8 PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Hazardous Materials 420 (2021) 126686

Contents lists available at ScienceDirect

Journal of Hazardous Materials


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

Facile construction of dual heterojunction CoO@TiO2/MXene hybrid with


efficient and stable catalytic activity for phenol degradation with
peroxymonosulfate under visible light irradiation
Mingmei Ding a, b, c, Wang Ao a, c, Hang Xu a, c, *, Wei Chen a, c, Lin Tao a, c, Zhen Shen a, c,
Huihong Liu a, c, Chunhui Lu d, Zongli Xie b, **
a
Ministry of Education Key Laboratory of Integrated Regulation and Resource Development on Shallow Lakes, Hohai University, Nanjing 210098, PR China
b
CSIRO Manufacturing, Private Bag 10, Clayton South, VIC 3169, Australia
c
College of Environment, Hohai University, Nanjing 210098, PR China
d
State Key Laboratory of Hydrology-Water Resources and Hydraulic Engineering, Hohai University, Nanjing 210098, PR China

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

Editor: Dr. B. Lee Photocatalysis and peroxymonosulfate (PMS) based advanced oxidation processes (AOP) are emerging tech­
nology for the degradation of refractory organic pollutant in the field of water treatment. Here we report a novel
Keywords: CoO@TiO2/MXene (CTM) hybrid through a facile sonication-hydrothermal method for efficient degradation of
Dual heterojunction phenol via an integrated PMS activation-photocatalysis process. Benefiting from the proper position and superior
Photocatalysis
suitability of the band structure, a dual interfacial charge transport channel was constructed, boosting the
Peroxymonosulfate activation
separation of photo-generated charge carriers and generating sufficient potential difference for redox reaction.
MXene
Cobalt monoxide Accordingly, the CTM hybrid not only possessed the outstanding photocatalytic activity but also dramatically
accelerated PMS activation to generate considerable reactive radicals. As a result, over 96% phenol degradation
was achieved in the 10% CTM/PMS/Vis system within 15 min. The radical quenching test and EPR analysis
1
reveal that SO•−
4 , O2 and O2 were predominant reactive species involved in the catalytic process. Moreover, the
•−

damaged chemical structure of CoO during PMS activation could be healed by the photo-actuated Co(II)
regeneration to allow for continuous and stable catalytic process. This study offers a promising perspective for
the rational design of competent and stable hybrid heterojunction catalyst to construct PMS activation-
photocatalysis processes for the efficient degradation of organic contaminants.

1. Introduction activate PMS for the generation of SO•− 4 (Zhang et al., 2014; Liu et al.,
2018; Shao et al., 2017). However, the ion leaching in conventional
Advanced oxidation processes (AOPs) using highly reactive oxidizing homogeneous system causing the secondary pollution due to the toxicity
radicals such as hydroxyl and sulfate radicals have gained extensive and potential carcinogenicity prevents its practical application (Chen
research attention for the degradation of refractory pollutant in waste­ et al., 2019a). Encouragingly, the heterogeneous catalysts provide a new
water (Ahn et al., 2016; Oturan and Aaron, 2014). Due to the consid­ way for the PMS activation owing to the advantages of excellent stability
erable merits of sulfate radical (SO•− 4 ) including the higher redox and recyclability (Su et al., 2016; Ding et al., 2013). Among them, cobalt
potential, wide range of pH applicability and long half-life period, per­ monoxide (CoO) with abundant Co(II) active sites is considered as a
oxymonosulfate (PMS) based AOP have gained popularity as a prom­ promising PMS activator (Xie et al., 2020; Zhang et al., 2017b). How­
ising alternative to the traditional Fenton process recently (Chen et al., ever, the deactivation of CoO induced by the rapid transformation from
2019b; Xu et al., 2018; Zhang et al., 2014). It is well known that the Co(II) to inactive Co(III) during catalysis process is major challenge for
transition metal ions (such as Co (II) and Fe (II) ions) can efficiently its further practical application. Therefore, it is highly desirable to

* Corresponding author at: Ministry of Education Key Laboratory of Integrated Regulation and Resource Development on Shallow Lakes, Hohai University, Nanjing
210098, PR China.
** Corresponding author.
E-mail addresses: xuhang810826@hhu.edu.cn (H. Xu), zongli.xie@csiro.au (Z. Xie).

https://doi.org/10.1016/j.jhazmat.2021.126686
Received 12 April 2021; Received in revised form 26 June 2021; Accepted 16 July 2021
Available online 18 July 2021
0304-3894/© 2021 Elsevier B.V. All rights reserved.
M. Ding et al. Journal of Hazardous Materials 420 (2021) 126686

