Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Separation and Purification Technology 303 (2022) 122177

Contents lists available at ScienceDirect

Separation and Purification Technology


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

Singlet oxygen-dominated activation of peroxymonosulfate by 3D


hierarchical MnO2 nanostructures for degradation of organic pollutants in
water: Surface defect and catalytic mechanism
Yichuan Wang , Dezhi Chen *, Zhi-Xia Zhang , Tianlin Zhou , Jian-Ping Zou *
Key Laboratory of Jiangxi Province for Persistent Pollutants Control and Resources Recycle, School of Environmental and Chemical Engineering, Nanchang Hangkong
University, Nanchang 330063, China

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

Keywords: Manganese dioxide (MnO2) is attracting significant attention in the activation of persulfate for the degradation of
Manganese oxide organic pollutants in water. But the factor determining the catalytic efficiency of MnO2, and the evolution of
Peroxymonosulfate activation reactive oxygen species (ROS) remain equivocal and elusive. Herein, three-dimensional (3D) hierarchical MnO2
Non-radical pathway
nanostructures are synthesized via the calcination of hydrothermal product and used to active perox­
Degradation of organic pollutant
Wastewater treatment
ymonosulfate (PMS) for degradation of acid orange 7 (AO7) in water. The results show that the calcination
temperature can tune the morphology of 3D hierarchical MnO2 nanostructures from flower to urchin-like as well
as phase transformation from δ-MnO2 to α-MnO2. Owing to the presence of more surface defects and surface
adsorbed oxygen species, the urchin-like α-MnO2 nanostructures prepared at 500 ◦ C (MnO2-500) deliver the best
catalytic activity in activation of PMS, and 98.3 % of AO7 is degraded in 300 s. Additionally, the MnO2-500
shows wide pH applicability (3.5–9.6). Quenching tests and electron paramagnetic resonance spectra reveal that
the catalytic oxidation of AO7 in the MnO2-500/PMS system is mainly mediated by the non-radical pathway, and
the dominated ROS is 1 O2 . Besides, electrochemical experiment confirms the presence of electron transfer in the
catalytic process. Density functional theory calculations indicate that the C–N and C–S groups of AO7 with high
electron density are easily attacked by electrophilic ROS, which is further confirmed by liquid chromatography-
mass spectrometry. Finally, MnO2-500 exhibits gradually enhanced catalytic activity with the increased cycle
number due to the increased surface defects and the surface adsorbed oxygen. This study provides new insight
into the PMS activation by MnO2-based catalyst for the remediation of organic pollutants in water.

1. Introduction effective in removing refractory azo dyes in water due to their mild
operating conditions, high decolorization and mineralization efficiency
With the rapid industrialization, many organic pollutants such as azo [5,6]. Recently, sulfate radicals (SO•− 4 )-based AOPs are gradually
dyes, pesticide residues, and antibiotics, are dumped into the water, attracting widespread attention [7–9]. Compared with •OH (E0 = 1.9 ~
causing serious threats to the water quality, ecosystem, and public 2.7 V), SO•−
4 not only shows similar strong oxidation (E0 = 2.5 ~ 3.1 V)
health impacts [1,2]. Acid orange 7 (AO7) is a common azo dye used in but also possesses a longer half-life (30 ~ 40 μs) and better selectivity,
paper, printing, and other chemical industries. AO7 is highly toxic and which is conducive to the degradation of the target pollutant [10,11].
can cause cancerization of tissues and organs [3]. Because of its rela­ Generally, SO•−4 radicals are obtained from the activation of persulfate,
tively high stability and biological toxicity, it is challenging to treat AO7 including peroxymonosulfate (PMS) and peroxydisulfate (PDS), by
efficiently by biodegradation and conventional physicochemical pro­ external substances or energy.
cesses, such as adsorption, membrane separation, oxidation, and so on Manganese dioxide (MnO2) has demonstrated outstanding perfor­
[4]. mance and great potential in the activation of persulfate [12]. Saputra
Advanced oxidation processes (AOPs), mainly including Fenton and et al. [13], first reported that α-MnO2 nanowires can activate PMS to
Fenton-like by generating powerful ROS of hydroxyl radicals (•OH), are degrade phenol in aqueous solutions, which is superior to β-MnO2

* Corresponding authors.
E-mail addresses: chendz@nchu.edu.cn (D. Chen), zjp_112@126.com (J.-P. Zou).

https://doi.org/10.1016/j.seppur.2022.122177
Received 28 July 2022; Received in revised form 7 September 2022; Accepted 17 September 2022
Available online 23 September 2022
1383-5866/© 2022 Elsevier B.V. All rights reserved.
Y. Wang et al. Separation and Purification Technology 303 (2022) 122177

nanorods and γ-MnO2 nanofibers [14]. Zhu et al. [15] discovered that collected by vacuum filtration and washed six times with deionized
one-dimension α-MnO2 and β-MnO2 nanomaterials demonstrated much water and then dried at 60 ◦ C. The dried product was heat-treated at
high activity in the activation of PDS among the various manganese different temperatures for 2 h in air with a heating rate of 5 ◦ C/min. The
oxides for selective degradation of organic contaminants, and 1 O2 was obtained 3D hierarchical MnO2 nanostructures at 300, 400, 500, 600,
the primary ROS. Zhou et al. [16], found that α-MnO2 exhibited much 700 ◦ C were labeled as MnO2-300, MnO2-400, MnO2-500, MnO2-600,
better catalytic performance in activating PMS for the degradation of 4- and MnO2-700, respectively.
nitrophenol in water than δ-MnO2, and SO•− 4 and •OH were the main
active radicals. Khan et al. [17]. reported that ε-MnO2 could activate 2.3. Characterization
PMS more efficiently than α-MnO2, β-MnO2, and δ-MnO2, and the
1
generated ROS were SO•− 4 , •OH, O2 , and O2 . Lyu et al.[18] found that
•−
The as-prepared MnO2 samples were characterized using thermog­
ε-MnO2 can effectively enhance the catalytic performance of β-FeOOH in ravimetric/differential thermal analyzer (TG/DTA), scanning electron
PMS activation due to the synergistic effect Fe3+/Fe2+ and Mn4+/Mn3+ microscope (SEM), transmission electron microscopy (TEM), X-ray
as well as the generated FeOH+ and MnOH2+. The above results indicate powder diffractometer (XRD), Fourier-transform infrared spectrometer
that the crystalline phase of MnO2 nanomaterials can influence their (FTIR), Raman spectrometer, EPR spectrometer, surface area & pore size
catalytic activity for the activation of persulfate. Besides, it was reported analyzer, X-ray photoelectron spectrometer (XPS). Details of charac­
that the morphology of MnO2 nanomaterials also plays essential roles in terization are shown in Text S2.
activating persulfate. Wang et al. [19] found that α-MnO2 nanowires
exhibited higher activity than the α-MnO2 nanorods and nanotubes 2.4. Catalytic degradation of AO7
because of the higher specific surface area and larger amount of hy­
droxyl radical generation. In addition, α-MnO2 nanowires could effi­ In a typical degradation process, 50 mg of MnO2 catalyst was added
ciently activate persulfate for 2,4-dichlorophenol (DCP) oxidation, and to a 250 mL brown beaker containing 100 mL of AO7 solution with the
SO•−4 and •OH radicals were the dominated ROS in the α-MnO2/per­ initial concentration (C0) of 50 mg/L with continuous stirring at 400
sulfate system [20]. Furthermore, Compared with these non- rpm. After uniform dispersion of the catalyst, PMS was put in the reac­
hierarchical nanostructures, three-dimensional (3D) hierarchical archi­ tion solution to trigger the AO7 degradation. At the set time interval, 2.5
tectures are of much interest in heterogeneous catalysis because of their mL of the reaction solution was collected and separated using a pipette
high performance and easy separation [21–23]. Deng et al.[24] verified equipping with 0.22 μm filter membrane, 0.5 mL of methanol was used
that the highly catalytic activity of α-MnO2 nanoflowers in activating to terminate the degradation reaction. All experiments were carried out
PMS mainly relied on their high specific surface area and high crystal­ in triplicate. Liquid chromatography (LC) and LC-mass spectrometry
linity. 3D-hierarchically structured corolla-like δ-MnO2 assembled by (LC-MS) were used to detect the AO7 concentration and the in­
nanosheets showed higher activity for activating PMS in the degradation termediates (Detail in Text S3), respectively. The radicals in the MnO2/
of phenol than the urchin-shaped α-MnO2 with sub-blocks of nanorods, PMS system were determined by EPR spectrum (Detail in Text S4). TOC-
and SO•− 4 played a major role in phenol oxidation [21]. Although so L CPH/CPN analyzer (Shimadzu) was used to analyze the total organic
many meaningful results have been achieved, the activation of persul­ carbon (TOC) during the degradation of AO7 in water. Finally, to
fate by MnO2 with different 3D hierarchical nanostructures prepared evaluate the reproducibility of the catalyst, the used MnO2 was collected
under different calcination temperature for the degradation of pollut­ and washed using water and methanol alternately-three times, and then
ants have not been described, and the intrinsic mechanisms are required dried at 80 ◦ C. The leached metal ions were detected using a Bremen
to be further studied. Hanna-Kunath-Sir.11, 28,199 inductively coupled plasma mass spec­
In this paper, 3D hierarchical MnO2 nanostructures with different trometer (ICP-MS).
morphologies and phases were prepared by a simple calcination of hy­
drothermal products, and their catalytic activities in activating PMS for 3. Results and discussion
the AO7 degradation were evaluated. In addition, the influence of
calcination temperature on the catalytic efficiency of 3D hierarchical 3.1. Characterization
MnO2 nanostructures for activating PMS was investigated. The influence
of experimental parameters, including initial pH, PMS dosage, MnO2 3.1.1. SEM
dosage, reaction temperature, coexisting inorganic ions, and organic Fig. 1 shows the typical SEM images of as-prepared MnO2 hierar­
matters, on the degradation of AO7 in the MnO2/PMS system was chical structures. Pristine MnO2 exhibits the hydrangea-like structures
explored. Finally, the degradation pathway of AO7 and its intermediates (Fig. 1a), assembled by nanosheets (Fig. 1b). After heat-treatment at 300
in the MnO2/PMS system were proposed, and the metal leaching and the and 400 ◦ C, the obtained MnO2-300 (Fig. 1c-d) and MnO2-400 (Fig. 1e-f)
cycle stability of catalysts were evaluated to prove the application po­ remain the hydrangea-like structures, but the thickness of nanosheet
tential of the MnO2/PMS system in the actual wastewater treatment. units is increased. When the temperature of heat-treatment increased to
500 ◦ C, the hydrangea-like structures are transformed into urchin-like
2. Materials and methods structures composed by nanorods (Fig. 1g-h), implying the presence of
crystal transformation between 400 and 500 ◦ C. From SEM images of
2.1. Materials Fig. 1i-j at 600 ◦ C, we can observe that these nanorods in urchin-like
structures become thin and long. When the temperature increased to
All reagents (see Text S1) were analytical and used without further 700 ◦ C, the size of urchin-like structures decreases and the length of
purification. Deionized water (18 MΩ cm) was used in the experiment. nanorod units gets shorter and parts of nanorods detach from the hier­
archical architecture (Fig. 1k-l), which may be due to the structure
collapse at high temperature.
2.2. Preparation of 3D hierarchical MnO2 nanostructures
3.1.2. XRD
In the synthesis process, 6 mol of KMnO4 and 1 mol of MnSO4 were XRD patterns were used to determine the crystalline phase of as-
dissolved in 70 mL of deionized water and the obtained solution was prepared MnO2 samples. As shown in Fig. 2a, we can see three main
transferred to a 100 mL Teflon-lined stainless steel autoclave. The XRD peaks in the pristine MnO2 at 2θ = 12.35, 24.70 and 37.22◦ , which
autoclave reactor was sealed and placed in an oven at 120 ◦ C for 12 h match well with the (1 0 0), (0 0 2) and (0 0 3) crystal plane of
and then naturally cooled to room temperature. The product was K0.46Mn1.54Mn0.46O4(H2O)1.4 (JCPDF No. 80–1098) namely δ-MnO2,

