You are on page 1of 10

Chemical Engineering Journal 371 (2019) 404–413

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

The enhanced catalytic potential of sulfur-doped MgO (S-MgO) T


nanoparticles in activation of peroxysulfates for advanced oxidation of
acetaminophen

Farzaneh Fanaeia, Gholamreza Moussavia, , Varsha Srivastavab, Mika Sillanpääb
a
Department of Environmental Health Engineering, Faculty of Medical Sciences, Tarbiat Modares University, Tehran, Iran
b
Department of Green Chemistry, School of Engineering Science, Lappeenranta University of Technology, Sammonkatu 12, FI-50130 Mikkeli, Finland

H I GH L IG H T S

• PMS was efficiently activated with S-MgO nanoparticles.


• S-MgO/PMS process generated a high amount of hydroxyl and sulfate radicals.
• AS-MgO/PMS
very high ACT oxidation rate was obtained in S-MgO/PMS process.
• S-MgO/PMS process transformed ACT molecules to harmless compounds.
• process is an efficient AOP for removing emerging contaminants.

A R T I C LE I N FO A B S T R A C T

Keywords: The peroxymonosulfate (PMS) and peroxydisulfate (PDS) were activated using plain and S-doped MgO (S-MgO)
AOP for oxidation of acetaminophen (ACT) in the contaminated water. The findings indicated that S-MgO was more
Catalytic activation efficient than plain MgO in activating the oxidants and that PMS was much better activated than PDS using S-
Peroxymonosulfate MgO. The complete degradation and significant mineralization (up to 62.4%) of 50 mg/L ACT could be achieved
Peroxydisulfate
in the S-MgO/PMS process using 1.4 mM PMS within a reaction time of 30 min and 60 min, respectively. Both
Emerging contaminants
SO·4− and HO· species were simultaneously generated in the S-MgO/PMS process with the contributions of 73.8%
Pharmaceuticals
and 26.2%, respectively, in the ACT removal attained in the process. ACT removal in the S-MgO/PMS process
was not considerably affected by the solution pH between 4 and 10 and the maximum ACT removal of 93.4%
was achieved at the neutral condition and optimum PMS to catalyst ratio of 0.07 mM L/g. The performance of S-
MgO/PMS process for removal of ACT was not affected by the presence of conventional anions in natural waters.
The pathway of ACT degradation in the S-MgO/PMS process was proposed based on the Liquid chromato-
graphy–mass spectrometry (LC/MS) analysis of the effluent. Accordingly, the S-MgO is a very active catalyst to
activate the PMS for simultaneous generation ofSO·4− and HO· reactive species hence S-MgO/PMS process is an
emerging AOP for high rate degrading the pharmaceutical compounds present in contaminated waters.

1. Introduction sulfate radical (SO·4− , E0 = 2.6 V ) based AOPs, mainly due to the high
oxidation potential of SO·4−, the higher lifetime of SO·4− compared to
Pharmaceuticals are one of the main groups of recalcitrant com- hydroxyl radicals (%OH) [2], simultaneously generation of both SO·4−
pounds released to the environment through the municipal wastewater and •OH as well as high efficiency in the degradation of diverse water
and the pharmaceutical manufacturing plants. Advanced oxidation contaminants [3–5]. SO·4− is generated mainly through activation of
processes (AOPs) are an emerged class of advanced technologies widely peroxysulfates (peroxymonosulfate (PMS) or peroxydisulfate (PDS)) by
used for the elimination of recalcitrant and synthetic organic com- heat [6,7], UV radiation [8,9], metals [10–14], metal oxides [15–18].
pounds from the contaminated water, the wastewater and the industrial Based on the published studies, minerals and metal oxides are het-
effluents [1]. Considerable attention has been paid recently on the erogeneous catalysts capable of efficiently activating PMS and PDS to


Corresponding author.
E-mail address: moussavi@modares.ac.ir (G. Moussavi).

https://doi.org/10.1016/j.cej.2019.04.007
Received 8 January 2019; Received in revised form 31 March 2019; Accepted 2 April 2019
Available online 02 April 2019
1385-8947/ © 2019 Elsevier B.V. All rights reserved.
F. Fanaei, et al. Chemical Engineering Journal 371 (2019) 404–413