develop a suitable approach for the regeneration of the chemical 2. Materials and methods
structure of CoO to hinder the deterioration of its catalytic activity.
Photocatalysis, as a sustainable and eco-friendly technology, also 2.1. Chemicals and reagents
offers a promising platform for the organic contaminant degradation by
energy conversion (Zhang et al., 2017b, 2017a; Tang et al., 2020). It is Ti3AlC2 (MAX) powder (purity ≥ 98%, 300 mesh) was obtained
generally recognized that separation and transportation of the light from Mingshan Company (Nanjing, China). PMS (KHSO5⋅0.5KHSO4⋅0.5
absorption-induced electron–hole pairs are the core step dominating the K2SO4), 5,5-Dimethyl-1-pyrroline N-oxide (DMPO), benzoquinone (p-
photocatalysis process. However, the conventional semiconductors BQ) and 2,2,6,6-Tetramethyl-4-piperidone (4-oxo-TEMP) were pur­
(such as TiO2, C3N4 and BiVO4) usually suffer from the rapid recombi­ chased from Sinopharm Chemical Reagent Co., Ltd. (Beijing, China).
nation of photo-generated charges, leading to the low photon conversion NaN3 was bought from Aladdin, China. Hydrofluoric acid (HF, 48%),
efficiency (Gao et al., 2018; Golshan et al., 2018; Yan et al., 2016; Zeng methanol (MeOH), tertbutyl alcohol (TBA), disodium ethylenedi­
et al., 2020). MXene, as a new class of two dimensional (2D) materials, aminetetraacetate (EDTA-2Na), Co(NO3)2⋅6H2O, CoO and TiO2 were
has gained enormous attention since its first discovery by Yury Gogotsi’s purchased from Sigma-Aldrich (Beijing, China). All chemical reagents
team (Anasori et al., 2017; Zeng et al., 2021). Due to its excellent metal are of analytical grade without further purification.
conductivity, hydrophilic surface, good stability and superior bending
stiffness, MXene has been widely investigated in the field of water pu­ 2.2. Synthesis of MXene, TiO2/MXene and CoO@TiO2/MXene catalyst
rification (Ding et al., 2020; Guo et al., 2015), electromagnetic shielding
(Han et al., 2016), electrochemical energy storage (Hu et al., 2016; Ling The Ti3C2Tx MXene was prepared by using Ti3AlC2 as precursor and
et al., 2014) and catalysis (Luo et al., 2019; Ran et al., 2017). It has been 48% HF as etchant, following the procedures as reported in our previous
reported that Ti3C2 MXene can be partially converted into crystalline study (Ding et al., 2020). The hierarchical CoO@TiO2/MXenes (CTM)
TiO2 under oxidizing condition owing to the considerable exposure of hybrid were synthesized by a facial sonication - hydrothermal method.
titanium atoms on the MXene surface (Su et al., 2016; Huang et al., Briefly, 0.5 g of Ti3C2Tx MXene particles were dispersed in 25 mL DI
2019). Benefiting from the tunable Fermi level and the outstanding water under ultrasonication for 30 min to form a homogeneous solution.
electrical conductivity (Naguib et al., 2014), MXene could serve as the 1 mL Co(NO3)2⋅6H2O solution with the concentration of 0.2 M was then
electron trapper to boost the separation and transport of slowly added into above suspension under stirring and sonicated for
photo-generated carriers in TiO2 by the interfacial heterojunctions, another 30 min. Subsequently, the above mixture was subjected to hy­
which could significantly improve the visible-light photocatalytic ac­ drothermal treatment at 100 ◦ C for 12 h. After the reaction, the
tivity (Wang et al., 2020b, 2020a; Li et al., 2018a). Thus, the TiO2/M­ as-prepared materials were washed with DI water by centrifugation for 3
Xene composite has attracted increasing attention in the field of water times, and dried at 80 ◦ C for 24 h. The products are labeled as x CTM,
treatment for photocatalytic degradation of contaminant (Miao et al., where x represents 5%, 10% and 20 wt% according to the designed ratio
2020; Shahzad et al., 2018). Although some exciting results have been of CoO relative to the whole catalyst. Additionally, TiO2/MXene (TM)
reported, further enhancing the catalytic performance is still imperative was synthesized using the similar method without addition of Co
to advance the wide application of the photocatalysis. (NO3)2⋅6H2O.
In recent years, the strategy of consuming photo-generated electrons
by PMS molecules has been identified an effective approach to simul­ 2.3. Characterization
taneously suppress the recombination of charge carriers and activate
PMS, which would accelerate the generation of reactive species (Wang The X-ray diffractometer (XRD, Bruker, AXS D8-Focus, Germany)
et al., 2018; Lim et al., 2018). In addition, constructing binary hetero­ was used to determine the crystal phase of the samples within the 2θ
junctions by coupling appropriate semiconductors also demonstrates a range of 5–70◦ at a scanning step of 1 ◦ C/min. High resolution trans­
great prospect in the design of efficient photocatalyst, in which the mission electron microscopy (HR-TEM, FEI Tecnai G2 F20, Japan) and
photo-induced carriers are efficiently separated via the inter Field emission-scanning electron microscopy (FE-SEM, TecnaiF30,
cross-sectional charge transfer, thus leading to the high redox potential Hitachi, Japan) were employed to characterize the morphology and
for controlling reduction and oxidation reactions (Wang et al., 2020a, texture of the as-prepared samples. Energy Dispersive Spectrometer
2020b). In view of these, we hypothesized that incorporating CoO into (EDS) equipped on the HR-TEM performed to analyze the elemental
TiO2/MXene composite to construct a dual charge transport channel components of samples. N2 adsorption– desorption apparatus (Quan­
may provide an ideal platform to simultaneously realize the integration tachrome Autosorb-1 Analyser) was applied to determine the pore size
of PMS activation and photocatalysis to achieve in-situ regeneration of distribution and Brunauer-Emmett-Teller (BET) surface area of the cat­
active sites of Co(II), leading to the fast reaction kinetics and stable alysts. A zeta sizer (Nano Z, Malvern) was employed to determine the
catalytic process. zeta potential of the catalysts. X-ray photoelectron spectroscopy (XPS)
Herein, we reported for the first time the fabrication of a new hier­ was carried out using a ESCALAB 250Xi X-ray photoelectron spec­
archical CoO@TiO2/MXene hybrid catalyst with the dual heterojunction trometer with a 100-eV Al-Kα X-ray radiation to evaluate the chemical
through a facile sonication-hydrothermal approach. The prepared components and valence states of sample surface. The diffuse reflectance
CoO@TiO2/MXene catalyst featured significant self-healing ability from UV–Vis spectra of catalysts was recorded on a UV–Vis spectrophotom­
the destructed chemical structure and exhibited excellent and stable eter (Hitachi 3010, Japan) by employing BaSO4 as the standard refer­
catalytic performance for the degradation of phenol by integrating the ence. Ultraviolet photoelectron spectroscopy (UPS) analysis was
photocatalysis and the PMS activation process. The effect of CoO con­ performed on an unfiltered He I (21.22 eV) gas discharge lamp with total
tent, catalyst and PMS dosage, phenol concentration, initial pH value instrumental energy resolution of 100 meV. Electrochemical impedance
and the coexisting anions on the phenol degradation were explored in spectroscopy (EIS) and the transient photocurrent responses were tested
detail. Moreover, the possible catalytic mechanism was proposed based on a Gamry instrument with a standard three-electrode system, where
on systematic investigation of the photoelectrochemical tests, determi­ the catalyst dip-coated FTO glass was used as working electrode, Pt wire
nation of reactive radicals and band gap structure analysis. was used as the counter electrode, and Ag/AgCl electrode was used as
the reference electrode. 0.1 M Na2SO4 solution was served as the elec­
trolyte with an adjusted pH of 7, and the visible light (> 420 nm) was
provided by a 300 W Xenon lamp. The involved free radicals were
detected by the electron paramagnetic resonance (EPR, Bruker EMX 10/
12 spectrometer), where the 5,5-Dimethyl-1-pyrroline N-oxide (DMPO)

