Aula 6
Aula 6
Aula 6
A R T I C L E I N F O A B S T R A C T
Keywords: In this study, the degradation of highly toxic pollutants acrylonitrile (ACN) and acrylamide (ACM) by perox
Heterogeneous catalysis ymonosulfate (PMS) activation with series perovskite-like catalyst SrCuxNi1-xO3 (x = 0, 0.2, 0.4, 0.6, 0.8, 1) were
Peroxymonosulfate activation studied systematically in the batch experiments. Synthesized catalysts were characterized by various analytical
Acrylonitrile
techniques e.g., XRD, FTIR, FESEM with EDAX, TEM, XPS and EPR spectroscopy. Subsequently, effects of key
Acrylamide
Reactive oxygen species
parameters viz., PMS dosage, catalyst dosage, pH and reaction temperature were evaluated on the degradation
Catalyst characterization efficiency and simultaneously evaluated the PMS utilization efficiencies along with PMS consumption. Catalyst
0.6SCN exhibited the highest catalytic activity and maximum removal of acrylonitrile (96%) and acrylamide
(81%) along with 86% PMS consumption were observed at optimum operatizing conditions. In addition, stability
and recyclability of 0.6SCN catalyst were assessed and removal efficiency of acrylonitrile suppressed 6% and
acrylamide suppressed 11% over the fourth cycle of experiments along with very low leaching of Cu and Ni
observed. Furthermore, electron paramagnetic resonance (EPR) and quenching experimental studies reveal that
in-situ generated both reactive species (SO4•– and •OH) were accountable for the removal of acrylonitrile and
acrylamide synergistically in 0.6SCN/PMS system. Reaction pathways of acrylamide and acrylonitrile degra
dation were proposed based on the intermediate detected through GC–MS analysis and literature studies. Finally,
the operating cost of treatment process was assessed 50.62$/m3 of wastewater in 0.6SCN/PMS system.
1. Introduction L with extremely high COD 355,000 mg/L [5,6]; during ABS resin
production released the acrylamide ~ 75 mg/L, acrylonitrile 30–414
In the recent decades, extensive industrial usage of acrylonitrile and mg/L along with very high COD 1100–1300 mg/L [7-9]. In view of the
acrylamide have led to contaminate the surface water as well as ground strict regulatory standards for water safety and emissions, Environ
water and received an increasing concern on environmental pollution mental and Water Bureau, Michigan, United States has established
[1]. Acrylonitrile and acrylamide both compounds are very highly toxic allowable limits for acrylonitrile 0.0026 mg/L and acrylamide 0.0003
even at low concentration levels, because they exhibit characteristics of mg/L in groundwater [10]. The chronic exposure of acrylamide leads to
neurotoxicity, mutagenicity, carcinogenic and reprotoxic nature [2,3]. damage the central and nervous peripheral system and ataxia in legs
Acrylamide has been classified Group B2, probable potential human [11] and chronic exposure to acrylonitrile causes asphyxia, irritation to
carcinogen by United States Environmental Protection Agency (USEPA) the respiratory system and digestive tract [11,12]. Consequently,
and acrylonitrile has been identified as the third highest priority toxic effective treatment methods are urgently needed to mitigate acryloni
pollutant in the 129 priority pollutants list [4]. Acrylonitrile and trile and acrylamide from industrial wastewater in order to prevent the
acrylamide are generally discharged from acrylonitrile butadiene sty detrimental impacts on flora and fauna. Recently, various authors have
rene (ABS) resin manufacturing, nitrile rubber, petrochemical industry, been treated the acrylamide and acrylonitrile containing wastewater by
acrylic fiber, adhesive and flocculent manufacturing etc. The petro various techniques e.g., photocoagulation and electrocoagulation pro
chemical industry wastewater discharged the acrylonitrile 2000 mg/L cess [13], photocatalytic process [14], advanced UV/H2O2 post treat
and their family compounds acetonitrile 178 mg/L, adiponitrile 68 mg/ ment [15], Ferrate(VI) Oxidation [16], adsorption process [5],
* Corresponding author.