2
Y. Wang et al. Separation and Purification Technology 303 (2022) 122177

Fig. 1. SEM images of a-b) pristine MnO2, c-d) MnO2-300, e-f) MnO2-400, g-h) MnO2-500, i-j) MnO2-600, k-l) MnO2-700.

respectively. The shift to low angel for the three XRD peaks is common in α-MnO2 is conducive to the oxygen species adsorbed on the surface of
the layered δ-MnO2 because of the intercalation of water molecules and MnO2, and the infrared activity of OH combined with Mn atoms in­
K ions. The obtained MnO2-300 displays the same characterization creases with the increase of calcination temperature.
peaks as the pristine MnO2. Meanwhile, the increased 2θ to 24.82◦ for
the (0 0 2) peak indicates the decrease of interplanar spacing in MnO2 3.1.4. Raman and EPR
caused by the removal of intercalated water molecules, which was Fig. 2c demonstrates the Raman spectra of the as-prepared MnO2
confirmed by TG curves shown in Fig. S1. With the increase of calci­ nanostructures after calcination. Raman spectroscopy is sensitive for
nation temperature, apart from the characterization peaks for δ-MnO2, probing the short-range environment of oxygen coordination around
we can observe new diffraction peaks at 2θ = 12.60, 18.04, 28.76, transition metal cations in oxide lattices [32]. The band at 642 cm− 1
37.47, 41.79, 49.67, 56.14, 59.99, and 65.10◦ in MnO2-400, corre­ corresponds to the symmetric stretching vibration of Mn-O bond in
sponding to the (1 1 0), (2 0 0), (3 1 0), (2 1 1), (3 0 1), (4 1 1), (6 0 0), MnO6 groups [33]. Generally, the intensity of Raman band at 642 cm− 1
(5 2 1) and (0 0 2) crystal plane of α-MnO2 (JPCDS No. 44–0141), is used to evaluate the state of Mn-O bond. More high intensity repre­
respectively. It indicates that part of the metastable δ-MnO2 was trans­ sents the more perfect Mn-O bond. MnO2-500 shows a barely visible
formed into thermostable α-MnO2 at 400 ◦ C, which can be confirmed by band at 642 cm− 1, suggesting its weakest Mn-O bond in MnO6 groups in
TG/DTA results (Fig. S1). More water molecules were removed from the all the heat-treated samples. The weak Mn-O bond implies the presence
crystal structure of MnO2-400. When the temperature was increased to of structure defects on the surface of MnO2. Therefore, the room-
500 ◦ C and more, it can be seen only α-MnO2 diffraction peaks, agreeing temperature solid-state EPR spectroscopy was used to determine the
well with the morphological evolution observed in SEM images (Fig. 1). oxygen vacancy of all the heat-treated MnO2 samples. As shown in
In addition, the enhanced diffraction peak intensity demonstrates that Fig. 2d, the signal at g = 2.002 indicates the presence of oxygen va­
the crystallinity of α-MnO2 increased gradually with the increase of cancies on the surface of the as-prepared MnO2 samples. The stronger
calcination temperature. signal for MnO2-500 suggests its more oxygen vacancies than other
MnO2 samples.
3.1.3. FTIR
Fig. 2b shows the FTIR spectra of as-prepared MnO2 samples. The 3.1.5. XPS
bands between 400 and 800 cm− 1 correspond to the Mn–O vibrations of To further determine the valence state of Mn and the chemical
MnO6 octahedra in MnO2 [25,26]. With the increase of calcination environment of O on the surface of the as-calcined MnO2 samples, XPS
temperature, the bands centered at 472, 590 and 717 cm− 1 appear and spectra were obtained and displayed in Fig. 3. In the XPS survey spectra
become gradually sharp, indicating the presence of phase trans­ shown in Fig. 3a, Mn, O and K signals can be observed in all the heat-
formation from δ-MnO2 to α-MnO2 [25,27] and the increased degree of treated MnO2 samples. K is originated from the KMnO4. The Mn
crystallinity for α-MnO2. The bands located at 1637 and 3440 cm− 1 can average oxidation state (AOS) on the surface of MnO2 plays a vital role in
be attributed to the O–H stretching and bending vibrations of water, the activation of PMS. The AOS was obtained by the Mn 3 s multiple
respectively [28]. The band at 1051 cm− 1 corresponds to the bending splitting spectra method based on the following equation: AOS = 8.95 –
vibrations of OH in Mn-OH. The strong band at 1384 cm− 1 for MnO2 1.13 × ΔE, where ΔE is the energy difference between the main peak
nanostructures, prepared at 400 ◦ C and more-higher calcination tem­ and its satellite in Mn 3 s [34]. According to the Mn 3 s XPS spectra
2 ) adsorbed on
perature, belongs to the bending of peroxide species (O•− presented in Fig. 3b, the calculated Mn surface AOS (Table S1) of MnO2-
the lattice surface of MnO2 [29,30]. When the heat-treatment temper­ 300, MnO2-400, MnO2-500, MnO2-600, and MnO2-700 is 3.74, 3.7,
ature increased to 600 ◦ C and more, the band at 1117 cm− 1 appears and 3.63, 3.82, and 3.87, respectively. Previous literature reported that the
can be attributed to the bending vibrations of OH in Mn-OH [31]. The low AOS is beneficial to obtaining high catalytic activity for crystalline
FTIR results suggest that the phase transformation from δ-MnO2 to MnO2 [34,35]. The valence state of Mn on the surface of the as-calcined

3
Y. Wang et al. Separation and Purification Technology 303 (2022) 122177

Fig. 2. a) XRD pattens, b) FTIR, c) Raman and d) EPR spectra of as-prepared MnO2 samples.

