugated system massive adsorbed TC (66.7% adsorptive removal) on FCCN2 (prepared
and the confined transfer of electrons in the two–dimensional plane under N2) during degradation process, which indicates that the syner­
[21,51]. Benefiting from the improvement of surface aperture channels gistic doping of Fe, Co and O plays an important role in enhancing
for reactant, i.e., PS and TC mass transfer diffusion, and the increase of catalytic activity of g–C3N4. By coupling of Fe, Co and O into CN, more
accessible active sites, the degradation efficiency of TC is remarkably active centers on surface with improved pore structure could be ach­
advanced in doped g − C3N4/PS, especially in FCOCN2/PS system. ieved for PS activation, reactive species generation and TC

5
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

Fig. 4. The XPS spectra of FCOCN2 sample: (a) survey spectrum, (b) C 1 s, (c) N 1 s, (d) O 1 s, (e) Fe 2p and (f) Co 2p.

disintegration. In order to quantitatively understand the PS utilization amount of produced free radical species. As shown in Fig. 6b, effects of
over catalysts, the parameter of reaction stoichiometric efficiency (RSE) oxidant dosage on catalytic degradation of TC over FCOCN2 were
was calculated by equation of RSE = Δn(TC)/Δn(PS), where Δn(TC) and investigated by adding different amount of PS from 100 mg to 300 mg. It
Δn(PS) was the mole change of TC and PS after reaction [53]. Accord­ can be observed that the degradation efficiency of TC is firstly increased
ingly, the RSE are 3.1%, 1.9%, 1.9% and 5.2% for CN, FCN, CCN and and then slightly decreased. When added amount of oxidant is 250 mg,
FCOCN2, respectively. The higher RSE of FCOCN2 compared with the the maximal TC degradation efficiency is obtained. As presented in
other three suggests that FCOCN2 is an efficient catalyst for PS activa­ Fig. 7b, the corresponding apparent degradation rate firstly increases
tion. As summarized in Table S1 (Supporting Information), the FCOCN2 from 0.013 min− 1 to 0.016 min− 1 and then marginally decreases to
has the comparable and/or even better TC removal performance than 0.015 min− 1. Generally, the increasing dosage of PS would provide more
Fe0 [2], γ–Fe2O3/CeO2 [53], Fe3O4 [54], magnetic biochar [55], opportunities for collision between PS and surface active sites, which
Co–biochar [56] and few–layer g–C3N4/Vis–light [57] in terms of allows more reactive oxygen species (i.e., SO4•− , •OH) to be produced
degradation efficiency, reaction rate constant and TOC removal for TC degradation. However, when the added PS is excessive, the
(68.6%), demonstrating that it may be a desirable catalyst to activate PS generated SO4•− and •OH would be further quenched by extra PS
for TC mineralization in wastewater treatment. forming the SO5•− with relatively weak oxidation ability by Eqs. (1) and
The dosage of oxidant was a key direct factor in determining the (2) [2]. Besides, the superfluous SO4•− also have self − quenching effect

6
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

significantly decreases from 85.1% to 73.3% with initial concentration


of TC increases. Also, the corresponding degradation rate constant
(Fig. 7d) is reduced from 0.020 min− 1 to 0.012 min− 1, suggesting that
the degradation process is seriously depended on the initial concentra­
tion of TC. Theoretically, under a fixed dosage of catalyst and PS, the
total yields of radicals are at constant. When initial TC concentration is
low, the higher ratio of radicals to TC would make the greater proba­
bility for radicals attacking TC molecules, and thus the higher TC
degradation efficiency can be obtained. Conversely, the strong
competitive reaction between TC molecules with limited free radicals
and the coverage of redox active sites by adsorbed TC could cause less
removal of TC at a specific reaction time [6,11]. Therefore, at lower
initial TC concentration, the consumption rate of radicals could be lower
than their generation rate, which is beneficial to the degradation of TC.
As pH was one of most important factors affecting the decomposition
of organic pollutants in PS/catalysts system, the effect of pH on removal
efficiencies and degradation rate constant of TC over FCOCN2 by PS are
investigated in Fig. 6e and Fig. 7e. Apparently, the removal process of
TC is significantly depended on initial solution pH, in which the
degradation efficiency drastically decreases from 91.7% to 74.3% with
pH increasing from 3.0 to 11.0. Correspondingly, the degradation ki­
netic constant of TC decreases from 0.026 min− 1 to 0.011 min− 1. which
is extremely consistent with the result that acidic condition favors the
degradation of TC reported by Pi et al. [9]. Generally, under acidic
condition, SO4•− is the dominant active species that may be produced in
large quantities for TC degradation through the extra acid–catalytic
pathway via Eqs. (5) and (6) [28,59]. However, when under alkaline
condition, SO4•− is prone to react with OH− forming •OH as Eqs. (8)
[32]. The •OH radical has a relative lower oxidative potential than
SO4•− , 1.8 – 2.7 V vs. 2.5 – 3.1 V for destruction of chemical bond of TC.
Also, as the concentration of OH− increasing, the •OH can further
interact with OH− to form H2O, causing the annihilation of free radicals
and weakened degradation efficiency of TC over FCOCN2 [59]. Besides,
as presented in Fig. S5 (Supporting Information), the pH of point of zero
Fig. 5. The N2 ad-desorption isotherms (a) and pore-size-distribution (b) charge (pHPZC) of FCOCN2 is estimated to be 9.1 by zeta potential
of samples. measurement. It means that the surface of FCOCN2 is more positive
charge when pH gets lower and lower than pHPZC, and vice versa. Thus,
as Eqs. (3) and (4) [25]. the oxidant of HSO5– could be effectively adsorbed on positive charged
FCOCN2 by electrostatic attraction for radical generation at pH < 9.1,
HSO−5 + SO4 ∙− → SO5 ∙− + H + + SO−4 (1)
while that the opposite due to the presence of electrostatic repulsion
between HSO5– and negative charged FCOCN2 at pH > 9.1 [50]. Simi­
HSO−5 + ∙OH→ SO5 ∙− + H2 O (2)
larly, since the major species of TC in solution are H4TC+ (pH < 3.4),
H3TC (3.4 < pH < 7.6), H2TC– (7.6 < pH < 9.0) and HTC2– (pH > 9.0)
SO4 ∙− + SO4 ∙− → S2 O2−8 (3)
[60], the electrostatic attraction between TC and FCOCN2 would be
become stronger at lower pH. This can be confirmed by the fact that the
S2 O2−8 + SO4 ∙− → S2 O8 ∙− + SO2−4 (4)
adsorptive removal efficiency of TC on FCOCN2 is higher at lower pH
Effect of added catalyst dosage of FCOCN2 on the removal process of except pH 3.0 (Fig. S5, Supporting Information). The special case may be
TC by PS is illustrated in Fig. 6c and Fig. 7c. Apparently, after reaction due to that dominant adsorption driving forces i.e., π – π interactions, are
for 120 min, the degradation process with amount of FCOCN2 at 60 mg greatly weakened at pH 3.0. Therefore, the influence of pH on PS acti­
achieves the best catalytic performance towards TC degradation effi­ vation for TC removal is achieved by affecting the generation of free
ciency of 83.6% and kinetic constant of 0.014 min− 1. Under the a con­ radicals and the adsorptive diffusion of reactants.
stant of PS addition amount(250 mg) and the relatively low dosage
S2 O2−8 + H +
→HS2 O−8 (5)
range of catalyst (20 mg to 60 mg), the increase amount of FCOCN2 can
provide more redox active sites for PS activation. Numerous previous
HS2 O−8 →SO4 ∙ + SO2−4 + H+ (6)
studies have declared that the increase in active sites with increasing
catalysts dosage is favorable to accelerate the production of free radicals Since temperature was another crucial factor affecting chemical re­
for pollutants decomposition [32,36]. However, as the dosage of action, the influence of temperature on PS activation for TC degradation
FCOCN2 further increase to 80 mg, the degradation efficiency of TC is was investigated. As depicted in Fig. 6f, the degradation performance of
slightly reduced, which may be attributed to that a large amount of TC via PS activation over FCOCN2 is extremely depended on reaction
FCOCN2 provides excessive divalent metal ion, resulting in the strong temperature. In detail, the degradation efficiency of TC in PS/FCOCN2
scavenging effect of SO4•− [58]. Moreover, Exorbitant density of C≡N system is 86.3% (30 ◦ C), 93.9% (40 ◦ C) and 96.9% (50 ◦ C) after 120 min
group in FCOCN2 could induce scavenging effect on SO4•− (C≡N + catalytic reaction. Furthermore, the degradation rate constant (Fig. 7g)
SO4•− →C = N•+ + SO42− ) [33]. increases from 0.018 min− 1 to 0.034 min− 1, which indicates endo­
The influence of initial TC concentration in the range of 5 mg/L to 25 thermic feature of PS activation initiated by FCOCN2 [61]. Thus, tem­
mg/L on its degradation efficiency in FCOCN2/PS system was investi­ perature increasing can directly promote the fission of O − O bond in PS
gated. As presented in Fig. 6d, the removal percentage of TC