2
M. Ding et al. Journal of Hazardous Materials 420 (2021) 126686

or 2,2,6,6-Tetramethyl-4-piperidone (4-oxo-TEMP) was used as spin- pristine MXene, TiO2/MXene (TM) composite showed weaker XRD
trapping agent. spectra of MXene centered at 8.6◦ and 18.2◦ , while new peaks located at
25.2◦ , 37.7◦ and 47.6◦ appeared, corresponding to the (101), (004) and
2.4. Catalytic performance testing (200) planes of anatase TiO2. This confirms the partial oxidization of
MXene into nanocrystalline anatase in TM sample. For Co incorporated
The catalytic performance of CTM and TM catalysts was evaluated in 10%CTM sample, the XRD spectra centered at 36.2◦ and 41.9◦ which
the integrated PMS activation-photocatalysis process for phenol degra­ corresponds to the (111) and (200) planes of CoO were detected, sug­
dation. In a typical catalytic activity test, 20 mg catalyst was added into gesting the presence of crystalline CoO nanoparticles. The XPS survey
100 mL of phenol solution (20 mg L− 1). The mixture was continuously spectra of 10%CTM hybrid in Fig. S2 showed the appearance of Co peak
stirred under dark condition for 30 min to attain the adsorption - and the increased intensity of O peak compared to pristine MXene,
desorption equilibrium. Next, 20 mg of PMS was added and the obtained indicating the successful introduction of Co and the implantation of
mixture was immediately irradiated by a 300 W Xenon lamp with a UV additional oxygen species. Furthermore, the high-resolution Co 2p
cutoff filter (λ > 420 nm) to trigger the degradation. At each sampling spectra of 10%CTM (Fig. 2b) could be fitted into two peaks at 781.5 and
interval, 1 mL of aqueous aliquot was withdrawn and filtered by a 0.22 787.3 eV, which were assigned to Co (II) and the satellite peak (Savio
µm polyethersulfone syringe filter prior to the analysis. et al., 2016; Yan et al., 2018; Chang et al., 2017; Liu and Shao, 2017;
The phenol concentration was measured by the ultra-high- Meng et al., 2017), confirming the existence of CoO.
performance liquid chromatography (HPLC, Agilent) equipped with a Raman spectroscopy was employed to validate the partial conversion
reversed phase Diamonsil C18 column under UV wavelength of 254 nm. of MXene into TiO2 and amorphous carbon (Fig. 2c). The A1g modes of
The mobile phase was a mixture of methanol and ultrapure water (vol/ MXene at 189 cm− 1 (ω2) and 711 cm− 1 (ω3) could be assigned to the
vol, 65/35) with a flow rate of 1.0 mL/min. 20 μL volume of solution was out-of-plane vibrations of Ti and C atoms. After the sonication-
injected and measured at 30 ℃. Total organic carbon (TOC) was eval­ hydrothermal process, both TM and 10%CTM catalysts showed
uated by the micro TOC analyzer (Elementar, Germany) to determine distinct Raman bands of anatase TiO2 at 156, 399, 515 and 633 cm− 1,
the mineralization of phenol. which are corelated to Eg(1), B1g(1), A1g & B1g(2) and the Eg(3) vibrational
modes, respectively (Huang et al., 2019; Gao et al., 2015). Moreover,
two carbon-related broad peaks between 1350 cm− 1 and 1600 cm− 1 are
3. Results and discussion
the characteristic for the D band and G band. The intensity ratios of D-
and G-band (ID/IG) for TM and 10%CTM were 0.866 and 0.922,
3.1. Characterization of materials
respectively, revealing the formation of highly disordered carbon layer.
Fig. 2d shows N2 adsorption-desorption isotherms of MXene, TM and
Fig. 1 shows the synthesis strategy of hierarchical CTM hybrid.
10%CTM catalysts. All catalysts exhibited a typical type IV curve with a
Firstly, the precursor of Ti3AlC2 phase was etched by HF to eliminate the
hysteresis loop for mesoporous material. When compared with the pure
aluminum layer and form the Ti3C2Tx MXene bulk with the typical
MXene (10.4 m2 g− 1), the BET specific surface areas of TM and 10%CTM
accordion-like multi-layer structure. The successful formation of MXene
catalysts were increased to 41.3 m2 g− 1 and 47.6 m2 g− 1, respectively.
was confirmed by a slight shift of the (002) plane to lower angle and the
The high surface area could benefit the exposure of adsorption/reaction
disappearance of the diffraction peaks around 39◦ in the XRD patterns
sites for PMS activation and photocatalysis. The corresponding pore size
(Fig. S1) (Natarajan et al., 2018). The CTM catalyst was then fabricated
distribution was obtained by the BJH method using the desorption data.
by the sonication-hydrothermal process, in which the MXene bulk was
The pore size of 10% CTM is distributed in the range of 1.2–21.5 nm
partially converted to TiO2 nanoparticles, and the CoO nanoparticles
with two distinct pore sizes of ~2.6 nm and ~20 nm (insert of Fig. 2d).
were simultaneously generated and uniformly decorated on the MXene
The morphology of CoO@TiO2/MXene catalyst was further analyzed
surface, forming a hierarchical structure. It was noted that the TiO2
by the electron microscopy. As shown in Fig. 1 and Fig. S3, SEM images
nanoparticles were closely connected to the unoxidized MXene nano­
indicates the TM, 5%CTM and 10%CTM catalyst maintained the layered
sheets via the amorphous carbon“binder”. The ICP-OES analysis iden­
structure and exhibited a rougher surface compared to the pristine
tifies that the actual content of CoO in 5% CTM, 10%CTM and 20%CTM
MXene owing to the formation of small particles on the nanosheet sur­
catalysts is 2.21, 5.39 and 11.42 wt%, respectively.
face. However, when the CoO loading was further increased in 20%
XRD pattern was used to investigate the structure and components of
CTM, the aggregation of large particles and fragmentized nanosheets
CTM composite and the results are shown in Fig. 2a. Comparing with the

Fig. 1. Schematic illustration for the synthesis strategy of CoO@TiO2/MXene hybrid.

3
M. Ding et al. Journal of Hazardous Materials 420 (2021) 126686

Fig. 2. (a) XRD pattern of pure MXene, TM and 10%CTM catalysts, (b) High-resolution Co 2p XPS spectra of the 10%CTM. (c) Raman spectra of pure MXene and 10%
CTM. (d) N2 adsorption-desorption isotherms of TM, 10% CTM and pure MXene catalysts (insert shows the pore size distribution of 10% CTM catalyst).

were observed. Fig. S4 displays the serious aggregation of pure CoO respectively (Fig. S6). Furthermore, MXene alone exhibited weak cata­
particles. TEM images manifest that abundant nanoparticles were lytic activity for the phenol degradation in the constructed MXene/PMS,
decorated on the nanosheet surface 10%CTM (Fig. 3a), illustrating the MXene/Vis and MXene/PMS/Vis systems, wherein the corresponding
structure evolution compared to pure MXene. EDS elemental mapping of removal efficiencies of phenol after 15 min were 16.4%, 3.6% and
10%CTM in Fig. 3c displays that Ti, C, O and F species are well 17.3%, respectively, which demonstrates the limited activity of pristine
distributed in the whole area, while the Co element are aggregated as the MXene for the PMS activation and photocatalysis. After partial oxida­
uniformly dispersed clusters with the size of 10–20 nm. As revealed by tion, the TM catalyst showed improved degradation efficiency of phenol
the HRTEM images (Fig. 3b), the interlayer spacing of 0.35 nm could be under visible light irradiation (27.6%). The degradation efficiency was
ascribed to the exposure of (101) facets of anatase TiO2, which was further enhanced in the integrated TM/PMS/Vis catalytic system
interconnected with the MXene substrate by the amorphous carbon. This (38.2%) (Fig. S6). However, almost no phenol degradation was observed
will favor the migration of photoinduced electrons and the formation of in the commercial TiO2/Vis system owing to the rapid recombination of
the Schottky junction. Additionally, the inter-planar distance of 0.19 nm photo-induced carriers (Fig. S7). The higher photocatalytic property of
corresponding to the (200) plane of CoO, further confirms the presence TM composite compared to commercial TiO2 could be ascribed to the
of CoO. These findings validate the hierarchical nanostructure with CoO “bridge” effect of carbon layer, whereby the photo-generated electrons
uniformly anchoring on TiO2/MXene composite, benefiting construction could be fast captured and transported from the conduction band of TiO2
of the dual charge transport channel. to MXene nanosheets, and thereby forming the Schottky junction and
hindering the charge-recombination process (Wang et al., 2020b, 2020a;
Li et al., 2018b, 2018a; Peng et al., 2018; Zhang et al., 2016). Addi­
3.2. Catalytic performance and stability
tionally, the TM catalyst displayed limited catalytic activity for PMS
activation without visible light irradiation, leading to lower phenol
3.2.1. Catalytic performance
removal (18.8%). By contrast, the degradation efficiency of phenol was
In this work, to evaluate the catalytic activity of CTM hybrid in
significantly improved when the 10%CTM hybrid was used as the
conjunction with photocatalysis and PMS activation for phenol degra­
catalyst. In 15 min, the 10%CTM/PMS, 10%CTM/Vis and 10%
dation, a series of control experiments were conducted initially. As
CTM/PMS/Vis systems achieved the phenol degradation of 67.2%,
shown in Fig. 4a, negligible degradation towards phenol was observed
47.7% and 96.6%, respectively. This result indicates that the 10%CTM
when only PMS or visible light was used, revealing that PMS or visible
possessed dual catalytic activity for both the photocatalysis and PMS
light irradiation alone is limited in generating reactive radicals. Simi­
activation, which results in the considerable generation of reactive
larly, when the catalyst alone was added into the phenol solution
radicals for organic contaminant degradation. The superior catalytic
without PMS and visible light, there was negligible change in the phenol
performance of 10%CTM could be explained as follows: Firstly, the
concentration, suggesting a negligible adsorption of catalysts for phenol
higher BET surface area is beneficial to the exposure of considerable
(Fig. S5). As a comparison, the integration of photocatalysis and PMS
active sites for the dual catalysis process. Secondly, the formation of
activation shows enhancing effect for phenol degradation. When using
heterojunction in the catalyst would serve as the charge transfer chan­
CoO alone as catalyst, the phenol removal efficiencies in the CoO/PMS,
nels for the efficiently separation and migration of photogenerated
CoO /Vis and CoO /PMS/Vis systems were 41.9%, 22.6% and 56.6%,

4
M. Ding et al. Journal of Hazardous Materials 420 (2021) 126686

Fig. 3. (a) TEM image, (b) HRTEM image and (c) STEM image and EDS elemental mapping of 10%CTM.