MnO2 samples was further analyzed by Mn 2p spectra (Fig. 3c). Mn 2p3/ samples. Fig. 4a shows that the hydrangea-like MnO2 structures in the
2 can be fitted into three peaks around 640.4, 641.5, and 643.2 eV, as-prepared MnO2-400 are assembled by nanosheets, which are curled
corresponding to Mn3+ and Mn4+ [36]. The content of Mn3+ and Mn4+ and stacked together (Fig. 4b). HRTEM image of Fig. 4c displays mul­
on the surface of all the as-prepared MnO2 samples was calculated from tiple groups of lattice fringes, including ordered and disordered lattice
the peak area, and the ratio of Mn3+/Mn4+ is displayed in Table S1. fringes. Two kinds of lattice fringes with the lattice spacing of 2.53 and
Generally, the high ratio of Mn3+/Mn4+ is good for the catalytic activity 2.39 Å match well with the (2 0 0) plane of δ-MnO2 and (2 1 1) plane of
of MnO2 because the electron transfer between Mn3+ and Mn4+ can α-MnO2, respectively. It indicates that the MnO2-400 contains both
promote the PMS activation for generating ROS [17,37]. Moreover, metastable δ-MnO2 and thermally stable α-MnO2, matching with the
oxygen species on the surface of MnO2 can affect its catalytic perfor­ XRD results shown in Fig. 2a. The high-angle annular dark-field scan­
mance in activating PMS. Fig. 3d demonstrates the O 1 s narrow scan ning transmission electron microscopy (HAADF-STEM) and energy
spectra, which can be deconvoluted into two peaks centered at 529.6 dispersive X-ray spectroscopy (EDX) results in Fig. 4d demonstrate that
and around 532 eV, corresponding to surface lattice oxygen (OL) and the presence of O, Mn, and K elements. Fig. 4e displays the TEM image of
surface adsorbed oxygen or surface hydroxy species (OA). The center of MnO2-500, indicating the urchin-like structure constructed by MnO2
OA peaks is shifted from 531.9 eV of MnO2-300 to 532.3 eV of MnO2- nanorods. The HRTEM image in Fig. 4f shows the highly ordered lattice
500, suggesting the change of OA. According to the XPS and the FTIR, structure of MnO2 nanorods, and the corresponding selected area elec­
we hold that the shift is mainly attributed to the adsorbed oxygen (O•−2 ) tron diffraction pattern (SAED) (inset of Fig. 4f) confirms that the MnO2
formatted at the calcination temperature of 400 ◦ C and more. It indicates nanorod is a single crystal. The clearly ordered lattice fringes (Fig. 4g)
that the phase transformation from δ-MnO2 to α-MnO2 is conducive to with the plane spacing of 6.99 Å are assigned to the (1 1 0) plane of
the formation of O•− 2 on the surface of MnO2 samples. The previous α-MnO2. The HADDF-STEM and EDX results in Fig. 4h exhibit the
works had found that the increased OA showed a positive effect on the presence of O, Mn, and K elements.
activation of PMS by MnO2 [34,38–40].
3.1.7. N2 adsorption/desorption isotherm
3.1.6. TEM Fig. S2a shows the nitrogen adsorption/desorption isotherms of all
TEM images (Fig. 4) of the as-calcined MnO2-400 and MnO2-500 the as-prepared MnO2 samples. The typical type IV isotherms suggest the
were selected to observe the microstructure of the as-prepared MnO2 presence of IUPAC-graded slit-like pores. The obtained isotherms and

4
Y. Wang et al. Separation and Purification Technology 303 (2022) 122177

Fig. 3. XPS spectra of all heat-treated MnO2 sample. a) survey, b) Mn 3 s, c) Mn 2p, d) O 1 s.

hysteresis loop are almost overlapping, indicating no obvious difference much higher catalytic activity than the pristine δ-MnO2, δ-MnO2 at
in the specific surface area (SBET) and pore volume (VBJH) of all the as- 300 ◦ C and the δ-MnO2 plus α-MnO2 at 400 ◦ C. In addition, the MnO2-
prepared MnO2 samples before and after calcination, though the 500 demonstrates the highest catalytic activity in all as-prepared sam­
morphology of MnO2 units change significantly from nanosheets to ples, suggesting that the suitable calcination temperature is essential for
nanorods. Table S2 presents the calculated SBET and VBJH. Fig. S2b the catalytic activity of the as-prepared MnO2 to activate PMS. To reveal
demonstrates a similar pore size distribution for all MnO2 samples and the effect of calcination temperature on the catalytic activity, the rela­
the presence of mesopores is conducive to the transfer of mass [41]. tionship between AO7 degradation rate constants and structure char­
acteristics of the as-prepared MnO2 samples are plotted. Generally, the
higher SBET of catalyst provide more active sites for the activation of
3.2. Catalytic performance of MnO2
PMS [24]. Fig. 5c displays that the high SBET of the as-prepared MnO2
samples is advantageous to the AO7 degradation in the MnO2/PMS
3.2.1. Effect of calcination temperature
system except for the MnO2-300 because of its crystalline phase of
Fig. 5a shows the effect of calcination temperature on the catalytic
δ-MnO2 and low crystallinity [13,16]. However, the low correlation
degradation of AO7. AO7 is stable and almost no self-degradation in
value (R2 = 0.66) indicates the SBET is not the main reason for control­
aqueous solution. When MnO2-500 was put alone into AO7 aqueous
ling the catalytic performance of MnO2. In addition, the surface defect is
solution, negligible degradation of AO7 indicates that both the adsorp­
a key to the catalytic activity of MnO2 because the surface defects are
tion capacity and the degradation efficiency of AO7 is low in the pres­
usually used as reaction sites, and higher defect density is conducive to
ence of MnO2-500. Furthermore, only 12.6 % of AO7 is degraded under
the PMS adsorption on the surface of MnO2 [42]. Fig. 5d shows the
the dosing of PMS alone. However, the degradation of AO7 is fast in the
correlation between degradation rate of AO7 and the intensity of the
MnO2/PMS system, and the degradation efficiency of AO7 is from 56.0
Raman band at 642 cm− 1, which represents the state of the Mn-O bond
% to 98.3 % in 300 s, suggesting that the oxidant PMS can be effectively
on the surface of MnO2. From MnO2-300 to MnO2-700, we can see that
activated by the as-prepared MnO2 samples. Fig. 5b shows the corre­
the MnO2 sample with a weaker Mn-O bond exhibit a higher degradation
sponding pseudo-first-order kinetic rate constant for the MnO2/PMS
rate of AO7 no matter what δ-MnO2 or α-MnO2. It means that the weak
system, and the obtained α-MnO2 samples at 500, 600 and 700 ◦ C show

5
Y. Wang et al. Separation and Purification Technology 303 (2022) 122177

Fig. 4. TEM, HRTEM, HADDF-STEM and EDX mapping images of a-d) MnO2-400, e-f) MnO2-500.

Fig. 5. a) Effect of calcination temperature on the degradation of AO7 (Reaction condition: [catalyst] = 0.5 g/L, [PMS] = 1.2 g/L, [AO7] = 50 mg/L, T = 30 ◦ C,
initial pH = 5.60); b) the corresponding pseudo-first order kinetic rate constant; AO7 degradation rate constant versus c) SBET, d) intensity of band at 642 cm− 1, e)
intensity of Ov, f) AOS, g) content of surface Mn3+, and h) content of surface OA.

