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

Environmental Technology & Innovation 35 (2024) 103710

Contents lists available at ScienceDirect

Environmental Technology & Innovation


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

Facile synthesis of zeolite catalyst from lithium slag efficiently


activates peroxymonosulfate for tetracycline degradation: •O-2 and
the electron transfer
Yi Chen a, * , Haolan Zhang a, * , Siyu Hu a , Jiaxin Zhang b , Shiruo Zhang a , Qing He a ,
Qingyue Luo a , Jinchuan Gu a, *
a
School of Food and Bioengineering, Civil Engineering and Architecture and Environment, Xihua University, Chengdu 610039, PR China
b
College of Environmental Sciences, Sichuan Agricultural University, Chengdu 611130, PR China

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

Keywords: Iron is the most suitable candidate for activating peroxymonosulfate (PMS) for organic degra­
Peroxymonosulfate dation. To realize the effective separation and recycling of iron catalysts, it is urgent to explore a
Iron catalyst low-cost and excellent supporter. In this study, a zeolite catalyst (CAN) with both supporter and
Zeolite
catalyst functions was synthesized in one step using lithium slag (LS, containing iron), and used to
Lithium slag
degrade tetracycline (TC) wastewater. The results showed that the electron transfer of iron and
Tetracycline
sulfur in CAN promotes the decomposition of PMS. The CAN/PMS system exhibited great removal
efficiency (89.5 %, 0.083 min− 1) and relatively lower activation energy (15.31 kJ/mol) toward
TC. And •O-2 plays a crucial role in the degradation. The CAN also represented great catalytic
stability and reusability. This study revealed a new approach for the synthesis of high efficiency
and low-cost catalysts and the tetracycline degradation.

1. Introduction

Emerging pollutants (EPs) in water pose a significant threat to Earth’s ecosystems and public health, having drawn attention by
environmental scientists, policymakers, and the public. Antibiotics are a category of EPs and frequently detected in water resources
due to the abuse and improper discharge. The presence of antibiotics in water is a potential risk to humans, as it can induce antibiotic
resistance of pathogenic bacteria, leading to the emergence of superbugs, which poses a major challenge to the medicinal usefulness of
antibiotics(Silva et al., 2023). In the Drinking Water Contaminant Candidate List of Environmental Protection Agency (EPA), anti­
biotics are including (Vargas-Berrones et al., 2020). Recently, antibiotics has also been included in the list of EPs under Key Control in
China. Tetracycline (TC) is a typical refractory antibiotic and has attracted much attention due to its high detection in water.
Therefore, it is imperative to finding an effective method for the TC remove from wastewaters.
The conventional treatment methods of antibiotic wastewater encompass biological, physical, and chemical approaches(Gahrouei
et al., 2024; Hacıosmanoğlu et al., 2022; Huidobro et al., 2024; Wang et al., 2024). However, the presence of biological toxicity in
wastewater hampers the efficacy of antibiotic removal through biological means(Wang et al., 2024). The physical method, although
effective in rapidly eliminating antibiotics, does not fundamentally achieve their complete degradation(Gahrouei et al., 2024).
Recently, advanced oxidation processes (AOPs) systems based on peroxymonosulfate (PMS, HSO-5) have been widely used in the

* Corresponding authors.
E-mail addresses: evechen1@163.com (Y. Chen), 18081972165@163.com (H. Zhang), gu6471@163.com (J. Gu).

https://doi.org/10.1016/j.eti.2024.103710
Received 31 March 2024; Received in revised form 19 May 2024; Accepted 9 June 2024
Available online 11 June 2024
2352-1864/© 2024 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC license
(http://creativecommons.org/licenses/by-nc/4.0/).
Y. Chen et al. Environmental Technology & Innovation 35 (2024) 103710

investigated of antibiotic removal in water due to the high redox potential and reaction efficiency over a broad pH range(Huang et al.,
2021; Xiong et al., 2021; Yang et al., 2019). Previous studies have shown that the degradation efficiency of individual PMS is low, but it
can be significantly improved through ultraviolet irradiation(Li, H. et al., 2022; Lou et al., 2024), thermal activation(Wang et al., 2023)
and transition metal activation(Lv et al., 2022; Xu et al., 2023; Yu et al., 2024). Among them, transition metal activation has been
extensive concerned owing to the advantages of simplicity of use, mild reaction conditions and low energy consumption. Among the
many transition metals, iron is considered to be the most suitable candidate for activating PMS due to its wide source, low cost, high
activation efficiency and environmentally friendly(Li et al., 2021). To realize the effective separation and recycling of iron catalysts,
the preparation of heterogeneous solid catalysts which supported iron on suitable supporter has become the focus of research.
Zeolite is a typical inorganic crystal material with developed pore structure and thermal stability. Moreover, it has catalytic activity
owning to their acidic sites, and is the preferred material as the catalyst supporter of iron series active components. At present, it has
been widely used in the fields of separation, adsorption and catalysis. Guo et al.(2017) synthesized composite by nano zero valent iron
(NZVI) modified MCM-41-zeolite A through precipitation method, which showed outstanding adsorption capacity (526.32 mg/g) and
reuse performance for TC. Chi et al.(Chi et al., 2020) prepared ZSM-5-(C@Fe) by grafting Fe and doping C on ZSM-5, which was able to
completely remove norfloxacin within 15 min through five cycles, and the total iron leaching concentration was very low. Lv et al.
(2022) synthesized CuO-Fe-MFI zeolite via an impregnation procedure, which the largest TC degradation rate constant is 3.4 times that
of FeS-1 zeolite and 1.8 times that of CuO, respectively. Zafar et al.(2022) loaded Fe-Zn onto zeolite 5 A, which combined with
Electroflocculation and O3 has highest elimination efficiency for pharmaceutical wastewater. The above studies show that the catalyst
prepared by zeolite as a supporter has excellent catalytic property and stability for antibiotic wastewater. In order to achieve the high
efficiency and reuse stability of the catalyst, synthetic zeolite is generally used. However, the high cost of synthesis has hindered its
extensive application. Therefore, exploring a low-cost synthesis process is greatly expected.
Currently, lithium slag (LS) output sharply increased with the demand of lithium batteries, etc. The effective reuse way becomes
very important to solve the problem of LS accumulation. It is reported that LS not only has a high content of silicon and aluminum, but
also contains active components such as iron that can activate PMS (Chen et al., 2020; Pu et al., 2020). Thus, we proposed to synthesis
zeolite using LS as raw material, and applied to active PMS for degraded TC in the wastewater. On the one hand, it not only reduced the
cost, but also simplifies the synthesis process. On the other hand, the study not only provides a low-cost and excellent catalyst for TC
removal, also provides a new avenue for LS recycling.

2. Experimental

2.1. Materials

The raw material LS (mainly consists of: 64.84 % O, 19.51 % Si, 10.15 % Al, 0.11 % Na, 2.54 % S, 2.40 % Ca, 0.16 % Fe, etc.),
obtained from Sichuan Green Ruisang Renewable Resources Utilization Co. Ltd (China). Sodium hydroxide (NaOH), sodium aluminate
(NaAlO2) and sodium silicate (Na2SiO3⋅9 H2O) were purchased from Chengdu Kelong Chemicals Co., Ltd. TC was obtained from
Shanghai Aladdin Bio-Chem Technology Co., Ltd. Potassium peroxymonosulfate (PMS, KHSO5‧0.5KHSO4‧0.5 K2SO4, 47 %) was pur­
chased from Shanghai Yuanye Bio-Technology Co., Ltd. All the chemicals were used as received without further purification.