Fig. 4. (a) Degradation efficiency of phenol and (b) the values of degradation kinetics constants of phenol in different systems over 15 min. Conditions: [phenol] =
20 mg L− 1, [PMS] = 0.2 g L− 1, [catalyst] = 0.2 g L− 1, initial pH 7, λ > 420 nm, and T = 25 ◦ C.

5
M. Ding et al. Journal of Hazardous Materials 420 (2021) 126686

carriers, thus facilitating the integrated PMS activation-photocatalysis photocatalysis.


process.
It is worth noting, despite achieving 96.6% of phenol degradation, 3.2.2. Effect of operating parameters and the catalyst stability
the TOC removal in the 10%CTM/PMS/Vis system was only 50.7% Fig. 5a shows that the degradation efficiency of phenol gradually
within 15 min (Fig. S8), indicating the formation of the reaction in­ increased from 79.1% to 96.6% when the CoO loadings in CTM
termediates in the solution. Furthermore, the kinetic study reveals that increased was from 2.21% (5%CTM) to 5.39% (10%CTM), which could
the phenol degradation follow the pseudo-first-order model, and the be ascribed to the increased active sites from CoO, boosting the PMS
corresponding k values of different systems are displayed in Fig. 4b. The activation and offering more heterojunction sites for charge transfer.
k values for phenol degradation in the 10%CTM /PMS system are higher However, there was a slight decline in the degradation efficiency of
than those in the TM /PMS system, suggesting that CoO plays important phenol with the CoO content further increasing to 11.42% (20%CTM).
role in the PMS activation to produce free radicals. In addition, the This is probably because the excessive addition of Co species resulted in
higher k value of the 10%CTM/Vis system compared to that of TM/Vis the agglomeration of CoO nanoparticles, the coverage of TiO2 and
system also confirms the improved photocatalytic activity of 10%CTM, fragmentized MXene nanosheets (Fig. S3), resulting in the suppression
which is mainly contributed to the rapid interfacial charge transfer at of catalytic activity. 10%CTM was therefore chosen as the optimized
the hybrid heterojunction. Remarkably, 10%CTM/PMS/Vis system catalyst for the subsequent studies.
exhibited outstanding catalytic performance for the phenol degradation, A further evaluation for the catalytic performance of the 10%CTM
even surpassing other catalytic system reported in the previous studies was conducted by varying the operating parameters including 10%CTM
(Table S1). These findings highlight the significant reactivity of CTM for dosage, PMS dosage, phenol concentration, initial pH and coexisting
the integrated catalytic process combining PMS activation and anions. Fig. 5b shows that the degradation efficiency of phenol exhibited

Fig. 5. Effect of the (a) Co content, (b) catalyst (10% CTM) dosage, (b) PMS dosage, (d) phenol concentration (e) pH and (d) coexisting anions on the removal of
phenol. Conditions: [phenol] = 20 mg L− 1, [PMS] = 0.2 g L− 1, [catalyst] = 0.2 g L− 1, initial pH 7, λ > 420 nm, T = 25 ◦ C, [ions] = 10 mM, unless other­
wise specified.

6
M. Ding et al. Journal of Hazardous Materials 420 (2021) 126686

positive correlation with the dosage of 10%CTM catalyst. This is mainly To evaluate the stability of 10% CTM catalyst in the simultaneous
because that the increased catalyst amount improved the availability of PMS/visible light system, recycle tests were performed for the phenol
redox-active centers for photon adsorption and PMS activation, which in degradation. As displayed in Fig. 6a, the removal efficiency of phenol,
turn benefiting the generation of reactive radical species and conse­ can still maintain 87.3% after five successive cycles, highlighting that
quently the catalytic activity. the 10% CTM catalyst has a good stability in the combined catalytic
Generally, the PMS concentration would directly influence the process. During the catalysis process, the concentration of leached co­
amount of generated •OH and SO•− 4 species. As depicted in Fig. 5c, the balt ions was detected to be 0.52 μg/L (Fig. S9), which was well below
degradation efficiency of phenol gradually accelerated with the PMS the maximum permissible concentration (1 mg L− 1) on the basis of the
concentration increasing from 0.1 g L− 1 to 0.5 g L− 1, while decreased Chinese National Standard (GB 25467-2010) (Li et al., 2018a, 2018b).
when the PMS concentration was further increased to 0.8 g L− 1. This The slight decline in the catalytic activity of 10%CTM during the recycle
phenomenon could be explained by three reasons: (i) the appropriate test was probably attributed to the coverage of active sites by the organic
PMS could act as the charge acceptor to enhance the catalytic process, compounds. Moreover, the XRD spectra in Fig. 6b demonstrate that the
(ii) the photogenerated electrons and Co(II) active sites are not sufficient intensity of the diffraction peaks of CoO slightly declined after the
to activate the excessive PMS and (iii) the excessive PMS would react recycle test. This could be due to the partial conversion of the CoO
with the SO•− 4 radicals to generate the SO5 species with weaker
•−
component and the adsorption of intermediate products on the material
oxidative ability, reducing the removal efficiency (Eq. (1)) (Golshan surface.
et al., 2018). Fig. 5d illustrates the effect of phenol concentration on the
degradation efficiency. As expected, the degradation efficiency of 3.3. Proposed catalysis mechanism
phenol decreased from 99.8% to 96.6%, 76.3% and 56.7% at 15 min
when the phenol concentration was increased from 10 to 20, 50 and In order to identify the active radicals in the 10%CTM/PMS/Vis
100 mg L− 1, respectively. Higher amount of phenol would increase system, a series of EPR experiments using DMPO and TEMP as spin-
coverage of active sites on the catalyst surface and hinder the penetra­ trapping reagents were carried out. As depicted in Fig. 7, the distinct
tion of visible light, thus inhibiting the generation of radical species for DMPO- •OH and DMPO- SO•− 4 signals were observed in the dark con­
the elimination of the phenol. dition with the existence of PMS, which was derived from the PMS
activation by CoO nanoparticles. After the visible light irradiation, the
SO•− − •− 2− +
4 + HSO5 → SO5 + SO4 + H /OH

(1)
intensity of DMPO signals gradually strengthened with the increasing of
Fig. 5e shows the influence of the initial pH of the solution on phenol illumination time, suggesting that visible light could accelerate the PMS
degradation. It was found that the 10%CTM/PMS/Vis system main­ activation for the generation of •OH and SO•− 4 radicals. Moreover, no
tained a good catalytic performance over a wide pH range from 3 to 11 characteristic peaks of TEMP-1O2 and DMPO-O•− 2 were detected in the
and the removal efficiency of phenol improved with the increasing pH. 10%CTM/PMS system. The corresponding signals were only observed
In acidic conditions, the interaction between H+ and HSO−5 become after the illumination and increased over time, which demonstrates that
1
significant and could hinder the catalysis process (Lim et al., 2018). At both O•− 2 and O2 signals were produced during the photocatalysis
the same time, the consumption of free radicals including •OH and SO•− 4
process.
by excess H+ (Eqs. (2 and 3)) could also be responsible for the declined Radical quenching experiments were further performed to evaluate
degradation efficiency (Wang et al., 2017). When under the alkaline the contribution of the involved reactive species in the 10%CTM/PMS/
condition, OH− could participate in the PMS activation process, accel­ Vis system. In this work, five scavengers including MeOH, TBA, p-BQ,
erating the generation of reactive radical species (Chen et al., 2020). EDTA-2Na and NaN3 were applied to capture the •OH, SO•− 4 , O2 , h ,
•− +
1
and O2, respectively, owing to the corresponding rapid reaction rate
•OH + H+ + e− → H2O (2) (Lim et al., 2018; Gong et al., 2018). Specially, MeOH was used as the
quenching agent for both ⋅OH and SO•- 4 while TBA was scavenger of ⋅OH.
SO•− +H +e → + −
HSO•− (3)
It is clear from Fig. 8 that all scavengers exhibited inhibition perfor­
4 4