Mn-O bond is beneficial for the high catalytic activity of MnO2 [37,43]. concentration of catalysts, suggesting that the OA on the surface of as-
Meanwhile, the weak Mn-O band suggests the presence of oxygen va­ calcined MnO2 samples, including both adsorbed oxygen and surface
cancies (Ov) and Mn3+. Fig. 5e and Fig. 5f demonstrate that the presence hydroxy species, can improve the activation of PMS. Furthermore, the
of Ov and Mn3+ is good for the activation of PMS, respectively, and the correlation value (R2 = 0.78) indicates that the OA plays a crucial role in
high density of both Ov and Mn3+ can boost the AO7 degradation in the the PMS activation of the heat-treated MnO2 samples. Finally, the
MnO2/PMS system. Subsequently, the influence of Ov and Mn3+ on the leaching of Mn was determined by ICP-MS, Table S3 presents the
degradation rate of AO7 is further confirmed by the AOS of the as- leaching of Mn in these MnO2/PMS systems, the extremely low con­
calcined MnO2 samples shown in Fig. 5g, which is consistent with pre­ centration of Mn ions indicates that as-calcined MnO2 samples are stable
vious results that the low AOS of MnO2 was conducive to activating PMS and the influence of leached Mn ions on the degradation of AO7 can be
[37,44]. Besides, the surface adsorbed oxygen species (OA) are impor­ almost negligible. The highest leaching concentration of Mn ions is 167
tant for the catalytic activity of MnO2 [45]. Fig. 5h shows a positive μg/L for MnO2-700, which is much lower than the requirements (2 mg/
linear relationship between the degradation rate of AO7 and the OA L) of the Effluent Standards (GB8978-1996) in China.

6
Y. Wang et al. Separation and Purification Technology 303 (2022) 122177

3.2.2. Effect of experimental parameters


3.2.2.3. Effect of pH. As shown in Fig. 6c, the degradation efficiency of
AO7 in the MnO2-500/PMS system under the initial solution pH value
3.2.2.1. Effect of the catalyst dosage. The effect of catalyst dosage on the
from 3.6 to 9.6 reach more than 90 % in 300 s, demonstrating the high
degradation of AO7 in the MnO2-500/PMS system is presented in
efficiency of the MnO2-500/PMS system in a wide pH range. Fig. S3
Fig. 6a. The degradation efficiency of AO7 gradually increases from
displays the Zeta potential of MnO2-500. As the increase of pH value, the
91.87 % to 98.32 % as the increased concentration of MnO2-500 from
Zeta potential of MnO2-500 decrease, and the PZC of MnO2-500 is 3.08,
0.3 to 0.5 g/L in 300 s, indicating that increasing the dosage of MnO2-
suggesting that the surface charge of MnO2-500 is positive under acidic
500 can effectively improve the degradation efficiency of AO7 due to
condition. The pKa1 of PMS and pKa2 of PMS are 0 and 9.442, respec­
more MnO2-500 can provide more active sites for the PMS activation.
tively [49]. Therefore, at pH 5.6–9.6, HSO−5 , which is the main active
However, at 0.7 g/L, almost the same degradation kinetics and degra­
species of PMS, is easily activated by MnO2-500 to produce SO•− 4 radi­
dation efficiency of AO7 indicate that the excessive catalyst dosage does
not improve the degradation reaction because of the limited ROS. cals, leading to the rapid degradation rate of AO7 [50]. In addition, in
strongly acidic media, 1 O2 is produced in the system much less than in
other media. Moreover, SO•− 4 and •OH may react with H , thus reducing
+
3.2.2.2. Effect of PMS dosage. The effect of PMS dosage on the AO7
degradation is shown in Fig. 6b. With the increase of PMS concentration the activity of the oxidant in acidic media [51,52]. At high alkali con­
from 1 to 1.2 g/L, the degradation efficiency of AO7 increases in the centration with an initial pH of 11.6, the degradation efficiency is
MnO2-500/PMS system. It is mainly attributed to that more ROS are significantly reduced to 70.26 %, probably due to the fact that the
generated with the increase of PMS dose [46]. However, when the PMS adsorption of PMS on the surface of MnO2 is suppressed by excess alkali
concentration is 1.4 g/L, both the degradation rate and degradation [42].
efficiency of AO7 decrease, which is mainly due to the self-scavenging
effect in the presence of excess PMS (Eqs. (1)–(5)) [47,48]. 3.2.2.4. Effect of reaction temperature. Fig. 6d shows the influence of
reaction temperature on the AO7 degradation in the MnO2-500/PMS
2−
S2 O8 + SO•−4 →SO2−4 + S2 O•−8 (1) system. It indicates that high reaction temperature is conductive to the
degradation of AO7. The relationship between the reaction rate constant
(2)

S2 O28 + • OH→OH− + S2 O•−8 and reaction temperature was illustrated through the variant of the
Arrhenius Equation:
SO•−4 + SO•−4 →2SO2−4 (3)
Ea
In k = − + In A (6)
•OH + • OH→H2 O2 (4) RT
where k (min− 1) is the reaction rate constant and Ea (J/mol) is the
SO•−4 +• OH→HSO−4
1
+ O2 (5) activation energy for the degradation of AO7, T (K) is the absolute
2 temperature, R (8.314 J/(mol⋅K)) is the universal gas constant, and A is

Fig. 6. Effect of key experimental parameters on the degradation of AO7 in the MnO2-500/PMS system. a) MnO2-500 dosage, b) PMS concentration, c) initial pH of
AO7 solution, d) reaction temperature, e) inorganic anions (10 mM), f) humic acid. Control condition: [catalyst] = 0.5 g/L, [PMS] = 1.2 g/L, [AO7] = 50 mg/L, T =
30 ◦ C, initial pH = 5.60.

7
Y. Wang et al. Separation and Purification Technology 303 (2022) 122177

the frequency or pre-exponential factor. The influence NO−3 on the degradation of AO7 is very slight. Besides,
The correlation between the rate constant and temperature is fitted the SO2− 2−
4 can react with SO4 and •OH to form S2 O8 , OH , and SO4
•− − •−

by Eq. (6), as shown in the inset of Fig. 6d. The calculated Ea of AO7 (Eqs. (8) and (9)), declining the degradation efficiency of AO7 [58].
degradation in the MnO2-500/PMS system is 12.72 kJ/mol [14], which
is much lower than the reported systems, including Mn3O4/rGO/PMS SO2−4 + SO•−4 →S2 O2−8 + e− (8)
(49.5 kJ/mol) [53], MFe2O4/rGO/PMS (25.7 kJ/mol) [54], Fe3O4@C/
MnCo2O4/PMS (52.62 kJ/mol) [55], and MnOx/PDS (36.614–46.772 SO2−4 + • OH→OH− + SO•−4 (9)
kJ/mol) [56]. It indicates that the MnO2-500/PMS system exhibits a Fig. 6f displays the effect of humic acid on the degradation of AO7 in
high reactivity in the degradation of AO7. MnO2-500/PMS system. Generally, the adsorption competition between
humic acid and PMS as well as the adsorption and oxidation competition
3.2.2.5. Effect of organic and inorganic anions. To explore the applica­ between humic acid and target pollutant is not conducive to the
tion potential of catalyst in actual wastewater treatment, the effect of degradation of target pollutant in water [59]. Under the presence of
these common anions (Cl-, SO2− 4 and NO3 ) and natural organic matter

humic acid from 10 to 50 mg/L, no apparent change can be seen. At 100
(NOM) in actual water body on the MnO2-500/PMS system was evalu­ mg/L humic acid, the degradation efficiency of AO7 is slightly reduced.
ated. As shown in Fig. 6e, the addition of 10 mM of Cl- exhibits an Fig. S4 shows the degradation of AO7 by the MnO2-500/PMS in two
enhancement on the degradation rate of AO7, which may be due to the kinds of natural water samples. Compared with deionized water, the
reaction of Cl- and SO•−
4 to form Cl (Eq. (7)). Moreover, the produced

degradation kinetics and degradation efficiency of AO7 in tap water and
reactive Cl can catalyze the propagation reactions to produce more

river water are slightly decreased because of the inhibitory effect caused
4 , resulting in improving the degradation performance [57].
SO•− by the presence of inorganic anions and organic matter [60], indicating
that the MnO2-500/PMS system has potential in actual application of
SO•−4 + Cl− →SO2−4 + Cl• (7)
wastewater treatment.