7
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

Fig. 6. The catalytic degradation performance of TC by different catalysts (a), and the effects of PS dosage (b), FCOCN2 dosage (c), initial TC concentration (d), pH
(e) and temperature on TC degradation in FCOCN2/PS system.

for generation radicals of SO4•− , •OH, etc., and accelerate electron − (MeOH), tert–butyl alcohol (t–BuOH), phenol, p − benzoquinone (p −
transfer for TC degradation over FCOCN2 [2,30]. Besides, the activation BQ) and furfuryl alcohol (FFA) as scavengers. Generally, MeOH can
energy (Ea) was calculated by according Arrhenius equation of lnk1 = lnA quench both SO4•− (kSO4•− = 1.0 × 107 M− 1s− 1) and •OH (k•OH = 1.0 ×
– Ea/RT, where A was the pre–exponential factor (min− 1), T was the 109 M− 1s− 1,) radicals, while t–BuOH can only trap •OH (k•OH = 5.2 ×
temperature (K) and R was universal gas constant (8.314 J.mol− 1.K− 1) 108 M− 1s− 1) [25]. The phenol could simultaneously capture SO4•−
[18]. As presented Fig. 7f, the value of Ea is computed to be 23.88 kJ. (kSO4•− = 8.8 × 109 M− 1s− 1), •OH (k•OH = 6.6 × 109 M− 1s− 1,) and
mol− 1, which is lower than that of Co@N/S − doped carbon (44.5 kJ. SO5•− (kSO5•− = 1.0 × 1010 M− 1s− 1) [62]. The p − BQ and FFA were
mol− 1) [27], NiFe2− xCoxO4 − reduced graphene oxide (39.62 kJ.mol− 1) commonly considered as quencher for •O2− (k•O2− = 1.0 × 109 M− 1 s− 1)
[24], S − doped g − C3N4 (38.6 kJ.mol− 1) [13] and Pd/g − C3N4 (33.7 and single oxygen (1O2, k1O2 = 1.0 × 108 M− 1s− 1), respectively [63]. As
kJ.mol− 1) [30] in PS activation. This lower activation energy suggests presented in Fig. 8a and Fig. S6 (Supporting Information), the degra­
that the PS activation for TC degradation by FCOCN2 is favorable. dation efficiency and pseudo − first − order kinetic constant of TC are
suppressed to different degrees after introduction of scavengers. In
detail, the addition of MeOH and t–BuOH makes the TC removal effi­
3.3. Mechanism of catalytic degradation
ciency decrease from 90.1% to 68.2% and 75.6%, respectively, and the
kinetic constant reduce from 0.021 min− 1 to 0.009 min− 1 and 0.013
To get insights into the generation of active species, quenching ex­
min− 1, respectively. This result suggests that both SO4•− and •OH are
periments were performed by using chemical agents of methanol

8
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

Fig. 7. The observed pseudo-first-order rate constant of TC degradation with different catalysts (a), PS dosage (b), FCOCN-2 dosage (c), TC concentration (d), pH (e)
and temperature (g): inset is the Arrhenius plot (f).

generated and involved in the TC degradation process by FCOCN − 2/PS Therefore, SO5•− was not detected by ESR in this study. During ESR tests
system. Furthermore, the phenol exhibits a greater inhibitory effect on for SO4•− , •OH, •O2− and 1O2 species detection, the radical spin trap­
TC removal than MeOH, in which the removal efficiency and kinetic ping agent of 5–dimethyl–1–pyrroline–N–oxide (DMPO) is used for the
constant decreases to 27.1% and 0.003 min− 1, respectively, indicating detection of SO4•− , •OH and •O2− radicals, and 2, 2, 6, 6 − tetrame­
that SO5•− is also generated over FCOCN2 for TC degradation. Obvi­ thylpiperidine (TEMP) is applied for 1O2 specie observation. As shown in
ously, the oxidation of TC is almost completely terminated by addition of Fig. 8b − d, in the absence of FCOCN (PS alone, 0 min), no noticeable
p − BQ, implying that •O2− plays the most important role in TC signal can be observed, indicating that the generation of radicals by self
degradation by FCOCN − 2/PS system. Besides, the presence of FFA − decomposition of PS is negligible. In contrast, very clear signals were
significantly inhibits the degradation of TC with 48.4% reduction in detected in presence of both PS and FCOCN catalyst, and intensities of
removal efficiency and the corresponding kinetic constant is drastically signals were significantly enhanced with the prolongation of reaction
reduced to 0.004 min− 1, which suggests that 1O2 makes a significant time. The characteristic signals with intensity ratio of 1:2:2:1 could be
contribution to TC degradation. As previously reported, the nonradical ascribed to DMPO–•OH (with hyperfine couplings of αN = αβ− H = 14.9
of 1O2 could be generated from the self − decomposition of PS and the G), and other peaks with much lower intensity than DMPO–•OH are
reaction of PS with carbonyl moieties on skeleton of g–C3N4 [34,37,64]. belonged to DMPO–SO4• (with hyperfine splitting constants of αN =
More convincing evidence for the production of SO4•− , •OH, •O2− 13.2 G, αH1 = 9.6 G, αH2 = 1.48 G and αH3 = 0.78 G) (Fig. 8b) [25]. As
and 1O2 species were supported by electron spin resonance (ESR) mea­ presented in Fig. 8c, four characteristic signals of DMPO–•O2− adducts
surements. Previous literature have reported that the SO5•− is hardly be (with hyperfine splitting constants of αN = 14.3 G, αH1 = 11.2 G, and αH2
detected by ESR technology due to extremely weak signal [62]. = 1.3 G) could also be observed in ethanol [28]. In addition, the strong

9
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

Fig. 8. (a) Trapping experiment for the catalytic degradation of TC over FCOCN2, (b) the ESR spectra of DMPO-•OH and DMPO-SO4•− , (c) DMPO− •O2− and
(d) TEMP-1O2 in FCOCN-2/PS system.