generate the reactive oxidizing species of SO·4− and •OH. The previous composition (pH and ACT concentration) was transferred to the reactor
studies have examined PDS activation with different metals and metal into which the predetermined amount of PMS or PDS and of catalyst
oxides including Fe2O3-Cu2O [10], Fe0 [11,12], Fe2+ [14], iron-based was added. Then the suspension was magnetically stirred at 150 rpm for
bimetals and trimetals [13], reduced graphene oxide–elemental silver/ the specific time period. At the end of the reaction period, the sus-
magnetite nanohybrids [18], etc. for the degradation of pharmaceutical pension was sampled and 1 mL of sodium thiosulfate solution (1 M) was
compounds. In addition, PMS was activated with perovskite [15], Fe3O4 immediately added to the sample to quench the oxidation reaction.
[16] and MnFe2O4 and CoFe2O4 [17] for the degradation of pharma- Finally, the quenched sample was filtered through a 0.22 µm filter and
ceutical compounds. These studies indicate that the activation of PMS the filtrate was analyzed for the target parameters. The effect of initial
and PDS with metal oxides for the generation of SO·4− is a promising solution pH, type of oxidant (PDS and PMS), PMS to S-MgO, reaction
AOP efficient for decomposition of water toxic contaminants. time, and inorganic (chloride, bicarbonate, phosphate, nitrate) and
In order to improve the SO·4−-based AOPs in terms for mineralization organic (tert-butanol and ethanol acid) radical scavengers was in-
extent of such recalcitrant compounds as pharmaceuticals, sustain- vestigated on ACT degradation. The concentration molar ratio of sca-
ability, and reaction stoichiometric efficiency (% RSE), finding a more venger to ACT in the radical scavenging tests was kept 1:1. The control
active catalyst is required and thereby the activation potential of other experiments were conducted with the catalyst and the oxidants (PMS
agents needs to be investigated. Due to its unique features, recently a and PDS) alone under the same reaction conditions. Initial solution pH
considerable focus has been paid on using magnesium oxide (MgO) as a was adjusted at the desired value using 0.1 M NaOH or HCl. In order to
catalyst in the water and wastewater treatment field. For instance, re- avoid scavenging the reactive radicals by buffer species such as bi-
cently a new sulfur-doped MgO (S-MgO) was prepared which showed carbonate and phosphate, no buffer was used in the preparation of re-
high catalytic activity in the ozonation of acetaminophen (ACT) in action solutions. All experiments were conducted in duplicate and the
aqueous solution [19]. Due to the formation of surface defects in the results were reported as mean value ± SD (standard deviation).
MgO structure upon doping with S [19], the S-MgO may be an attrac-
tive metal oxide catalyst for the activation of PMS and PDS and thus for 2.2. Analysis
the generation of an increased amount of reactive oxidative species
mainly SO·4− and %OH. The X-ray photoelectron spectroscopy (XPS) analysis of MgO and S-
Accordingly, the present study aimed at comparing the efficacy of MgO was examined by (ESCALAB 250) with an Al-K x-ray source of
plain MgO and S-MgO in activating PMS and PDS for the degradation 1486.6 eV for the investigation of binding energies of elements and
and the mineralization of ACT as a model of emerging water con- surface composition of catalysts. The photoluminescence (PL) mea-
taminants. ACT (C8H9NO2) is an antipyretic pharmaceutical, which is surements of the MgO and S-MgO nanoparticles were conducted via
easily accessible around the world. Over 50% of ingested ACT is ex- FL3-TCSPC. The aqueous samples taken from the reactor were filtered
creted unchanged from the body upon ingestion and find its way to the through 0.2 μm cellulose acetate syringe filter and the filtrate was
wastewater resulted in the contamination of the receiving water bodies. analyzed for the target parameters. The concentration of ACT in the
The PS activation was examined by using plain MgO and S-MgO na- reaction solutions before and after the reaction was determined by
noparticles under different experimental conditions of the solution pH, HPLC (Agilent Co.) equipped with a UV detector at 242 nm. A reversed-
the catalyst to PS ratio, ACT concentration, and the reaction time for phase C18 column (Eclipse plus C18 column; 3.5 µm, 4.6 × 100 mm)
the degradation and mineralization of ACT. The generation of radicals was used in the HPLC. The mobile phase was a mixture of phosphate
was indicated using the related scavengers and confirmed by the EPR buffer (pH = 4.8) and acetonitrile at a volumetric ratio of 85:15 in-
analysis. The kinetics of degradation and mineralization of ACT was jected at the flow rate of 1 mL/min. The concentration of total organic
evaluated in the S-MgO/PMS process under their optimum experi- carbon (TOC) residual was measured with a TOC analyzer (Shimadzu,
mental conditions. TOC analyzer –VCSH model) under optimum experimental conditions.
The sample volume required for each TOC analysis was 25 mL, and two
2. Materials and methods measurements were performed for each sample and the average was
reported. The PDS and PMS residual concentrations were measured by
2.1. Materials and experimental procedures the iodometric titration with sodium thiosulfate as titrant with a de-
tection limit below 0.1 mg/L. The concentration of sulfate ions was
The pure ACT powder was obtained from a local company. The determined using the turbidimetric method as described in standard
stock solution of ACT was prepared by dissolving 0.5 g ACT powder in methods [20]. The pH was determined by a Jenway Co. selective
per liter of bidistilled water and stored at 4 °C. The reaction solution electrode. The concentration of PMS was measured using the iodo-
was prepared by dissolving the aliquots of stock solution by distilled metric titration with standardized sodium thiosulphate. The main in-
water. All chemicals used in the present study were of the analytical termediates of ACT degradation in the S-MgO/PMS process were
grade. PMS (2KHSO5·KHSO4·K2SO4) was purchased from Aldrich Co. identified using liquid chromatography–mass spectrometry (LC/MS,
and used as received. The amount of PMS used in the tests was calcu- 2010 A, Shimadzu) equipped with an Eclipse Atlantis T3, C18 column
lated based on(H3 O18 S4 )5 −. The plain MgO and S-MgO catalysts used in (2.1 × 100 mm, 3 µ particle size) in the ambient temperature using the
this work were prepared as detailed in our previous study [19]. Sodium Quadrupole mass analyzer. The mobile phase consisted of 60% acet-
dodecyl sulfate (SDS) used as a capping agent and the sulfur source for onitrile (with 0.1% Formic Acid) and 40% water (with 0.1% formic
S-MgO. Various amount of SDS was used to prepare MgO doped with acid) injected at a flow rate of 0.3 mL/min. The quadrupole mass
different percentages of S between 0 and 3%. The S-MgO with S content spectrometer was operated under the following conditions: Gas nebu-
of 2% as an optimum catalyst was a semi-spherical and flaky-shaped lizer: N2, Probe Volt: 4 kV, Detection gain: 1.8 kV, CDL Volt: 25 V, CDL
mesoporous nanoparticle, had a BET surface area of 257.3 m2/g, an and Block temperature: 250 °C and Flow gas: 1.2 L/min. The spectro-
average crystallite size of 23.6 nm and an average particle size of meter scanning was conducted for m/z ranging between 50 and 500.
13.6 nm, a pHpzc of 10.9, with a high density of active surface func- The injection sample volume for the chromatographic separation was
tional groups [19]. It has been shown that the S-MgO is a stable and 20 µL.
durable material that its catalytic potential is relatively preserved The reactive oxidative species (ROSs) generated by the MgO and S-
during the several reuse times. MgO catalysts were determined by electron paramagnetic resonance
The ACT oxidation tests by S-MgO/PMS and S-MgO/PDS were (EPR) technique (EPR- CMS 8400). Two different spin trapping agents
conducted in batch mode in a 100 mL glass column as reactor at room 5,5-Dimethyl-1-Pyrroline N-Oxide (DMPO) and 2, 2, 6, 6-tetra-
temperature. For each test, 50 mL of ACT reaction solution with specific methylpiperidinyloxyl (TEMP) spin trapping agents were used for ROS

405
F. Fanaei, et al. Chemical Engineering Journal 371 (2019) 404–413

detection [15]. The EPR analysis was carried out by adjusting Power
attenuation: 20 dB, Sweep time: 80 s, Sweep width: 150,00 G, Mod.
amplitude: 1000 mG and Microwave frequency: 9438,80 MHz using
TEMP (100 mM) and DMPO (100 mM).

