Applied Catalysis A, General

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Applied Catalysis A, General 623 (2021) 118246

Contents lists available at ScienceDirect

Applied Catalysis A, General


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

Synergistic effect for efficient catalytic persulfate activation in conducting


polymers-hematite sand composites: Enhancement of chemical stability
Abdellah Ait El Fakir a, b, *, Zakaria Anfar a, b, c, Abdallah Amedlous d, Asma Amjlef a,
Salaheddine Farsad a, Amane Jada b, c, *, Noureddine El Alem a
a
Laboratory of Materials & Environment (LME), Ibn Zohr University, Agadir, 80000, Morocco
b
Institute of Materials Science of Mulhouse (IS2M-CNRS), Haute Alsace University, Mulhouse, 68100, France
c
Strasbourg University, Strasbourg, 67081, France
d
Laboratory of Materials, Catalysis &Valorization of Natural Resources (LaMaCaVa), Hassan II University, Casablanca, 20650, Morocco

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

Keywords: This study explores new chemical strategies for coating the hematite sand (HS) by conductive polymers to
Hematite sand (HS) enhance the persulfate activation performance. The highest catalytic activity was due to synergistic effect be­
Conductive polymers tween the HS and the conductive polymers. In fact, conjugated polymers could act as an activator and electronic
Synergistic effect
mediator to promote the conversion of Fe (III) to Fe (II). Hence, both conductive polymers used (polyaniline and
Persulfate activation
Chemical stability
polypyrrole) have been found to provide the necessary electron transfer, thanks to their nitrogen atoms located in
their conjugated chains. The radical scavenging experiments assessed by electronic paramagnetic resonance
indicated that the studied pathways nature was both a radical and non-radical type. Further, it was found that the
protective polymer layers prevented metal ion leaching, mitigate catalyst deactivation, enabled pH insensitive
persulfate activation, and increased the chemical stability. This work provides a new innovative way to remove
multiple pollutants in wastewater.

1. Introduction inherent drawbacks of iron leaching and the slow conversion of Fe (III)
to Fe (II) have resulted in poor catalytic stability of iron oxide, and
Advanced oxidation processes technologies based on the use of sul­ severely hamper their widespread applications [10]. Furthermore,
fate radicals (SR-AOPs) have received much attention in the last decades carbon-based materials or carbocatalysts such as graphene oxide,
for wastewater treatment [1]. In fact, sulfate radicals (SO4− ), has shown reduced graphene oxide [12–14], nanodiamond carbon nanotubes [15,

more efficient performance in the degradation of recalcitrant pollutants, 16], were considered as promising carbocatalysts for the PS activation
due to their high oxidative potentials (2.5–3.1 V) [1–3]. In addition, for organic pollutants degradation. In fact, the unique properties of these
persulfates (PS) or peroxymonosulfate (PMS), are considered the main metal-free catalysts (e.g., large surface area, superior conductivity,
sources to generate the SO4− radical [4,5]. The activation processes of non-toxicity, high catalytic activity, and superior biocompatibility)

PS and PMS include many physicochemical approaches that increase the make them promising candidates in the AOPs [6,7,17]. Further, the
oxidative reactivity of persulfate, as a moderate oxidant [6,7]. There­ nitrogen doping in the carbocatalyst structures increases the charge
fore, the PS activation involves the formation of highly reactive sulfate density of adjacent carbon atoms, which promotes the interaction be­
radicals (SO4− ) by the O–O bond dissociation, and persulfate conversion tween the catalyst and the PS, facilitating hence the O–O bond cleav­

to SO4 by electrons transfer [7]. Among the various activation methods age, and also inducing a non-radical degradation pathway of the
• −

including heterogeneous catalysis (electro or photo), UV photolysis and pollutant [17–19]. It is supposed that the N atoms can effectively create
the use of metal oxide-based catalysts [8]. In addition, iron-based ma­ point defects, which could generate 1O2, while oxygen groups in the
terials have been widely studied as activators of PS, due to the Fe carbon matrices network are the main active sites contributing to the
element abundant reserve, their low toxicity, and their very high electrons transfer [6,7]. Recently, nitrogen doped carbon-based cata­
achievable catalysts in the activation of PS [9–11]. Unfortunately, the lysts-encapsulated metal particles have been shown to promote the

* Corresponding authors at: Laboratory of Materials & Environment (LME), Ibn Zohr University, Agadir, 80000, Morocco; Institute of Materials Science of Mul-
house (IS2M-CNRS), Haute Alsace University, Mulhouse, 68100, France; Strasbourg University, Strasbourg, 67081, France.
E-mail addresses: [email protected] (A. Ait El Fakir), [email protected] (A. Jada).

https://doi.org/10.1016/j.apcata.2021.118246
Received 20 April 2021; Received in revised form 4 June 2021; Accepted 6 June 2021
Available online 15 June 2021
0926-860X/© 2021 Elsevier B.V. All rights reserved.
A. Ait El Fakir et al. Applied Catalysis A, General 623 (2021) 118246