Considering the natural environment and practical application, it is mance for the removal efficiency of phenol and followed the order of
necessary to evaluate the interference of inorganic anions. Fig. 5f MeOH > p-BQ > NaN3 > EDTA-2Na > (TBA), revealing that •OH,
1
demonstrates that the introduction of inorganic anions apparently SO•−4 , O2 , h , and O2 all contributed to the degradation of phenol.
•− +

retarded the degradation efficiency of phenol, following the order of Remarkably, the presence of MeOH, p-BQ and NaN3 showed more sig­
HCO−3 > NO−3 > Cl− > SO2− 4 . This result was mainly derived from the nificant influence on phenol degradation with smaller reaction rate
1
scavenging effect of anions for the photo-generated holes and free rad­ constants, indicating that SO•− 4 , O2 and O2 are the predominant
•−

icals (such as •OH and SO•− 4 ), as described in Eqs. (4–11) (Jorfi et al., reactive species in the oxidation process.
2017; Wang et al., 2016), which would lead to the production of radical The light-harvesting capability is one of the crucial factors governing
species with a lower oxidation ability. Moreover, it is clear that SO2− 4 the combined catalysis process. The UV–visible diffuse reflectance
also exhibited some minor inhibition on the phenol degradation, which spectroscopy was conducted to investigate the light absorption proper­
is still ambiguous to explain and deserves further research. ties of the catalysts and the results are shown in Fig. 9a. Pure MXene
showed the highest light absorption from 230 to 800 nm owing to its
Cl− + SO•− 2−
4 → SO4 + •Cl (4) characteristic of black color, while no obvious absorption edge was
Cl •−
+ OH → ClOH •−
(5) observed, revealing the metallic nature of MXene. By contrast, TM
exhibited the distinct absorption edge at 449 nm mainly due to the

Cl + h → •Cl +
(6) formation of TiO2. With the addition of CoO, 10% CTM showed
improved optical adsorption throughout the measured UV–vis region
NO−3 + SO•− → SO2− + •NO3 (7)
4 4
with the absorption edges at 575 nm, which would benefit the utiliza­
NO−3 + •OH → •NO3 + OH− (8) tion of visible light for the photocatalytic reaction. According to the
empirical equation (Chen et al., 2020), the band gap energy (Eg) of TM
NO−3 +
+ h → •NO3 (9) and 10% CTM catalyst was calculated to be 2.76 eV and 2.15 eV,
HCO−3 + SO•− → SO2− +
+H + CO•− (10) respectively. The narrower band gap of 10% CTM is probably derived
4 4 3
from the synergistic effect between CoO and TM through significant
HCO−3 + •OH → CO•−
3 + H2O (11) interfacial coupling interaction. It has been reported that the band gap
energy of commercial TiO2 is about 3.1 eV (Wang et al., 2013), which is

7
M. Ding et al. Journal of Hazardous Materials 420 (2021) 126686

Fig. 6. (a) Recycle test of 10% CTM/PMS/Vis system for the degradation of phenol. (b) XRD pattern of 10% CTM before and after recycle test.

1
Fig. 7. EPR spectra of 10%CTM/PMS/Vis system for (a) DMPO- •OH and DMPO- SO•−
4 , (b) DMPO- O2 and (c) TEMP- O2.
•−

Fig. 8. (a) Degradation efficiency of phenol and (b) the values of degradation kinetics constants of phenol with the presence of different scavengers. Conditions:
[phenol] = 20 mg L− 1, [PMS] = 0.2 g L− 1, [catalyst] = 0.2 g L− 1, [MeOH] = [TBA] = 2 M, [p-BQ] = [EDTA-2Na] = [NaN3] = 12 mM, initial pH 7,
λ > 420 nm, and T = 25 ◦ C.

higher than that of TM. This is probably ascribed to the “bridge” effect of relative negative shift of C1s state in the binding energy of TM was
the carbon layer, significantly boosting the interfacial electron transfer. observed, which manifests that the photogenerated electrons were
In addition, the band gap of CoO was evaluated to be 2.2 eV (Fig. S10), moved from TiO2 to MXene nanosheets, leading the formation of
which is in accordance with the previous reports (Mao et al., 2017). Schottky junction between TiO2 and MXene.
Heterojunction offers a special channel for the transport of photo- To gain deep insight into the coupling interaction between TM and
generated charge carriers in hybrids, which could be identified by the CoO, ultraviolet photoelectron spectroscopy (UPS) was further applied
interfacial interaction of each composition and the surface chemical to evaluate the energy band structures of these two materials. As shown
state. The Fermi level of pure MXene (− 0.04 eV vs. NHE) was lower than in Fig. 9b, the valence band (VB) energy (Ev) of TM and CoO was esti­
the conduction band (CB) of the TiO2 (− 0.25 eV vs. NHE) (Low et al., mated to be − 8.18 eV and − 5.52 eV by subtracting the width of the He
2018), suggesting that the photogenerated electrons from TiO2 could be I UPS spectra from the excitation energy (21.22 eV) (Wang et al., 2013;
easily captured by MXene. As the photocatalyst could be in-situ excited Shi et al., 2017). Based on the formula of Eg = Ev − Ec, the corre­
by the X-ray irradiation, the XPS spectrum was used to further confirm spondingly conduction band (CB) energy (Ec) was thus evaluated to be
the migration of electrons (Fig. S11). Compared with pure MXene, a − 5.42 eV and − 3.32 eV for TM and CoO, respectively, equivalent to

8
M. Ding et al. Journal of Hazardous Materials 420 (2021) 126686

Fig. 9. (a) UV–vis diffuse reflectance spectrum of MXene, TM and 10%CTM. (b) UPS spectra and (c) the energy level diagrams of TM and CoO. (d) High-resolution
XPS spectrum of Co2p in 10%CTM/PMS/Vis and 10%CTM/PMS system.

0.72 V and − 1.12 V versus normal hydrogen electrode (NHE). In effectively favor the recovery of Co(II) and restrict the passivation of
principle, the position and compatibility of the conduction band and CoO. As revealed by the high-resolution XPS spectrum of Co2p after
valence band levels of catalysts are essential for the formation of het­ catalytic reaction in Fig. 9d, the appearance of Co(III) in the 10%
erojunction and redox reaction. The energy level diagrams of TM and CTM/PMS system demonstrates the Co(II) as the active site was partially
CoO is demonstrated in Fig. 9c. It is reasonable to deduce that photo­ oxidized during the PMS activation process. It is obvious that the rela­
generated electrons in the CB of the TM would easily migrate to the VB tive content of Co(II) in the 10%CTM/PMS/Vis system (36.8%) is higher
of the CoO and combine with the holes to form the direct Z-scheme than that in 10%CTM/PMS system (32.4%), confirming the self-healing
heterojunction, contributing to the effective separation of charge carries effect of visible light irradiation for the continuous replenishment of Co
and strong oxidation and reduction potentials. In this case, abundant (II).
reactive species including •OH, O•−
2 and SO4 could be produced due to
•−
Transient photocurrent responses and electrochemical impedance
their lower production potentials (•OH /OH- = +2.73 eV, O2/ spectroscopy (EIS) were conducted to verify the separation and migra­
2−
O•−
2 = − 0.33 eV and SO4 / SO4 = 2.5–3.1 V vs NHE, respectively)
•−
tion properties of photo-generated carriers in the CTM hybrid. By
(Wang et al., 2020a, 2020b), which is in accordance with the results of contrast with the TM/Vis system, 10% CTM/Vis system exhibited
EPR. Remarkably, the CB level after band alignment is more negative enhanced photocurrent density and maintained outstanding photo-
than the redox potential of Co(III)/Co(II) in weak acid solution (1.95 V stability after five-cycle successive on-off operation (Fig. 10a), sug­
vs NHE), meeting the thermodynamic requirements for the reduction of gesting that recombination of photo-generated charge carriers was
Co(III) by the interfacial charge transport from TM, which would effectively restricted by the introduction of CoO. After adding PMS, a

Fig. 10. (a) Transient photocurrent responses and (b) EIS Nyquist plots of different systems.