2.3. Performance indices

The oxidation of ACT in the S-MgO/PS process was evaluated based


on ACT degradation (Eq. (1)) and mineralization (Eq. (2)) percentages.
ACT0 − ACTt
ACT degradation (%) = × 100
ACT0 (1)
where ACT0 and ACTt demonstrate the concentrations of ACT before
and after the reaction, respectively.
The degree of ACT mineralization in the S-MgO/PMS process was Fig. 2. PL of MgO and S-MgO.
determined from Eq. (2) by measuring the concentration of TOC in the
solution before (TOC0) and after (TOCt) the reaction. The degree of ACT comparison, 43.8% and 95.6% of ACT could be removed in the S-MgO/
mineralization was then calculated from Eq. (2). PDS and S-MgO/PMS processes, respectively. It was found that both
TOC0 − TOCt PMS and PDS were better activated by S-MgO than plain MgO.
ACT mineralization (%) = × 100 Fig. 2 depicts the PL spectra of MgO and S-MgO nanoparticles. A
TOC0 (2)
broad PL peak observed for MgO and S-MgO nanoparticles at the wa-
The kinetics of ACT degradation and mineralization were evaluated velength around 550 nm is related to the defects as oxygen vacancy
by fitting the pseudo-first order (PFO) reaction model (Eqs. (3) and (4)) [22]. It is seen in Fig. 2 that the peak intensity of MgO increased con-
with the related experimental results. siderably upon doping with S indicating the development of defect
C centers as oxygen vacancies in the S-MgO. The formation of surface
ln ⎛ t ⎞ = −k obst
⎜ ⎟
defects favor the valence band and charge separation and enhanced the
⎝ C0 ⎠ (3)
transfer rates of photogenerated electron and hole resulting in the more
dC degradation of ACT [23]. Therefore, the greater PDS/PMS activation
robs = − = k obsC
dt (4) potential of S-MgO than that of plain MgO is related to the increased
surface defects as oxygen vacancies in the S-MgO upon S-doping due to
where C0 and Ct are the ACT or TOC concentrations at the beginning
the substitution of O by S atoms in MgO lattice upon doping with sulfur
and the time t of the reaction, respectively, kobs is the PFO reaction rate
[19,22]. The effect of different S content between 0 and 3% in the MgO
constant and robs is the PFO reaction rate. The RSE defined as the
was evaluated on the catalytic activity of the prepared material in ac-
number of ACT moles degraded over the number of PMS moles con-
tivating PMS and the results showed that the ACT degradation was
sumed during a specific reaction time [21] was calculated for the S-
21.5, 68.3, 100 and 80.5% for 0, 1, 2 and 3% of S-doped content (Fig.
MgO/PMS process conducted under the optimum experimental condi-
S1). Therefore the S-MgO contained 2% S was used as a catalyst in the
tion.
rest of the experiments and characterization.
Figs. 3a,b shows the XPS analysis results of MgO and S-MgO cata-
3. Results and discussion
lysts. In MgO (Fig. 3a), binding energies at 1306.6 eV and 531.7 eV are
attributable to Mg1s and O1s respectively. XPS clearly showed the
3.1. Catalytic potential of MgO and S-MgO in PDS and PMS activation
presence of S in modified catalyst S-MgO (Fig. 3b). No sulfur specific
XPS signal was observed in MgO catalyst. For both catalysts, the range
Fig. 1 shows the ACT removal in the different processes under the
of binding energy for O1s spectra was 531.7–532.3 eV, which is the
selected experimental conditions. The ACT removal in the presence of
characteristic of metal oxides. C1s (289.8 eV) is taken as the reference.
MgO and S-MgO was around 1% and 2.9%, respectively. When PDS and
The presence of another band of C1s in S-MgO sample indicates carbon
PMS were added into the reaction medium containing MgO, the ACT
contamination due to CO32− species [24]. The binding energies
removal efficiency increased to 8.2% and 21.5%, respectively. In
1303.0 eV for Mg ls is in good agreement with the reported studies
[25,26]. Presence of sulfur was evidenced by S2p peak at 169.4 eV,
which can be attributed to the positive oxidation state of sulfur as a
sulfate (SO42–) groups [27]. It is reported that the broad peak of O1s at
∼528 eV represent the lattice oxygen, ∼530.2 eV for surface oxygen,
∼531.1 eV for defect oxygen and ∼532.3 eV for the adsorbed oxygen
[15]. In S-MgO it was observed that O1s peak position is 532.3 eV is due
to oxygen defect on the surface. Oxygen vacancy acted as active centers
giving S-MgO a high catalytic activity [19,22,28] for activation of PDS/
PMS.
Higher ACT removal in the S-MgO/PMS than that in the S-MgO/PDS
indicates that PMS could be faster oxidized than PDS by S-MgO, which
is attributed to the asymmetric molecular structure of PMS [29]. In-
deed, PDS is a symmetric compound and SO·4− is generated (Eq. (5))
when activated with S-MgO, whereas PMS is asymmetrical around the
peroxide bond which is broken down by S-MgO yielding both SO·4− and
HO· oxidative radicals (Eq. (6)) [30] work together for efficient de-
gradation of ACT molecules.
Fig. 1. Removal of ACT in different processes (ACT = 50 mg/L; catalyst = 1 g/
L; PDS = 0.5 g/L; PMS = 0.5 g/L; solution pH = 7; reaction time = 30 min).

406
F. Fanaei, et al. Chemical Engineering Journal 371 (2019) 404–413

Fig. 3a. XPS analysis of MgO catalyst.

Fig. 3b. XPS analysis of S-MgO catalyst.

407
F. Fanaei, et al. Chemical Engineering Journal 371 (2019) 404–413

S−MgO
S2 O82 − → 2SO·4− (5) ethanol, respectively. The reduction in PFO reaction rate constant of
ACT removal in the presence of both tert-butanol and ethanol clearly
HSO−
S−MgO
→ SO·4− + HO· (6) indicates that both SO·4− and HO· were generated in the S-MgO/PMS
5
process but with the different contribution in the removal attained. The
In addition, (H3 O18 S4 )5 −
used a source of PMS is dissociated in water reaction rate constant of tert-butanol and ethanol with HO· are
to two molecules of HSO− −
5 and a molecule of HSO4 (Eq. (7)). When S- (3.8–7.6) × 108 M−1s−1 and (1.2–2.8) × 109 M−1s−1, respectively,
MgO was added to the reaction medium, these molecules were de- while those react with SO·4− at the rate constants of
composed to SO·4−. Therefore, a greater amount of oxidative radicals was (4.0–9.4) × 105 M−1s−1 and (1.6–7.7) × 107 M−1s−1, respectively [2].
generated per molecule of (H3 O18 S4)−5 (Eqs. (6) and (8))) than that of These reaction rate constants indicate that tert-butanol selectively re-
S2 O−8 2 resulted in achieving higher ACT degradation percentages [4]. acts with HO· around 1000 times faster than with SO·4−, whereas ethanol
(H3 O18 S4 )5 − → 2HSO5− + HSO−4 + SO24− (7) reacts with both SO·4− and HO· at the relatively close rate. Accordingly,
much greater reduction of ACT removal in the presence of ethanol than
HSO−4 + HO· → SO·4− + H2 O (8) that at the presence of tert-butanol clearly reveals that although both
SO·4− and HO· were generated and involved in the ACT degradation re-
The ACT molecules were then reacted with the generated radicals
actions, SO·4− played a dominant role in ACT degradation under the
(SO·4− and/or HO·) and were degraded into the products as simply shown
selected conditions. The contributions of SO·4− (Eq. (11)) [10] and HO·
in Eqs. ((9) and (10)).
(Eq. (12)) in the ACT removal process were calculated to be 80.3% and
SO·4− + ACT → products (9) 19.7%, respectively. Zhang et al. [10] reported a contribution of 79.9%
for SO·4− in the degradation of ACT in the Fe2O3@Cu2O/PDS process.
HO· + ACT → products (10)
k obs with HO· scavenger
Normally, SO·4−
reacts with organic compounds through the electron Contribution of SO·4− (%) = × 100
k obs without scavenger (11)
transfer process while HO∙ do so via the hydrogen-atom abstraction and
electron transfer mechanisms [31]. Therefore, the degradation of ACT Contribution of HO· (%) = 100 − contribution of SO·4− (12)
in the S-MgO/PMS is more efficient than that in the S-MgO/PDS process
due to the greater number of reaction mechanisms involved in the de- Based on Fig. 5, the EPR spectrum taken with DMPO did not show
gradation process. Accordingly, the high oxidation potential of SO·4− and any peaks which revealed that catalysts themselves or in the presence of
HO· along with the high stability of PMS made the S-MgO/PMS process PMS do not give HO· as no peaks for DMPO-OH adduct was observed
very efficient for the degradation of recalcitrant organic compounds [32,33]. When TEMP spin trapping agent was used, a triplet pattern of
[31] such as ACT. equal intensity was recorded for both MgO and S-MgO catalysts, which
In order to confirm the formation SO·4− and HO· and to indicate their is characteristic for TEMP-1O2 adducts and confirm the singlet oxygen
relative contribution in the degradation of ACT in the S-MgO/PMS in both catalysts [32]. The ACT degradation in the S-MgO/PMS process
process, the ACT removal was evaluated in the absence and presence of was compared in the presence of aeration and N2 blowing. It was found
ethanol and tert-Butanol as radical scavengers with a considerably that the ACT degradation in the process with aeration was 17% higher
different rate constant with SO·4− and HO·. The complete degradation of than that with N2 blowing; confirming that dissolved oxygen played a
50 mg/L ACT was obtained in the S-MgO/PMS process within 30 min role in producing ROSs in solution. It can be seen that intensity of tri-
reaction time in the absence of scavengers (Figs. 3a,b). plet peaks is higher for MgO in comparison of S-MgO, which also
However, when tert-butanol was added to the reaction medium, the confirmed the surface composition modification due to the presence of
degradation rate decreased and it took 40 min to complete removal of S. EPR of MgO and S-MgO samples was also collected in presence of
50 mg/L ACT. Moreover, the ACT degradation reached to 86.3% within PMS by using DMPO [33]. EPR spectrum clearly indicates the presence
60 min in the S-MgO/PMS process in the presence of ethanol. The inset of sulfate radicals in S-MgO/PMS, which is in good agreement with a
Fig. 4 depicts the PFO plots of ACT removal in the S-MgO/PMS process seven-line pattern for sulfate radicals as reported by Lin et al. [34].
in the absence and in the presence of the selected radical scavengers. As However, no peak for sulfate radicals was observed in MgO/PMS pro-
shown in inset Fig. 4, the PFO rate constant of ACT degradation reac- cess. This study clearly confirms that SO·4− was the main reactive specie
tion was 0.289 min−1 in the absence of radical scavengers, whereas it involved in degradation of ACT in the S-MgO/PMS process and in-
decreased to 0.232 and 0.037 min−1 in the presence of tert-butanol and dicates that the S doping enhanced the generation of sulfate radicals
and hence higher degradation of ACT molecules was obtained once S-
MgO was used as a catalyst.