Fig. 7. Influence of various quenching agents on AO7 degradation. a) MeOH, b) TBA, c) p-BQ, d) L-His. Reaction condition: [catalyst] = 0.5 g/L, [PMS] = 1.2 g/L,
[AO7] = 50 mg/L, T = 30 ◦ C, initial pH = 5.60.

8
Y. Wang et al. Separation and Purification Technology 303 (2022) 122177

3.3. Mechanism of AO7 catalytic degradation by the MnO2-500/PMS the electron transfer process on the surface of MnO2-500 was further
system validated electrochemically (Text S5). Cyclic voltammetry (CV) curves
(Fig. S5a) demonstrate that the current increases when PMS and AO7
To investigate the mechanism of AO7 degradation in the MnO2-500/ was dosed into the reaction system. According to the linear sweep vol­
PMS system, the ROS during the reaction were identified by quenching tammetry (LSV) curves in Fig. S5b, when PMS and AO7 existed at the
experiments. Methanol is usually considered as an effective scavenger of same time, the oxidation current increases significantly, further proving
4 and •OH, while tert-butyl alcohol (TBA) is only effective for •OH
SO•− that the charge transfer in the degradation of AO7 is accelerated.
[61]. Fig. 7a shows that the inhibition effect of methanol on the Furthermore, Fig. S5c shows the i-t curve in which displays that the
degradation reaction of AO7 is weak even the concentration of methanol current slightly increases after the input of PMS at 200 s and AO7 at 400
increased to 1000 mM, suggesting that SO•− 4 or •OH is not the dominant s. It implies that the electron reaction and electron transfer ability are
ROS in the degradation of AO7 by the MnO2-500/PMS system. Fig. 7b better in the presence of both PMS and AO7, suggesting that the current
demonstrates that the inhibitory effect of TBA is slightly stronger than flows from the AO7 molecule to the catalyst surface [68]. The conduc­
that of methanol, probably because TBA is more hydrophobic than tivity of catalyst plays an important role in the catalytic process of
methanol, thus preventing the free radicals generated near the catalyst producing active 1 O2 [69]. Electrochemical impedance spectroscopy
surface by AO7 [62]. With the increase of TBA from 10 mM to 1000 mM, (EIS) experiment was performed to analyze the conductivity of catalyst.
the slightly reduced degradation rate and almost invisible difference in Fig. S5d displays the obtained EIS Nyquist plots of MnO2-500 and pure
the degradation efficiency of AO7 further indicate that the presence of • glassy carbon electrode (GCE). In the high frequency range, the half
OH is not the dominant ROS in the degradation of AO7 by MnO2-500/ circles indicate the charge transfer resistance (Rct). In the low frequency
PMS system. With the introduction of different doses of p-benzoquinone range, the linear branches correspond to the diffusion process of the
(p-BQ) (5, 10, and 20 mM), no inhibition in the degradation of AO7 reactants. It is clear that the MnO2-500 shows low Rct and good diffu­
reveals that the concentration of O•− 2 radicals is very low or its contri­
sivity, suggesting its good electrical conductivity and good electron
bution is very little to the degradation reaction in the MnO2-500/PMS transfer rate as well as easy diffusion of reactants [70]. Thus, the
system (Fig. 7c). L-histidine (L-His) was used as a specific probe for the development of non-radical catalytic pathways has been greatly
generation of 1 O2 . As shown in Fig. 7d, the inhibition for the degrada­ facilitated.
tion of AO7 is greater with increasing L-His concentration, indicating According to the above analysis, the possible catalytic mechanism for
that 1 O2 is generated and possibly dominate the degradation reaction of the AO7 degradation in the MnO2-500/PMS system is shown in Scheme
AO7 in the MnO2-500/PMS system. 1. The defective structure and vacancies of the MnO2-500 catalyst can
EPR spectra were obtained to further analyze the involvement of generate more electronic contact sites, which is conducive to improving
these ROS in the reaction solution. Fig. 8a shows the signals of DMPO- the catalytic activity. PMS can be activated by the mutual trans­
4 and DMPO-•OH during the degradation of AO7 by the MnO2-500/
SO•− formation between Mn3+ and Mn4+ on the surface MnO2. Mn3+ provides
PMS system. It indicates that the presence of SO•− electrons for reducing PMS to generate SO•− 4 and is converted to higher
4 and •OH species, but
the signals with moderate intensity for the DMPO- SO•− 4 and DMPO-•OH
valence states. Meanwhile, Mn4+ accepts electrons from PMS to develop
3+
suggests that the SO•−4 and •OH species are not the dominant ROS in the
SO•−5 with lower activity and is reduced to Mn . The reaction processes
MnO2-500/PMS system. Fig. 8b displays the EPR spectra of DMPO-O•− 2 ,
are shown in Eqs. 10–11 [71]. Owing to its high reactivity, O•− 2 radicals
weak signals after the introduction of PMS are observed due to the can rapidly react with H+ to produce the hydroperoxyl radical (HO2 •)
autolytic decomposition of PMS [63,64], which suggests that the according to Eq. (12) [72]. Then, the produced HO2 • is recombined with
1 4+
contribution of O•− 2 can be ignored in the degradation reaction of AO7 2 to generate O2 (Eq. (13)) [73,74]. In addition, Mn
O•− species on the
by MnO2-500/PMS system. Fig. 8c demonstrates the signals of TEMP- surface of the MnO2 catalyst can oxide O2 to product 1 O2 (Eq. (14)),
•−
1
O2 at different reaction times in the degradation of AO7 by MnO2-500/ which is thermodynamically feasible (E0(Mn4+ /Mn3+ ) = 0.95 VNHE and
PMS system. The observed strong peaks of 1 O2 further reveal that the E0O•−2 = –0.34 VNHE ). The redox cycles between Mn3+ and Mn4+ occur on
generated 1 O2 dominates the degradation of AO7 in the MnO2-500/PMS the surface of the MnO2 catalyst and facilitated electron transfer,
system, which matches well with the quenching experiment (Fig. 7) and resulting in the formation of ROS during MnO2-500/PMS oxidation
previous literature report [63,65,66]. [17].
Based on the above results, the quenching agents could not
completely inhibit the degradation reaction of AO7, indicating the Mn3+ + HSO−5 →Mn4+ + SO•−4 + OH− (10)
presence of another degradation path. MnO2 was reported to undergo
valence changes during degradation and used as a catalyst for electron Mn4+ + HSO−5 →Mn3+ + SO•−5 + H+ (11)
transfer to achieve high valence oxidation properties [67]. Therefore,

1
Fig. 8. EPR spectra of a) •OH and SO•−
4 , b) O2 and c) TEMP- O2 at different reaction times in the degradation of AO7 by MnO2-500/PMS system.
•−

9
Y. Wang et al. Separation and Purification Technology 303 (2022) 122177

Scheme 1. The proposed reaction mechanism for the MnO2-500/PMS system.

O•−2 + H+ →HO2 • (12) 3.4. AO7 degradation pathway in the MnO2-500/PMS system

O•−2 + HO2 • →1 O2 + HO2 (13) The electron density of atoms in the AO7 molecule was calculated by
density functional theory (DFT). As shown in Fig. S6, the C–N and C–S
[ ] [ ]
Mn4+ − OSO3 − OH + 3O•−2 → Mn3+ − OH + SO2−4 + 31 O2 (14) groups of AO7 possess high electron densities, implying that these
functional groups are easily attacked by electrophilic ROS. Furthermore,
the intermediate products during the AO7 degradation in the MnO2-

Scheme 2. Possible intermediates and degradation pathways of AO7 in the MnO2-500/PMS system.