2.2. Synthesis of cancrinite-type zeolite

LS was sieved and dried at 105 ◦ C. 3 g LS was mixed and ground with 5 g NaOH, pour into the nickel crucible and calcined at 600 ◦ C
under air flow for 2 h. Then, the sample was ground to powder after cooling down to room temperature. The powder was mixed with
40 mL of deionized water for 2 h, resulting in an initial gel mixture. After that, the gel mixture was placed in the autoclave and then
hydrothermally treated under 100 ◦ C for 8 h. Finally, the product was washed to neutral with deionized water and dried at 105 ◦ C to
obtain cancrinite-type zeolite (CAN).

2.3. Characterization

The surface morphology of CAN was observed by scanning electron microscope (SEM, Apreo 2 C, American Thermo Scientific,
USA) and transmission electron microscope (TEM, Talos F200S G2, American Thermo Scientific, USA). The specific surface area and
porosity of CAN was determined using N2 adsorption isotherms (JW-BK122W, JWGB, CN). CAN X-ray diffraction (XRD) pattern was
obtained using Rigaku Ultima IV equipped with Cu-Kα radiation at 40 kV/40 mA with scan step of 0.02 ◦ /s. The Fourier translation
infrared spectrum (FT-IR) was performed on a Perkin Elmer Spectrum Two Li10014 spectrometer. The surface elemental analysis was
obtained using AXIS Ultra DLD with Al Kα radiation (1486.6 eV) at beam power of 150 W. The reactive oxygen species (ROS) electron
spin resonance (ESR) spectra were investigated by an EMXplus.

2.4. TC degradation experiments

Batch experiments were conducted in a 250 mL Erlenmeyer flask containing 150 mL TC solution with a concentration of 20 mg/L.
0.15 g CAN (1 g/L) was added to the TC solution and carried out at 25 ◦ C in a thermostatic shaker (160 r/min) until reach the
adsorption-desorption equilibrium (35 min). Then, the oxidation reaction was initiated after adding 0.015 g PMS (0.15 mM). After a
certain interval, 4 mL of the solution was collected with a 5 mL syringe and filtered through 0.22 μm PTFE membrane, and then

2
Y. Chen et al. Environmental Technology & Innovation 35 (2024) 103710

immediately measured by an ultraviolet spectrophotometer (λ=355 nm, UV-2400).

2.5. Analytical method

The previous literature shows that the degradation of TC followed pseudo-first-order kinetics, The kinetics equation is as follows
(Eq. (1))(Nguyen et al., 2023; Xie, T. et al., 2023):
Ct
ln = − kobs • t (1)
C0

Where Ct and C0 are the concentrations of TC at t reaction time (min) and in the solution after the physical adsorption for 35 min (mg/
L), kobs is the kinetic rate constant (min− 1).
The degradation activation energy (Ea) of TC in the system was calculated by Arrhenius equation (Eq. (2))(Yu et al., 2023):
Ea
lnkobs = lnA − (2)
RT

Where A is the prefactor, and R is the universal gas constant (8.314 J mol− 1 K− 1).
The intermediates of TC during the degradation process were detected on a liquid chromatography-mass spectrometry (Agilent
6550 Q-TOF), equipped with the waters BEH C18 (2.1*100 mm,1.7 µm) column. The mobile phase consisted of acetonitrile and 0.1 %
methanoic acid aqueous solution (85:15, v/v) at a flow rate of 0.3 mL/min with a gradient elution.

3. Results and discussion

3.1. Effect of the synthesis parameters

3.1.1. Calcination activation


As seen from Fig. 1, the diffraction peaks of LS at 19.624◦ , 24.164◦ , 25.764◦ , 26.914◦ , 28.354◦ , 41.683◦ and 48.348◦ are

Fig. 1. XRD of calcination activation samples prepared at different m(NaOH):m(LS) (a), calcination temperature (b) and calcination time (c).

3
Y. Chen et al.
4

Environmental Technology & Innovation 35 (2024) 103710


Fig. 2. TC removal capacity of zeolite synthesized with different hydrothermal time (a), hydrothermal temperature (b) and S/A ratio (c) and corresponding XRD patterns (d-f). (Conditions: T = 25 ◦ C,
pH = 6.6, C0 = 20 mg/L, MCAN = 1 g/L, MPMS = 0.15 mM).
Y. Chen et al.
5

Environmental Technology & Innovation 35 (2024) 103710


Fig. 3. The removal capacity of TC (a) and PMS residues (b) in various systems. The impact of the initial TC solution concentration (c), CAN dosage (d), initial pH (e) and reaction temperature (f).
(Conditions: T = 25 ◦ C, pH = 6.6, C0 = 20 mg/L, MCAN = 1 g/L, MPMS = 0.15 mM).
Y. Chen et al. Environmental Technology & Innovation 35 (2024) 103710

characteristic peaks of spodumene (JCPDS No. 39–0049). The peaks at 20.859◦ and 26.640◦ are diffraction peaks of quartz (JCPDS No.
33–1161). The results showed that the LS mainly consists of spodumene (LiAlSi2O6) and quartz (SiO2). To take full advantage of the
effective components in LS, it is activated by calcinating. With the increased of m(NaOH):m(LS) (Fig. 1(a)), calcination temperature
(Fig. 1(b)) and time (Fig. 1(c)), the characteristic peaks of spodumene and quartz became weak and even disappeared, which indicated
that the two crystal substances in LS transformed into amorphous. When the m(NaOH):m(LS) is 5:3 (Fig. 1(a)), the calcination
temperature is 600 ◦ C (Fig. 1(b)), and the calcination time is 2 h (Fig. 1(c)), the characteristic peaks of spodumene and quartz are no
longer obvious and basically transform completely. Therefore, the LS was calcinated in the above conditions during the subsequent
experiments.