3.2. Effect of solution pH

The effect of initial solution pH between 4 and 10 was evaluated on


the degradation of 50 mg/L ACT in the S-MgO/PMS process. The results
are shown in Fig. 6 indicating that ACT degradation in the S-MgO/PMS
process had the same trend as that in the S-MgO/PDS process, although
with a higher level of efficiency. As observed in Fig. 6, ACT degradation
was not affected considerably by the solution pH between 4 and 10.
Since the pH of natural waters is around neutral condition and that the
ACT removal in the selected processes was highest at the neutral con-
dition, the remained tests were conducted at solution pH of 7. It should
be noted that the adsorption of ACT (absence of PMS) onto the catalyst
(MgO or S-MgO) was below 2% over the selected range of the initial pH
(4–10), suggesting that adsorption had not any significant contribution
in removing ACT and thus is negligible.
Fig. 4. Effect of radical scavengers on ACT degradation in the S-MgO/PMS Since the natural waters are almost neutral in pH, obtaining the
process (ACT = 50 mg/L, S-MgO = 1 g/L, PMS = 1 g/L, solution pH = 7, re- maximum ACT degradation in the S-MgO/PMS process at the neutral
action time = 0–60 min). solution pH reveals that the process can be appropriately applied for

408
F. Fanaei, et al. Chemical Engineering Journal 371 (2019) 404–413

Fig. 6. Effect of initial solution pH on ACT degradation in the S-MgO/PMS


process (ACT concentration = 50 mg/L, S-MgO = 1 g/L, PMS = 1 g/L, reaction
time = 10 min).

the surface charge of the catalyst. S-MgO has a zero point charge pH
(pHzpc) of 10.9 [19], causing the increase of the final solution pH to the
alkaline range around 9, the surface of the catalyst was positively
charged at all the selected solution pHs. Moreover, considering its
pKa1 = 0 and pKa2 = 9.2, HSO5− was the dominant species of PMS
under the acidic and neutral solution pHs below 9.2 whereas SO52−
would be dominated at the solution pHs over 9.2 [16,35]. Therefore, at
solution pH < 9.2 the negative HSO5− anions were attracted toward
the positively charged surface of S-MgO resulted in its becoming cata-
lyzed into SO·4− and HO· (Eq. (8)). The lower ACT removal efficiency at
the acidic pHs can be related to the significant inhibition effect of so-
lution pH on the formation of H-bond between H+ and the O − O group
of HSO5− which caused the attachment of positive charge to HSO5−
and thus impeding the reaction rate between HSO5− and positively
charged surface of S-MgO [16,35]. The effect of solution pH on the ACT
removal efficiency in the S-MgO/PMS is in accordance with the related
literature on catalytic activation of PMS with metal oxides. For in-
stance, the maximum ACT degradation efficiency was obtained in the
Fe3O4/PMS [16] and CuFe2O4/PMS [35] sulfate-based AOPs at the
neutral condition.

3.3. Effect of PMS to catalyst ratio

The effect of PMS concentrations between 0.2 and 1 g/L (0.025 and
0.12 mM) with a constant S-MgO concentration of 1 g/L as activator
catalyst corresponding to PMS to S-MgO ratio between 0.025 and
0.12 mM.L/g was investigated on the degradation of 50 mg/L ACT in
the S-MgO/PMS process. Fig. 7 shows the ACT degradation at different
PMS to S-MgO ratio as a function of reaction time, generation of radical
species as well as for providing greater time for the reaction between
ACT molecules and radical species.
It can be seen in Fig. 7 that at a given concentration of PMS (PMS to
S-MgO ratio), the ACT degradation improved with the increase in re-
action time, which can be related to the increased interaction time
between the PMS and the catalyst resulted in forming the greater
amount of reactive species and thereby in increasing the degradation of
ACT molecules. In addition, at a given reaction time, the ACT de-
Fig. 5. Detection of reactive species by EPR spectra collected with (a) TEMP for gradation increased with the increase in the PMS concentration (PMS to
MgO and S-MgO samples, (b) DMPO for MgO and S-MgO samples and (c) DMPO S-MgO ratio). For instant, the ACT degradation in the process was
for MgO/PMS and S-MgO/PMS. around 45.7, 65.2, 75.3, 88.2, 90.2 and 91.7% at ratio of 0.025, 0.04,
0.05, 0.07, 0.1 and 0.12 mM.L/g, respectively, within a reaction time of
10 min. It is worth noting that the ACT removal both in the absence of
efficient treatment of the contaminated water without needing pH ad-
PMS and S-MgO was below 3% at a reaction time of 60 min. It indicates
justment [16,35]. The observed trend of ACT degradation as a function
that the adsorption onto S-MgO as well as oxidation by single PMS was
of solution pH in the selected processes can be explained considering
ineffective in ACT removal and thus that ACT was removed in the
the effect of solution pH on the PMS dissociation in the solution and on
process through degradation by reaction with the generated radicals