1:1:1 triplet signal corresponding to TEMP− 1O2 compounds (with hy­ could serve as electron acceptor to interact with PS for SO5•− and •O2−
perfine coupling constants of αN = 17.24 G) are observed in Fig. 8d, radical production (Eqs. (12) and (13)), in which SO5•− and •O2− can
demonstrating the production of abundant 1O2 [64]. Combining the further transferred to 1O2 via direct oxidation or recombination (Eqs.
quenching experiment and ESR test, it can be confirmed that SO4•− , (14) and (15)) [34,37,65]. Moreover, due to structural defects of C − O
•OH, •O2− , 1O2 as well as SO5•− radicals are responsible for TC oxida­ − Fe − N (and/or C − O − Co − N) covalency in matrix, the intervalence
tion in FCOCN/PS system. electron transfer of Eqs. (16) could be promoted with redox potential
As known, PS can be heterogeneously activated by transition metal (E0) of Co3+/Co2+ and Fe3+/Fe2+ at + 1.82 V and + 0.77 V, respectively,
through O–O (SO4 − SO4) bond–breaking [23,61]. The XPS analysis has which is facilitate to maintain the stable electronic state redistribution of
identified that iron and cobalt doped in g–C3N4 are in the form of Fe(III)/ Fe and Co for high activity [22,33]. Finally, the TC could be progres­
Fe(II) and Co(III)/Co(II), which plays an indispensable role in PS acti­ sively degraded to small molecules and even CO2 by oxidation of SO4•− ,
vation. As shown in Fig. 9a, the HSO5− produced from Eqs. (7) could SO5•− , •OH, •O2− and 1O2 active species [62].
interact with Fe(III)/Co(III) doped in g–C3N4 yielding radicals of •O2−
S2 O2−8 + H2 O→HSO−5 + HSO−4 (7)
and SO5•− as Eqs. (12) and (13), and meanwhile the Fe(III)/Co(III) is
reduced to Fe(II)/Co(II) [34,64]. Then the Fe(II)/Co(II) would further
Fe(II)/Co(II) − OER − g − C3 N4ER + HSO−5 →Fe(III)/Co(III) − O − g − C3 N4
react with HSO5− as Eqs. (8) and (11) to generate radicals (•O2− and
SO4•− ) and Fe(III)/Co(III) [10,11,52]. Recent reports have shown that + SO−4 ∙ + OH− (8)
the pure g − C3N4 has the ability to activate PS in an electron − transfer
mediated non − radical pathway, in which the pyridinic N with lone pair SO−4 ∙ + H2 O → ∙OH + SO2−4 + H+ (9)
electrons can act as electron donors for reduction of PS, and C atoms
adjacent to N atoms can play as electron acceptors to oxide PS [31]. In SO−4 ∙ + OH− → SO2−4 + ∙OH (10)
present work, the co − doping of iron, cobalt and oxygen could alter the
electronegativity of substrate C, which may boost PS activation via Fe(II)/Co(II) − OER − g − C3 N4ER + 2HSO−5 →Fe(III)/Co(III) − O − g
synergistic effect of radical and non − radical way [37]. During poly­ − C3 N4 + 2HSO−4 + ∙O−2 (11)
merization process, the O was firstly doped into g − C3N4 by replacing 2
− coordinated N, then Fe and Co were bonded with C and finally evolved
ED
Fe(III)/Co(III) − O − g− C3 N4 + HSO−5 + H2 O→Fe(II)/Co(II) − O − g
into C − O − Fe − N and C − O − Co − N bond [39]. Thus, the catalytic − C3 N4 + SO2−4 + 3H + + ∙O−2 (12)
centers of electron − rich O (OER) and adjacent electron–deficient C
(EDC) are formed. The electron − rich O and pyridinic N (NER) with high Fe(III)/Co(III) − O − g− ED
C3 N4 + HSO−5 →Fe(II)/Co(II) − O − g
electron density could act as electron donor for SO4•− , •OH and •O2− − C3 N4 + SO5 ∙ + H +

(13)
radicals production (Eqs. (7)–(11)), and the electron–deficient C (EDC)

10
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

Fig. 9. (a) Schematic depiction of PS activation over FCOCN2 for TC removal; (b) electrochemical impedance spectroscopy (EIS) response and (c) linear-sweep
voltammograms (LSV) under different conditions.

2∙O−2 + 2H2 O → 1 O2 + H2 O2 + 2OH− (14)


Table 1
Electrochemical parameters of equivalent circuit elements for the samples.
2SO−5 ∙ + H2 O→ 2HSO−4 1
+ 1.5 O2 (15)
Parameters CN FCN CCN FCOCN2

Co(III) + Fe(II)→ Fe(III) + Co(II) ΔE = 1.05 V (16) R1(Ω) 11.9 11.28 11.74 12.06
R2(Ω) 637.6 563.9 635.4 364.0
The promotion of electron − transfer over FCOCN2 for PS activation CPE1-T 2.9746e-5 2.3441e-5 3.0819e-5 1.9682e-5
was further identified by ultraviolet photoemission spectroscopy (UPS) CPE1-P 0.87847 0.89423 0.87127 0.89156
and electrochemical measurements including electrochemical imped­ Ws1-R(Ω) 1218 1013 1176 1245
Ws1-T 0.046017 0.04468 0.046022 0.049862
ance spectroscopy (EIS), cyclic voltammograms (CV) and linear sweep
Ws1-P 0.83084 0.8414 0.83353 0.75735
voltammetry (LSV). Previous reports have clarified that the oxygenated Chi-squared 0.00106 0.00083 0.00105 0.00062
species could withdraw electrons from C–N = C skeleton and then in­
crease the thickness of space–charge region for charges migration
[51,57]. As presented in Fig. S7 (Supporting Information), the work resistance (R2) value of FCOCN2 is estimated to be 364.0 Ω based on the
function (Φ) related to minimum energy for inner electrons escaping of fitting of Nyquist plots to equivalent circuit diagram, which is much
CN, FCN, CCN and FCOCN2 are investigated by UPS on the basis of Φ = lower than that of CN (637.6 Ω), FCN (563.9 Ω) and CCN (635.4 Ω). The
hν − |Ecutoff – EF|, where hν denotes the energy of excited photon applied lowest R2 value of FCOCN2 also suggests that FCOCN2 possesses a better
for detection (He I: 21.21 eV), Ecutoff represents the secondary electron conductivity for electron − transfer, which is consistent with the
cut–off energy, EF is the Fermi levels that is set at 0 eV [66]. It can be calculation result of the work function. Fig. S8 (Supporting Information)
found that the Ecutoff of CN, FCN, CCN and FCOCN2 is 16.27 eV, 16.20 shows the CV curves of CN, FCN, CCN and FCOCN2 electrodes within the
eV, 16.17 eV and 16.36 eV, respectively. Therefore, the corresponding voltage region of − 1.5 to 1.5 V. Apparently, the FCOCN2 presents an
values of Φ are calculated to be 4.94 eV, 5.01 eV, 5.04 eV and 4.85 eV, enhanced current intensity compared to the other three electrodes,
respectively. The lower work function of FCOCN2 suggests that elec­ indicating that FCOCN2 retains a higher reversible redox capacity for PS
trons of FCOCN2 have a lower excitation energy barrier and are more activation due to the fast electron circulation [23]. Furthermore, to
likely to be motivated than the other three due to the electrons accu­ understand the interaction of FCOCN2 with reactants of PS and TC, the
mulation of O species after Fe, Co and O co–doping [67]. Fig. 9b presents linear sweep voltammetry (LSV) curves were measured in the presence
EIS Nyquist plots with a simulated equivalent circuit, in which R1 is and/or absence of PS. As can be found in Fig. 9c, the FCOCN2 has a
overall electrolyte resistance, R2 is charge transfer resistance connected larger current density due to the preferable conductivity. With the
in series with the Warburg impedance (Ws1), CPE1 is constant phase addition of PS, the current of the FCOCN2 electrode is significantly
angle element [68]. It can be found that the arc diameter of FCOCN2 is enhanced, demonstrating the promoted electron transfer between PS
smaller than that of other samples, indicating the superior charge and FCOCN2 in presence of redox shuttles, electron − enriched and
transfer ability of FCOCN2. As listed in Table 1, the charge − transfer deficient sites [29,49]. After injection of TC, the current of is further