3.1.2. Hydrothermal synthesis


Previous studies have found that the main factors affecting zeolite crystallization are not only alkalinity, but also hydrothermal
time, hydrothermal temperature and S/A(SiO2/Al2O3) ratio(Chen et al., 2020; Đặng et al., 2021).
As shown in Fig. 2(a), the removal rate of TC by zeolite first decreased and then increased with the increase of hydrothermal time. It
has been reported that hydrothermal time is one of the key factors affecting the crystallinity of zeolite (Selim et al., 2018; Zhang et al.,
2019; Zheng et al., 2021). The XRD patterns of zeolite prepared at different hydrothermal times are shown in Fig. 2(d), all samples
show significant diffraction peaks at 2θ of 14.024◦ , 19.071◦ , 21.249◦ , 24.413◦ , 27.686◦ , 32.749◦ , 34.896◦ , 37.241◦ and 42.965◦ , which
is a typical peak of CAN (JCPDS No.73–0540, Na7.6Ca4(Si6Al6O24)(CO3)(H2O)2.2). The results show that the synthesized zeolites are
pure CANs. By comparing the XRD peaks of different samples, it is found that the diffraction peaks of CAN intensified gradually with
the hydrothermal time increased, indicating that the crystallinity of zeolite is enhanced.
The effect of hydrothermal temperature on the zeolite was investigated by fixing the hydrothermal time at 8 h. As shown in Fig. 2
(b), with the increase of hydrothermal temperature, the removal rate of TC by zeolite first increased and then gradually decreased.
Among them, the difference in the removal rate of TC is slight between the samples synthesis at 100 ◦ C and 80 ◦ C. The characteristic
diffraction peak of CAN strengthened with the hydrothermal temperature, indicating that the crystallinity of the CAN increased and
the crystallite dimension enlarged gradually. However,the removal rate of TC by CAN decreased, which may be attributed to the
decrease of catalytic activity due to the larger crystallite dimension covering the effective active components (iron, etc.).
The depolymerization and rearrangement of Si and Al species are continuously carried out in the synthesis process of zeolite, and
the adjustment of S/A ratio will cause differences in the crystallite dimension and morphology, which is an important way to determine
the final crystal structure of zeolite(Chen et al., 2020; Gusti Wibowo et al., 2023; Si et al., 2021). It can be seen from Fig. 2(c) that the
adsorption and catalytic removal rates of TC by CAN first increased and then decreased with the increase of S/A ratio. When the S/A
ratio is 4:1 (no additional Si or Al source is added), the CAN has the best adsorption capacity (32.7 %) and catalytic capacity (89.5 %).
Fig. 2(f) shows that with the increase of S/A ratio, the characteristic diffraction peaks of CAN increased and then enhanced gradually.
When the S/A ratio is lower than 4:1, the growth of CAN is incomplete and some crystal faces do not grow due to the lack of silicon
source. When the S/A ratio is 4:1, the characteristic diffraction peak of CAN is relatively complete, and it also has great
adsorption-catalytic capacity for TC, which may be attributed to the fact that the fully grown crystal structure is conducive to
increasing the active site on the CAN surface. When the S/A ratio is greater than 4:1, the characteristic peaks of CAN were intensified,
the crystallite dimension enlarged gradually, and the removal rate of TC decreased. Consequently, the CAN hydrothermally synthe­
sized under 100 ◦ C for 8 h was applied during the subsequent experiments.

3.2. Degradation of TC

3.2.1. Effect of various catalyst


In the presence of PMS, the CAN showed high removal efficiency of 89.5 % and kobs of 0.083 min− 1 toward TC (Fig. 3(a) and
(Fig. S1 (a)). Meanwhile, LS only resulted in 33.4 % and 0.015 min− 1 of TC removal. Residual PMS was determined by iodometry (Text
S1). The curve of PMS residue in the above various systems illustrates that the concentration of PMS decreases significantly after the
reaction in CAN/PMS system (Fig. 3(b)). It suggests that PMS is effectively activated in CAN/PMS system compared with LS/PMS. This
is attributed to the fact that the specific surface area increases after LS is prepared into CAN, which significantly improving the
adsorption capacity of TC, and the contact area between TC and the active site.

3.2.2. Effect of initial TC solution concentration


Fig. 3(c) depicts the degradation of TC at various initial concentrations, and it was evident that the greater the concentration, the
lower the adsorption and degradation efficiency. The kobs value decreased from 0.098 min− 1 to 0.026 min− 1 (Fig. S1 (b)), with the
initial TC concentration increased range from 10 mg/L to 100 mg/L.

3.2.3. Effect of CAN dosage


Fig. 3(d) and Fig. S1 (c) showed the TC degradation efficiency improved to 89.5 % when the CAN dose was raised from 0.1 g/L to
1.5 g/L, and the maximum kobs was obtained as 0.136 min− 1. It is mainly attributed to that the more ROS for TC degradation are
generated with the increase of the CAN dosage through the interaction with PMS(Nguyen et al., 2023; Qian et al., 2023; Yu et al.,
2023). However, when the CAN dose was raised to 2.0 g/L, the degradation efficiency and the kobs value slightly dropped to 88.6 % and
0.129 min− 1, respectively. Most likely, this is due to the excessive CAN dosage is hard to dispersion, and the self-quenching reaction
occurs between excess ROS(Nguyen et al., 2023; Song et al., 2023; Zhang et al., 2022; Zhou et al., 2022).
The TC removal efficiency had little change when the PMS dosage was range from 0.15 to 0.76 mM (Fig. S2). The previous reports

6
Y. Chen et al.
7

Environmental Technology & Innovation 35 (2024) 103710


Fig. 4. The influence of inorganic anions on the degradation of TC in the CAN/PMS system. Cl- (a), NO-3 (b), SO2- - -
4 (c), H2PO4 (d), HCO3 (e) and TC removal in actual water (f). (Conditions: T = 25 C, pH

= 6.6, C0 = 20 mg/L, MCAN = 1 g/L, MPMS = 0.15 mM).


Y. Chen et al. Environmental Technology & Innovation 35 (2024) 103710

stated that may be due to the self-destruction of sulfate ions (Eq. (3)) and scavenging of ROS (Eqs. (4)-(5)), reducing the degradation
efficiency(Nguyen et al., 2023; Xu et al., 2023; Zhou et al., 2022).
2-
SO•-
4 + SO4 → S2O8
•-
(3)

HSO-5 + SO•-
4 → SO•-
5 + SO2-
4 +H +
(4)

HSO-5 + •OH → SO•-


5 + H2O (5)

3.2.4. Effect of initial pH


The initial pH of the solution can significantly affect the surface charge of CAN, the type of adsorbate and the generate of ROS in the
system, thus promoting or inhibiting the adsorption process and catalytic reaction process. According to the findings in Fig. 3(e) and
Fig. S1 (e), the degradation efficiency of TC and the kobs value at pH 1 were low (5.2 % and 0.001 min− 1) due to the abundance of H+
might quench •OH and SO•- 4 (Eqs. (6)-(7))(Gu et al., 2022; Nguyen et al., 2023). The adsorption capacity of CAN on TC first increased
and then decreased with the pH value increased. It is reported that at pH < 3.3, TC exists in the form of TCH+ 3 , in the form of TCH2
±
- 2-
when pH is 3–8, and in the form of TCH and TC at pH > 8(Kanmaz et al., 2023; Xie, X. et al., 2023). Since the surface of CAN has a
negative charge potential (Fig. S3, pH > 1), the increased anionic form of TC with the increased pH value decreased the removal rate.
The TC decomposition efficiency rose from 5.2 % to 89.5 % when the pH value was raised from 1 to 6, and then it dropped to 81.2 %
when the pH raised to 11. Under alkaline conditions, the slight decrease of TC degradation efficiency may be originated from the
decomposing of PMS (Eq. (8)), which was detrimental to the generation of ROSs(Gu et al., 2022; Lv et al., 2022; Nguyen et al., 2023;
Zou et al., 2023). In general, CAN/PMS have shown great removal efficiency on TC in a broad pH range (3− 11), which exhibits
significant potential for TC degradation.
-
SO•-
4 + H + e → HSO4
+ •-
(6)
-

OH + H + e → H2O
+
(7)

2HSO-5 + 2OH → -
2SO2-
4 + O2 + H2O (8)

3.2.5. Effect of reaction temperature


The reaction temperature has great influence on adsorption and catalysis of TC by CAN/PMS. As observed in Fig. 3(f), when the
reaction temperature ranges from 15 to 45 ℃, the adsorption capacity of TC on the CAN surfaces significantly enhanced. However, as
the temperature increase to 55 ℃, the adsorption capacity sharply decreased. The TC degradation rate of CAN/PMS was promoted as
the temperature increased. When the reaction temperature increased from 15 to 45 ℃, the kobs values increased from 0.055 to
0.134 min− 1 accordingly. This suggests that the higher temperature is conducive to accelerating the activation of PMS to produce more
ROS, thus promoting the removal rate of TC. Based on the Arrhenius plots, the activation energy Ea of CAN was calculated to be
15.31 kJ/mol as seen in the inset (Fig. 3(f)). Obviously, CAN presents a relatively lower activation energy compared with the before
study (Table S1). These results indicate that the degradation of TC in CAN/PMS system can be easily achieved, which may be beneficial
to the practical application.