activating ability of persulfate in a synergistic and significant manner control catalyst [31].
(Table S1). In fact, transition metals (e.g. iron) coated with nitrogen
doped carbon structures could increase the charge density of adjacent 2.2. Catalysts characterization methods
carbon atoms. This option favours the interfacial electronic exchange by
creating a synergistic effect between the core (transition metals) and the The composition and the surface chemical states of the prepared
shell (carbon matrix), enhancing hence the electrons exchange between catalysts were analysed by X-ray photoelectrons spectroscopy (XPS),
the catalyst and the persulfate and also facilitating the mass recovery of using XPS SES-2002 (VG SCIENTA). Transmission Electron Microscopy
the catalyst after use [20–27]. Moreover, the carbon coated metals (TEM) type Philips CM 200, 20–200 kV, was used to assess the catalysts
catalysts attenuate or probably prevent the metal ions leaching, during nanostructures, whereas their morphologies were analysed by using
the catalytic reaction, and improves hence the chemical stability of Scanning Electron Microscopy (SEM) FEI, Model Quanta 400. RAMAN
metal catalysts [22,28–30]. On the other hand, the determination of the Spectroscopy, type Horiba model Labram BX40, with CCD detector,
dominant pathways during the PS activation, by using these kinds of operating at laser line of 532 nm, was used in this work, to determine the
catalysts, needs to be elucidated (i.e., the incorporation of metal into the molecular composition and the catalysts external structures. The pore
carbon matrix may involve the production of either radical or size distributions and the specific surface areas were measured using
non-radical species, or both). Therefore the origin of the synergistic ef­ ASAP 2420. To get further insights on the catalysts surface structures,
ficiency occurring between both materials and resulting in the catalytic Fourier Transform Infrared (FTIR) spectra were obtained in the mid
performance improvement and the stability during the recycling tests, infrared region (400-4000cm− 1) using a Shimadzu 4800S. The catalyst
must be studied in depth and confirmed [20,22,28]. structural analyzes were assessed by X-ray diffraction (XRD), and the
The present study was carried out in order to fill the knowledge gap data were collected on a PANalytical MPD X ’diffractometer, Pert Pro,
regarding the use of metal-carbon composites for persulfate activation. operating with Cu Kα radiation (λ=0.15418 nm), equipped with a real-
Thus, in our work, nitrogen-rich conductive polymers, such as poly­ time multi-band detector X’Celerator (active length ¼ 2.12 2θ).
aniline (PANI), and polypyrrole (PPy), were used for the functionaliza­
tion of rich iron oxide hematite sand (HS) particles. It should be noted 2.3. Experimental procedure and analytical methods
that the location of nitrogen atoms in the conjugate chains, improves the
polymer stability and prevents its oxidation. Furthermore, PANI and The catalytic activation of the PS was carried out in a 100 mL beaker
PPy, are electrically conductive polymers, and their conjugated chains containing 50 mL of Orange G (OG) solution (50 mg L− 1), with constant
can provide the electrons transfer, necessary to create a synergistic effect agitation, by using known concentration of the PS, and the catalyst
with the iron metal species. On the other hand, the hematite sand (HS) doses. The initial solution pH was adjusted with aqueous solutions of
rich in iron oxide were chosen for their low cost and their abundant in H2SO4 (0.1 M) and NaOH (0.1 M). The experiments in real conditions
nature. The prepared catalysts HS@PANI and HS@PPy, were used as were carried out in tap water (TW) under the same conditions. Further,
heterogeneous catalysts to activate the PS for the degradation of organic other pollutants were tested, such as Rhodamine B (RhB) and Bisphenol
pollutants to CO2 and H2O. More, we have used various characterization A (BPA), under the same conditions as OG, but with different pollutant
techniques to verify and to confirm the expected functionalized catalyst concentrations (20 mg L− 1 for RhB and 10 mg L− 1 for BPA). After the
structures, and to elucidate the synergistic effect occurring between the degradation reaction, the residual organic contaminants concentrations
hematite sand and the conductive polymers. of OG, RhB and BPA, were analysed by using UV–vis spectroscopy model
6705 UV / Vis JENWAY, at wavelengths of, 478, 554 and 278 nm,
2. Experimental section respectively. To study the mineralization of the different pollutants, the
total organic carbon (TOC) was measured, using TOC-L SHIMADZU
2.1. Catalysts preparation analyser. In addition, to determine the nature of reactive species pro­
duced during the catalytic conversion process, various radical scaven­
The chemicals and the materials used in this study are listed in Text gers such as ethanol (EtOH) and tert-butyl-alcohol (TBA), with well-
S1, Supporting Information. The PANI functionalized HS (HS@PANI) determined molar ratio EtOH/PS or TBA/PS, were added in the reac­
catalyst was prepared following in-situ polymerization strategy, and by tion media to trap SO4− and OH . Further, the sodium azide (NaN3) was
• •

using common oxidation method of aniline monomer with a known ratio used as a deactivator for 1O2. It should be noted that the experimental
(aniline/ammonium persulfate (APS): 1/1.5). Thus, 1 g of HS was firstly tests were performed in triplicate, so the error bars were added in figures
crushed and screened at 50 μm (HS particle size < 50 μm), then dealing with the contaminant degradation. Further, in order to identify
dispersed in 50 mL of HCl (0.1 M). Subsequently, 0.5 mL of aniline was the presence of the ROS in the degradation process, in situ electron
added to the HS aqueous dispersion, and the resulting mixture was paramagnetic resonance (EPR) analyzes were conducted. Thus, for each
sonicated for 30 min. In the later step, APS (1.85 g dispersed in 50 mL of analysed solution, a volume of 50 μL of the reaction medium was
HCl (0.1 M)) was slowly added to the HS-aniline suspension for 2 h with withdrawn and then it was transferred in a glass capillary sealed with
constant stirring. A dark green color resulted from the HS-aniline-APS Crit-O-seal TM and finally placed in an ESR-tube with an outside
mixture which was due to the oxidative polymerization process. At the diameter of 5 mm. N-tert-butyl-α-phenylnitrone (PBN) and 2,2,6,6-tera­
end of the polymerization reaction, the obtained composite was washed methypiperidine (TEMP) were used as spin trap agents. Continuous-
with distilled water and ethanol, then dried at 60 ◦ C under vacuum for wave EPR spectra were recorded at room temperature and in aerated
12 h. By using the same in-situ polymerization strategy, the polyaniline medium using an EMX-plus X-band spectrometer (Bruker). For the sta­
(PANI) alone was prepared as a control catalyst [31]. bility tests, the catalysts were magnetically separated after each run,
Thus, a weighted of HS (HS particle size < 50 μm) amount of 1 g was thereafter it was washed with ethanol /distilled water three times, and
dispersed in 50 mL of distilled water. Thereafter, 0.5 mL of pyrrole was finally dried at 333◦ K for their reuses.
added to the HS aqueous dispersion, and the mixture was sonicated for
30 min. In the later step, FeCl3, 6 H2O (4.38 g dispersed in 50 mL of 3. Results and discussions
distilled water) was slowly added to the HS-pyrrole suspension for 2 h
with constant stirring. As it was expected, the resulting HS-pyrrole-APS 3.1. Catalysts characterization
mixture took a black color due to the oxidative polymerization process.
Finally, the resulted HS@PPy catalyst was washed with distilled water XPS analyzes, of the prepared HS@PANI and HS@PPy, were per­
and ethanol, and then dried in an oven at 60 ◦ C under vacuum for 12 h. formed in order to highlight the catalyst surface chemical functional­
In the same context, the Polypyrrole (PPy) alone was also prepared, as a ities. Thus, after the insertion of HS particles in the polymer’s matrix, the