9
M. Ding et al. Journal of Hazardous Materials 420 (2021) 126686

further significant improvement in the photocurrent density was ob­ photocatalytic reduction process (Eq. (19)). Along with these reactions,
tained, indicating that the PMS molecules were prone to serve as the the phenol in the solution would be degraded by the generated reactive
charge acceptor to undergo the activation and benefit the separation species into smaller intermediates, CO2 and H2O.
efficiency of photo-generated electrons and holes pairs. The arc radius
on the EIS Nyquist plot is related to the interfacial charge transport TM/CoO + hv → h+ + e− (12)
ability, wherein the smaller arc size corresponds to the lower charge O2 + e− → O•−
2 (13)
transfer resistance and fast charge migration property. As shown in
Fig. 10b, the EIS Nyquist plot of 10% CTM/PMS/Vis system showed the H2O + h+ → •OH + H+ (14)
smallest arc radius compared to other systems, confirming the acceler­
HSO−5 −
+e → SO•− + OH -
(15)
ated kinetics in the interfacial electron shuttling and highly efficient
4

separation of photo-generated electron-hole pairs, which was in good SO•−


4 + OH → -
SO2−
4 + •OH (16)
agreement with the result of transient photocurrent. Data were fitted
according to an equivalent electrical circuit with two constant-phase O•−
2 + •OH → O2 + OH 1 -
(17)
elements, where the high frequency region is related to surface Co(II) + HSO−5 → Co(III) + SO•− −
(18)
4 + OH
porosity (CPE1-Rp) and the low frequency region reflects the charge
transfer process at the catalyst/electrolyte interface (CPE2-Rct). The

Co(III) + e → Co(II) (19)
fitted values of Rct in 10% CTM/PMS/Vis system was 1695 ohm, which
is lower than that of the other system (10% CTM/Vis for 3709 ohm, TM/
PMS/Vis for 2410 ohm and Mxene/PMS/Vis for 12,319 ohm), suggest­
ing the rapid charge transfer in 10% CTM/PMS/Vis system. 4. Conclusions
On the basis of the above analysis, a possible mechanism for the
stable synergistic effect of photocatalysis and PMS activation in the In summary, a hierarchical CoO@TiO2/MXene hybrid consisting of
CTM/PMS/Vis system was proposed and depicted in Fig. 11. Upon the well-designed Schottky junction and Z scheme heterojunction was pre­
visible light illumination, the electrons excited in the VB of TiO2 firstly pared using a facile sonication-hydrothermal treatment method. By
transported to the conductive MXene through the carbon layer binder, virtue of the improved light harvesting ability, effective interfacial
forming the Schottky junction in the TM composite. Then, photo- carrier separation, rapid recovery of active site and large surface area,
induced electron-hole pairs were generated in both TM composite and the CTM hybrid exhibited superior and continuous reactivity for the
CoO (Eq. (12)). Owing to the proper position and superior suitability of synergistic photocatalysis and PMS activation process. over 96% phenol
the band structure, the electrons in the CB of the TM would recombine degradation was achieved in the 10%CTM/PMS/Vis system within
with the holes in the VB of the CoO to form the direct Z-scheme heter­ 15 min. Based on radical quenching experiment and EPR analysis, it is
1
ojunction, thus leading to the enhanced photocatalytic activity. The concluded that ⋅OH, SO•−4 , O2 , h , and O2 are involved in the oxidation
•− +
1
photogenerated electrons in the CB of CoO were inclined to be trapped process, while SO4 , O2 and O2 are the predominant reactive species
•− •−

by O2 to generate O•− for the phenol degradation. This work would pave an alternative path
2 (Eq. (13)), while the holes in the VB of TM could
directly oxidize organic contaminants or combine with the adsorbed for the rational design of efficient heterostructured catalysts for envi­
H2O to form •OH (Eq. (14)). In the presence of PMS, the photo- ronmental remediation.
generated electrons would further react with PMS molecules, facili­
tating the separation of charge carriers and the production of SO•− 4
CRediT authorship contribution statement
radicals and •OH (Eqs. (15 and 16)). The subsequent reaction between
O•− 1
2 and •OH could generate O2 (Eq. (17)). Besides, PMS also could be
Ding Mingmei: Methodology, Data curation, Writing – original
activated by Co(II) active sites in CoO to form SO•−
4 (Eq. (18)). It is worth
draft. Wang Ao: Conceptualization, Methodology, Software. Xu Hang:
noting that Co(II) could be immediately regenerated by the Supervision, Funding acquisition. Chen Wei: Software, Validation. Lin
Tao: Software, Validation. Shen Zhen: Software, Validation. Liu

Fig. 11. Proposal of catalytic degradation mechanism in CTM/PMS/Vis system.