10
Y. Wang et al. Separation and Purification Technology 303 (2022) 122177

500/PMS system were analyzed by LC-MS. According to the spectra CRediT authorship contribution statement
shown in Fig. S7 and the previous DFT result, the possible degradation
pathways of AO7 in the MnO2-500/PMS systems are proposed and Yichuan Wang: Data curation, Writing – original draft. Dezhi Chen:
shown in Scheme 2. The initial AO7 molecule (m/z = 329) is mainly Conceptualization, Supervision, Funding acquisition, Writing – review
degraded at the azo double bond. In pathway I, the ROS attacks the two & editing. Zhi-Xia Zhang: Validation. Tianlin Zhou: Data curation.
C–N bonds in the azo dye concurrently, resulting in the decomposition Jian-Ping Zou: Supervision, Project administration.
of AO7 into p-phenol sulfonic acid (m/z = 175), 1, 2-naphthoquinone
(m/z = 159), and N2. 1, 2-naphthoquinone is transformed into the
Declaration of Competing Interest
product (m/z = 175), which is further oxidated to the hydroxylated
product (m/z = 191) and the ring-opening product (m/z = 193). In
The authors declare that they have no known competing financial
pathway II, the sulfonic group of AO7 is removed and transformed into
interests or personal relationships that could have appeared to influence
the product (m/z = 265), which is further decomposition to p-diphenol,
the work reported in this paper.
1, 2-naphthoquinone, and N2 because the two C–N bonds are simulta­
neously attacked by ROS. The following degradation path is consistent
Data availability
with the pathway I. These intermediates are degraded into small mole­
cules of carboxylic acids such as HCOOH, CH3COOH, etc., which are
No data was used for the research described in the article.
further oxidized to generate carbon dioxide, water, and other inorganic
ions. Fig. S8 presents that 40 % TOC is removed, supporting the above
results. Acknowledgements

This work was supported by the National Natural Science Foundation


3.5. Stability and recyclability of the catalyst (51720105001, 51968049) of China, the Natural Science Foundation
(20192ACB21031, 20192BCB23012) of Jiangxi Province, China.
The catalyst was used repeatedly to evaluate its stability and recy­
clability in activating PMS for degradation of AO7. Fig. S9a demon­ Appendix A. Supplementary material
strates the catalytic degradation curves of AO7 by the MnO2-500/PMS
system, and the corresponding degradation rate constant kobs of AO7 Supplementary data to this article can be found online at https://doi.
was calculated using the pseudo-first-order kinetic model and shown in org/10.1016/j.seppur.2022.122177.
Fig. S9b, which indicates that the degradation rate of AO7 is increased
with the increase of cycle number in the MnO2-500/PMS system. To References
analyze the enhanced catalytic activity of MnO2-500 after cycle, the XPS
spectra for the used catalyst after 2, 5 and 9 cycles are presented in [1] J.H. Qu, M.H. Fan, The current state of water quality and technology development
for water pollution control in China, Crit. Rev. Environ. Sci. Technol. 40 (2010)
Fig. S9c. With the increase of cycle, the content of lattice oxygen on the
519–560.
surface of MnO2-500 decreases but the surface adsorbed oxygen in­ [2] A. Tkaczyk, K. Mitrowska, A. Posyniak, Synthetic organic dyes as contaminants of
creases, which suggests that the decreased lattice oxygen is transformed the aquatic environment and their implications for ecosystems: A review, Sci. Total
to surface adsorbed oxygen for improving the contact between PMS and Environ. 717 (2020), 137222.
[3] Y. Yan, Z.Q. Geng, K. Dai, X. Guo, R.J. Zeng, F. Zhang, Decoupling mechanism of
catalyst and promoting the PMS activation. In addition, the increased Acid Orange 7 decolorization and sulfate reduction by a Caldanaerobacter
density of Mn3+ (Fig. S9d) and the decreased AOS (Table S1) for the used dominated extreme-thermophilic consortium, J. Hazard. Mater. 419 (2021),
MnO2-500 after cycle further verify the above results. Moreover, the Mn 126498.
[4] M.C. Collivignarelli, A. Abba, M.C. Miino, S. Damiani, Treatments for color
ion leaching of the MnO2-500 catalyst during each cycle was determined removal from wastewater: State of the art, J. Environ. Manage. 236 (2019)
by ICP-MS. As shown in Table S4, the maximum leaching concentration 727–745.
of Mn ions is 199 μg/L at the 2nd and 4th cycle, which is lower than the [5] J. Su, L. Zhu, P. Geng, G. Chen, Self-assembly graphitic carbon nitride quantum
dots anchored on TiO2 nanotube arrays: An efficient heterojunction for pollutants
requirements (2 mg/L) of the Effluent Standards (GB8978-1996). The degradation under solar light, J Hazard Mater 316 (2016) 159–168.
gradually reduced leaching concentration of Mn ions with the increase [6] Y. Xiang, H.L. Liu, E.R.Y. Zhu, K. Yang, D.L. Yuan, T.F. Jiao, Q.R. Zhang, S.F. Tang,
of cycle indicates the excellent potential of the MnO2-500 in the actual Application of inorganic materials as heterogeneous cocatalyst in Fenton/Fenton-
like processes for wastewater treatment, Sep. Purif. Technol. 295 (2022), 121293.
application for wastewater treatment.
[7] G. Nie, L. Xiao, J.X. Bi, S.B. Wang, X.G. Duan, New insight to piezocatalytic
peroxymonosulfate activation: The critical role of dissolved oxygen in mediating
4. Conclusion radical and nonradical pathways, Appl. Catal. B Environ. 315 (2022), 121584.
[8] Z.K. Xiong, Y.N. Jiang, Z.L. Wu, G. Yao, B. Lai, Synthesis strategies and emerging
mechanisms of metal-organic frameworks for sulfate radical-based advanced
In conclusion, 3D hierarchical MnO2 nanostructures with different oxidation process: A review, Chem. Eng. J. 421 (2021), 127863.
morphologies and phases were prepared through the calcination of hy­ [9] J. Li, X. Li, X. Wang, L. Zeng, X. Chen, J. Mu, G. Chen, Multiple regulations of Mn-
based oxides in boosting peroxymonosulfate activation for norfloxacin removal,
drothermal product and used as heterogeneous catalysts for PMS acti­ Appl Catal A-Gen. 584 (2019), 117170.
vation. The heat-treatment temperature can control the catalytic activity [10] W.J. Tian, S. Chen, H.Y. Zhang, H. Wang, S.B. Wang, Sulfate radical-based
of as-calcined MnO2 nanostructures. Because of the more surface defects advanced oxidation processes for water decontamination using biomass-derived
carbon as catalysts, Curr. Opin. Chem. Eng. 37 (2022), 100838.
and OA, the urchin-like α-MnO2 nanostructures (MnO2-500) delivers a [11] J. Li, G. Gou, H.L. Zhao, C. Liu, N.W. Li, L.G. Li, B. Tan, B. Lai, Efficient
greater catalytic performance in activating PMS than other MnO2 sam­ peroxymonosulfate activation by CoFe2O4-CeO2 composite: Performance and
ples. The quenching experiment and EPR spectra indicate that the non- catalytic mechanism, Chem. Eng. J. 435 (2022), 134840.
[12] X. Zheng, X. Niu, D. Zhang, M. Lv, X. Ye, J. Ma, Z. Lin, M. Fu, Metal-based catalysts
radical path dominate the catalytic degradation of AO7 in the MnO2- for persulfate and peroxymonosulfate activation in heterogeneous ways: A review,
500/PMS system. Furthermore, the intermediate products in the Chem. Eng. J. 429 (2022), 132323.
degradation of AO7 were identified, and the possible degradation [13] E. Saputra, S. Muhammad, H. Sun, A. Patel, P. Shukla, Z.H. Zhu, S. Wang, α-MnO2
activation of peroxymonosulfate for catalytic phenol degradation in aqueous
pathway was proposed. More importantly, the MnO2-500 demonstrates
solutions, Catal. Commun. 26 (2012) 144–148.
superior recyclability because of the gradually increased OA and surface [14] E. Saputra, S. Muhammad, H. Sun, H.M. Ang, M.O. Tade, S. Wang, Different
defects. Besides, the as-prepared 3D hierarchical MnO2 nanostructures crystallographic one-dimensional MnO2 nanomaterials and their superior
possessed a very low leaching amount. These impressive results provide performance in catalytic phenol degradation, Environ Sci Technol. 47 (11) (2013)
5882–5887.
a new insight into the activation of PMS by MnO2-based catalyst for [15] S. Zhu, X. Li, J. Kang, X. Duan, S. Wang, Persulfate activation on crystallographic
wastewater treatment. manganese oxides: Mechanism of singlet oxygen evolution for nonradical selective