2
A. Ait El Fakir et al. Applied Catalysis A, General 623 (2021) 118246

wide scan XPS spectra show the presence of the C, N and O chemical polymer composite.
elements on the PANI and PPy surfaces, and the Fe element has appeared To study deeply the structural and the chemical functionalities of the
with a low percentage as resulting from the PANI and PPy layer depo­ prepared catalysts (HS@PANI and HS@PPY), FTIR analyzes were per­
sition on the HS particle surface (Fig. S1). Further, the deconvolved C 1s, formed as depicted in Fig. 3a, b. Hence, the spectrum of PPy alone
N 1s and Fe 2p spectra are represented in Fig. 1. The deconvolved Fe 2p (Fig. 3a) shows that characteristic bands appearing at 1554 and 1466
spectrum for HS@PANI (Fig. 1a) shows two peaks centred mainly at cm− 1 which are attributed to the C=C stretch of the PPy ring. The bands
710.4 and 724.6 eV, corresponding to Fe 2p3/2 and Fe 2p1/2, respec­ located at 1316 and 1180 cm− 1 are due to the =C–H bond. In addition,
tively, and a satellite peak at 719.2 eV corresponds to Fe2O3 [32–34]. the band centered at 910 cm− 1 is due to the asymmetric stretching of the
Further, the C 1s spectrum of HS@PANI (Fig. 1b) can be deconvolved C–N bond, and the bands located at 1045 and 780 cm− 1 are due to the
into four principal components at 283.87, 285.01, 286.29 and 287.76 N–H movement [37,38]. After the modification of HS particles by PPy
eV, which are mainly due to C–C/C–H, C–N/C=N, C–O and C=O polymer, the characteristic peaks of PPy are almost the same with some
respectively [33,34]. The N 1s spectrum of HS@PANI (Fig. 1c) can be differences on the peak position and intensity. Further, the disappear­
deconvoluted into four peaks at 398.70, 399.96, 401.74 and 403.96 eV, ance of the –OH bond after the addition of the polymer was seen, in
which are mainly due to the =N– (quinoid imine group), –NH– (benzoic which there is a strong chemical interaction between the residual –OH in
amine group), =NH+– (charged amine) and –NH+ 2 – (imine) respectively the sand surface and the positively charged amine group of PPy. In
[33,34]. The deconvolution of Fe 2p of HS@PPy is presented in Fig. 1d, addition, upon covering the HS particles by the polymer, the peak in­
and it shows the same peaks of HS@PANI (710.4 and 724.6 eV, corre­ tensity attributed to Fe–O bond in the composite becomes very weak, as
sponding to Fe 2p3/2 and Fe 2p1/2 and the satellite peak at 719.2). In compared to the same bond in the pristine HS. The overall FTIR results
addition, the deconvolution of C 1s for HS@PPy (Fig. 1e) consists mainly confirm the occurrence of chemical interactions between the PPy active
of three components at 288.43, 286.80 and 284.87 eV, corresponding, sites and the Fe2O3 present on the HS surface [39]. On the other hand,
respectively, to C–O, C–N and C–C bonds [35,36]. The N 1s spectrum of The FTIR spectrum of PANI, as presented in Fig. 3b, shows many bands
HS@PPy (Fig. 1f) can be deconvolved into three peaks centred at 398.2, occurring at 1500 and 1555 cm− 1, which are due to the C=C stretching
399.8 and 400.786 eV, and corresponding, respectively, to =Nδ+/C–Nδ+ vibrations of the benzenoid and quinoid group, respectively. The bands
(positively charged nitrogen group), pyrrolic N (–N–H) and nitrogen at 1300 and 1153 cm− 1 are attributed to the stretching C–N bond in the
bond (=N–) [35,36]. aromatic amine group, and the band located at 1123 cm− 1 is due to
To observe the morphologies and the nanostructure natures of HS, bending of the C–H bond located at the plane of the aromatic ring.
PANI, PPy, HS@PANI and HS@PPy materials, we carried out SEM and Moreover, the band centred at 810 cm− 1 is due to the deformation of the
TEM analyzes. Thus, as can be observed in Fig. S2, the HS particles are C–H bond located in the out plane of the disubstituted benzene ring
irregularly shaped, and their sizes show polydispersity. On the other [31]. Likewise, the HS@PANI spectrum shows almost the same peaks
hand, the TEM images show the PANI and PPy morphologies and reveal characteristic of PANI. Further, the disappearance of the –OH bond after
the presence of, respectively, irregular, and spherical nanometric par­ the addition of the polymer is clearly seen in the HS@PANI spectrum, in
ticles. However, upon covering of HS particles by PANI and PPy poly­ which there is a strong chemical interaction between the residual –OH in
mers, it can be seen in Fig. 2a-d, that small HS particle sizes are indeed the surface of sand and the positively charged amine group of PANI.
encapsulated inside the polymer matrix. Moreover, PANI and PPy layers These results confirm the successful incorporation of hematite sand into
attached on the surface of HS particles have been clearly seen. The the polymer matrix. The structures of the prepared catalysts were
polymer layers were thinner. In fact, HS particles are visible as darker further investigated by Raman spectroscopy. The obtained Raman
dots within the brighter polymer chains. The SEM and TEM results spectra of HS@PANI and HS@PPy are shown in Fig. 3c and d. As can be
confirm the successful HS encapsulation by the polymer and the for­ seen, the HS@PPy spectrum (Fig. 3c) indicates peaks occurring at 1575
mation of the molecularly engineered hematite sand/conductive and 1325 cm− 1 which are attributed to the stretching of the backbone

Fig. 1. High-resolution XPS spectra Fe 2p(a), C 1s (b)and N 1s(c) of HS@PANI; Fe 2p(d), C 1s (e) and N 1s(f) of HS@PPy.

3
A. Ait El Fakir et al. Applied Catalysis A, General 623 (2021) 118246

Fig. 2. SEM and TEM images of (a, c) PPy and HS@PPy; (b, d) PANI and HS@PANI.

Fig. 3. FTIR spectra (a, b), Raman spectra (c, d), XRD spectra (e) and TG spectra(f) of HS, HS@PANI and HS@PPy.

C=C, and the C–N, bonds, respectively. In addition, the peak located at nature of the sand used in our work. Hence, the XRD results show
1055 cm− 1 is due to the bending of the C–H bond in the plane. Finally, diffraction peaks occurring at 2θ = 24.19◦ , 33.20◦ , 35.67◦ , 40.90◦ ,
the peak occurring at 971 cm− 1 is due to the bending of the C–H bond 49.49◦ , 57.63◦ corresponding to the iron oxide (Fe2O3) (JCPDS No.
[40]. Likewise, The Raman spectrum of HS@PANI (Fig. 3d) shows the 33-0664) [34,39,42]. The overall XRD data indicate that PPy and PANI
characteristic main peaks of PANI due to the stretching C–N bond cen­ polymers, covering the HS particles, are amorphous, contrary to the HS
tred in 1339 cm− 1, the bending C–H bond in the benzene ring centred mineral which is highly crystallized. To assess the mass content in
1259 cm-1, the bending C–H bond in the quinoid ring centred at 1169 percent, of PANI and PPy in the HS@PANI and HS@PPy catalysts, the
cm− 1, and two characteristic peaks appear at 1500 cm-1 and 1600 cm-1 TGA curves were determined as shown in Fig. 3f. Thus, as can be seen in
corresponding to the stretching of C–C bond located in the benzene ring this figure, the HS@PANI and HS@PPy composites have two major
[41]. In addition, Raman bands characteristic of HS species, having low stages of mass losses. The first mass loss which appears in the temper­
intensities, were also observed with at 610 and 655 cm-1, which may be ature range 50 ◦ C–150 ◦ C, corresponds to the loss of the adsorbed water,
due to the deposition of PANI and PPy layer on the surface of HS par­ and it is expected to be 2 % for both composites. The second mass loss is
ticles. The phase nature and the crystallinity of the HS@PANI and the observed in the ranges 250− 500 ◦ C, and it is indicative of the break­
HS@PPy composites were confirmed by XRD analyzes (Fig. 3e). The down of PANI and PPy chains and the decomposition of the polymer
results indicate that the diffraction patterns of the HS particles, before matrix backbone. This second mass loss is about 15 % for HS@PPy and
and after their modification by PPy and PANI, remain the same, with 19 % for HS@PANI [34,42,43]. After 505 ◦ C, a plateau has been reached
small difference in the peaks Intensities. Therefore, most HS diffraction corresponding to the remaining HS material. Fig. S3 shows the N2
peaks can be perfectly indexed to the Fe2O3 structure, considering the adsorption-desorption isotherms of pure HS, and HS@PANI and