10
M. Ding et al. Journal of Hazardous Materials 420 (2021) 126686

Huihong: Visualization, Investigation. Lu Chunhui: Software, Valida­ on FeIIFe2IIIO4@C as an efficient activator of peroxymonosulfate. Appl. Catal. B
Environ. 219, 216–230.
tion. Xie Zongli: Writing – reviewing & editing.
Lim, J., Kwak, D.-y, Sieland, F., Kim, C., Bahnemann, D.W., Choi, W., 2018. Visible light-
induced catalytic activation of peroxymonosulfate using heterogeneous surface
complexes of amino acids on TiO2. Appl. Catal. B Environ. 225, 406–414.
Declaration of Competing Interest Ling, Z., Ren, C.E., Zhao, M.Q., Yang, J., Giammarco, J.M., Qiu, J., Barsoum, M.W.,
Gogotsi, Y., 2014. Flexible and conductive MXene films and nanocomposites with
high capacitance. Proc. Natl. Acad. Sci. USA 111, 16676–16681.
The authors declare that they have no known competing financial Liu, Y., Luo, R., Li, Y., Qi, J., Wang, C., Li, J., Sun, X., Wang, L., 2018. Sandwich-like
interests or personal relationships that could have appeared to influence Co3O4/MXene composite with enhanced catalytic performance for Bisphenol A
degradation. Chem. Eng. J. 347, 731–740.
the work reported in this paper.
Liu, G., Shao, J., 2017. Pomegranate-like CoO@nitrogen-doped carbon microspheres
with outstanding rate behavior and stability for lithium storage. J. Mater. Chem. A 5,
Acknowledgements 9801–9806.
Li, Y., Deng, X., Tian, J., Liang, Z., Cui, H., 2018. Ti3C2 MXene-derived Ti3C2/TiO2
nanoflowers for noble-metal-free photocatalytic overall water splitting. Appl. Mater.
This work was supported by National Key Research and Develop­ Today 13, 217–227.
ment Project of China (2016YFC0400803), the National Natural Science Li, X., Huang, X., Xi, S., Miao, S., Ding, J., Cai, W., Liu, S., Yang, X., Yang, H., Gao, J.,
Wang, J., Huang, Y., Zhang, T., Liu, B., 2018. Single cobalt atoms anchored on
Foundation of China (No. 51738013 and No. 51978239), the 111 Project porous N-doped graphene with dual reaction sites for efficient Fenton-like catalysis.
under grant number B20044 and a project funded by the Priority Aca­ J. Am. Chem. Soc. 140, 12469–12475.
demic Program Development of Jiangsu Higher Education Institutions Low, J., Zhang, L., Tong, T., Shen, B., Yu, J., 2018. TiO2/MXene Ti3C2 composite with
excellent photocatalytic CO2 reduction activity. J. Catal. 361, 255–266.
(PAPD). Luo, Y., Chen, G.-F., Ding, L., Chen, X., Ding, L.-X., Wang, H., 2019. Efficient
electrocatalytic N2 fixation with MXene under ambient conditions. Joule 3, 279–289.
Appendix A. Supporting information Mao, Z., Chen, J., Yang, Y., Wang, D., Bie, L., Fahlman, B.D., 2017. Novel g-C3N4/CoO
nanocomposites with significantly enhanced visible-light photocatalytic activity for
H2 evolution. ACS Appl. Mater. Interfaces 9, 12427–12435.
Supplementary data associated with this article can be found in the Meng, C., Ling, T., Ma, T.Y., Wang, H., Hu, Z., Zhou, Y., Mao, J., Du, X.W., Jaroniec, M.,
online version at doi:10.1016/j.jhazmat.2021.126686. Qiao, S.Z., 2017. Atomically and electronically coupled Pt and CoO hybrid
nanocatalysts for enhanced electrocatalytic performance. Adv. Mater. 29, 1604607.
Miao, Z., Wang, G., Zhang, X., Dong, X., 2020. Oxygen vacancies modified TiO2/Ti3C2
References derived from MXenes for enhanced photocatalytic degradation of organic pollutants:
the crucial role of oxygen vacancy to schottky junction. Appl. Surf. Sci. 528, 146929.
Naguib, M., Mashtalir, O., Lukatskaya, M.R., Dyatkin, B., Zhang, C., Presser, V.,
Ahn, Y.Y., Yun, E.T., Seo, J.W., Lee, C., Kim, S.H., Kim, J.H., Lee, J., 2016. Activation of
Gogotsi, Y., Barsoum, M.W., 2014. One-step synthesis of nanocrystalline transition
peroxymonosulfate by surface-loaded noble metal nanoparticles for oxidative
metal oxides on thin sheets of disordered graphitic carbon by oxidation of MXenes.
degradation of organic compounds. Environ. Sci. Technol. 50, 10187–10197.
Chem. Commun. 50, 7420–7423.
Anasori, B., Lukatskaya, M.R., Gogotsi, Y., 2017. 2D metal carbides and nitrides
Natarajan, T.S., Thampi, K.R., Tayade, R.J., 2018. Visible light driven redox-mediator-
(MXenes) for energy storage. Nat. Rev. Mater. 2, 16098.
free dual semiconductor photocatalytic systems for pollutant degradation and the
Chang, L., Wang, K., Huang, L.-a, He, Z., Zhu, S., Chen, M., Shao, H., Wang, J., 2017.
ambiguity in applying Z-scheme concept. Appl. Catal. B Environ. 227, 296–311.
Hierarchical CoO microflower film with excellent electrochemical lithium/sodium
Oturan, M.A., Aaron, J.-J., 2014. Advanced oxidation processes in water/wastewater
storage performance. J. Mater. Chem. A 5, 20892–20902.
treatment: principles and applications. A review. Crit. Rev. Environ. Sci. Technol. 44,
Chen, F., Huang, G.X., Yao, F.B., Yang, Q., Zheng, Y.M., Zhao, Q.B., Yu, H.Q., 2020.
2577–2641.
Catalytic degradation of ciprofloxacin by a visible-light-assisted peroxymonosulfate
Peng, C., Wei, P., Li, X., Liu, Y., Cao, Y., Wang, H., Yu, H., Peng, F., Zhang, L., Zhang, B.,
activation system: performance and mechanism. Water Res. 173, 115559.
Lv, K., 2018. High efficiency photocatalytic hydrogen production over ternary Cu/
Chen, Y., Zhang, G., Ji, Q., Liu, H., Qu, J., 2019a. Triggering of low-valence molybdenum
TiO2@Ti3C2Tx enabled by low-work-function 2D titanium carbide. Nano Energy 53,
in multiphasic MoS2 for effective reactive oxygen species output in catalytic Fenton-
97–107.
like reactions. ACS Appl. Mater. Interfaces 11, 26781–26788.
Ran, J., Gao, G., Li, F.T., Ma, T.Y., Du, A., Qiao, S.Z., 2017. Ti3C2 MXene co-catalyst on
Chen, Y., Zhang, G., Liu, H., Qu, J., 2019b. Confining free radicals in close vicinity to
metal sulfide photo-absorbers for enhanced visible-light photocatalytic hydrogen
contaminants enables ultrafast Fenton-like processes in the interspacing of MoS2
production. Nat. Commun. 8, 13907.
membranes. Angew. Chem. Int. Ed. Engl. 58, 8134–8138.
Savio, A.K.P.D., Fletcher, J., Smith, K., Iyer, R., Bao, J.M., Robles Hernández, F.C., 2016.
Ding, M., Chen, W., Xu, H., Shen, Z., Lin, T., Hu, K., Lu, C.H., Xie, Z., 2020. Novel
Environmentally effective photocatalyst CoO–TiO2 synthesized by thermal
α-Fe2O3/MXene nanocomposite as heterogeneous activator of peroxymonosulfate
precipitation of Co in amorphous TiO2. Appl. Catal. B Environ. 182, 449–455.
for the degradation of salicylic acid. J. Hazard. Mater. 382, 121064.
Shahzad, A., Rasool, K., Nawaz, M., Miran, W., Jang, J., Moztahida, M., Mahmoud, K.A.,
Ding, Y., Zhu, L., Wang, N., Tang, H., 2013. Sulfate radicals induced degradation of
Lee, D.S., 2018. Heterostructural TiO2/Ti3C2Tx (MXene) for photocatalytic
tetrabromobisphenol A with nanoscaled magnetic CuFe2O4 as a heterogeneous
degradation of antiepileptic drug carbamazepine. Chem. Eng. J. 349, 748–755.
catalyst of peroxymonosulfate. Appl. Catal. B Environ. 129, 153–162.
Shao, H., Zhao, X., Wang, Y., Mao, R., Wang, Y., Qiao, M., Zhao, S., Zhu, Y., 2017.
Gao, Y., Wang, L., Zhou, A., Li, Z., Chen, J., Bala, H., Hu, Q., Cao, X., 2015. Hydrothermal
Synergetic activation of peroxymonosulfate by Co3O4 modified g-C3N4 for enhanced
synthesis of TiO2/Ti3C2 nanocomposites with enhanced photocatalytic activity.
degradation of diclofenac sodium under visible light irradiation. Appl. Catal. B
Mater. Lett. 150, 62–64.
Environ. 218, 810–818.
Gao, H., Yang, H., Xu, J., Zhang, S., Li, J., 2018. Strongly coupled g-C3N4 nanosheets-
Shi, W., Guo, F., Zhu, C., Wang, H., Li, H., Huang, H., Liu, Y., Kang, Z., 2017. Carbon dots
Co3O4 quantum dots as 2D/0D heterostructure composite for peroxymonosulfate
anchored on octahedral CoO as a stable visible-light-responsive composite
activation. Small 14, 1801353.
photocatalyst for overall water splitting. J. Mater. Chem. A 5, 19800–19807.
Golshan, M., Kakavandi, B., Ahmadi, M., Azizi, M., 2018. Photocatalytic activation of
Su, C., Duan, X., Miao, J., Zhong, Y., Zhou, W., Wang, S., Shao, Z., 2016. Mixed
peroxymonosulfate by TiO2 anchored on cupper ferrite (TiO2@CuFe2O4) into 2,4-D
conducting perovskite materials as superior catalysts for fast aqueous-phase
degradation: process feasibility, mechanism and pathway. J. Hazard. Mater. 359,
advanced oxidation: a mechanistic study. ACS Catal. 7, 388–397.
325–337.
Tang, M., Ao, Y., Wang, C., Wang, P., 2020. Rationally constructing of a novel dual Z-
Gong, Y., Zhao, X., Zhang, H., Yang, B., Xiao, K., Guo, T., Zhang, J., Shao, H., Wang, Y.,
scheme composite photocatalyst with significantly enhanced performance for
Yu, G., 2018. MOF-derived nitrogen doped carbon modified g-C3N4 heterostructure
neonicotinoid degradation under visible light irradiation. Appl. Catal. B Environ.
composite with enhanced photocatalytic activity for bisphenol A degradation with
270, 118918.
peroxymonosulfate under visible light irradiation. Appl. Catal. B Environ. 233,
Wang, Y., Cao, D., Zhao, X., 2017. Heterogeneous degradation of refractory pollutants by
35–45.
peroxymonosulfate activated by CoOx-doped ordered mesoporous carbon. Chem.
Guo, J., Peng, Q., Fu, H., Zou, G., Zhang, Q., 2015. Heavy-metal adsorption behavior of
Eng. J. 328, 1112–1121.
two-dimensional alkalization-intercalated MXene by first-principles calculations.
Wang, Y.-F., Hsieh, M.-C., Lee, J.-F., Yang, C.-M., 2013. Nonaqueous synthesis of CoOx/
J. Phys. Chem. C 119, 20923–20930.
TiO2 nanocomposites showing high photocatalytic activity of hydrogen generation.
Han, M., Yin, X., Wu, H., Hou, Z., Song, C., Li, X., Zhang, L., Cheng, L., 2016. Ti3C2
Appl. Catal. B Environ. 142–143, 626–632.
MXenes with modified surface for high-performance electromagnetic absorption and
Wang, F., Lai, Y., Fang, Q., Li, Z., Ou, P., Wu, P., Duan, Y., Chen, Z., Li, S., Zhang, Y.,
shielding in the X-band. ACS Appl. Mater. Interfaces 8, 21011–21019.
2020. Facile fabricate of novel Co(OH)F@MXenes catalysts and their catalytic
Huang, H., Song, Y., Li, N., Chen, D., Xu, Q., Li, H., He, J., Lu, J., 2019. One-step in-situ
activity on bisphenol A by peroxymonosulfate activation: the reaction kinetics and
preparation of N-doped TiO2@C derived from Ti3C2 MXene for enhanced visible-
mechanism. Appl. Catal. B Environ. 262, 118099.
light driven photodegradation. Appl. Catal. B Environ. 251, 154–161.
Wang, Q., Shao, Y., Gao, N., Chu, W., Shen, X., Lu, X., Chen, J., Zhu, Y., 2016.
Hu, M., Li, Z., Hu, T., Zhu, S., Zhang, C., Wang, X., 2016. High-capacitance mechanism
Degradation kinetics and mechanism of 2,4-Di-tert-butylphenol with UV/persulfate.
for Ti3C2Tx MXene by in situ electrochemical Raman spectroscopy investigation. ACS
Chem. Eng. J. 304, 201–208.
Nano 10, 11344–11350.
Wang, Y., Wang, H., Li, J., Zhao, X., 2020. Facile synthesis of metal free perylene imide-
Jorfi, S., Kakavandi, B., Motlagh, H.R., Ahmadi, M., Jaafarzadeh, N., 2017. A novel
carbon nitride membranes for efficient photocatalytic degradation of organic
combination of oxidative degradation for benzotriazole removal using TiO2 loaded