11
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

strengthened, in which the TC acts as an electron donor, PS acts as an The last one could be further transformed to P17 (m/z = 397) with
electron acceptor, and the electrode of FCOCN2 with high conductivity fracture of double bond at C10 − 10a, and then due to the deamidation
serves as an electron transport channel allowing active species genera­ and oxidative ring opening of quinone group, the P18 (m/z = 329)
tion for oxidative degradation of TC [19]. would be obtained [78]. Finally, all reaction intermediates are decom­
posed into CO2, H2O and NH4+ under the strong oxidation of radicals.
Previous reports have been proved that many intermediates gener­
3.4. Possible degradation pathway and products toxicity
ated in degradation process might have higher potential toxicity than it
itself [14,70]. Therefore, the acute toxicities of TC and transformation
To gain insight into reaction pathway in TC degradation process,
products were predicted by using toxicity estimation software tool (T.E.
high performance liquid chromatography mass spectrometry (HPLC −
S.T.) based on the quantitative structure − activity relationship (QSAR)
MS) measurements are employed to detect intermediates (Fig. S9, Sup­
method, and the aquatic toxicities of treated water were evaluated by of
porting Information). It can be found that the characteristic peak of TC
luminescence change of V. fisheri, growth inhibition of E. coli and
(m/z = 445) is disappeared after reaction, and eighteen major degra­
B. subtilis. According to Globally Harmonized System of Classification
dation transient products with m/z values of 510, 499, 467, 443, 431,
and Labeling of Chemicals (GHS), the chemicals are divided into four
397, 375, 363, 329, 307, 301, 295, 284, 261, 227, 165, 159, 132 are
categories of very toxic, toxic, harmful and not harmful. As presented in
detected during TC decomposition. Under the attack of SO4− •, •OH,
Fig. 11a, the lethal concentration 50% (LC50, 48 h) of TC is 0.17 mg/L
O2− •, SO5− • and 1O2 species, the double bond, amine group, and
towards Daphnia magna, indicating that TC has the very toxic property,
phenolic groups of TC are most vulnerable. As presented in Fig. 10, the
which may cause potential adverse effects on Daphnia magna. Also, it
probable structures of intermediates and degradation pathway of TC are
can be seen that the acute toxicities of degradation intermediates are
proposed. By demethylation of an amino group connected the C4 atom,
lower than that of TC reflecting from the higher LC50 values. Especially,
the intermediate P1 with m/z of 431 is generated, and then it can be
the products of P5, P11, P15, P16 and P17 are classified as not harmful
evolved into P2 (m/z = 363) and P6 (m/z = 375) through de–amidation
category. Except for P6, the developmental toxicities (Fig. 11b) of other
[20,69,70]. After decarboxylation, C–N bond cleavage, ring opening,
degradation intermediates is equivalent to and/or even lower than that
and/or hydroxylation reaction, the P2 is gradually decomposed into P3
of TC, and P5 is identified as developmental non − toxicant. Fig. 11c
(m/z = 307), P4 (m/z = 295) and P5 (m/z = 227) [71,72]. Meanwhile,
exhibits the bioaccumulation factor of intermediate products, in which
the P7 (m/z = 307) that generated by the loss of hydroxyl, amino and
more than half of the degradation intermediates have lower bio −
methyl groups, and the fracture of C = O bond of P6, could be converted
concentration ability than TC ascribing to hydroxylation reaction.
to P8 (m/z = 132) via ring opening [73]. Furthermore, TC can also be
Similarly, the mutagenicity of overwhelming majority of intermediates
decomposited into P9 (m/z = 443) as a result of the dehydroxylation of
have been reduced and even altered to negative status as shown in
C6 atom, and then compounds of P10 (m/z = 301) and P11 (m/z = 165)
Fig. 11d, indicating the lower genotoxicity of these products than TC.
could be generated from P9 (m/z = 443) due to the destruction of cyclic
The actual effect of TC aqueous solutions at different degradation time
hydrocarbon structure [40,74]. After that, the P12 (m/z = 261) and P13
on the luminescence inhibition of V. fisheri is shown in Fig. S10 (Sup­
(m/z = 159) are successively generated via ring opening reaction and
porting Information). It can be found that the luminescence inhibition
methyl/hydroxyl groups loss, respectively [75,76]. In addition to the
rate of solutions for V. fisheri is gradually increased from 66.8% to 90.4
above reactions, when the C = C bond at C6a – C7 position of TC is
% in the first 40 min due to some toxic intermediates, e.g. P8, P13, etc.,
broken for ketone group and carboxylic group formation, the P14 (m/z
was generated in the early stage as predicted by T.E.S.T Software.
= 465), P15 (m/z = 510) and P16 (m/z = 499) could be formed [1,77].

Fig. 10. Proposed degradation pathways and intermediates of TC by FCOCN2.

12
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

Fig. 11. Prediction of acute toxicity for daphnia magna (a), developmental toxicity (b), bioaccumulation factor (c) and mutagenicity (d) of main TC degradation
intermediates by T.E.S.T. based on consensus method.