3.3. Practical application

Influence of different anions (as Cl-, NO-3, SO2- -


4 and HCO3, etc.) which are widely detected in natural water on TC removal were
- -
investigated. Cl and NO3 slightly suppressed TC degradation at a certain concentration (Fig. 4(a-b)), which was resulted from the
reaction of them with the radicals (SO•- 4 and OH) to produce lower reactive radicals (Cl , Cl2 , ClOH , NO3 and NO2) (Eqs. (9)-(15))
• • •- •- • •

(Chen, D. et al., 2022; Jiang et al., 2023; Song et al., 2023). The TC degradation rates (Fig. 4(c)) and the kobs values (Fig. S4 (c)) were
obviously decreased with the presence of SO2- 2-
4 , which can due to the SO4 scavenge of the OH and h (Eqs. (16)-(17)) (Lv et al., 2022;
• +

Song et al., 2023). Adding a certain amount of H2PO-4 had little influence on TC removal (Fig. 4(d)). With the varied HCO-3 concen­
trations, the adsorption and catalysis removal rate decreased significantly (Fig. 4(e)). The existence of HCO-3 can increase the pH value
of the solution, which not only inhibits the adsorption of TC by CAN but also can scavenge the radicals (SO•- 4 and OH) (Eqs. (18)-(19))

(Mao et al., 2022; Song et al., 2023; Zhang et al., 2023). Nevertheless, Fig. 4(f) show that the natural water background has a positive
effect on the TC degradation by CAN/PMS. The results indicated that CAN showed brilliant adaptability to purification of wastewater
in different natural water background.

NO-3 + SO•- 2-
4 → SO4 + NO3

(9)

NO-3 + •OH → OH- + NO•3 (10)

NO-3 + OH + e → 2OH +
• - -
NO•2 (11)
-
Cl + SO•-
4 → Cl + •
SO2-
4 (12)

8
Y. Chen et al.
9

Fig. 5. The degradation of different pollutants in the CAN/PMS system (a), recycling study of TC degradation by CAN (b), Fe leaching concentration in each recycling process (c). (Conditions: T = 25 ◦ C,
pH = 6.6, C0 = 20 mg/L, MCAN = 1 g/L, MPMS = 0.15 mM).

Environmental Technology & Innovation 35 (2024) 103710


Y. Chen et al.
10

Environmental Technology & Innovation 35 (2024) 103710


Fig. 6. SEM (a) and TEM (b1)- (b3) images of CAN, N2 adsorption-desorption isotherms (c), XRD patterns (d) and FTIR spectra (e) of fresh and used CAN.
Y. Chen et al. Environmental Technology & Innovation 35 (2024) 103710

Cl- + Cl• → Cl•-


2 (13)

Cl- + •OH → ClOH•- (14)

ClOH•- + H+ → Cl• + H2O (15)

SO2- • •- -
4 + OH → SO4 + OH pH < 3.0 (16)

SO2-
4 +h → +
SO•-
4 (17)

SO•-
4 + HCO-3 → SO2-
4 + HCO•3 (18)

OH + HCO-3 → H2O + CO•-
3 (19)

In the same reaction system, the degradation efficiency of CAN on other organics (Fig. 5(a)) was also investigated. The results
showed that the CAN had better removal efficiency for TC than other organics, and CAN showed better selectivity for TC.

3.4. Catalyst stability

The reusability of TC degradation by CAN was shown in Fig. 5(b), the removal rate of TC by CAN gradually decreased with the
increase of regeneration times, which may be related to the leaching of active components and the decrease of specific surface area of
CAN. The iron in the CAN is one of the main active components in the catalytic process. The concentration of iron ion in the solution of
the recycling experiment was determined, and the results are shown in Fig. 5(c). There is a minute quantity of iron leaching in each
recycling process, but its concentration is far lower than the limit concentration of 0.3 mg/L in "Standards for drinking water quality"
(GB 5749–2022).
The adsorption capacity of N2 by CAN decreased slightly after regeneration (Fig. S5 (b)). The XRD pattern of the regenerated CAN
(Fig. S5 (a)) shown that the diffraction peaks of cancrinite became weaker slightly with the increase of regeneration times. However,
the characteristic peaks are more significant, and no impure peaks appear, which indicates that the structure of CAN is stable in the
process of reuse. Overall, the catalyst prepared in this study has great catalytic stability and reusability.

3.5. Degradation mechanisms of TC in the CAN/PMS system

3.5.1. Textural properties


The SEM and TEM images in Fig. 6 shows that CAN is well-developed with prismatic crystals. The morphology of the used CAN after
the TC degradation experiment (Fig. S6 (a)) was not obviously changed, indicating that the structure of CAN is stable. The elemental
distribution of CAN (Fig. S6 (b)) showed that CAN contains Al, Si, Mn, Fe, Ca, Na, O and S, and these elements are evenly distributed.
The N2 adsorption-desorption isotherms of CAN (Fig. 6(c)) showed IV isotherm with H3-type hysteresis loop and textural properties
(Table 1) indicating that the CAN has abundant mesopores. The BET surface area (SBET) and mesopore volume (Vmes) values of used
CAN were decreased, indicating the occupation or destruction of pores following the adsorption and degradation of TC. The change of
SBET and Vmes after multiple cycles also confirmed this (Table S2).

3.5.2. Chemical properties


The XRD characteristic peaks of CAN were significant both after the reaction (Fig. 6(d)) and after four times of reuse (Fig. S5 (a)),
and no impure peaks appeared, which indicated that the structure of the CAN was stable during the reuse process.
The FTIR spectra of CAN before and after reaction was shown in Fig. 6(e), the band around 1633 cm− 1 and 3449 cm− 1 assigned to
the bending vibration of absorbed H2O and O-H bond stretching of internal silanol groups (Chen et al., 2022; Wu et al., 2023; Xie, et al.,
2023; Zhu et al., 2023). The bands in the ranges of 1380 cm− 1 to 1530 cm− 1 corresponds to carbonate anions occluded inside the
cancrinite structure(Chukanov et al., 2023; Chukanov et al., 2022; Zhang et al., 2019). However, the band at 1479 cm− 1 disappeared
after the TC degradation reaction, which may due to the adsorption of TC by CAN. The peaks appearing at 1200 cm− 1-800 cm− 1
correspond to the asymmetric stretching of tetrahedral TO4, where T is Si or Al (Đặng et al., 2021; Hua et al., 2023; Limlamthong et al.,
2021; Yu et al., 2023). The bands at 686 cm− 1, 622 cm− 1 and 572 cm− 1 could be assigned to symmetric stretching of 4- or 6-membered
rings and the vibration of double rings on CAN(Đặng et al., 2021; Hua et al., 2023). The absorption peaks at 460 cm− 1 and 435 cm− 1
indicated the bending vibrations of tetrahedral Al-O and Si-O(Đặng et al., 2021; Hua et al., 2023; Zhu et al., 2023). The bands observed

Table 1
Textural properties of the fresh and used CAN.
Samples SBET Vtot Vmic Vmes Dp
m2/g cm3/g cm3/g cm3/g nm

LS 6.75 0.039 0.000 0.039 15.96


Fresh CAN 82.66 0.216 0.003 0.213 10.30
Used CAN 62.46 0.178 0.000 0.178 10.31

SBET: BET surface area; Vtot: total pores volume; Vmic: micropore volume; Vmes: mesopore volume; DP: average pore diameter.