409
F. Fanaei, et al. Chemical Engineering Journal 371 (2019) 404–413

cause the reduction of the ACT degradation percentages. The ACT re-
moval rate decreased to 14.95 and 13.40 mg/L.min with further in-
crease in PMS concentration to 0.8 and 1 g/L (PMS to catalyst ratio of
0.1 and 0.12 mM.L/g) likely due to the recombination of SO·4− (Eq. (13))
or HO· (Eq. (14)) reactive species of the reaction between the radical
species and the extra PMS (Eqs. (15)–(17)) [16,36,37].
SO·4− + SO·4− → S2 O−8 2 (13)

HO· + HO· → H2 O2 (14)

SO·4− + HO· → HSO−4 + 0.5O (15)

SO·4− + HSO5− → SO24− + HSO·5− (16)

HO· + HSO− ·−
5 → SO5 + H2 O (17)
The optimum mass ratio of PMS to catalyst was therefore found to
be 0.07 mM.L/g for S-MgO, which is much lower than that reported for
Fig. 7. Effect of PMS to S-MgO ratio on ACT degradation in the S-MgO/PMS
Fe3O4 [16] and for CoFe2O4 and MnFe2O4 [17]. It is clearly deduced
process (ACT = 50 mg/L, solution pH = 7, reaction time = 2.5–30 min).
that the S-MgO is a more efficient catalyst than the other tested ones for
activation of PMS. Table 1 compares the performance of the S-MgO/
rather than by adsorption onto S-MgO. The ACT degradation was faster PMS process as the most efficient process developed in this study with
in the first 10 min-time mainly due to the greater amount of PMS mo- different SO·4−-based AOPs for the degradation of ACT. As given in
lecules available for participating in the reaction with S-MgO and thus Table 1, the S-MgO/PMS process had much greater performance than
the generation of the greater amount of radicals, as well as to the the other SO·4−-based AOPs for the degradation of ACT. It shows that the
greater amount of ACT molecules available for interaction with the S-MgO has a high activity for activation of the PMS for the increased
generated radicals. Tan et al. [16] also reported the same trend for ACT generation of oxidative radicals.
removal in the Fe3O4/PMS process. The ACT removal efficiency in
MnFe2O4/PMS and CoFe2O4/PMS processes increased with the increase 3.4. Effect of water anions
in PMS to catalyst ratio from 0.25 to 1.0 [17]. It clearly indicates that
the optimum ratio of the PMS oxidant to catalyst greatly depends on the Contaminated waters contain different anions, which may affect the
type of the catalyst. The highest value of RSE was calculated during the performance of AOPs used for the decontamination. Fig. 8 shows the
first 10 min of the reaction of the S-MgO/PMS process conducted under results of the effect of main water anions (concentration of each anion
the optimum conditions (PMS = 0.6 g/L, S-MgO = 1 g/L) in which was 100 mg/L) including chloride, carbonate, nitrate, sulfate, and
95.6% of ACT was degraded to be 43.6% (Fig. S2). The value of RSE for phosphate the performance of S-MgO/PMS process for the degradation
the ACT removal in the activated PMS process has not been reported to of ACT. The degradation of ACT in the contaminated tap water was also
compare with the S-MgO. tested to find out the effect of a mixture of anions. Fig. 8 indicates that
In order to better illustrate the effect of PMS concentration on the the presence of anions at the selected concentrations either individually
performance of S-MgO/PMS process in degradation of ACT, the PFO or as a mixture did not affect considerably the ACT degradation in the
reaction kinetic model was fitted with the experimental data shown in process under the examined conditions.
Fig. 7. The kinetic information of ACT degradation in the S-MgO/PMS Although the selected anions may act as the radical scavengers, the
process at different PMS concentrations is given in Table S1. As ob- finding revealed that they could not compete with the generated SO·4−
served in Table S1, the rate of ACT degradation in the S-MgO/PMS and HO· at the selected concentrations. Generally, the degradation rate
process increased from 3.1 to 15.7 mg/L.min due to an increased in the of organic pollutant is retarded in the presence of Cl- ions because it can
generated SO·4− and HO· (Eq. (8)) when the PMS concentration was in- consume HO· radicals. But, in the presence of PMS, the direct reaction
creased from 0.2 to 0.6 g/L corresponding to the increase of PMS to between PMS and Cl- could lead to the formation of active chlorine
catalyst ratio of 0.025 to 0.07 mM L/g. With considering the constant species (Cl2/HClO), which are capable of degrading pollutants [40]. In
amount of S-MgO used in the tests, the increase of ACT removal rate this case, there may be mainly SO·4−, HO·, Cl·2−, and HClO in the aqueous
with the increase in PMS to S-MgO ratio up to 0.07 mM L/g is attributed system [41]. All of these strong oxidizers might participate in the de-
to the generation of greater amount of radical species due to the in- gradation of organic contamination. Further, the active chlorine species
crease in availability of the precursor which limited the radical gen- (HClO/Cl2) are dominant oxidants under neutral/alkaline conditions.
eration in the process [16,17]. Indeed, it shows that when the PMS was PMS is more stable at pH < 4.0 in comparison to alkaline conditions
further increased, the recombination of the generated radicals might (pH > 7.0) hence faster degradation rates of ACT were expected under

Table 1
Summary of ACT removal in the sulfate radical-based AOPs with metal/metal oxides as the activator.
Oxidant Catalyst pH ACT (mg/L) Oxidant to catalyst (mM L/g) Removal efficiency kobs (min−1) robs (mg/L.min) Ref.

PDS Fe2O3-Cu2O 6.5 100 10.4 ∼90% @ 40 min 0.045 4.50 [10]
PDS Fe0 3 10 1 > 90% @ 180 min 0.023 0.23 [11]
PDS Fe2+ 3 7.5 14.3 ∼70% @ 30 min – – [14]
PMS Fe3O4 7 10 0.25 – 0.012 0.12 [16]
PMS MnFe2O4 7 10 1 100% @ 60 in 0.121 1.21 [17]
PMS CoFe2O4 9 10 1 100% @ 120 min 0.053 0.53 [18]
PDS rGO-Ag0 /Fe3O4 7 1.5 10 > 90% @ 60 min 0.008 0.01 [19]
PDS Bicarbonate 8.3 1.5 6.6 ∼70% @ 420 min 0.003 0.0045 [38]
PDS TiO2-x/rGO-Vis 10 28.6 ∼100% @ 40 min 0.452 4.52 [39]
PMS S-MgO 7 50 0.07 100% @ 30 min 0.314 15.70 Present work

410
F. Fanaei, et al. Chemical Engineering Journal 371 (2019) 404–413

time was 16.2% and 3.5%, respectively, at PMS and catalyst con-
centrations of 0.2 mM and 0.2 g/L, respectively [17]. Yang et al. [39]
reported a 53% mineralization efficiency for 0.07 mM ACT in the TiO2-
x/rGO-PDS-Vis process using 1 g/L catalyst and 2 mM PDS. The higher
degree of ACT mineralization in the S-MgO/PMS process as compared
to other processes is attributed again to the greater activity of the S-
MgO, due to the creation of oxygen vacancies in its structure [15,42],
than the other tested materials for activation of peroxysulfates and thus
greater generation of oxidative radicals.