E-mail address: akumar6@ch.iitr.ac.in (A. Kumar).
https://doi.org/10.1016/j.seppur.2021.119677
Received 22 July 2021; Received in revised form 3 September 2021; Accepted 4 September 2021
Available online 13 September 2021
1383-5866/© 2021 Elsevier B.V. All rights reserved.
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
biological process [3,17,18]. However, aforementioned technologies Entire reagents and aqueous solution were preserved at 4 ◦ C to avoid the
have some drawbacks such as fenton process produce the large volume biodegradation and microorganism’s growth. Initial concentration of
of sludge which create the secondary pollution, very slow pollutant acrylonitrile and acrylamide was taken composition similar to real ABS
degradation in biological process [19,20]. In the recent years, sulfate resin wastewater as previous study reported [8,39].
radical based advanced oxidation processes (AOPs) have gained
increasing attention in wastewater treatment due to its attractive char 2.2. Catalyst synthesis
acteristics viz., stronger redox potential (Eθ = 2.5–3.1 V) and longer half-
life (30–40 µs) for SO4•– radical as compare to •OH radical (Eθ = 1.9–2.7 A series of SrCuxNi1-xO3 (x = 0, 0.2, 0.4, 0.6, 0.8, 1) perovskite-like
V) and 20 µs half-life [21,22]. Many authors have been reported that catalysts were synthesized using sol–gel method according to previously
peroxymonosulfate (PMS) is very high potential oxidant that could be reported method [39,40]. Schematic diagram of SrCuxNi1-xO3 series
activated with transition metals [1,22], ultraviolet irradiation [23,24], perovskite like catalyst synthesis as shown in Fig. 1(a). All the details of
ultrasound [25], heat [26], and carbon materials [27]. Sulfate radical catalyst synthesis are explained in Supplementary material file (Text
based AOPs have some drawback such as generation of sulfate ion. To S2).
protect the aquatic life, sulfate ion is typically regulated to be less than
500 mg/L [28,29]. In the present study, the residual concentration of 2.3. Catalytic testing
sulfate ion in treated water is within this limit. Moreover, the
enhancement of PMS activation efficiency with perovskite-like catalyst In this study, all the experimental runs were conducted in 250 mL
is still needed for degradation of toxic pollutants. Recently, perovskite three-neck glass reactor with binary solution (100 mL, 100 mg/L
have emerged a new class of material in the mixed oxide family to acrylamide and 100 mg/L acrylonitrile) inserted in water bath at the
effectively activation of PMS for the degradation of toxic pollutants from desired temperature and continuous stirring by magnetic stirrer at 300
aqueous media due to its attractive characteristics viz., excellent redox rpm. A certain concentration of catalyst was dispersed into aqueous
property, high hydrothermal stability and high electron mobility etc., solution prior to attaining the desired temperature and followed by
[30]. Perovskite-like materials have been used in various areas i.e., adding the known amount of oxidant dosage (PMS) to initiate the
catalysis with PMS activation [1,31], electrochemical [32], chemical degradation reaction. Initial pH of the solution was regulated by using 1
sensor [30], fuel cells material as a solid oxide [33]. The general N sulfuric acid or 1 N sodium hydroxide acid as accordingly before the
structure formula of perovskite oxides is ABO3, where A is typical experimentation. Periodically 1 mL of aliquot were taken and immedi
occupied by lanthanum, alkaline earth metal cation having large ionic ately quenched with equal volume of methanol to quench the residual
radius and B is occupied by transition metal ions [34,35]. Authors radicals and filtered with 0.22 µm syringe membrane filters prior to
Hammouda et al., synthesized the various types of perovskite catalysts analysis. All the experiments conducted duplicate and produced data
A-CoO3 (A = Ba, La, Ce and Sr) and found that catalyst SrCoO3 exhibited were taken average and reported with the standard deviations.