11
Y. Chen et al.
12

Environmental Technology & Innovation 35 (2024) 103710


Fig. 7. High-resolution XPS spectra and fitting peaks of Fe 2p (a), S 2p (a) and C 1 s (c), the degradation of TC in the CAN/PMS system under different quenching conditions (d-e), EPR spectra of PMS
and CAN/PMS catalytic processes using DMPO and TEMP as the trapping agents (f).
Y. Chen et al. Environmental Technology & Innovation 35 (2024) 103710
Fig. 8. Degradation pathways of TC in the CAN/PMS system.
13
Y. Chen et al. Environmental Technology & Innovation 35 (2024) 103710

at 876 cm− 1 disappeared, and the other bands range from 1105 cm− 1 to 435 cm− 1 have occurred a smaller shift after the TC degra­
dation reaction, indicating that the reaction resulted in a slight distortion of the CAN skeleton.
In the Fe 2p spectra (Fig. 7(a)), the peaks at 710.1–710.6 eV and 712.5–712.9 eV corresponded to Fe2+ and Fe3+(Cai et al., 2022;
Feng et al., 2023; Zhang et al., 2023), respectively. Fig. 7(a) revealed that the content of Fe2+ decreased and Fe3+ increased signifi­
cantly after the TC degradation reaction, indicating that the conversion between the two Fe species on the CAN surface (Eqs. (20)-(22))
(Cai et al., 2022; Feng et al., 2023; Gu et al., 2022; Zhang et al., 2023). The XPS wide-scan spectra (Fig. S5 (c)) and the XRF analysis
(Table S3) of the CAN shown the non-negligible content of S, which was attributed to the composition of raw material (LS). For the S 2p
spectrum (Fig. 7(b)), the characteristic peaks at 168.6–168.7 eV and 169.8–169.9 eV could be assigned to SO2- 2-
3 and SO4 (Feng et al.,
2023; Yin et al., 2023; Zeng et al., 2023), respectively. The percentage of SO2- 2-
3 and SO4 changed after the TC degradation reaction
(Fig. 7(b)) indicating the participation of S species in the reaction (Eqs. (23)-(25)) (Wang, S. et al., 2022; Zeng et al., 2023). In C 1 s
spectra (Fig. 7(c)), peaks centered at 284.8 eV, 286.0 eV and 288.1 eV can be attributed to C-C, C-O and O-C– –O (Zhang et al., 2023,
2024; Zuo et al., 2023), respectively. The increased content of C-C is related to the incomplete decomposition of organic matter on CAN
surface. In addition, the new peaks appeared at 293.8 eV and 296.2 eV after the reaction belongs to K 2p3/2 and K 2p1/2, respectively.
This indicates that K exists on the surface of CAN after reaction,and it also can be confirmed by the XPS wide-scan spectra (Fig. S5 (c),
Table S4). The appearance of K is mainly due to the decomposition of PMS. To sum up, the above results show that the electron transfer
of iron and sulfur occurs on the surface of CAN and promotes the decomposition of PMS, thus a lot of active substances are produced.

Fe2+ + 2HSO-5 → Fe3+ + •O-2 + 2HSO-4 (20)


2+
Fe + HSO-5 → Fe 3+
+ OH + -
SO•-
4 (21)
3+
Fe + HSO-5 → Fe 2+
+H + +
SO•-
5 (22)

SO2-
3 + Fe 3+
→ SO•-
3 + Fe 2+
(23)

SO•-
3 + O2 → SO5
•-
(24)

SO•-
5 + SO2-
3 → SO•-
4 + SO2-
4 (25)

3.5.3. ROS
1
Typically, EtOH was the scavenger of •OH and SO•- 4 (Xie et al., 2023), FFA was used to explore the presence of O2(Gu et al., 2022),
TBA was used as an effective quenching agent for OH only(Cheng et al., 2024), and p-BQ was applied for detecting the existence of

• -
O2(Silva et al., 2023). As displayed in Fig. 7(d), compared with no scavenger group, the removal efficiency of TC slightly decreased
(Control-89.5 %, 3 M EtOH-85.7 %, 3 M TBA-83.7 %). The result demonstrating that •OH and SO•- 4 were present in the system, and

OH serves as a more significant role than SO•-4 . However, they were not the main active species. The inhibition of TC degradation by
FFA was significant than above, indicated that 1O2 was one of the dominant reactive species (Eqs. (26− 29))(Feng et al., 2023). When
p-BQ was added into the system, the degradation efficiency sharply reduced to 56.5 % (1 mM), and it continuously decreased to
18.0 % with the p-BQ increased to 3 mM (Fig. 7(e)). Which implying that •O-2 plays a crucial role in the degradation of TC.
EPR spectra in Fig. 7(f) presents the typical characteristic peaks of TEMP-1O2, DMPO-•O-2, DMPO-•OH and DMPO-SO•- 4 in CAN/PMS
catalytic processes. This finding suggested the presence of 1O2, •O-2, •OH and SO•- 4.

HSO-5 + SO2- - 2- 1
5 → HSO4 + SO4 + O2 (26)
2+
Fe + HSO-5 + SO2-
5 → Fe 3+
+ HSO-4 + SO2-
4
1
+ O2 (27)

2SO•-
5 + H2O → 2HSO-4 + 1/2 O2 1
(28)

OH + •O-2 → 1O2 + OH- (29)

3.5.4. Intermediates and pathway of TC degradation


To explore the degradation pathway of TC, the intermediates of TC degradation were detected by LC-MS (Fig. S7). According to the
LC-MS spectra, four possible degradation pathways of TC over CAN are proposed (Fig. 8). Pathway I: P1 (m/z = 461) is produced
through the TC (m/z = 445) molecules occurs hydroxylation reaction by ROSs attacked(Wang, M. et al., 2022). Afterwards, P2 (m/z =
277) are generated via the deamination and the loop opening process of P1(Wang, L. et al., 2022). And then, P2 is further oxidized to
P8 (m/z = 213). In Pathway II, TC separated one hydroxyl to obtain P4 (m/z = 427). In pathway III, TC molecules were easily losing
N-methyl groups and transformed into P5 (m/z = 417) under ROSs attack, due to the lower bond energy of the connected N-C bond(Li
et al., 2022; Wang et al., 2022). Pathway IV: P6 (m/z = 433) is generated due to the detachment of N-methyl group and hydrogenation
reaction occur at the carbonyl group of TC(Li et al., 2022; Wang et al., 2022). Subsequently, P6 is transformed into P7 by ring break and
oxidation of ROSs. With the process of degradation, the above all intermediate products are transformed into low molecular weight
organic compounds by further oxidation ring opening and functional groups dissociation. Finally, these low molecular weight organic
compounds are further completely mineralized into CO2, H2O, and inorganic ions.