3.6. ACT degradation intermediates and pathway

The intermediates of ACT degradation (reaction time = 60 min,


catalyst concentration = 1 g/L, ACT concentration = 50 mg/L, solution
pH = 7, PMS concentration = 0.6 g/L) in the S-MgO/PMS process was
identified using LC/MS analysis. The LC/MS analysis data compared
Fig. 8. Effect of water anions on ACT degradation in the S-MgO/PMS process with the NIST mass spectrometry data center revealed the formation of
(ACT = 50 mg/L, S-MgO = 1 g/L, PMS = 0.6 g/L, initial solution pH = 7, re- some intermediates during ACT degradation (Table S2). Table S2
action time = 30 min). summarizes the reaction intermediates along with their molecular
structure identified in the LC/MS analysis. Fig. S3 shows the signals and
neutral or alkaline conditions due to the availability of more reactive abundances of a scan LC/MS analysis of ACT degradation intermediates
species. S-MgO has a zero point charge pH (pHzpc) of 10.9, which re- in [M + H+] mode at different m/z values. Based on intermediates
sults in an increase of the final solution pH to the alkaline range. Tan products found in the LC/MS analysis (Fig. S3), Fig. 10 proposes the
et al. [16] found slightly increased ACT degradation rate in the reaction pathways of ACT degradation in the S-MgO/PMS process,
MnFe2O4/PMS and CoFe2O4/PMS processes in the presence of bi- showing that two groups of reactions through mainly SO·4− and HO·
carbonate concentration of 2 mM and related the enhancement to in- species were possibly involved in ACT degradation. The SO·4− interacted
crease in solution pH rather than inhibition caused by scavenging effect with ACT molecules and its degradation intermediates through trans-
of bicarbonate. They [16] also report the reduction in ACT degradation ferring the single electron to ACT/intermediates and through producing
rate at the presence of 2 mM chloride. It is deduced that due to the of HO· species resulted in cleaving the bonds and finally in mineralizing.
generation of a high amount of radicals in the S-MgO/PMS process, the N-(2,4-dihydroxyphenyl)acetamide (TP3), N-(3,4-dihydroxyphenyl)
scavenging effect of anions was insignificant. Accordingly, it implies acetamide (TP3), 4-aminophenol (TP4), N-(1-hydroperoxy-4-ox-
that the S-MgO/PMS process can be efficiently applied for treating the ocyclohexa-2,5-dienyl)acetamide (TP10), N-(3,4,5-trihydroxyphenyl)
contaminated real streams. acetamide (TP5), N-(1,4-dihydroxycyclohexa-2,4-dienyl)acetamide
(TP6) and ACTdimer (TP1) were the intermediates produced through
3.5. Mineralization of ACT in the S-MgO/PMS process paths I, II, III, IV, V and VI respectively, during the degradation of ACT
in the S-MgO/PMS process under the selected conditions. Single elec-
Fig. 9 depicts the mineralization of 50 mg/L ACT (32 mg/L TOC) in tron transfer and oxidation could attack the weak bonds such as hy-
the S-MgO/PMS process under optimum experimental conditions. As droxyl, oxygen or amide groups. Because of unpaired electrons, the
can be seen in Fig. 9, the ACT mineralization increased from 8% at the amide and hydroxyl groups resonated to the respective conjugated sites
reaction time of 10 min to 62.4% when the reaction time was increased at the benzene ring. It leads to different suitable sites for connecting
to 60 min. The PFO reaction rate constant of ACT mineralization in the toHO·, the formation of multiple radical-radical complexes and finally
S-MgO/PMS process obtained from fitting the PFO reaction kinetic several isomers.
model with the mineralization experimental results (Fig. 7) was Further, the oxidation of these intermediates and hydroxylation
0.025 min−1 under the selected conditions. The information on the products by SO·4− and HO· led to breaking down the amide bonds and
mineralization of ACT in sulfate radical-based AOPs is limited. Deng aromatic structures and thereby being converted to some linear pro-
et al. [11] reported an ACT (10 mg/L) mineralization of 38.9% in the ducts like 3-hydroxypropanoic acid (TP20), malonic acid (TP19) and
Fe0/PDS process within 180 min. The degree of ACT mineralization in Other compounds, finally mineralized to CO2 and H2O. It should be
CoFe2O4/PMS and MnFe2O4/PMS processes within 120 min reaction mentioned that the m/z value for TP15 and TP17, TP9 and TP19, and
TP14 and TP20 (Fig. 10) were the same indicating that these pairs of
products could be probably formed during the degradation pathway of
ACT. The measurement of nitrate as an inorganic intermediate in the
effluent (TP15) clearly reveals the mineralization of nitrogen-con-
taining bonds in ACT and the degradation intermediates.

3.7. Toxicity bioassay

The acute 48-h toxicity of the ACT solution before and after treat-
ment in the S-MgO/PMS process was bio-assessed using Daphnia magna.
The LC50 of 50 mg/L ACT solution before treatment was determined to
be 30.7% or 18.1 mg/L, which is within the previously reported range
[43,44]. In comparison, the solution treated in the S-MgO/PMS process
under the optimum condition had the LC50 of 265.5% meaning the
toxicity of CAT solution decreased by 8.6 times upon treating in the S-
Fig. 9. Degradation and mineralization of ACT in the S-MgO/PMS process as a MgO/PMS process. TU values of ACT solution before and after treat-
function of the reaction time (ACT = 50 mg/L, S-MgO = 1 g/L, PMS = 0.6 g/L, ment in the S-MgO/PMS process was calculated to be 3.26 and 0.38,
solution pH = 7, reaction time = 0–60 min). respectively. Considering the classification made by Persoone et al.

411
F. Fanaei, et al. Chemical Engineering Journal 371 (2019) 404–413

Fig. 10. Pathway of ACT degradation in the S-MgO/PMS process (ACT = 50 mg/L, S-MgO = 1 g/L, PMS = 0.6 g/L, solution pH = 7, reaction time = 60 min).

[45], the bioassay findings indicate that the toxicity of ACT solution which along with the mineralization findings are evidence of complete
decreased from ‘acute toxicity’ class for the raw solution to ‘no acute transformation of ACT molecules to harmless compounds. Treatment in
toxicity’ class for the S-MgO/PMS process-treated solution. It implies the S-MgO/PMS process considerably decreased the toxicity of the ACT
that the effluent from the S-MgO/PMS process has no hazard to be solution from acute toxicity to no acute toxicity. In conclusion, the
discharged to the water environment. present work is of great importance for developing the novel advanced
oxidation technologies for efficient degradation of emerging con-
4. Conclusion taminants in the contaminated waters.