outstanding catalytic activity and stability among the remaining catalyst
during the degradation of phenol [36]. In our previous work, partially 2.4. Analytical techniques and instrumentation
substitution of Sr in A site cations showed the excellent catalytic activity
for acrylic acid degradation [31]. Generally, transition metals (Cu, Ni, The concentration of ACN and ACM were monitored by high per
Fe, Mn, Co) were well employed in B site cations to activate PMS for formance liquid chromatography (HPLC, Waters India, Ltd.) system
contamination abatement from wastewater [1,31,37] In the last few equipped with C18 column (3.9 mm × 150 mm) and an ultraviolet (UV)
years, various types of pollutants e.g., phenol, atrazine, Rhodamine B detector at 207 nm wavelength. The mobile phase was used pure Milli-Q
dye and herbicides has been efficiently removed from aqueous solution water with 0.5 mL/min flowrate and injection volume 20 µL (manually
by perovskite-like catalysts with employing the transition metals (Cu, injected) in isocratic mode. Inductively coupled plasma mass spec
Ni, Fe, Mn, Co) in B sites cations [22,35,38,39]. To our best knowledge, trometry (ICP-MS) (Perkin Elmer SCIEX, model ELAN-DRC-e) was used
no study has been conducted for simultaneous degradation of acryl to detect the metal leaching in aqueous phase. The •OH radical species
amide and acrylonitrile by heterogeneous catalysis with series perov was detected by electron paramagnetic resonance (EPR) using 10 mM
skite SrCuxNi1-xO3 catalyst so far. Therefore, in this study a series 5,5-Dimethyl-1-pyrroline N-oxide (DMPO) as a spin-trapping agent. The
perovskite SrCuxNi1-xO catalysts were synthesized by citric sol–gel intermediates formed during the reaction were identified by gas chro
method and characterized by various analytical techniques viz., TEM, matography coupled with mass spectrometer (GC–MS, Varian 3900,
FE-SEM, XPS, FTIR, BET, XRD and EPR. The aim of present study was Saturn 2100 T) with a capillary column DB-5 MS (30 m × 0.25 mm, 0.25
comprehensively evaluation of SrCuxNi1-xO/PMS system, evaluation of m) and Saturn Chemstation software. Prior to GC–MS analysis, sample
operating parameters influences on acrylonitrile and acrylamide was extracted with dichloromethane and dehydrated by sodium sulfate
removal efficiency, and PMS stoichiometric efficiencies along with (anhydrous). The resulting solution was filtered through 0.22 µm
estimation of PMS consumption were systematically investigated. membrane syringe filter and analyzed by GC–MS. The obtained data was
analyzed by Nist library. The concentration of PMS was measured using
2. Materials and methods the KI colorimetric method with a Visible Spectrophotometer
(HANNA™ HI801 iris, USA) at ƛmax 352 nm [41].
2.1. Chemicals used and preparation of synthetic aqueous solution
2.5. Physiochemical characterization
Acrylonitrile, sulfuric acid and tert-butyl alcohol were procured from
SRL Laboratory, Mumbai (India). Acrylamide, peroxymonosulfate Details of physiochemical characterization techniques employed in
(PMS), strontium nitrate, nickel (II) nitrate hexahydrate, and sodium the present study are available in Text S3 in Supporting Information file.
hydroxide, L-histidine were procured from Himedia, Mumbai, India.
Potassium dichromate, silver sulphate and mercury sulphate were ob 3. Results and discussions
tained from Ranbaxy Fine Chemicals, New Delhi, India. All the chem
icals and solvents used for the experimental purpose were of analytical 3.1. Characterization of SrCuxNi1-xO3 (x = 0, 0.2, 0.4, 0.6, 0.8, 1)
reagent (AR) grade. All the aqueous solutions were prepared in Milli-Q catalysts
water throughout the study. A stock solution of binary compounds
acrylonitrile (100 mg/L) and acrylamide (100 mg/L) were prepared 3.1.1. X-ray diffraction (XRD) analysis
synthetically by dissolving its known concentration in Milli-Q water. The crystalline phase and chemical structure of series perovskite-like
2
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
Fig. 1. (a) Schematic diagram of SrCuxNi1-xO3 series perovskite like catalyst synthesis by solgel method, (b) XRD spectra of, 0.2SCN, 0.4SCN, 0.6SCN and
0.8SCN catalysts.