Fortunately, with further mineralization of such intermediates, the in­ activation of PS for TC degradation. Also, the signal intensity of O 1 s for
hibition rate is drastically decreased to 34.5%, which is very consistent recycled FCOCN2 is stronger than that for fresh one, which confirms that
with the results reported by Shen et al [5]. It also can be confirmed by some oxygen containing degradation intermediates of TC still remains
comprehensive toxicity of treated solution to E. coli and B. subtilis on the surface of FCOCN2. From high resolution spectrum of Fig. 12b −
growth. As exhibited in Fig. S11 (Supporting Information), the number f, it can be found that all species contained C, N, O, Fe and Co belonging
of colonies of E. coli and B. subtilis grown on agar plate containing TC to fresh FCOCN2 can be identified, implying that FCOCN2 has no
degradation solution (60 min) are comparable with that of blank sample qualitative change in the composition after reaction. Furthermore, the
without TC. Therefore, it can be concluded that although some toxic metal leaching concentration detected by inductively coupled plasma
intermediates were produced in the degradation process, the compre­ mass spectrometer (ICP–MS) after catalytic reaction is also very low. The
hensive toxicity of all intermediates is significantly decreased with the leaching concentration Fe and Co is only 0.075 mg/L and 0.032 mg/L,
extension of degradation time, indicating the great ability of FCOCN2 respectively, demonstrating that Fe and Co are firmly integrated into
for efficient decomposition and detoxification of TC. FCOCN2. To reveal the practical application potential of FCOCN2 in
wastewater treatment, the tap water, river water, municipal wastewater
and medical wastewater were applied as background matrices of TC
3.5. Reusability and application potential of FCOCN2 solution for degradation. As illustrated in Table 2, the TOC removal
proportion by FCOCN2/PS system in tap water, river water, municipal
The reusability of the catalyst was an important feature to reflect the wastewater and medical wastewater is 19.0 %, 34.2 %, 6.9 % and 4.7 %,
durability of catalyst for sustainable application. Therefore, the reuse respectively. Compared with that in ultrapure water (Table S1, Sup­
experiment of FCOCN2 for TC degradation in presence of PS was con­ porting Information), the drastic reduction of TOC removal ratio in
ducted, and the changes of structural components before and after re­ actual water matrices may be caused by the complex impact of high
action were detected to evaluate the catalytic stability of FCOCN2. From background TOC level, organic matter competition, ions interference
Fig. 12a, it can be found that the catalytic activity of FCOCN2 for TC and so on [53]. Thus, more reaction time, PS dosage and catalysts usage
decomposition has a slight decrease after five consecutive cycles, in may be needed in terms of efficient removal of TOC from actual
which the degradation efficiency of TC is 72.7% in fifth cycle. This wastewater. Gratifyingly, the corresponding TC degradation efficiency
unremarkable reduction may be attributed to the residual adsorption of is 90.0%, 88.6%, 74.0%, 81.5%, respectively, which demonstrates the
intermediates on surface of FCOCN2 retarding the interaction between very powerful ability of FCOCN2/PS for disintegration of TC structure in
PS and active sites [36,42,65]. As presented in Fig. 12b, the XRD actual wastewater. Therefore, the FCOCN2/PS system has great appli­
diffraction pattern of recycled FCOCN2 has the similar peaks to fresh cation potential in pretreatment and/or advanced treatment of anti­
one, indicating that no obvious change occurs in the phase and structure biotic contained wastewater with acceptable catalyst stability and
of FCOCN2 after catalytic reaction. Further information is supported by reusability.
XPS investigation. The survey scan spectra (Fig. 12c) demonstrates that
the recycled FCOCN2 contains the same elements of C 1 s, N 1 s, O 1 s, Fe 4. Conclusions
2p and Co 2p as fresh FCOCN2. However, the corresponding binding
energies are slightly shifted to 288.2 eV, 399.1 eV, 531.1 eV, 709.1 eV Porously thin layered Fe, Co and O co − doped g − C3N4 catalysts
and 786.1 eV, respectively, indicating that all elements are involved in

13
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

Fig. 12. The repeated catalytic experiments of TC in FCOCN2/PS system (a) and the XRD pattern (b), survey XPS spectrum (c), C 1 s (d), N 1 s (e), O 1 s (f), Fe 2p (g)
and Co 2p (h) of FCOCN2 after catalytic reaction.

14
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

Table 2 [4] K. Li, A. Yediler, M. Yang, S. Schulte-Hostede, M.H. Wong, Ozonation of