An emerging water contaminant, ACT, was selected to evaluate the


catalytic activity of S-MgO nanoparticles as compared to plain MgO in Acknowledgments
activating the PMS and PDS compounds. It was found that the S-doped
MgO was more active than the plain MgO in activating the selected This study was funded by Iran National Institute for Medical
oxidants and that higher amount of radical species were generated Research Development (NIMAD) under grant No. 963379. The authors
when using PMS rather than PDS. High degradation efficiency was are also grateful to Tarbiat Modares University, Tehran, Iran for pro-
achieved in the case of using S-MgO as catalyst and PMS as oxidant; viding technical support. Meghdad Karimi is acknowledged for his help
ACT removal efficiencies in MgO/PMS, MgO/PDS, S-MgO/PDS, and S- in interpreting LC-MS data and in preparing the degradation pathway.
MgO/PMS were 8.0, 21.5, 43.8 and 95.6%, respectively. ACT de- Santtu Heinilehto, (University of Oulu-Finland) is also acknowledged
gradation was evaluated in the S-MgO/PMS process under various ex- for providing XPS analysis of catalyst samples.
perimental conditions. A high rate of ACT degradation and miner-
alization could be achieved in the S-MgO/PMS process under neutral
pH. It was found that although both SO·4− and HO· species were gener- Appendix A. Supplementary data
ated in the S-MgO process, SO·4−played the main role in the achieved
ACT degradation in the process. A few simple intermediates were Supplementary data to this article can be found online at https://
identified by LC/MS analysis in the effluent of the S-MgO/PMS process, doi.org/10.1016/j.cej.2019.04.007.

412
F. Fanaei, et al. Chemical Engineering Journal 371 (2019) 404–413

References J. Mol. Catal. A: Chem. 265 (2007) 90–97.