4
A. Ait El Fakir et al. Applied Catalysis A, General 623 (2021) 118246

HS@PPy composites, as measured at − 196 ◦ C, after degassing the the PS to some extent. Surprisingly, when the HS particles were modified
samples at 80 ◦ C for 24 h. The calculated BET surface areas were found by PANI and PPy, increases in the catalytic oxidation efficiency of
to increase from 1.4087 m2/g (bare HS particles) to 4.7078 m2/g and HS@PPy/PS and HS@PANI/PS systems were observed (reaching almost
3.2648 m2/g after in-situ, PANI and PPy polymerization, respectively. 100 % in 60 min). In addition, the corresponding reaction rate constant
is significantly increased from 0.0054 min− 1 for HS alone to 0.053 and
0.07 min− 1 for HS modified with PANI and PPy respectively. These
3.2. Catalytic performance better results are due to the synergistic effect and the combination of
both HS and polymer materials. Initially, the formed polymer layer
The catalytic oxidation performances of the prepared catalysts (PANI or PPy) on the HS surface allowed us a good fixation of the PS on
HS@PPy and HS@PANI, were evaluated for Orang G degradation by the composite surface, as confirmed in the characterization part, and
using PS as oxidant. The adsorption and catalytic performances are enhanced the production of the active species (SO4− , OH and 1O2).
• •

shown in Fig. 4a. The corresponding reaction rate constants are shown Afterwards, based on the oxidizing power of these reactive species, the
in Fig. 4b. Firstly, the adsorption performances of HS@PANI and OG molecules were then totally converted to CO2 and H2O. More pre­
HS@PPy were tested. The results show that only about 15 % and 20 % cisely, the use of conductive polymers, as it was characterized by
OG were removed from water by adsorption on the catalyst, within 60 structures rich in nitrogen atoms located in a conjugated system, pro­
min, by using HS@PPy and HS@PANI, respectively. Secondly, blank vides a good performance PS activation. In addition, the N atoms act as
experiments were carried out by using PS oxidant without catalyst to active sites for the persulfate activation, as resulting from their higher
confirm its principal role, as well as, the resulting poor OG degradation electronegativity in comparison to carbon atoms (χC = 2.55 < χN =
rate. Such experiments have indicated that the PS alone was unable to 3.04). This difference in the electronegativity between the nitrogen and
oxidize OG in aqueous solution. These preliminary experiments allowed the carbon atoms increases the electrons transfer capacity, and the
us to confirm the negligible contribution of the PS alone and the charge density of the positively charged carbon adjacent, which behaves
adsorption catalysts alone during the catalytic performance. After these as an active site for adsorbing persulfate, cleaving hence the O–O bond,
control tests, the OG degradation efficiency, by using the HS/PS system, and generating the ROS [12,44–47]. In addition, Rhodamine B (RhB)
was found to be 30 % indicating that HS particles alone might activate

Fig. 4. Degradation of OG, and corresponding reaction rate constants, by using HS@PANI and HS@PPy (a, b) catalysts; Degradation of BPA and RhB: with HS@PANI
and HS@PPy (c). Effect of tap water matrix, on the OG degradation (d). [Catalyst] =0.2 g L− 1, [OG] =50 mg L− 1, [BPA] =10 mg L− 1, [RhB] =20 mg L− 1, [PS]
=4 mM.

5
A. Ait El Fakir et al. Applied Catalysis A, General 623 (2021) 118246

and Bisphenol A (BPA) were selected as other organic contaminants to performances, the effects of experimental factors on the OG degradation
confirm the performance of catalysts under the same experimental were evaluated in Text S2 and Fig. S5.
conditions. In Fig. 4c the catalytic activity and the reaction rate con­
stants are shown and indicate that both HS@PPy/PS and HS@PANI/PS
3.3. ROS involved in the OG degradation
systems can highly degrade the BPA and RhB after 120 min of contact
time. These results confirm that HS@PANI and HS@PPy can be highly
The radical quenching tests are performed to investigate the main
cleaning up wastewater from a wide range of toxic organic pollutants. In
reactive oxygen species (ROS) during the PS activation by using
addition, to study the OG degradation under real conditions, the effect of
different catalysts. Ethanol (EtOH) and tert-butanol are used as selective
tap water was performed for HS@PANI and HS@PPy systems. As shown
radical scavengers to distinguish the contribution of SO4− and OH

in Fig. 4d, within 60 min, the OG degradation amounts decreased from


radicals in the PS system activation. Moreover, EtOH can deactivate

100 % to 82 % and to 76 %, for HS@PANI and HS@PPy, respectively.