14
Y. Chen et al. Environmental Technology & Innovation 35 (2024) 103710

4. Conclusions

In this work, mesoporous CAN was synthesized from LS by calcination activation and hydrothermal. The abundant mesopores not
only provides a channel for the mass transfer of TC, but also increases the exposed area of the active site. Thus, the CAN displayed
excellent catalytic activity and lower activation energy (Ea, 15.31 kJ/mol) to TC by activating PMS. The CAN/PMS system also shows
significant potential for degrading TC in a broad pH range (3− 11) and different natural water background. The electron transfer of iron
and sulfur from raw materials on the surface of CAN promotes the decomposition of PMS, thus producing a lot of active substances. TC
is finally mineralized into CO2, H2O and inorganic ions through a series of ring opening and functional groups dissociation under the
attack of ROS. And •O-2 plays a crucial role in the degradation. Additionally, CAN also exhibited satisfying catalytic stability and
reusability. This study not only provides a green and efficient catalyst for the removal of TC, but also provides a new avenue for the
reuse of solid waste.

Authors statement

We confirm that the manuscript has been read and approved by all named authors and that there are no other persons who satisfied
the criteria for authorship but are not listed. We further confirm that the order of authors listed in the manuscript has been approved by
all of us.

CRediT authorship contribution statement

Qing He: Validation. Shiruo Zhang: Validation. Jiaxin Zhang: Validation. Siyu Hu: Validation, Investigation. Jinchuan Gu:
Supervision, Resources, Methodology. Qingyue Luo: Validation. Haolan Zhang: Validation, Investigation. Yi Chen: Writing –
original draft, Supervision, Methodology, Funding acquisition.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data Availability

Data will be made available on request.

Acknowledgements

The authors acknowledge the funding support from the Guiding Plan of Sichuan for the Transfer and Transformation of Scientific
and Technological Achievements (24ZHSF0062), and Key Laboratory of Special Waste Water Treatment, Sichuan Province Higher
Education System (No. SWWT2023-6). The authors would like to thank Ceshigo Research Service (www. ceshigo.com) for the XRD
analysis and Shiyanjia Lab (www.shiyanjia.com) for the ESR analysis.

Appendix A. Supporting information

Supplementary data associated with this article can be found in the online version at doi:10.1016/j.eti.2024.103710.

References

Cai, A., Ling, X., Wang, L., Sun, Q., Zhou, S., Chu, W., Li, X., Deng, J., 2022. Insight into UV-LED/PS/Fe(III) and UV-LED/PMS/Fe(III) for p-arsanilic acid degradation
and simultaneous arsenate immobilization. Water Res. 223, 118989.
Chen, D., Bai, Q., Ma, T., Jing, X., Tian, Y., Zhao, R., Zhu, G., 2022. Stable metal–organic framework fixing within zeolite beads for effectively static and continuous
flow degradation of tetracycline by peroxymonosulfate activation. Chem. Eng. J. 435, 134916.
Chen, Y., Armutlulu, A., Sun, W., Jiang, W., Jiang, X., Lai, B., Xie, R., 2020. Ultrafast removal of Cu(II) by a novel hierarchically structured faujasite-type zeolite
fabricated from lithium silica fume. Sci. Total Environ. 714, 136724.
Chen, Y., Dong, S., Wei, S., Zhang, S., Zhang, J., Gu, J., 2022. Green and efficient synthesis of mesoporous sodalite assisted by persulfate or peroxymonosulfate.
Microporous Mesoporous Mater. 346, 112321.
Cheng, H., Liu, H., Huang, C., Xu, J., Tian, H., Yang, J., Wang, P., Cai, J., Cheng, M., Liu, Z., 2024. Tungsten carbide induced acceleration of Fe3+/Fe2+ cycle in Fe2+/
PMS process for rapid degradation of tetracycline hydrochloride. Sep. Purif. Technol. 330, 125311.
Chi, H., Wan, J., Ma, Y., Wang, Y., Huang, M., Li, X., Pu, M., 2020. ZSM-5-(C@Fe) activated peroxymonosulfate for effectively degrading ciprofloxacin: In-depth
analysis of degradation mode and degradation path. J. Hazard. Mater. 398, 123024.
Chukanov, N.V., Aksenov, S.M., Pekov, I.V., 2023. Infrared spectroscopy as a tool for the analysis of framework topology and extra-framework components in
microporous cancrinite- and sodalite-related aluminosilicates. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 287, 121993.
Chukanov, N.V., Vigasina, M.F., Shendrik, R.Y., Varlamov, D.A., Pekov, I.V., Zubkova, N.V., 2022. Nature and isomorphism of extra-framework components in
cancrinite- and sodalite-related minerals: New data. Minerals 12, 729.