11
Y. Wang et al. Separation and Purification Technology 303 (2022) 122177

degradation of aqueous contaminants, Environ. Sci. Technol. 53 (1) (2019) [41] P. Li, S. Zhan, L. Yao, Y. Xiong, S. Tian, Highly porous α-MnO2 nanorods with
307–315. enhanced defect accessibility for efficient catalytic ozonation of refractory
[16] Z.-G. Zhou, H.-M. Du, Z. Dai, Y. Mu, L.-L. Tong, Q.-J. Xing, S.-S. Liu, Z. Ao, J.- pollutants, J. Hazard Mater. 437 (2022), 129235.
P. Zou, Degradation of organic pollutants by peroxymonosulfate activated by MnO2 [42] Y. Wang, M. Liu, C. Hu, Y. Xin, D. Ma, M. Gao, H. Xie, Enhanced MnO2/
with different crystalline structures: Catalytic performances and mechanisms, peroxymonosulfate activation for phthalic acid esters degradation: Regulation of
Chem. Eng. J. 374 (2019) 170–180. oxygen vacancy, Chem. Eng. J. 433 (2022), 134048.
[17] A. Khan, K. Zhang, A.K.A. Taraqqi, X. Wang, Y. Chen, Y. Zhang, Degradation of [43] D.M. Robinson, Y.B. Go, M. Mui, G. Gardner, Z. Zhang, D. Mastrogiovanni,
antibiotics in aqueous media using manganese nanocatalyst-activated E. Garfunkel, J. Li, M. Greenblatt, G.C. Dismukes, Photochemical water oxidation
peroxymonosulfate, J. Colloid. Interface. Sci. 599 (2021) 805–818. by crystalline polymorphs of manganese oxides: structural requirements for
[18] C. Lyu, D. He, Z. Mou, X. Yang, Synergetic activation of peroxymonosulfate by catalysis, J. Am. Chem Soc. 135 (2013) 3494–3501.
MnO2-loaded beta-FeOOH catalyst for enhanced degradation of organic pollutant [44] F. Wang, J. Deng, S. Impeng, Y. Shen, T. Yan, G. Chen, L. Shi, D. Zhang, Unraveling
in water, Sci. Total Environ. 693 (2019), 133589. the effects of the coordination number of Mn over α-MnO2 catalysts for toluene
[19] Y. Wang, S. Indrawirawan, X. Duan, H. Sun, H.M. Ang, M.O. Tadé, S. Wang, New oxidation, Chem. Eng. J. 396 (2020), 125192.
insights into heterogeneous generation and evolution processes of sulfate radicals [45] B. Chen, B. Wu, L. Yu, M. Crocker, C. Shi, Investigation into the catalytic roles of
for phenol degradation over one-dimensional α-MnO2 nanostructures, Chem. Eng. various oxygen species over different crystal phases of MnO2 for C6H6 and HCHO
J. 266 (2015) 12–20. oxidation, ACS Catal. 10 (2020) 6176–6187.
[20] Y. Zhao, Y. Zhao, R. Zhou, Y. Mao, W. Tang, H. Ren, Insights into the degradation [46] Q. Qin, T. Liu, J. Zhang, R. Wei, S. You, Y. Xu, Facile synthesis of oxygen vacancies
of 2,4-dichlorophenol in aqueous solution by α-MnO2 nanowire activated enriched alpha-Fe2O3 for peroxymonosulfate activation: A non-radical process for
persulfate: catalytic performance and kinetic modeling, RSC Adv. 6 (2016) sulfamethoxazole degradation, J. Hazard. Mater. 419 (2021), 126447.
35441–35448. [47] D. Yang, P. Hong, Y. Hu, Y. Li, C. Wang, J. He, B. Sun, S. Zhu, L. Kong, J. Liu,
[21] Y. Wang, H. Sun, H.M. Ang, M.O. Tadé, S. Wang, 3D-hierarchically structured Carbon framework-encapsulated copper oxide particles to activate
MnO2 for catalytic oxidation of phenol solutions by activation of peroxymonosulfate for the efficient degradation of tetracycline, Appl. Surf. Sci. 552
peroxymonosulfate: Structure dependence and mechanism, Appl. Catal. B Environ. (2021), 149424.
164 (2015) 159–167. [48] S. Adil, W.S. Kim, T.H. Kim, S. Lee, S.W. Hong, E.J. Kim, Defective, oxygen-
[22] H. Huang, K. Xiao, T. Zhang, F. Dong, Y. Zhang, Rational design on 3D hierarchical functionalized multi-walled carbon nanotubes as an efficient peroxymonosulfate
bismuth oxyiodides via in situ self-template phase transformation and phase- activator for degradation of organic pollutants, J. Hazard. Mater. 396 (2020),
junction construction for optimizing photocatalysis against diverse contaminants, 122757.
Appl. Catal. B Environ. 203 (2017) 879–888. [49] S. Xiao, M. Cheng, H. Zhong, Z. Liu, Y. Liu, X. Yang, Q. Liang, Iron-mediated
[23] M. Halhouli, J. Kieninger, O. Yurchenko, G. Urban, Rational design of activation of persulfate and peroxymonosulfate in both homogeneous and
morphological characteristics to determine the optimal hierarchical nanostructures heterogeneous ways: A review, Chem. Eng. J. 384 (2020), 123265.
in heterogeneous catalysis, ChemCatChem. 9 (2017) 354–364. [50] D. Sun, C. Li, S. Lu, Q. Yang, C. He, Magnetic Fe3O4@CoFe-LDH nanocomposite
[24] J. Deng, Y.J. Ge, C.Q. Tan, H.Y. Wang, Q.S. Li, S.Q. Zhou, K.J. Zhang, Degradation heterogeneously activated peroxymonosulfate for degradation of azo-dye AO7,
of ciprofloxacin using alpha-MnO2 activated peroxymonosulfate process: Effect of RSC Adv. 11 (2021) 20258–20267.
water constituents, degradation intermediates and toxicity evaluation, Chem. Eng. [51] N. Jaafarzadeh, F. Ghanbari, M. Moradi, Photo-electro-oxidation assisted
J. 330 (2017) 1390–1400. peroxymonosulfate for decolorization of acid brown 14 from aqueous solution,
[25] H. Wang, Z. Lu, D. Qian, Y. Li, W. Zhang, Single-crystal α-MnO2 nanorods: Korean. J. Chem. Eng. 32 (2015) 458–464.
synthesis and electrochemical properties, Nanotechnology. 18 (11) (2007), [52] X. Zhao, X. Yi, J. Song, X. Yuan, S. Yu, Y. Nie, J. Zhang, G. Cao, Mesoporous and
115616. flexible polyimide aerogel as highly active catalytic membrane for AO7
[26] T. Xu, P. Zhang, H. Zhang, Ultrathin δ-MnO2 nanoribbons for highly efficient degradation by peroxymonosulfate activation, Chem. Eng. J. 431 (2022), 134286.
removal of a human-related low threshold odorant-acetic acid, Appl. Catal. B [53] Yunjin, Yao, Chuan, Xu, Shaoming, Yu, Facile Synthesis of Mn3O4-reduced
Environ. 309 (2022), 121273. graphene oxide hybrids for catalytic decomposition of aqueous organics, Ind. Eng.
[27] V. Sannasi, K. Subbian, Influence of Moringa oleifera gum on two polymorphs Chem. Res. 2013; 52(10): 3637-3645.
synthesis of MnO2 and evaluation of the pseudo-capacitance activity, J. Mater. Sci. [54] Y. Yao, Y. Cai, F. Lu, F. Wei, X. Wang, S. Wang, Magnetic recoverable MnFe2O4 and
Mater. Electron. 31 (2020) 17120–17132. MnFe2O4-graphene hybrid as heterogeneous catalysts of peroxymonosulfate
[28] D.P. Dubal, D.S. Dhawale, T.P. Gujar, C.D. Lokhande, Effect of different modes of activation for efficient degradation of aqueous organic pollutants, J. Hazard. Mater
electrodeposition on supercapacitive properties of MnO2 thin films, Appl. Surf. Sci. 270 (2014) 61–70.
257 (2011) 3378–3382. [55] J. Lu, Q. Liu, Z. Xiong, Z. Xu, Y. Cai, Q. Wang, Activation of peroxymonosulfate
[29] G. Zhu, J. Zhu, W. Li, W. Yao, R. Zong, Y. Zhu, Q. Zhang, Tuning the K+ with magnetic and recyclable Fe3O4@C/MnCo2O4 nanocomposites for the
Concentration in the Tunnels of alpha-MnO2 To Increase the Content of Oxygen decolorization of Acid Orange II, J. Chem. Technol. Biot. 92 (7) (2017) 1601–1612.
Vacancy for Ozone Elimination, Environ. Sci. Technol. 52 (2018) 8684–8692. [56] D.Q. Liu, Q.W. Li, J.B. Hou, H.Z. Zhao, Mixed-valent manganese oxide for catalytic
[30] S. Sivakumar, L. Nelson Prabu, Synthesis and characterization of α-MnO2 oxidation of Orange II by activation of persulfate: heterojunction dependence and
nanoparticles for Supercapacitor application, Mater. Today: Proc. 47 (2021) mechanism, Catal. Sci. Technol. 11 (11) (2021) 3715–3723.
52–55. [57] C. Qi, X. Liu, C. Lin, X. Zhang, J. Ma, H. Tan, W. Ye, Degradation of
[31] R. Yang, Z. Wang, L. Dai, L. Chen, Synthesis and characterization of single- sulfamethoxazole by microwave-activated persulfate: Kinetics, mechanism and
crystalline nanorods of α-MnO2 and γ-MnOOH, Mater. Chem. Phys. 93 (2005) acute toxicity, Chem. Eng. J. 249 (2014) 6–14.
149–153. [58] Z. Zhao, J. Zhao, C. Yang, Efficient removal of ciprofloxacin by
[32] C. Julien, M. Massot, S. Rangan, M. Lemal, D. Guyomard, Study of structural peroxymonosulfate/Mn3O4 -MnO2 catalytic oxidation system, Chem. Eng. J. 327
defects in γ-MnO2 by Raman spectroscopy, J. Raman Spectrosc. 33 (2002) (2017) 481–489.
223–228. [59] H. Wang, W. Guo, B. Liu, Q. Wu, H. Luo, Q. Zhao, Q. Si, F. Sseguya, N. Ren, Edge-
[33] L. Yan, C. Shen, L. Niu, M.C. Liu, J. Lin, T. Chen, Y. Gong, C. Li, X. Liu, S. Xu, nitrogenated biochar for efficient peroxydisulfate activation: An electron transfer
Experimental and theoretical investigation of the effect of oxygen vacancies on the mechanism, Water. Res. 160 (2019) 405–414.
electronic structure and pseudocapacitance of MnO2, ChemSusChem. 12 (2019) [60] W.-D. Oh, Z. Dong, Z.-T. Hu, T.-T. Lim, A novel quasi-cubic CuFe2O4–Fe2O3 catalyst
3571–3581. prepared at low temperature for enhanced oxidation of bisphenol A via
[34] J. Wan, L. Zhou, H. Deng, F. Zhan, R. Zhang, Oxidative degradation of peroxymonosulfate activation, J. Mater. Chem. A. 3 (2015) 22208–22217.
sulfamethoxazole by different MnO2 nanocrystals in aqueous solution, J. Mol. [61] Y. Liu, J. Luo, L. Tang, C. Feng, J. Wang, Y. Deng, H. Liu, J. Yu, H. Feng, J. Wang,
Catal. A-Chem. 407 (2015) 67–74. Origin of the enhanced reusability and electron transfer of the carbon-coated
[35] F. Nawaz, H. Cao, Y. Xie, J. Xiao, Y. Chen, Z.A. Ghazi, Selection of active phase of Mn3O4 nanocube for persulfate activation, ACS Catal. 10 (2020) 14857–14870.
MnO2 for catalytic ozonation of 4-nitrophenol, Chemosphere. 168 (2017) [62] P. Hu, H. Su, Z. Chen, C. Yu, Q. Li, B. Zhou, P.J.J. Alvarez, M. Long, Selective
1457–1466. degradation of organic pollutants using an efficient metal-free catalyst derived
[36] X. Peng, Double-exchange effect in two-dimensional MnO2 nanomaterials, J. Am. from carbonized polypyrrole via peroxymonosulfate activation, Environ. Sci.
Chem. Soc. 139 (14) (2017) 5242–5248. Technol. 51 (2017) 11288–11296.
[37] J. Huang, Y. Dai, K. Singewald, C.-C. Liu, S. Saxena, H. Zhang, Effects of MnO2 of [63] P. Liang, C. Zhang, X. Duan, H. Sun, S. Liu, M.O. Tade, S. Wang, An insight into
different structures on activation of peroxymonosulfate for bisphenol A metal organic framework derived N-doped graphene for the oxidative degradation
degradation under acidic conditions, Chem. Eng. J. 370 (2019) 906–915. of persistent contaminants: formation mechanism and generation of singlet oxygen
[38] A. Khan, H. Wang, Y. Liu, A. Jawad, J. Ifthikar, Z. Liao, T. Wang, Z. Chen, Highly from peroxymonosulfate, Environ. Sci. Nano 4 (2017) 315–324.
efficient α-Mn2O3@α-MnO2-500 nanocomposite for peroxymonosulfate activation: [64] Y. Zhao, H. An, J. Feng, Y. Ren, J. Ma, Impact of crystal types of AgFeO2
comprehensive investigation of manganese oxides, J. Mater. Chem. A. 6 (2018) nanoparticles on the peroxymonosulfate activation in the water. Environ, Sci.
1590–1600. Technol. 53 (2019) 4500–4510.
[39] X. Bi, Y. Huang, X. Liu, N. Yao, P. Zhao, X. Meng, D. Astruc, Oxidative degradation [65] Y. Li, J. Li, Y. Pan, Z. Xiong, G. Yao, R. Xie, B. Lai, Peroxymonosulfate activation on
of aqueous organic contaminants over shape-tunable MnO2 nanomaterials via FeCo2S4 modified g-C3N4 (FeCo2S4-CN): Mechanism of singlet oxygen evolution for
peroxymonosulfate activation, Sep. Purif. Technol. 275 (2021), 119141. nonradical efficient degradation of sulfamethoxazole, Chem. Eng. J. 384 (2020),
[40] S.-L. Chiam, S.-Y. Pung, F.Y. Yeoh, M. Ahmadipour, Highly efficient oxidative 123361.
degradation of organic dyes by manganese dioxide nanoflowers, Mater. Chem. [66] M. Pu, J. Niu, M.L. Brusseau, Y. Sun, C. Zhou, S. Deng, J. Wan, Ferrous metal-
Phys. 280 (2022), 125848. organic frameworks with strong electron-donating properties for persulfate
activation to effectively degrade aqueous sulfamethoxazole, Chem. Eng. J. 394
(2020), 125044.