both OH (k = 9.7 × 108 M− 1 s− 1) and SO4− (k = 2.5 × 107 M− 1 s− 1),
• •

These decreases were also confirmed by the degradation rate constants


whereas TBA is used as a radical scavenger for OH (k = 3.8 − 7.6 × 108

which decreased from 0.053 to 0.025 min− 1 for HS@PANI, and from − 1 − 1
M s ) [1,6,7,17,48–50]. As shown in Fig. 5a, the catalytic efficiency
0.070 to 0.027 min− 1 for HS@PPy. The negative influence of the tap
of HS@PANI decreases slightly upon the addition of EtOH and TBA,
water medium on the OG removal rate could be due to the presence of
from 100 % to 69 % and to 88 %, respectively. The reaction rate con­
chloride ions (Cl-) which deplete the radicals with a high level, and also
stants decreased also from 0.053 min− 1 (Control) to 0.017 min− 1 (EtOH)
to the presence of organic matter which decreases the OG degradation
and to 0.03 min− 1 (TBA). However, the catalytic efficiency of HS@PPy
efficiency. The total organic carbon (TOC) analyzes were used to
was sharply reduced when EtOH and TBA were added. Thus, the reac­
confirm the OG mineralization efficiency by using the HS@PPy/PS and
tion rate constants decreased from 0.07 min− 1 (Control) to 0.017 min− 1
HS@PANI/PS systems. As shown in Fig. S4, the TOC removals are about
(EtOH) and to 0.034 min− 1 (TBA), respectively (Fig. 6a). These results
60 % and 57 %, by using HS@PANI and HS@PPy, respectively. These
suggest that SO4− and OH play the primary role in the OG oxidation in
• •

results confirm the catalytic degradation of OG into CO2 and H2O, but
the catalysts/PS systems. In addition to OH and SO4− , the nonradical
• •

also the presence of some intermediate molecules of OG in the medium. 1


specie ( O2) also plays a crucial role with a nonradical oxidation. The
To understand the origin of these excellent Fenton-like catalytic 1
O2 quenching experiments were conducted using NaN3 as a selective

Fig. 5. Radical quenching results by using EtOH, TBA, and NaN3 as the scavengers in (a) HS@PANI/PS and (b) HS@PPy/PS. EPR spectra using (c) PBN and (d)
TEMP, as the spin-trapping agents for the catalyst/PS processes. (Experimental conditions: [Catalyst] =0.2 g L− 1, [OG] =50 mg L− 1, [PS] =4 mM, [EtOH] = [TBA] =
1 M, [NaN3] =0.05 mM, [TEMP] = [BPN] =50 mM).

6
A. Ait El Fakir et al. Applied Catalysis A, General 623 (2021) 118246

Fig. 6. Efficiency of HS, HS@PANI and HS@PPy for the OG degradation as monitored over broad pH ranges (a, b). Stability and reusability of HS, HS@PANI and
HS@PPy (c). ([Catalyst] =0.2 g L− 1, [OG] =50 mg L− 1, [PS] =4 mM) [61–63].

scavenger [6,7]. As shown in Fig. 5a and b, the activities of HS@PANI species in the current systems are SO4− OH from radical pathway and
• •

1
and HS@PPy decrease slightly with the addition of NaN3, and the re­ O2 from nonradical pathway.
action rate decreases from 0.07 to 0.016 min− 1 for HS@PPy, and from
0.053 to 0.019 min− 1 for HS@PANI. It is therefore concluded that 1O2 is
an important ROS during the PS activation in addition to OH and SO4− .
• • 3.4. Identification of surface-active sites and derived reaction mechanism
Moreover, PBN and TEMP trapped EPR tests further confirm the main
reactive oxygen species (ROS). The PBN is an effective scavenger to To confirm the contributions of the HS particles rich in Fe2O3 and the
inhibit OH and SO4− radicals, while TEMP could be used as a scavenger
• • conductive polymers (PANI and PPy) rich in nitrogen, as well as, the
to probe the 1O2 signal [1,3,21,22,45,51–53]. As shown in Fig. 5c and d, synergistic effect during the PS activation, each material was studied
in the absence of the catalysts, and when only the PS is added to the separately.
reaction solutions, no characteristic signals were detected upon the
addition of the spin-trapping agents (PBN, TEMP). However, when 3.4.1. Role of HS core: Improving mediated electron transfer
HS@PANI and HS@PPy are added to the system, typical characteristic The Fe2O3 metallic particles of HS plays a crucial role in enhancing
EPR spectra are detected, which represent the adduct signal PBN-SO4−
• the activating capacity of the PS (i.e. improving the overall electron
and PBN-OH [6,7]. In addition, as shown in Fig. 5d, a characteristic
• transfer mediation capacity of the HS@Polymer). The results confirm
three-line signal as a 1: 1: 1 triplet of TEMP agent was observed in the that the incorporation of the Fe2O3 metallic particles of the HS into the
presence of the catalysts, which could be attributed to the TEMP-1O2 polymer matrices significantly reduced the charge transfer resistance.
adduct. Therefore, these results demonstrate that the main reactive The possible mechanism of the PS activation by the nucleus (HS) has
been proposed for the heterogeneous OG degradation in the

7
A. Ait El Fakir et al. Applied Catalysis A, General 623 (2021) 118246

HS@Polymer/PS system. As shown in Scheme 1, the generation of SO4− suitable for long-lasting and efficiency catalytic reactions. Based on the

radical by Fe2O3 was linked to the redox conversion of the ≡Fe(III)/≡Fe experimental results, the possible mechanisms of the PS activation in the
(II) ring as well as to the involvement of groups − OH on the surface (Eq. HS@PANI and HS@PPy, are directly linked by synergetic effect occur­
1). Subsequently, Fe (II) or Fe(III) from Fe2O3 from the HS could activate ring between the PANI and the PPy shells and the HS as a metal core. The
S2O2− 2−
8 to produce SO4 and SO4 via the equations below. Newly formed PS activation by HS@PANI and HS@PPy was expressed via radical and
• −

Fe(II)− OH co-activates PS with the inherent Fe(II) of Fe2O3 to produce non-radical pathway as confirmed in the previous deactivation results
SO4− and Fe(III)− OH (Eqs. 2 and 3) [54–57]. These reactions result in (Scheme 1). In fact, conducting polymers are the main sources of

the production of OH radicals in two pathways (Eqs. 4 and 5) [54–57]. non-radical species during the catalytic reactions, as resulting from the

As indicted, the role of HS in the HS@PANI and HS@PPy hybrid ma­ electronegativity difference occurring between nitrogen and carbon
terials was observed to facilitate the electron transfer from the conju­ atoms (χC = 2.55 < χN = 3.04). Further, the N sites located in the
gated polymers to the PS. Therefore, the integration of metallic conjugated system facilitate and increase the electrons transfer from the
materials within the conductive nitrogen-rich polymers synergistically adjacent carbon atoms, creating hence positively charged carbons and
determines the chemical potential of the polymer surface, which effec­ leading to strongest active site for adsorption. Furthermore, the presence
tively improve the PS activation. of the N atoms in the polymeric conjugated system creates a strong af­
finity towards the PS facilitating hence the O–O bond cleavage and
Fe(II) + 2H2O → Fe(II)− OH + H3O+ (1) generating the ROS for organic pollutants degradation. In more details,
Fe(II)− OH + S2O2−
8

→ Fe(III)− OH + SO4 + −
SO2−
4 (2) the hydrogen atoms transfer, occurring in the conjugated PANI and PPy
systems, generates sufficient electrons that can highly cleave the O–O
Fe(III)− OH + S2O2− − 2−
(3) bond of S2O2− 8 and release the SO4 radicals. Such reaction processes
• • −
8 → Fe(II)− OH + SO4 + SO4
convert the conjugated –NH2 bond of PANI to =NH+– (Eq. 6), and the
SO4− + H2O → SO2− +
(4)
• •
4 + OH + H
pyrrolic bond (–NH–) of PPy to =Nδ+/=N– (Eq. 7).
− −
SO2− (5)
• •
SO4 + OH → + OH
(6)
4 2− 2−
− NH+ + − •
2 − + S2O8 → =NH − + SO4 + SO4

− NH− + S2O2− δ+ − 2−
(7)

8 → =N /=N− + SO4 + SO4

3.4.2. Effect of polymers coating: Efficiency improvement versus pathway On the other hand, the generated SO4− radicals can react with H2O

switching molecules to yield the hydroxyl radical OH (Eq. 8).