15
Y. Chen et al. Environmental Technology & Innovation 35 (2024) 103710

Đặng, T.-H., Nguyễn, X.-H., Chou, C.-L., Chen, B.-H., 2021. Preparation of cancrinite-type zeolite from diatomaceous earth as transesterification catalysts for biodiesel
production. Renew. Energy 174, 347–358.
Feng, M., Xu, Z., Lin, K., Xie, H., Zhang, M., 2023. Heterogeneous activation of peroxymonosulfate by sulfur-doped FexMn3-xO4 (x = 1, 2) for trichloroethylene
degradation: non-radical and radical mechanisms. Chem. Eng. J. 459, 141627.
Gahrouei, A.E., Vakili, S., Zandifar, A., Pourebrahimi, S., 2024. From wastewater to clean water: Recent advances on the removal of metronidazole, ciprofloxacin, and
sulfamethoxazole antibiotics from water through adsorption and advanced oxidation processes (AOPs). Environ. Res. 252, 119029.
Gu, J., Yin, P., Chen, Y., Zhu, H., Wang, R., 2022. A natural manganese ore as a heterogeneous catalyst to effectively activate peroxymonosulfate to oxidize organic
pollutants. Chin. Chem.Lett. 33, 4792–4797.
Guo, Y., Huang, W., Chen, B., Zhao, Y., Liu, D., Sun, Y., Gong, B., 2017. Removal of tetracycline from aqueous solution by MCM-41-zeolite A loaded nano zero valent
iron: synthesis, characteristic, adsorption performance and mechanism. J. Hazard. Mater. 339, 22–32.
Gusti, Wibowo, Sudibyo, Y., Rianjanu, A., Taher, T., Susanti, N., Karo, Karo, Suprihatin, P., Iman Supriyatna, Y., Prasetyo, E., Bahfie, F., Amin, M., Syarifuddin, H.,
Safitri, H., Khairurrijal, K., 2023. Synthesis, Characterization, and performance test of modified Zeolite-Si/Al nanocomposite from pumice and waste cans using
slow pyrolysis process for removal pollutant parameters from hard water. Environ. Nanotechnol., Monit. Manag. 20, 100853.
Hacıosmanoğlu, G.G., Mejías, C., Martín, J., Santos, J.L., Aparicio, I., Alonso, E., 2022. Antibiotic adsorption by natural and modified clay minerals as designer
adsorbents for wastewater treatment: a comprehensive review. J. Environ. Manag. 317, 115397.
Hua, X., Gao, Z., Shi, Y., Hao, W., Liu, X., Li, R., 2023. Transforming industrial solid wastes into eco-friendly zeolite material for efficient heavy metal ion stabilization
through host-guest combination. Chem. Eng. Res. Des. 196, 656–670.
Huang, D., Zhang, G., Yi, J., Cheng, M., Lai, C., Xu, P., Zhang, C., Liu, Y., Zhou, C., Xue, W., Wang, R., Li, Z., Chen, S., 2021. Progress and challenges of metal-organic
frameworks-based materials for SR-AOPs applications in water treatment. Chemosphere 263, 127672.
Huidobro, L., Bautista, Q., Alinezhadfar, M., Gómez, E., Serrà, A., 2024. Enhanced visible-light-driven peroxymonosulfate activation for antibiotic mineralization
using electrosynthesized nanostructured bismuth oxyiodides thin films. J. Environ. Chem. Eng. 12 (3), 112545.
Jiang, M., Hu, L., Zhan, L., Wu, J., Fan, G., Wang, Y., 2023. Spatial self-confinement of heterogeneous metallic CoX (X = Cu, Ni, Fe) within laminated porous carbon
assemblies to efficiently activate peroxymonosulfate for tetracycline destruction. Appl. Surf. Sci. 639, 158202.
Kanmaz, N., Buğdaycı, M., Demirçivi, P., 2023. Investigation on structural and adsorptive features of BaO modified zeolite powders prepared by ball milling
technique: removal of tetracycline and various organic contaminants. Microporous Mesoporous Mater. 354, 112566.
Li, H., Yang, Y., Li, X., Zhou, Z., Feng, J., Dai, Y., Li, X., Ren, J., 2022. Degradation of sulfamethazine by vacuum ultraviolet-activated sulfate radical-advanced
oxidation: efficacy, mechanism and influences of water constituents. Sep. Purif. Technol. 282, 120058.
Li, J., Yang, L., Lai, B., Liu, C., He, Y., Yao, G., Li, N., 2021. Recent progress on heterogeneous Fe-based materials induced persulfate activation for organics removal.
Chem. Eng. J. 414, 128674.
Li, S., Wang, C., Cai, M., Liu, Y., Dong, K., Zhang, J., 2022. Designing oxygen vacancy mediated bismuth molybdate (Bi2MoO6)/N-rich carbon nitride (C3N5) S-scheme
heterojunctions for boosted photocatalytic removal of tetracycline antibiotic and Cr(VI): Intermediate toxicity and mechanism insight. J. Colloid Interface Sci.
624, 219–232.
Limlamthong, M., Lee, M., Jongsomjit, B., Ogino, I., Pang, S., Choi, J., Yip, A.C.K., 2021. Solution-mediated transformation of natural zeolite to ANA and CAN
topological structures with altered active sites for ethanol conversion. Adv. Powder Technol. 32 (11), 4155–4166.
Lou, J., An, J., Wang, X., Cheng, M., Cui, Y., 2024. A novel DBD/VUV/PMS process for efficient sulfadiazine degradation in wastewater: Singlet oxygen-dominated
nonradical oxidation. J. Hazard. Mater. 461, 132650.
Lv, G., Wang, T., Zou, X., Shen, J., Wang, J., Chen, Y., Wang, F., Zhang, X., 2022. Highly dispersed copper oxide-loaded hollow Fe-MFI zeolite for enhanced
tetracycline degradation. Colloids Surf. A Physicochem. Eng. Asp. 655, 130250.
Mao, W., Wang, D., Wang, X., Hu, X., Gao, F., Su, Z., 2022. Efficient cobalt-based metal-organic framework derived magnetic Co@C-600 Nanoreactor for
peroxymonosulfate activation and oxytetracycline degradation. Colloids Surf. A Physicochem. Eng. Asp. 648, 129234.
Masood, Z., Ikhlaq, A., Farooq, U., Qi, F., Javed, F., Aziz, H.A., 2022. Removal of anti-biotics from veterinary pharmaceutical wastewater using combined
Electroflocculation and Fe-Zn loaded zeolite 5A based catalytic ozonation process. J. Water Process Eng. 49, 103039.
Nguyen, T.-B., Nguyen, T.-K.-T., Chen, C.-W., Chen, W.-H., Bui, X.-T., Lam, S.S., Dong, C.-D., 2023. NiCo2O4-loaded sunflower husk-derived biochar as efficient
peroxymonosulfate activator for tetracycline removal in water. Bioresour. Technol. 382, 129182.
Pu, X., Yao, L., Yang, L., Jiang, W., Jiang, X., 2020. Utilization of industrial waste lithium-silicon-powder for the fabrication of novel nap zeolite for aqueous Cu(II)
removal. J. Clean. Prod. 265, 121822.
Qian, J., Mi, X., Chen, Z., Xu, W., Liu, W., Ma, R., Zhang, Y., Du, Y., Ni, B.-J., 2023. Efficient emerging contaminants (EM) decomposition via peroxymonosulfate
(PMS) activation by Co3O4/carbonized polyaniline (CPANI) composite: Characterization of tetracycline (TC) degradation property and application for the
remediation of EM-polluted water body. J. Clean. Prod. 405, 137023.
Selim, A.Q., Mohamed, E.A., Seliem, M.K., Zayed, A.M., 2018. Synthesis of sole cancrinite phase from raw muscovite: Characterization and optimization. J. Alloy.
Compd. 762, 653–667.
Si, D., Zhu, M., You, R., Li, Y., Wu, T., Gui, T., Hu, N., Kumakiri, I., Chen, X., Kita, H., 2021. Influences of alkali metal fluoride and Si/Al ratio on preparation
aluminum-rich ZSM-5 zeolite membrane without organic template. Microporous Mesoporous Mater. 324, 111286.
Silva, R.R.M., Ruotolo, L.A.M., Nogueira, F.G.E., 2023. Peroxymonosulfate activation by magnetic NiFe2O4/g-C3N4 for tetracycline hydrochloride degradation under
visible light. Chem. Eng. J. 476, 146621.
Song, J., Yuan, X., Sun, M., Wang, Z., Cao, G., Gao, K., Yang, C., Zhang, F., Dang, F., Wang, W., 2023. Oxidation of tetracycline hydrochloride with a photoenhanced
MIL-101(Fe)/g-C3N4/PMS system: Synergetic effects and radical/nonradical pathways. Ecotoxicol. Environ. Saf. 251, 114524.
Vargas-Berrones, K., Bernal-Jácome, L., Díaz de León-Martínez, L., Flores-Ramírez, R., 2020. Emerging pollutants (EPs) in Latin América: a critical review of under-
studied EPs, case of study -Nonylphenol. Sci. Total Environ. 726, 138493.
Wang, B., Xu, Z., Dong, B., 2024. Occurrence, fate, and ecological risk of antibiotics in wastewater treatment plants in China: A review. J. Hazard. Mater. 469, 133925.
Wang, L., Hu, A., Liu, H., Yu, K., Wang, S., Deng, X., Huang, D., 2022. Degradation of tetracycline hydrochloride (TCH) by active photocatalyst rich in oxygen
vacancies: performance, transformation product and mechanism. Appl. Surf. Sci. 589, 152902.
Wang, M., Kang, J., Li, S., Zhang, J., Tang, Y., Liu, S., Liu, J., Tang, P., 2022. Electro-assisted heterogeneous activation of peroxymonosulfate by g-C3N4 under visible
light irradiation for tetracycline degradation and its mechanism. Chem. Eng. J. 436, 135278.
Wang, S., Hu, J., Wang, J., 2022. Degradation of sulfamethoxazole using PMS activated by cobalt sulfides encapsulated in nitrogen and sulfur co-doped graphene. Sci.
Total Environ. 827, 154379.
Wang, Y., Lin, N., Xu, J., Jiang, H., Chen, R., Zhang, X., Liu, N., 2023. Construction of microwave/PMS combined dual responsive perovskite-MXene system for
antibiotic degradation: synergistic effects of thermal and non-thermal. Appl. Surf. Sci. 639, 158263.
Wu, Z., Wang, L., Li, Z., Wang, G., Li, C., 2023. Unveiling the promotion of Brønsted acid sites in Cs dispersion and consequential Si-O-Cs species formation for methyl
acrylate synthesis from methyl acetate and formaldehyde over Cs/Beta zeolite catalyst. Chem. Eng. J. 474, 145655.
Xie, K., Jiang, B., Wang, Z., Zuo, S., Wang, Q., 2023. Study on the performance and sulfur tolerance mechanism of Pd/beta zeolite in catalytic combustion of toluene.
J. Mater. Res. Technol. 25, 4543–4553.
Xie, T., Chen, B., Mei, Y., Feng, S., Tang, X., Xiang, W., Yang, J., He, J., Wang, J., Chen, H., Yang, J., Yang, F., 2023. Ultrafast degradation of tetracycline by PMS
activation over perfect cubic configuration MnCo2O4.5: New insights into the role of metal-oxygen bonds in PMS activation. Sep. Purif. Technol. 315, 123694.
Xie, X., Liu, Y., Li, Y., Tao, J., Liu, C., Feng, J., Feng, L., Shan, Y., Yang, S., Xu, K., 2023. Nitrogen-doped Fe-MOFs derived carbon as PMS activator for efficient
degradation of tetracycline. J. Taiwan Inst. Chem. Eng. 146, 104891.
Xiong, Z., Jiang, Y., Wu, Z., Yao, G., Lai, B., 2021. Synthesis strategies and emerging mechanisms of metal-organic frameworks for sulfate radical-based advanced
oxidation process: a review. Chem. Eng. J. 421, 127863.