12
Y. Wang et al. Separation and Purification Technology 303 (2022) 122177

[67] Y. Yang, P. Zhang, K. Hu, P. Zhou, Y. Wang, A.H. Asif, X. Duan, H. Sun, S. Wang, [71] M. Zhao, R. Xu, Z. Chen, Z. Gao, S. Zheng, H. Song, Kinetics and mechanisms of
Crystallinity and valence states of manganese oxides in Fenton-like polymerization diniconazole degradation by α-MnO2 activated peroxymonosulfate, Sep. Purif.
of phenolic pollutants for carbon recycling against degradation, Appl. Catal. B Technol. 281 (2022), 119850.
Environ. 315 (2022), 121593. [72] A. Mp, A. Jn, B. Mlb, A. Ys, A. Cz, D.A. Sheng, C. Jw, Ferrous metal-organic
[68] Y. Hu, D. Chen, R. Zhang, Y. Ding, Z. Ren, M. Fu, X. Cao, G. Zeng, Singlet oxygen- frameworks with strong electron-donating properties for persulfate activation to
dominated activation of peroxymonosulfate by passion fruit shell derived biochar effectively degrade aqueous sulfamethoxazole, Chem. Eng. J. 394 (2020), 125044.
for catalytic degradation of tetracycline through a non-radical oxidation pathway, [73] Shishu, Zhu, Xiaojie, Jian, Kang, Xiaoguang, Duan, Shaobin, Wang, Persulfate
J. Hazard. Mater 419 (2021), 126495. activation on crystallographic manganese oxides: Mechanism of singlet oxygen
[69] J. Yu, B. Hu, P. Nie, X. Shang, B. Zhang, Z. Tang, J. Yang, J. Liu, Co, Zn-MOF evolution for nonradical selective degradation of aqueous contaminants, Environ.
derived Co@C-NCNTs for boosting non-radical oxidation of norfloxacin by Sci. Technol. 2019,53(1):307-315.
peroxymonosulfate, Colloids Surf. A Physicochem. Eng. Asp. 647 (2022), 129037. [74] X. Liu, Y. Chen, Y. Yao, Q. Bai, Z. Wu, Iodine-doped carbon fibers as an efficient
[70] H. Dai, W. Zhou, W. Wang, Co/N co-doped carbonaceous polyhedron as efficient metal-free catalyst to activate peroxymonosulfate for the removal of organic
peroxymonosulfate activator for degradation of organic pollutants: Role of cobalt, pollutants, Catal. Sci. Technol 8 (21) (2018) 5482–5489.
Chem. Eng. J. 417 (2021), 127921.

13

You might also like