Efficiency of TC removal by PS activation with FCOCN2. oxytetracycline and toxicological assessment of its oxidation by-products,
Chemosphere 72 (3) (2008) 473–478.
Practical Background TC solution [5] Q. Shen, L. Wei, R. Bibi, K. Wang, D. Hao, J. Zhou, N. Li, Boosting photocatalytic
applications matrices degradation of tetracycline under visible light over hierarchical carbon nitride
microrods with carbon vacancies, J Hazard Mater 413 (2021) 125376–125391.
pH TOC TOC TC removal TOC removal [6] A. Jonidi Jafari, B. Kakavandi, N. Jaafarzadeh, R. Rezaei Kalantary, M. Ahmadi,
(mg/L) (mg/L) (%) (%) A. Akbar Babaei, Fenton-like catalytic oxidation of tetracycline by AC@Fe3O4 as a
heterogeneous persulfate activator: Adsorption and degradation studies, J Ind Eng
Tap water 7.6 1.86 10.78 90.0 19.0
Chem 45 (2017) 323–333.
River water 8.1 6.16 13.64 88.6 34.2
[7] X. Wang, B. Zhang, J. Ma, P. Ning, Novel synthesis of aluminum hydroxide gel-
Municipal 7.3 24.58 37.24 74.0 6.9 coated nano zero-valent iron and studies of its activity in flocculation-enhanced
Wastewater removal of tetracycline, J Environ Sci 89 (2020) 194–205.
Medical 7.8 25.54 31.25 81.5 4.7 [8] Z. Liu, M. Zhu, L. Zhao, C. Deng, J. Ma, Z. Wang, H. Liu, H. Wang, Aqueous
wastewater tetracycline degradation by coal-based carbon electrocatalytic filtration
membrane: Effect of nano antimony-doped tin dioxide coating, Chem Eng J 314
(2017) 59–68.
were successfully fabricated with enhanced specific surface area and [9] Z. Pi, X. Li, D. Wang, Q. Xu, Z. Tao, X. Huang, F. Yao, Y. Wu, L.i. He, Q.i. Yang,
Persulfate activation by oxidation biochar supported magnetite particles for
electron transfer ability for PS activation to degrade TC. The catalytic tetracycline removal: Performance and degradation pathway, J Clean Prod 235
activity towards TC degradation of all samples are in the order of CN < (2019) 1103–1115.
CCN < FCN < FCOCN1 < FCOCN3 < FCOCN2. The maximum TC [10] Z. Wu, Y. Liang, D. Zou, X. Yuan, Z. Xiao, Y. Deng, Y. Zhou, L. Jiang, P. Qin,
Enhanced heterogeneous activation of persulfate by NixCo3–xO4 for oxidative
removal efficiency of FCOCN2 was 90.1 % after 120 min reaction and
degradation of tetracycline and bisphenol A, J Environ Chem Eng 8 (6) (2020)
the kinetic constant of FCOCN2 is 2.67, 1.60, 2.18 times that of pure 104451, https://doi.org/10.1016/j.jece:2020.104451.
g–C3N4, Fe doped g–C3N4 and Co doped g–C3N4, respectively, due to the [11] R. Guan, X. Yuan, Z. Wu, L. Jiang, J. Zhang, Y. Li, G. Zeng, D. Mo, Efficient
degradation of tetracycline by heterogeneous cobalt oxide/cerium oxide
formation of electron − rich and − deficient catalytic centers for syn­
composites mediated with persulfate, Sep Purif Technol 212 (2019) 223–232.
ergetic electron transfer. The reaction stoichiometric efficiency of PS [12] S.S. Rezaei, B. Kakavandi, M. Noorisepehr, A.A. Isari, S. Zabih, P. Bashardoust,
and TOC removal efficiency is 5.2% and 68.1%, respectively. Except for Photocatalytic oxidation of tetracycline by magnetic carbon-supported TiO2
PS dose, the catalyst dosage, initial pollutant concentration, solution pH nanoparticles catalyzed peroxydisulfate: Performance, synergy and reaction
mechanism studies, Sep Purif Technol 258 (2021) 117936–117950.
and reaction temperature shown significant effect on TC degradation. In [13] K.-Y. Lin, Z.-Y. Zhang, Degradation of Bisphenol A using peroxymonosulfate
FCOCN2/PS system, the active species •O2− , 1O2, •OH, SO4•− and activated by one-step prepared sulfur-doped carbon nitride as a metal-free
SO5•− are responsible for TC oxidation as the edge defects, in − plane heterogeneous catalyst, Chem Eng J 313 (2017) 1320–1327.
[14] Y. Yang, X. Lu, J. Jiang, J. Ma, G. Liu, Y. Cao, W. Liu, J. Li, S. Pang, X. Kong, C. Luo,
cavities and metal (Fe, Co) redox shuttles in g − C3N4 as the active sites. Degradation of sulfamethoxazole by UV, UV/H2O2 and UV/persulfate (PDS):
According to the major intermediates detected in degradation process, Formation of oxidation products and effect of bicarbonate, Water Res 118 (2017)
the pathway of TC decomposition and the toxicity of intermediates are 196–207.
[15] A.A. Babaei, M. Golshan, B. Kakavandi, A heterogeneous photocatalytic sulfate
analyzed. The results suggests that the toxicity of intermediates de­ radical-based oxidation process for efficient degradation of 4-chlorophenol using
creases with the extension of degradation time. Considering the TiO2 anchored on Fe oxides@carbon, Process Saf Environ 149 (2021) 35–47.
considerable stability and reusability of FCOCN2 catalyst and the su­ [16] L. Liu, S. Lin, W. Zhang, U. Farooq, G. Shen, S. Hu, Kinetic and mechanistic
investigations of the degradation of sulfachloropyridazine in heat-activated
perior TC degradability of FCOCN2/PS system in real wastewater, the
persulfate oxidation process, Chem Eng J 346 (2018) 515–524.
FCOCN2/PS system may be a promising approach for pretreatment and/ [17] A.R. Rahmani, M. Salari, A. Shabanloo, N. Shabanloo, S. Bajalan, Y. Vaziri, Sono-
or advanced treatment of antibiotic contained wastewater. catalytic activation of persulfate by nZVI-reduced graphene oxide for degradation
of nonylphenol in aqueous solution: Process optimization, synergistic effect and
degradation pathway, J Environ Chem Eng 8 (2020) 104202–104218.
Declaration of Competing Interest [18] Y. Feng, J. Liu, D. Wu, Z. Zhou, Y.u. Deng, T. Zhang, K. Shih, Efficient degradation
of sulfamethazine with CuCo2O4 spinel nanocatalysts for peroxymonosulfate
The authors declare that they have no known competing financial activation, Chem Eng J 280 (2015) 514–524.
[19] S. Zhu, X. Huang, F. Ma, L. Wang, X. Duan, S. Wang, Catalytic Removal of Aqueous
interests or personal relationships that could have appeared to influence Contaminants on N-Doped Graphitic Biochars: Inherent Roles of Adsorption and
the work reported in this paper. Nonradical Mechanisms, Environ Sci Technol 52 (2018) 8649–8658.
[20] H. Huang, T. Guo, K. Wang, Y. Li, G. Zhang, Efficient activation of persulfate by a
magnetic recyclable rape straw biochar catalyst for the degradation of tetracycline
Acknowledgements hydrochloride in water, Sci Total Environ 758 (2021) 143957–143970.
[21] X. Duan, H. Sun, J. Kang, Y. Wang, S. Indrawirawan, S. Wang, Insights into
The authors gratefully acknowledge the National Natural Science Heterogeneous Catalysis of Persulfate Activation on Dimensional-Structured
Nanocarbons, ACS Catal 5 (8) (2015) 4629–4636.
Foundation of China (No. 51808215, 41877491). Innovation platform [22] Z.-Y. Guo, C.-X. Li, M. Gao, X. Han, Y.-J. Zhang, W.-J. Zhang, W.-W. Li, Mn-O
and talent plan of Hunan Province (2020RC2056), Provincial Natural Covalency Governs the Intrinsic Activity of Co-Mn Spinel Oxides for Boosted
Science Foundation of Hunan (No. 2019JJ50253). Peroxymonosulfate Activation, Angew Chem Int Ed Engl 60 (1) (2021) 274–280.
[23] Z. Dong, Q. Zhang, B.-Y. Chen, J. Hong, Oxidation of bisphenol A by persulfate via
Fe3O4-α-MnO2 nanoflower-like catalyst: Mechanism and efficiency, Chem Eng J
Appendix A. Supplementary data 357 (2019) 337–347.
[24] X. Xu, J. Qin, Y. Wei, S. Ye, J. Shen, Y. Yao, B.o. Ding, Y. Shu, G. He, H. Chen,
Heterogeneous activation of persulfate by NiFe2− xCoxO4-RGO for oxidative
Supplementary data to this article can be found online at https://doi.
degradation of bisphenol A in water, Chem Eng J 365 (2019) 259–269.
org/10.1016/j.cej.2022.134732. [25] L. Tang, Y. Liu, J. Wang, G. Zeng, Y. Deng, H. Dong, H. Feng, J. Wang, B.o. Peng,
Enhanced activation process of persulfate by mesoporous carbon for degradation of
References aqueous organic pollutants: Electron transfer mechanism, Appl Catal B: Environ
231 (2018) 1–10.
[26] X. Zhou, J. Li, X. Cai, Q. Gao, S. Zhang, S. Yang, H. Wang, X. Zhong, Y. Fang, In situ
[1] H. Wang, T. Chen, D. Chen, X. Zou, M. Li, F. Huang, F. Sun, C. Wang, D. Shu, H. Liu, photo-derived MnOOH collaborating with Mn2Co2C@C dual co-catalysts boost
Sulfurized oolitic hematite as a heterogeneous Fenton-like catalyst for tetracycline photocatalytic overall water splitting, J Mater Chem A 8 (33) (2020)
antibiotic degradation, Appl Catal B: Environ 260 (2020) 118203–118216. 17120–17127.
[2] J. Cao, L. Lai, B. Lai, G. Yao, X. Chen, L. Song, Degradation of tetracycline by [27] W. Tian, H. Zhang, Z. Qian, T. Ouyang, H. Sun, J. Qin, M.O. Tade, S. Wang, Bread-
peroxymonosulfate activated with zero-valent iron: Performance, intermediates, making synthesis of hierarchically Co@C nanoarchitecture in heteroatom doped
toxicity and mechanism, Chem Eng J 364 (2019) 45–56. porous carbons for oxidative degradation of emerging contaminants, Appl Catal B:
[3] Q. Yi, Y. Gao, H. Zhang, H. Zhang, Y.u. Zhang, M. Yang, Establishment of a Environ 225 (2018) 76–83.
pretreatment method for tetracycline production wastewater using enhanced [28] C. Liu, L. Liu, X. Tian, Y. Wang, R. Li, Y. Zhang, Z. Song, B. Xu, W. Chu, F. Qi,
hydrolysis, Chem Eng J 300 (2016) 139–145. A. Ikhlaq, Coupling metal–organic frameworks and g-C3N4 to derive Fe@N-doped
graphene-like carbon for peroxymonosulfate activation: Upgrading framework
stability and performance, Appl Catal B: Environ 255 (2019) 117763–117774.