[25] J.S. Corneille, J.-W. He, D.W. Goodman, XPS characterization of ultra-thin MgO
films on a Mo(100) surface, Surf. Sci. 306 (1994) 269–278.
[1] A. Sharma, J. Ahmad, S.J.S. Flora, Application of advanced oxidation processes and [26] A. Sternig, O. Diwald, Surface Decoration of MgO nanocubes with sulfur oxides:
toxicity assessment of transformation products, Environ. Res. 167 (2018) 223–233. experiment and theory, J. Phys. Chem. C Nanomater. Interfaces. 117 (2013)
[2] G.-D. Fang, D.D. Dionysiou, S.R. Al-Abed, D.-M. Zhou, Superoxide radical driving 7727–7735.
the activation of persulfate by magnetite nanoparticles: implications for the de- [27] J.F. Moulder, J. Chastain, Handbook of X-ray Photoelectron Spectroscopy. A
gradation of PCBs, Appl. Catal. B: Environ. 129 (2013) 325–332. Reference Book of Standard Spectra for Identification and Interpretation of XPS
[3] C. Chen, H. Feng, Y. Deng, Re-evaluation of sulfate radical based-advanced oxida- Data, Physical Electronics Division, Perkin-Elmer Corp., Eden Prairie, MN, 1992.
tion processes (SR-AOPs) for treatment of raw municipal landfill leachate, Water [28] N.M. Flores, U. Pal, R. Galeazzi, A. Sandoval, Effects of morphology, surface area,
Res. 153 (2019) 100–107. and defect content on the photocatalytic dye degradation performance of ZnO na-
[4] Y. Chen, J. Yan, D. Ouyang, L. Qian, L. Han, M. Chen, Heterogeneously catalyzed nostructures, RSC Adv. 4 (2014) 41099–41110.
persulfate by CuMgFe layered double oxide for the degradation of phenol, Appl. [29] Y.-H. Guan, J. Ma, Y.-M. Ren, Y.-L. Liu, J.-Y. Xiao, L-q Lin, et al., Efficient de-
Catal. A: Gen. 538 (2017) 19–26. gradation of atrazine by magnetic porous copper ferrite catalyzed perox-
[5] D. Xia, Y. Li, G. Huang, R. Yin, T. An, G. Li, H. Zhao, A. Lu, P.K. Wong, Activation of ymonosulfate oxidation via the formation of hydroxyl and sulfate radicals, Water
persulfates by natural magnetic pyrrhotite for water disinfection: efficiency, me- Res. 47 (2013) 5431–5438.
chanisms, and stability, Water Res. 112 (2017) 236–247. [30] T. Olmez-Hanci, I. Arslan-Alaton, Comparison of sulfate and hydroxyl radical based
[6] A. Ghauch, A.M. Tuqan, Oxidation of bisoprolol in heated persulfate/H2O systems: advanced oxidation of phenol, Chem. Eng. J. 224 (2013) 10–16.
kinetics and products, Chem. Eng. J. 183 (2012) 162–171. [31] A. Rastogi, S.R. Al-Abed, D.D. Dionysiou, Sulfate radical-based ferrous–perox-
[7] A. Ghauch, A.M.T. Tuqan, N. Kibbi, Ibuprofen removal by heated persulfate in ymonosulfate oxidative system for PCBs degradation in aqueous and sediment
aqueous solution: a kinetics study, Chem. Eng. J. 483 (197 2012) 492. systems, Appl. Catal. B: Environ. 85 (2009) 171–179.
[8] X. Zhang, J. Yao, Z. Zhao, J. Liu, Degradation of haloacetonitriles with UV/per- [32] T.D. Minh, M.C. Ncibi, V. Srivastava, S.K. Thangaraj, J. Jänis, M. Sillanpää,
oxymonosulfate process: degradation pathway and the role of hydroxyl radicals, Gingerbread ingredient-derived carbons-assembled CNT foam for the efficient
Chem. Eng. J. 364 (2019) 1–10. peroxymonosulfate-mediated degradation of emerging pharmaceutical con-
[9] A. Ghauch, A. Baalbaki, M. Amasha, R. El Asmar, O. Tantawi, Contribution of taminants, Appl. Catal. B: Environ. 244 (2019) 367–384.
persulfate in UV-254 nm activated systems for complete degradation of chlor- [33] S.B. Hammouda, C. Salazar, F. Zhao, D.L. Ramasamy, E. Laklov, S. Iftekhar, I. Babu,
amphenicol antibiotic in water, Chem. Eng. J. 317 (2017) 1012–1025. M. Sillanpää, Efficient heterogeneous electro -Fenton incineration of a contaminant
[10] Y. Zhang, Q. Zhang, J. Hong, Sulfate radical degradation of acetaminophen by novel of emergent concern-cotinine- in aqueous medium using the magnetic double per-
iron–copper bimetallic oxidation catalyzed by persulfate: mechanism and de- ovskite oxide Sr2FeCuO6 as a highly stable catalayst: degradation kinetics and
gradation pathways, Appl. Surf. Sci. 422 (2017) 443–451. oxidation products, Appl. Catal. B: Environ. 240 (2019) 201–214.
[11] J. Deng, Y. Shao, N. Gao, Y. Deng, C. Tan, S. Zhou, Zero-valent iron/persulfate (Fe0/ [34] K.-Y.A. Lin, Y.-C. Chen, C.-F. Huang, Magnetic carbon-supported cobalt prepared
PS) oxidation acetaminophen in water, Int. J. Environ. Sci. Technol. 11 (2014) from one-step carbonization of hexacyanocobaltate as an efficient and recyclable
881–890. catalyst for activating Oxone, Sep. Purif. Technol. 170 (2016) 173–182.
[12] A. Ghauch, G. Ayoub, S. Naim, Degradation of sulfamethoxazole by persulfate as- [35] T. Zhang, H. Zhu, J.-P. Croué, Production of sulfate radical from peroxymonosulfate
sisted micrometric Fe0 in aqueous solution, Chem. Eng. J. 228 (2013) 1168–1181. induced by a magnetically separable CuFe2O4 spinel in water: efficiency, stability,
[13] G. Ayoub, A. Ghauch, Assessment of bimetallic and trimetallic iron-based systems and mechanism, Environ. Sci. Technol. 47 (2013) 2784–2791.
for persulfate activation: application to sulfamethoxazole degradation, Chem. Eng. [36] G. Moussavi, H. Momeninejad, S. Shekoohiyan, P. Baratpour, Oxidation of acet-
J. 256 (2014) 280–292. aminophen in the contaminated water using UVC/S2O82− process in a cylindrical
[14] S. Wang, J. Wua, X. Lu, W. Xu, Q. Gong, J. Ding, B. Dan, P. Xie, Removal of acet- photoreactor: efficiency and kinetics of degradation and mineralization, Sep. Purif.
aminophen in the Fe2+/persulfate system: kinetic model and degradation path- Technol. 181 (2017) 132–138.
ways, Chem. Eng. J. 358 (2019) 1091–1100. [37] C. Tan, N. Gao, S. Zhou, Y. Xiao, Z. Zhuang, Kinetic study of acetaminophen de-
[15] P. Gao, X. Tian, Y. Nie, C. Yang, Z. Zhou, Y. Wang, Promoted peroxymonosulfate gradation by UV-based advanced oxidation processes, Chem. Eng. J. 253 (2014)
activation into singlet oxygen over perovskite for ofloxacin degradation by con- 229–236.
trolling the oxygen defect concentration, Chem. Eng. J. 359 (2019) 828–839. [38] M. Jiang, J. Lu, Y. Ji, D. Kong, Bicarbonate-activated persulfate oxidation of acet-
[16] C. Tan, N. Gao, Y. Deng, J. Deng, S. Zhou, J. Lia, X. Xin, Radical induced de- aminophen, Water Res. 116 (2017) 324–331.
gradation of acetaminophen with Fe3O4 magnetic nanoparticles as heterogeneous [39] L. Yang, L. Xu, X. Bai, P. Jin, Enhanced visible-light activation of persulfate by Ti3+
activator of peroxymonosulfate, J. Hazard. Mater. 276 (2014) 452–460. self-doped TiO2/graphene nanocomposite for the rapid and efficient degradation of
[17] C. Tan, N. Gao, D. Fu, J. Deng, L. Deng, Efficient degradation of paracetamol with micropollutants in water, J. Hazard. Mater. 365 (2019) 107–117.
nanoscaled magnetic CoFe2O4 and MnFe2O4 as a heterogeneous catalyst of perox- [40] B. Sheng, Y. Huang, Z. Wang, F. Yang, L. Aia, J. Liu, On peroxymonosulfate-based
ymonosulfate, Sep. Purif. Technol. 175 (2017) 47–57. treatment of saline wastewater: when phosphate and chloride co-exist, RSC Adv. 8
[18] C.M. Park, J. Heo, D. Wang, C. Su, Y. Yoone, Heterogeneous activation of persulfate (2018) 13865–13870.
by reduced graphene oxide–elemental silver/magnetite nanohybrids for the oxi- [41] P. Wang, S. Yang, L. Shan, R. Niu, X. Shao, Involvements of chloride ion in deco-
dative degradation of pharmaceuticals and endocrine disrupting compounds in lorization of Acid Orange 7 by activated peroxydisulfate or peroxymonosulfate
water, Appl. Catal. B: Environ. 225 (2018) 91–99. oxidation, J. Environ. Sci. 23 (2011) 1799–1807.
[19] A. Mashayekh-Salehi, G. Moussavi, K. Yaghmaeian, Preparation, characterization [42] K. Yaghmaeian, G. Moussavi, A. Mashayekh-Salehi, A. Mohseni-Bandpei, M. Satari,
and catalytic activity of a novel mesoporous nanocrystalline MgO nanoparticle for Oxidation of acetaminophen in the ozonation process catalyzed with modified MgO
ozonation of acetaminophen as an emerging water contaminant, Chem. Eng. J. 310 nanoparticles: effect of operational variables and cytotoxicity assessment, Process
(2017) 157–169. Saf. Environ. 109 (2017) 520–528.
[20] APHA, Standard Methods for the Examination of Water and Wastewater, 23th ed., [43] Y. Kim, K. Choi, J. Jung, S. Park, P.-G. Kim, J. Park, Aquatic toxicity of acet-
APHA, Washington, D.C, 2017. aminophen, carbamazepine, cimetidine, diltiazem and six major sulfonamides, and
[21] M. Amasha, A. Baalbaki, A. Ghauch, A comparative study of the common persulfate their potential ecological risks in Korea, Environ. Int. 33 (2007) 370–375.
activation techniques for the complete degradation of an NSAID: the case of keto- [44] J. Du, C.-F. Mei, G.-G. Ying, M.-Y. Xu, Toxicity Thresholds for Diclofenac,
profen, Chem. Eng. J. 350 (2018) 395–410. Acetaminophen and Ibuprofen in the Water Flea Daphnia magna, Bull. Environ.
[22] G. Moussavi, A. Mashayekh-Salehi, K. Yaghmaeian, A. Mohseni-Bandpei, The cat- Contam. Toxicol. 97 (2016) 84–90.
alytic destruction of antibiotic tetracycline by sulfur-doped magnesium oxide [45] G. Persoone, B. Marsalek, I. Blinova, A. Törökne, D. Zarina, L. Manusadzianas,
(S–MgO) nanoparticles, J. Environ. Manage. 210 (2018) 131–138. G. Nalecz-Jawecki, L. Tofan, N. Stepanova, L. Tothova, B. Kolar, A practical and
[23] J. Xu, Y. Teng, F. Teng, Effect of surface defect states on valence band and charge user-friendly toxicity classification system with microbiotests for natural waters and
separation and transfer efficiency, Sci. Rep. 6 (2016) 32457. wastewaters, Environ. Toxicol. 18 (2003) 395–402.
[24] B.M. Nagaraja, A.H. Padmasri, B. David Raju, K.S. Rama Rao, Vapor phase selective
hydrogenation of furfural to furfuryl alcohol over Cu–MgO coprecipitated catalysts,

413

You might also like