In the case of polymer materials containing nitrogen atoms, the


stability of N atoms is a very important parameter to improve the sta­ SO4− + H2O → SO2− +
(8)
• •
4 + OH + H
bility and in the non-radical pathway in the catalysts/PS systems [31].
Moreover, the hydrogen and electrons transfer reactions occurring,
The loss of nitrogen in the catalysts/PS systems was considered to be a at higher transient oxidation states, between the conjugated PANI and
primary reason for the catalyst stability slow down and catalysts deac­
PPy systems and the S2O28, lead to the break of the O–O bonds of the PS,
tivation [20,58,59]. Therefore, the conducting polymers (PANI and PPy) and the generation of the superoxide radicals O2− (Eqs. 9 and 10). It was

used in this work are characterized by their stable N sites, and are

Scheme 1. Mechanism of the PS activation, on the HS@PANI and HS@PPy surfaces.

8
A. Ait El Fakir et al. Applied Catalysis A, General 623 (2021) 118246

reported that the O2− radicals may be generated as intermediate product between the HS and the conducting polymers catalyst components,

and that the 1O2 may be produced from the oxidation of O2− (Eq. 11) generating hence various reactive oxygen species. The data indicated

[31]. that the nitrogen atoms, located in the conjugated chain of the con­
ducting polymers covering the HS particles, played the main role in the
=NH+− + S2O2− − + 2− − +
(9)

8 + 2OH → − NH2 − + 2SO4 + O2 + 2H adsorption and the activation of the persulfate, improving hence the
=Nδ+/=N− + S2O2− − 2− −
8 + 2OH → − NH− + 2SO4 + O2 + 2H
+ •
(10) organic pollutants degradation. Moreover, these polymeric systems
generate both radical and non-radical species such as SO4− , OH and 1O2.
• •

2 O2− + 2 H2O → 1O2 + H2O2 + 2OH− (11) The use, of PANI and PPy in coating HS particles, was found to prevent

iron leaching and polymer oxidation, avoiding hence secondary pollu­


tion, and offering an effective new strategy to develop green catalysts
with nitrogen stable sites. Finally, the prepared catalysts were efficiently
3.5. Catalyst chemical stability enhancement after the HS modification by recovered after their uses, without significant loss. This work provides
the conducting polymers new insights into the synergetic effect occurring at the interface, be­
tween the HS core and the conducting polymer shell, for the PS acti­
The solution pH is one of the important factors that affect the sta­ vation and the ROS generation. In summary, the approach used in the
bility of the prepared catalysts in the presence of PS, for the wastewater present work provides an innovative method to produce high-quality
treatment [53,57,60]. Therefore, the metal coating on the HS surface by and green catalysts in terms of environmental friendliness, cost effec­
polymers prevents their leaching in the aqueous redox reactions, which tiveness and comparable efficiency.
has been known to be an essential factor responsible for the deactivation
of metal catalysts. To confirm the protective role of these polymer (PANI CRediT authorship contribution statement
and PPy) we studied the activation capacity of PS with HS@PPy and
HS@PANI in a wide pH range (Fig. 6a). As shown in Figure, the OG was Abdellah Ait El Fakir: Writing - original draft, Conceptualization,
almost completely degraded within 60 min in the pH ranges 3–11, Methodology, Data curation, Writing - review & editing. Zakaria Anfar:
indicating that both catalysts showed good performance for the activa­ Methodology, Data curation, Writing - review & editing. Abdallah
tion of PS over a wide range of pH solution. However, when HS alone Amedlous: Methodology, Writing - review & editing. Asma Amjlef:
was used for the activation of PS, the higher removal rates were Writing - review & editing. Salaheddine Farsad: Writing - review &
observed under acidic conditions, in contrast, when applying alkaline editing. Amane Jada: Project administration, Supervision, Validation,
conditions (i.e. between pH 8 and 11), a drastic reduction in removal Writing - review & editing. Noureddine El Alem: Project administra­
efficiency was detected. These results suggest that the polymers act as tion, Supervision, Writing - review & editing.
protective layers, and/or they are related to the positive surface charge
of both polymers as well as the protonation of nitrogen groups at the Declaration of Competing Interest
surface, which facilitated the adsorption of the negatively charged of
S2O2−8 . The overall data indicate that the HS@PANI and HS@PPy cata­ The authors declare that they have no known competing financial
lysts exhibit stable catalytic capacities over a wide pH range from 3 to 11 interests or personal relationships that could have appeared to influence
(Fig. 6b), which is by far among the widest pH range achieved by het­ the work reported in this paper.
erogeneous systems.
To clearly show the role of our molecular modification strategies by Acknowledgments
conductive polymers, we monitored the decrease in the PS activation
efficiency of the pristine HS material, and the HS@PANI and the This work was supported by Franco-Moroccan cooperation frame­
HS@PPy catalysts during the reuse of 5 cycles during catalytic degra­ work under both grant research projects APUR 2019 and CEDocs 2018
dation process (Fig. 6c). The SEM images and infrared spectra of managed by Laboratory of Materials & Environment (LME), Ibn Zohr
HS@PANI and HS@PPy, after the recycling test remain almost the same University, Agadir – Morocco and the Institute of Materials Science of
with fresh HS@PANI and HS@PPy, which also prove that HS@PANI and Mulhouse (IS2M), Haute Alsace University, Mulhouse - France. We
HS@PPy are stable during the degradation process (Fig. S6). However, thank, VAULOT Cyril (IS2M), VIDAL Loïc (IS2M), FIOUX Philippe
the pristine HS showed a greater decrease in performance, in compari­ (IS2M), MORLET - SAVARY Fabrice (IS2M) and GREE Simon (IS2M) for
son to the HS@PANI and the HS@PPy catalysts due to iron leaching the analyses of samples by BET, TEM, SEM, XPS, EPR and FTIR-Raman,
after each cycle of use. Therefore, when using coated sand, the polymers respectively.
play a protective role in decreasing the leaching of iron during the
catalytic reaction. These results were well demonstrated in Fig. 6c. As
Appendix A. Supplementary data
can be seen in this figure, during the last cycle, the efficiency of
HS@PPy/PS and HS@PANI/PS decreased only 18 and 20 %, respec­
Supplementary material related to this article can be found, in the
tively. As the non-radical activation of persulfate involves insignificant
online version, at doi:https://doi.org/10.1016/j.apcata.2021.118246.
consumption of the material, the constant decrease in the catalytic ef­
ficiency of HS@PANI and HS@PPy during the catalyst reuse could be
References
attributed to the polymer surface and structural changes, including the
deactivation of nitrogen atoms, iron leaching, adsorption of in­ [1] W. Ren, G. Nie, P. Zhou, H. Zhang, X. Duan, S. Wang, Environ. Sci. Technol. 54
termediates during the degradation of organic molecules and coverage (2020) 6438–6447.
of active sites on the surface of the catalyst. [2] J. Yu, H. Feng, L. Tang, Y. Pang, G. Zeng, Y. Lu, H. Dong, J. Wang, Y. Liu, C. Feng,
J. Wang, B. Peng, S. Ye, Prog. Mater. Sci. 111 (2020), 100654.
[3] P. Shao, S. Yu, X. Duan, L. Yang, H. Shi, L. Ding, J. Tian, L. Yang, X. Luo, S. Wang,
4. Conclusions Environ. Sci. Technol. 54 (2020) 8464–8472.
[4] B. Xing, J. Dong, G. Yang, N. Jiang, X. Liu, J. Yuan, Appl. Catal. A Gen. 602 (2020),
117714.
In this study, we have investigated the effect of molecularly modified [5] X. Lin, Y. Ma, J. Wan, Y. Wang, Y. Li, Appl. Catal. A Gen. 589 (2020), 117307.
hematite sand/conductive polymer nanocomposites for ultrafast cata­ [6] Z. Anfar, A.A. El Fakir, M. Zbair, Z. Hafidi, A. Amedlous, M. Majdoub, S. Farsad,
lytic degradation of organic pollutants, by using persulfate as an acti­ A. Amjlef, A. Jada, N. El Alem, Chem. Eng. J. 405 (2021), 126660.
[7] Z. Anfar, A. Ait El Fakir, H. Ait Ahsaine, M. Zbair, S. Farsad, F. Morlet-Savary,
vator. The encapsulation of HS particles into nitrogen-rich polymer A. Jada, N. El Alem, New J. Chem. 44 (2020) 9391–9401.
(PANI and PPy) matrices has led to synergistic catalytic effect occurring [8] L.W. Matzek, K.E. Carter, Chemosphere 151 (2016) 178–188.