15
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

[29] Y. Gao, Y. Zhu, L. Lyu, Q. Zeng, X. Xing, C. Hu, Electronic Structure Modulation of [54] L. Hu, X. Ren, M. Yang, W. Guo, Facet-controlled activation of persulfate by
Graphitic Carbon Nitride by Oxygen Doping for Enhanced Catalytic Degradation of magnetite nanoparticles for the degradation of tetracycline, Sep Purif Technol 258
Organic Pollutants through Peroxymonosulfate Activation, Environ Sci Technol 52 (2021) 118014–118024.
(24) (2018) 14371–14380. [55] X. Luo, M. Shen, J. Liu, Y. Ma, B. Gong, H. Liu, Z. Huang, Resource utilization of
[30] Y. Wang, D. Cao, M. Liu, X. Zhao, Insights into heterogeneous catalytic activation piggery sludge to prepare recyclable magnetic biochar for highly efficient
of peroxymonosulfate by Pd/g-C3N4: The role of superoxide radical and singlet degradation of tetracycline through peroxymonosulfate activation, J Clean Prod
oxygen, Catal Commun 102 (2017) 85–88. 294 (2021) 126372–126383.
[31] C. Guan, J. Jiang, S. Pang, X. Chen, R.D. Webster, T.-T. Lim, Facile synthesis of [56] V.T. Nguyen, T.B. Nguyen, C.W. Chen, C.M. Hung, C.P. Huang, C.D. Dong, Cobalt-
pure g-C3N4 materials for peroxymonosulfate activation to degrade bisphenol A: impregnated biochar (Co-SCG) for heterogeneous activation of peroxymonosulfate
Effects of precursors and annealing ambience on catalytic oxidation, Chem Eng J for removal of tetracycline in water, Bioresour Technol 292 (2019)
387 (2020) 123726–123739. 121954–121962.
[32] M. Xie, J. Tang, L. Kong, W. Lu, V. Natarajan, F. Zhu, J. Zhan, Cobalt doped g-C3N4 [57] H. Guo, H.-Y. Niu, C. Liang, C.-G. Niu, Y. Liu, N. Tang, Y. Yang, H.-Y. Liu, Y.-
activation of peroxymonosulfate for monochlorophenols degradation, Chem Eng J Y. Yang, W.-J. Wang, Few-layer graphitic carbon nitride nanosheet with
360 (2019) 1213–1222. controllable functionalization as an effective metal-free activator for
[33] W.-D. Oh, V.W.C. Chang, Z.-T. Hu, R. Goei, T.-T. Lim, Enhancing the catalytic peroxymonosulfate photocatalytic activation: Role of the energy band bending,
activity of g-C3N4 through Me doping (Me = Cu, Co and Fe) for selective Chem Eng J 401 (2020) 126072–126086.
sulfathiazole degradation via redox-based advanced oxidation process, Chem Eng J [58] Y. Feng, D. Wu, Y. Deng, T. Zhang, K. Shih, Sulfate Radical-Mediated Degradation
323 (2017) 260–269. of Sulfadiazine by CuFeO2 Rhombohedral Crystal-Catalyzed Peroxymonosulfate:
[34] J. Fan, H. Qin, S. Jiang, Mn-doped g-C3N4 composite to activate peroxymonosulfate Synergistic Effects and Mechanisms, Environ Sci Technol 50 (2016) 3119–3127.
for acetaminophen degradation: The role of superoxide anion and singlet oxygen, [59] I. Hussain, M. Li, Y. Zhang, Y. Li, S. Huang, X. Du, G. Liu, W. Hayat, N. Anwar,
Chem Eng J 359 (2019) 723–732. Insights into the mechanism of persulfate activation with nZVI/BC nanocomposite
[35] H. Li, C. Shan, B. Pan, Fe(III)-Doped g-C3N4 Mediated Peroxymonosulfate for the degradation of nonylphenol, Chem Eng J 311 (2017) 163–172.
Activation for Selective Degradation of Phenolic Compounds via High-Valent Iron- [60] Y. Gao, Y. Li, L. Zhang, H. Huang, J. Hu, S.M. Shah, X. Su, Adsorption and removal
Oxo Species, Environ Sci Technol 52 (4) (2018) 2197–2205. of tetracycline antibiotics from aqueous solution by graphene oxide, J Colloid
[36] H. Li, C. Shan, B. Pan, Development of Fe-doped g-C3N4/graphite mediated Interf Sci 368 (1) (2012) 540–546.
peroxymonosulfate activation for degradation of aromatic pollutants via [61] K.-Y. Lin, Y.-C. Chen, Y.-F. Lin, LaMO3 perovskites (M=Co, Cu, Fe and Ni) as
nonradical pathway, Sci Total Environ 675 (2019) 62–72. heterogeneous catalysts for activating peroxymonosulfate in water, Chem Eng Sci
[37] S. Wang, Y. Liu, J. Wang, Iron and sulfur co-doped graphite carbon nitride (FeOy/ 160 (2017) 96–105.
S-g-C3N4) for activating peroxymonosulfate to enhance sulfamethoxazole [62] D. Li, D. Chen, Y. Yao, J. Lin, F. Gong, L. Wang, L. Luo, Z. Huang, L.i. Zhang, Strong
degradation, Chem Eng J 382 (2020) 122836–122847. enhancement of dye removal through addition of sulfite to persulfate activated by
[38] D. Yang, T. Jiang, T. Wu, P. Zhang, H. Han, B. Han, Highly selective oxidation of a supported ferric citrate catalyst, Chem Eng J 288 (2016) 806–812.
cyclohexene to 2-cyclohexene-1-one in water using molecular oxygen over [63] X. Li, Y. Jia, M. Zhou, X. Su, J. Sun, High-efficiency degradation of organic
Fe–Co–g-C3N4, Catal, Sci Technol 6 (1) (2016) 193–200. pollutants with Fe, N co-doped biochar catalysts via persulfate activation, J Hazard
[39] H. Song, Z. Guan, D. Xia, H. Xu, F. Yang, D. Li, X. Li, Copper-oxygen synergistic Mater 397 (2020) 122764–122775.
electronic reconstruction on g-C3N4 for efficient non-radical catalysis for [64] Y. Li, J. Li, Y. Pan, Z. Xiong, G. Yao, R. Xie, B. Lai, Peroxymonosulfate activation on
peroxydisulfate and peroxymonosulfate, Sep Purif Technol 257 (2021) FeCo2S4 modified g-C3N4 (FeCo2S4-CN): Mechanism of singlet oxygen evolution for
117957–117967. nonradical efficient degradation of sulfamethoxazole, Chem Eng J 384 (2020)
[40] T. Pan, D. Chen, W. Xu, J. Fang, S. Wu, Z. Liu, K. Wu, Z. Fang, Anionic 123361–123375.
polyacrylamide-assisted construction of thin 2D–2D WO3/g-C3N4 Step-scheme [65] H.A. Bicalho, R.D.F. Rios, I. Binatti, J.D. Ardisson, A.J. Howarth, R.M. Lago, A.P.
heterojunction for enhanced tetracycline degradation under visible light C. Teixeira, Efficient activation of peroxymonosulfate by composites containing
irradiation, J Hazard Mater 393 (2020) 122366–122380. iron mining waste and graphitic carbon nitride for the degradation of
[41] J. Mu, J. Li, X. Zhao, E.-C. Yang, X.-J. Zhao, Cobalt-doped graphitic carbon nitride acetaminophen, J Hazard Mater 400 (2020) 123310–123320.
with enhanced peroxidase-like activity for wastewater treatment, RSC Adv 6 (42) [66] Z. Dong, C. Niu, H. Guo, H. Niu, S. Liang, C. Liang, H. Liu, Y. Yang, Anchoring
(2016) 35568–35576. CuFe2O4 nanoparticles into N-doped carbon nanosheets for peroxymonosulfate
[42] W.-D. Oh, C.-Z. Ng, S.-L. Ng, J.-W. Lim, K.-H. Leong, Rapid degradation of organics activation: Built-in electric field dominated radical and non-radical process, Chem