11
M. Ding et al. Journal of Hazardous Materials 420 (2021) 126686

pollutants in the presence of peroxymonosulfate. Appl. Catal. B Environ. 278, Zeng, H., Deng, L., Yang, L., Wu, H., Zhang, H., Zhou, C., Liu, B., Shi, Z., 2021. Novel
118981. Prussian blue analogues@MXene nanocomposite as heterogeneous activator of
Wang, H., Wu, Y., Xiao, T., Yuan, X., Zeng, G., Tu, W., Wu, S., Lee, H.Y., Tan, Y.Z., peroxymonosulfate for the degradation of coumarin: the nonnegligible role of Lewis-
Chew, J.W., 2018. Formation of quasi-core-shell In2S3/anatase TiO2@metallic acid sites on MXene. Chem. Eng. J. 416, 128071.
Ti3C2Tx hybrids with favorable charge transfer channels for excellent visible-light- Zeng, H., Zhang, H., Deng, L., Shi, Z., 2020. Peroxymonosulfate-assisted photocatalytic
photocatalytic performance. Appl. Catal. B Environ. 233, 213–225. degradation of sulfadiazine using self-assembled multi-layered CoAl-LDH/g-C3N4
Xie, M., Tang, J., Fang, G., Zhang, M., Kong, L., Zhu, F., Ma, L., Zhou, D., Zhan, J., 2020. heterostructures: performance, mechanism and eco-toxicity evaluation. J. Water
Biomass Schiff base polymer-derived N-doped porous carbon embedded with CoO Process Eng. 33, 101084.
nanodots for adsorption and catalytic degradation of chlorophenol by Zhang, T., Chen, Y., Wang, Y., Le Roux, J., Yang, Y., Croue, J.P., 2014. Efficient
peroxymonosulfate. J. Hazard. Mater. 384, 121345. peroxydisulfate activation process not relying on sulfate radical generation for water
Xu, H., Wang, D., Ma, J., Zhang, T., Lu, X., Chen, Z., 2018. A superior active and stable pollutant degradation. Environ. Sci. Technol. 48, 5868–5875.
spinel sulfide for catalytic peroxymonosulfate oxidation of bisphenol S. Appl. Catal. Zhang, C.J., Kim, S.J., Ghidiu, M., Zhao, M.-Q., Barsoum, M.W., Nicolosi, V., Gogotsi, Y.,
B Environ. 238, 557–567. 2016. Layered orthorhombic Nb2O5@Nb4C3Txand TiO2@Ti3C2Tx hierarchical
Yan, X., Liu, Y., Lan, J., Yu, Y., Murowchick, J., Yang, X., Peng, Z., 2018. composites for high performance Li-ion batteries. Adv. Funct. Mater. 26, 4143–4151.
Crystalline–amorphous Co@CoO core–shell heterostructures for efficient electro- Zhang, G., Wu, Z., Liu, H., Ji, Q., Qu, J., Li, J., 2017. Photoactuation healing of
oxidation of hydrazine. Mater. Chem. Front. 2, 96–101. α-FeOOH@g-C3 N4 catalyst for efficient and stable activation of persulfate. Small
Yan, M., Zhu, F., Gu, W., Sun, L., Shi, W., Hua, Y., 2016. Construction of nitrogen-doped 13, 1702225.
graphene quantum dots-BiVO4/g-C3N4Z-scheme photocatalyst and enhanced Zhang, B., Zhang, Y., Xiang, W., Teng, Y., Wang, Y., 2017. Comparison of the catalytic
photocatalytic degradation of antibiotics under visible light. RSC Adv. 6, performances of different commercial cobalt oxides for peroxymonosulfate
61162–61174. activation during dye degradation. Chem. Res. Chin. Univ. 33, 822–827.

12

You might also like