9
A. Ait El Fakir et al. Applied Catalysis A, General 623 (2021) 118246

[9] G. Fang, T. Zhang, H. Cui, D.D. Dionysiou, C. Liu, J. Gao, Y. Wang, D. Zhou, [38] A. El Jaouhari, S. Ben Jadi, A. El Guerraf, M. Bouabdallaoui, Z. Aouzal, E.
Environ. Sci. Technol. (2020) acs.est.0c06091. A. Bazzaoui, J.I. Martins, M. Bazzaoui, Synth. Met. 245 (2018) 237–244.
[10] N. Jiang, H. Xu, L. Wang, J. Jiang, T. Zhang, Environ. Sci. Technol. 54 (2020) [39] F.A. Harraz, A.A. Ismail, S.A. Al-Sayari, A. Al-Hajry, J. Photochem. Photobiol. A:
14057–14065. Chem. 299 (2015) 18–24.
[11] G.-X. Huang, C.-Y. Wang, C.-W. Yang, P.-C. Guo, H.-Q. Yu, Environ. Sci. Technol. [40] A. El Jaouhari, A. El Asbahani, M. Bouabdallaoui, Z. Aouzal, D. Filotás, E.
51 (2017) 12611–12618. A. Bazzaoui, L. Nagy, G. Nagy, M. Bazzaoui, A. Albourine, D. Hartmann, Synth.
[12] X. Chen, X. Duan, W.-D. Oh, P.-H. Zhang, C.-T. Guan, Y.-A. Zhu, T.-T. Lim, Appl. Met. 226 (2017) 15–24.
Catal. B Environ. 253 (2019) 419–432. [41] P.D. Bui, H.H. Tran, F. Kang, Y.-F. Wang, T.M. Cao, S.-J. You, N.H. Vu, V. Van
[13] P. Sun, H. Liu, M. Feng, L. Guo, Z. Zhai, Y. Fang, X. Zhang, V.K. Sharma, Appl. Pham, ACS Appl. Nano Mater. 1 (2018) 5786–5794.
Catal. B Environ. 251 (2019) 335–345. [42] L. Wang, H. Yang, X. Liu, R. Zeng, M. Li, Y. Huang, X. Hu, Angew. Chemie Int. Ed.
[14] X. Duan, Z. Ao, H. Sun, S. Indrawirawan, Y. Wang, J. Kang, F. Liang, Z.H. Zhu, 56 (2017) 1105–1110.
S. Wang, ACS Appl. Mater. Interfaces 7 (2015) 4169–4178. [43] J. Liu, W. Zhou, L. Lai, H. Yang, S. Hua Lim, Y. Zhen, T. Yu, Z. Shen, J. Lin, Nano
[15] X. Duan, Z. Ao, H. Zhang, M. Saunders, H. Sun, Z. Shao, S. Wang, Appl. Catal. B Energy 2 (2013) 726–732.
Environ. 222 (2018) 176–181. [44] S. Liu, C. Lai, B. Li, C. Zhang, M. Zhang, D. Huang, L. Qin, H. Yi, X. Liu, F. Huang,
[16] P. Shao, J. Tian, F. Yang, X. Duan, S. Gao, W. Shi, X. Luo, F. Cui, S. Luo, S. Wang, X. Zhou, L. Chen, Chem. Eng. J. 384 (2020), 123304.
Adv. Funct. Mater. 28 (2018), 1705295. [45] K. Zhu, Q. Bin, Y. Shen, J. Huang, D. He, W. Chen, Chem. Eng. J. 402 (2020),
[17] X. Duan, H. Sun, Y. Wang, J. Kang, S. Wang, ACS Catal. 5 (2015) 553–559. 126090.
[18] P. Hu, H. Su, Z. Chen, C. Yu, Q. Li, B. Zhou, P.J.J. Alvarez, M. Long, Environ. Sci. [46] S. Ye, G. Zeng, X. Tan, H. Wu, J. Liang, B. Song, N. Tang, P. Zhang, Y. Yang,
Technol. 51 (2017) 11288–11296. Q. Chen, X. Li, Appl. Catal. B Environ. 269 (2020), 118850.
[19] X. Huo, P. Zhou, J. Zhang, Y. Liu, X. Cheng, Y. Liu, W. Li, Y. Zhang, J. Hazard. [47] L. Tang, Y. Liu, J. Wang, G. Zeng, Y. Deng, H. Dong, H. Feng, J. Wang, B. Peng,
Mater. 391 (2020), 122055. Appl. Catal. B Environ. 231 (2018) 1–10.
[20] H. Li, C. Shan, B. Pan, Environ. Sci. Technol. 52 (2018) 2197–2205. [48] D. Ding, S. Yang, X. Qian, L. Chen, T. Cai, Appl. Catal. B Environ. 263 (2020),
[21] S. Wang, Y. Liu, J. Wang, Environ. Sci. Technol. 54 (2020) 10361–10369. 118348.
[22] E.-T. Yun, S.-W. Park, H.J. Shin, H. Lee, D.-W. Kim, J. Lee, Appl. Catal. B Environ. [49] W. Ma, N. Wang, Y. Du, P. Xu, B. Sun, L. Zhang, K.-Y.A. Lin, ACS Sustain. Chem.
279 (2020), 119360. Eng. 7 (2019) 2718–2727.