16
Y. Chen et al. Environmental Technology & Innovation 35 (2024) 103710

Xu, R., Xiong, J., Shi, C., Liu, D., Yang, M., Wei, Z., Nie, Y., Li, J., Ming, Y.-A., 2023. Nickel-doped mesoporous manganese oxides for enhanced peroxymonosulfate-
based oxidation of tetracycline hydrochloride: Performance and mechanism. Microporous Mesoporous Mater. 360, 112702.
Yang, Q., Ma, Y., Chen, F., Yao, F., Sun, J., Wang, S., Yi, K., Hou, L., Li, X., Wang, D., 2019. Recent advances in photo-activated sulfate radical-advanced oxidation
process (SR-AOP) for refractory organic pollutants removal in water. Chem. Eng. J. 378, 122149.
Yin, Z., Zhong, Y., Lu, J., Liu, T., Song, Y., Qing, M., Wang, L., Wang, Y., Hu, S., Su, S., Xiang, J., 2023. Getting insight into the oxidation of SO2 to SO3 over P modified
VW/Ti catalyst: effect and inhibition mechanism. Chem. Eng. J. 459, 141614.
Yu, L., Xu, C., Zhou, Q., Fu, X., Liang, Y., Wang, W., 2023. Facile synthesis of hierarchical porous ZSM-5 zeolite with tunable mesostructure and its application in
catalytic cracking of LDPE. J. Alloy. Compd. 965, 171454.
Yu, T., Chen, H., Hu, T., Feng, J., Xing, W., Tang, L., Tang, W., 2024. Recent advances in the applications of encapsulated transition-metal nanoparticles in advanced
oxidation processes for degradation of organic pollutants: a critical review. Appl. Catal. B: Environ. 342, 123401.
Yu, Y., Quan, H., Zhang, Z.-X., Zhang, Q., Wang, H., Chen, D., Zou, J.-P., 2023. Nonradical pathway dominated activation of peroxymonosulfate by ZnFe2O4/C
composites to eliminate tetracycline hydrochloride: insight into the cycle of Zn/Fe and electron transfer. Sep. Purif. Technol. 322, 124336.
Zeng, H., Yang, B., Shi, W., Huang, K., Ye, C., Ma, X., Wang, Z., Huang, F., Li, X., Deng, J., 2023. Peroxymonosulfate activation by sulfur doped CoFe2O4 rod for
arsanilic acid removal: performance and arsenic enrichment. J. Environ. Chem. Eng. 11 (5), 111044.
Zhang, H., Tong, X., Xiao, H., Wang, H., Zhang, M., Lu, X., Liu, Z., Zhou, W., 2022. Promoting the performance of electrooxidation-PMS system for degradation of
tetracycline by introduction of MnFe2O4/CNT as a third-electrode. Sep. Purif. Technol. 294, 121171.
Zhang, M., Chen, Z., Shao, W., Tian, T., Wang, X., Chen, Z., Qiao, W., Gu, C., 2023. A confined expansion pore-making strategy to transform Zn-MOF to porous carbon
nanofiber for water treatment: Insight into formation and degradation mechanism. J. Colloid Interface Sci. 652, 69–81.
Zhang, P., Li, S., Zhang, C., 2019. Solvent-free synthesis of nano-cancrinite from rice husk ash. Biomass-.-. Convers. Biorefinery 9 (3), 641–649.
Zhang, W., Zhao, R., Bao, B., Liu, S., Hu, C., Ding, W., Zheng, H., 2023. Selective removal of phosphate by magnetic NaCe(CO3)2/Fe3O4 nanocomposites: performance
and mechanism. Sep. Purif. Technol. 325, 124741.
Zhang, X., Gu, W., Liu, D., Zhou, L., Huy, N.N., Wang, L., Zhang, J., Liu, Y., Lei, J., 2023. Fe(II) and Pyridinic N complex sites synergy to activate PMS for specific
generation of 1O2 to degrade antibiotics with high efficiency. Sci. Total Environ. 892, 164067.
Zhang, X., Tian, X., Song, W., Ma, B., Chen, M., Sun, Y., Chen, Y., Zhang, L., 2024. Adsorption of As(III) by microplastics coexisting with antibiotics. Sci. Total Environ.
907, 167857.
Zheng, R., Feng, X., Zou, W., Wang, R., Yang, D., Wei, W., Li, S., Chen, H., 2021. Converting loess into zeolite for heavy metal polluted soil remediation based on “soil
for soil-remediation” strategy. J. Hazard. Mater. 412, 125199.
Zhou, J., Li, X., Yuan, J., Wang, Z., 2022. Efficient degradation and toxicity reduction of tetracycline by recyclable ferroferric oxide doped powdered activated
charcoal via peroxymonosulfate (PMS) activation. Chem. Eng. J. 441, 136061.
Zhu, X., Ma, C., Yuan, X., Zhao, J., Hou, X., 2023. Synthesis of magnetic NaY zeolite for plasma proteomics application. Mater. Today Commun. 35, 106219.
Zou, X., Wang, J., Lv, G., Wang, T., Zhou, C., Chen, Y., Shen, J., Su, S., Liu, Z., 2023. Fe-MFI zeolite-encapsulated copper oxide with excellent re-usability for efficient
tetracycline degradation. Microporous Mesoporous Mater. 360, 112735.
Zuo, H., Wu, C., Du, H., Guo, Z., Cheng, Y., Yan, Q., 2023. Interfacial coupling of 3D nanoflower-like Bi2O2CO3 with PANI for tetracycline photocatalytic degradation
and intermediate toxicity analysis. Appl. Surf. Sci. 633, 157600.

17

You might also like