by peroxymonosulfate activated with ferric ions embedded in graphitic carbon Eng J 426 (2021) 130850–130863.
nitride, Sep Purif Technol 230 (2020) 115852–115860. [67] X. Cheng, H. Guo, W. Li, B. Yang, J. Wang, Y. Zhang, E. Du, Metal-free
[43] M. Noorisepehr, B. Kakavandi, A.A. Isari, F. Ghanbari, E. Dehghanifard, N. Ghomi, carbocatalysis for persulfate activation toward nonradical oxidation: Enhanced
F. Kamrani, Sulfate radical-based oxidative degradation of acetaminophen over an singlet oxygen generation based on active sites and electronic property, Chem Eng
efficient hybrid system: Peroxydisulfate decomposed by ferroferric oxide J 396 (2020) 125107–125121.
nanocatalyst anchored on activated carbon and UV light, Sep Purif Technol 250 [68] S. Chen, D. Huang, G. Zeng, W. Xue, L. Lei, P. Xu, R. Deng, J. Li, M. Cheng, In-situ
(2020) 116950–116971. synthesis of facet-dependent BiVO4/Ag3PO4/PANI photocatalyst with enhanced
[44] W. Yan, L. Yan, C. Jing, Impact of doped metals on urea-derived g-C3N4 for visible-light-induced photocatalytic degradation performance: Synergism of
photocatalytic degradation of antibiotics: Structure, photoactivity and degradation interfacial coupling and hole-transfer, Chem Eng J 382 (2020) 122840–122855.
mechanisms, Appl Catal B: Environ 244 (2019) 475–485. [69] Z. Pi, K. Hou, F. Yao, L. He, S. Chen, Z. Tao, P. Zhou, D. Wang, X. Li, Q. Yang, In-
[45] X. Zhou, J. Luo, B. Jin, Z. Wu, S. Yang, S. Zhang, Y. Tian, Y. Fang, Y. Hou, X. Zhou, situ regeneration of tetracycline-saturated hierarchical porous carbon by
Sustainable synthesis of low-cost nitrogen-doped-carbon coated Co3W3C@g-C3N4 peroxydisulfate oxidation process: Performance, mechanism and application,
composite photocatalyst for efficient hydrogen evolution, Chem Eng J 426 (2021) Chem Eng J 427 (2022) 131749–131759.
131208–131218. [70] G. Yang, Y. Liang, Z. Xiong, J. Yang, K. Wang, Z. Zeng, Molten salt-assisted
[46] Y. Yang, L. Geng, Y. Guo, J. Meng, Y. Guo, Easy dispersion and excellent visible- synthesis of Ce4O7/Bi4MoO9 heterojunction photocatalysts for Photo-Fenton
light photocatalytic activity of the ultrathin urea-derived g-C3N4 nanosheets, Appl degradation of tetracycline: Enhanced mechanism, degradation pathway and
Surf Sci 425 (2017) 535–546. products toxicity assessment, Chem Eng J 425 (2021) 130689–130701.
[47] W. Guo, J. Zhang, G. Li, C. Xu, Enhanced photocatalytic activity of P-type (K, Fe) [71] X. Huang, X. Zhang, K. Zhang, X. Xue, J. Xiong, Y. Huang, D. Zhang, J. Zhang,
co-doped g-C3N4 synthesized in self-generated NH3 atmosphere, Appl Surf Sci 470 Z. Zhang, F. Yan, Defect-mediated Z-scheme carriers’ dynamics of C-ZnO/A-CN
(2019) 99–106. toward highly enhanced photocatalytic TC degradation, J Alloy Compd 877 (2021)
[48] Z. Zhu, X. Tang, T. Wang, W. Fan, Z. Liu, C. Li, P. Huo, Y. Yan, Insight into the 160321–160333.
effect of co-doped to the photocatalytic performance and electronic structure of g- [72] X. Liu, L. Huang, X. Wu, Z. Wang, G. Dong, C. Wang, Y. Liu, L. Wang, Bi2Zr2O7
C3N4 by first principle, Appl Catal B: Environ 241 (2019) 319–328. nanoparticles synthesized by soft-templated sol-gel methods for visible-light-driven
[49] Y. Feng, C. Liao, L. Kong, D. Wu, Y. Liu, P.-H. Lee, K. Shih, Facile synthesis of catalytic degradation of tetracycline, Chemosphere 210 (2018) 424–432.
highly reactive and stable Fe-doped g-C3N4 composites for peroxymonosulfate [73] A. Serra, E. Gomez, J. Michler, L. Philippe, Facile cost-effective fabrication of Cu@
activation: A novel nonradical oxidation process, J Hazard Mater 354 (2018) Cu2O@CuO–microalgae photocatalyst with enhanced visible light degradation of
63–71. tetracycline, Chem Eng J 413 (2021) 127477–127490.
[50] A.A. Isari, S. Moradi, S.S. Rezaei, F. Ghanbari, E. Dehghanifard, B. Kakavandi, [74] F. Hu, W. Luo, C. Liu, H. Dai, X. Xu, Q. Yue, L. Xu, G. Xu, Y. Jian, X. Peng,
Peroxymonosulfate catalyzed by core/shell magnetic ZnO photocatalyst towards Fabrication of graphitic carbon nitride functionalized P-CoFe2O4 for the removal of
malathion degradation: Enhancing synergy, catalytic performance and mechanism, tetracycline under visible light: Optimization, degradation pathways and
Sep Purif Technol 275 (2021) 119163–119176. mechanism evaluation, Chemosphere 274 (2021) 129783–129796.
[51] Z. Teng, N. Yang, H. Lv, S. Wang, M. Hu, C. Wang, D. Wang, G. Wang, Edge- [75] B. Wang, C. Bozal-Ginesta, R. Zhang, B. Zhou, H. Ma, L. Jiao, L. Xu, E. Liu, C. Wang,
Functionalized g-C3N4 Nanosheets as a Highly Efficient Metal-free Photocatalyst Z. Li, A supramolecular H12SubPcB-OPhCOPh/TiO2 Z-scheme hybrid assembled via
for Safe Drinking Water, Chem 5 (3) (2019) 664–680. dimeric concave-ligand π-interaction for visible photocatalytic oxidation of
[52] R. Guan, X. Yuan, Z. Wu, H. Wang, L. Jiang, J. Zhang, Y. Li, G. Zeng, D. Mo, tetracycline, Appl Catal B: Environ 298 (2021) 120550–120562.
Accelerated tetracycline degradation by persulfate activated with heterogeneous [76] J. Pan, F. Guo, H. Sun, Y. Shi, W. Shi, Nanodiamonds anchored on porous ZnSnO3
magnetic NixFe3− xO4 catalysts, Chem Eng J 350 (2018) 573–584. cubes as an efficient composite photocatalyst with improved visible-light
[53] L. Niu, G. Zhang, G. Xian, Z. Ren, T. Wei, Q. Li, Y. Zhang, Z. Zou, Tetracycline
degradation by persulfate activated with magnetic γ-Fe2O3/CeO2 catalyst:
Performance, activation mechanism and degradation pathway, Sep Purif Technol
259 (2021) 118156–118167.

16
Z. Wu et al. Chemical Engineering Journal 434 (2022) 134732

photocatalytic degradation of tetracycline, Sep Purif Technol 263 (2021) [78] J. Wang, D. Zhi, H. Zhou, X. He, D. Zhang, Evaluating tetracycline degradation
118398–118407. pathway and intermediate toxicity during the electrochemical oxidation over a Ti/
[77] Y. Li, B. Yu, Z. Hu, H. Wang, Construction of direct Z-scheme SnS2@ZnIn2S4@ Ti4O7 anode, Water Res 137 (2018) 324–334.
kaolinite heterostructure photocatalyst for efficient photocatalytic degradation of
tetracycline hydrochloride, Chem Eng J 429 (2022) 132105–132117.

17

You might also like