[23] W. Ma, N. Wang, Y. Du, T. Tong, L. Zhang, K.-Y. Andrew Lin, X. Han, Chem. Eng. J. [50] H.-Y. Gao, C.-H. Huang, L. Mao, B. Shao, J. Shao, Z.-Y. Yan, M. Tang, B.-Z. Zhu,
356 (2019) 1022–1031. Environ. Sci. Technol. 54 (2020) 14046–14056.
[24] L. Xu, B. Fu, Y. Sun, P. Jin, X. Bai, X. Jin, X. Shi, Y. Wang, S. Nie, Chem. Eng. J. 400 [51] J. Yu, L. Tang, Y. Pang, G. Zeng, H. Feng, J. Zou, J. Wang, C. Feng, X. Zhu,
(2020), 125870. X. Ouyang, J. Tan, Appl. Catal. B Environ. 260 (2020), 118160.
[25] Y. Long, S. Li, Y. Su, S. Wang, S. Zhao, S. Wang, Z. Zhang, W. Huang, Y. Liu, [52] J. Miao, W. Geng, P.J.J. Alvarez, M. Long, Environ. Sci. Technol. 54 (2020)
Z. Zhang, Chem. Eng. J. 404 (2021), 126499. 8473–8481.
[26] Y. Zhou, Y. Zhang, X. Hu, J. Colloid Interface Sci. 575 (2020) 206–219. [53] C. Liu, L. Liu, X. Tian, Y. Wang, R. Li, Y. Zhang, Z. Song, B. Xu, W. Chu, F. Qi,
[27] N. Zhou, J. Zu, L. Yang, X. Shu, J. Guan, Y. Deng, D. Gong, C. Ding, M. Zhong, A. Ikhlaq, Appl. Catal. B Environ. 255 (2019), 117763.
J. Colloid Interface Sci. 563 (2020) 197–206. [54] F. Meng, M. Song, B. Song, Y. Wei, Q. Cao, Y. Cao, Chemosphere 243 (2020),
[28] P. Duan, Y. Qi, S. Feng, X. Peng, W. Wang, Y. Yue, Y. Shang, Y. Li, B. Gao, X. Xu, 125322.
Appl. Catal. B Environ. 267 (2020), 118717. [55] H. Zheng, J. Bao, Y. Huang, L. Xiang, Faheem, B. Ren, J. Du, M.N. Nadagouda, D.
[29] D. Ding, S. Yang, L. Chen, T. Cai, Chem. Eng. J. 392 (2020), 123725. D. Dionysiou, Appl. Catal. B Environ. 259 (2019), 118056.
[30] J. Ye, J. Dai, C. Li, Y. Yan, Chem. Eng. J. (2020), 127805. [56] S. Guo, H. Wang, W. Yang, H. Fida, L. You, K. Zhou, Appl. Catal. B Environ. 262
[31] A. Ait El Fakir, Z. Anfar, A. Amedlous, M. Zbair, Z. Hafidi, M. El Achouri, A. Jada, (2020), 118250.
N. El Alem, Appl. Catal. B Environ. 286 (2021). [57] L. Chen, Y. Huang, M. Zhou, K. Xing, W. Lv, W. Wang, H. Chen, Y. Yao,
[32] G. Sun, B. Dong, M. Cao, B. Wei, C. Hu, Chem. Mater. 23 (2011) 1587–1593. Chemosphere 250 (2020), 126300.
[33] X.-F. Lu, X.-Y. Chen, W. Zhou, Y.-X. Tong, G.-R. Li, ACS Appl. Mater. Interfaces 7 [58] Y. Gao, Y. Zhu, Z. Chen, C. Hu, ACS ES&T Eng. (2020), acsestengg.0c00039.
(2015) 14843–14850. [59] L. Peng, Y. Shang, B. Gao, X. Xu, Appl. Catal. B Environ. 282 (2021), 119484.
[34] Y. Zhang, D. Jiang, Y. Wang, T.C. Zhang, G. Xiang, Y.-X. Zhang, S. Yuan, Ind. Eng. [60] H. Chi, J. Wan, Y. Ma, Y. Wang, M. Huang, X. Li, M. Pu, J. Hazard. Mater. 398
Chem. Res. 59 (2020) 7554–7563. (2020), 123024.
[35] C. Wang, M. Yang, L. Liu, Y. Xu, X. Zhang, X. Cheng, S. Gao, Y. Gao, L. Huo, [61] P. Duan, T. Ma, Y. Yue, Y. Li, X. Zhang, Y. Shang, B. Gao, Q. Zhang, Q. Yue, X. Xu,
J. Colloid Interface Sci. 560 (2020) 312–320. Environ. Sci. Nano 6 (2019) 1799–1811.
[36] K. Le, M. Gao, D. Xu, Z. Wang, G. Wang, W. Liu, F. Wang, J. Liu, Dalton Trans. 49 [62] S.-F. Jiang, L.-L. Ling, W.-J. Chen, W.-J. Liu, D.-C. Li, H. Jiang, Chem. Eng. J. 359
(2020) 9701–9709. (2019) 572–583.
[37] J. Tabačiarová, M. Mičušík, P. Fedorko, M. Omastová, Polym. Degrad. Stab. 120 [63] S. Liu, Z. Zhang, F. Huang, Y. Liu, L. Feng, J. Jiang, L. Zhang, F. Qi, C. Liu, Appl.
(2015) 392–401. Catal. B Environ. 286 (2021), 119921.

10

You might also like