catalysts SrCuxNi1-xO3 (x = 0.2, 0.4, 0.6, 0.8) were investigated by XRD 0.6SCN, 0.8SCN and 1SCN) exhibited minor changes and no additional
patterns recorded in the range of 20◦ to 90◦ . As shown in Fig. 1(b), peaks were observed which suggested that structure and components of
catalysts 0SCN, 0.2SCN, 0.4SCN, 0.6SCN, 0.8SCN and 1SCN exhibited the catalysts were not destroyed after catalysis reaction with PMS. The
the various phases e.g., CuO, Sr2Ni2O5, Sr2Cu2O3, NiO and Cu0.2Ni0.8O above results proved that the structure of perovskite-like catalysts
present in the samples. Diffraction peaks located near 2θ values, 35.43◦ , SrCuxNi1-xO3 was very stable and having higher potential for reusability
35.53◦ , 38.73◦ , 38.94◦ , 48.74◦ and 68.17◦ indexed to plan (002), (111), of catalysts. The Scherrer equation (D = kʎ/ßcosθ) was used to assess the
(1 1 1), (2 0 0), (202) and (2 2 0) exhibited the monoclinic structure of average crystal sizes of catalysts, notation ʎ is X-ray wavelength, k is
CuO phase that was consistent with the diffraction data JCPDS card no. Scherrer constant, ß is the full width at half maxima of the peak in
(00–041-0254), space group (15/C2/c) and data reported by various radian, and θ is the Braggs angle in radian. Accordingly, the average
authors [7,42,43]. An additional peak at 2θ values 31.29◦ indexed to crystal sizes i.e., 35.71, 39.28 nm, 51.47 nm, 65.5 nm, 48.37 nm and
plan (1 3 1) exhibited orthorhombic structure of Sr2Cu2O3 phase and 47.29 nm were obtained for catalyst, 0SCN, 0.2SCN, 0.4SCN, 0.6SCN,
well accord with diffraction data (JCPDS card no. 01–080-0864, space 0.8SCN and 1SCN respectively.
group − 69/Fmm). The diffraction peaks at 2θ values, 32.65◦ and 44.25◦
corresponding to plane (1 1 0) and (0 0 2) appointed to Sr2Ni2O5 phase 3.1.2. Morphological, structural and compositional information of the
with hexagonal structure. These peaks were well corresponding to catalysts
JCPDS diffraction data card no. 00–028-1242 and data reported by au The surface morphology of series perovskite-like catalysts SrCuxNi1-
thors [44]. In addition, intense diffraction peaks 37.32◦ and 43.36◦ xO3 (x = 0, 0.2, 0.4, 0.6, 0.8, 1) were illustrated by FE-SEM micrographs
exhibited the cubic crystalline structure of NiO phase corresponding to and their corresponding metal compositions were evaluated by EDAX
plane (1 1 1) and (2 0 0) which was well accord with the diffraction data data as depicted in Fig. 2(a-d). Catalyst 0.2SCN showed granular-like
JCPDS card no. 01–085-0625, space group – 225/Fm3m) [45]. More irregular, non-uniformed, interlinked and randomly distributed nano
over, next diffraction peaks located near, 62.45◦ and 75.23◦ attributed sized particles having the size less than 100 nm in average diameter.
to cubic crystalline structure of Cu0.2Ni0.8O phase and well matched Almost similar like morphology has been reported by authors Chen
with diffraction data JCPDS card no. 00–025-1049, space group – 225/ et al., for SrCuxO and MgSrCuO catalysts [46,47]. While substituting the
Fm3m. The XRD patterns of used catalysts (0SCN, 0.2SCN, 0.4SCN, copper composition (x = 0.4), morphology of catalyst 0.4SCN showed
3
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
a b c d
e f g h
Fig. 2. FE-SEM images of (a) 0.2SCN, (b) 0.4SCN, (c) 0.6SCN and (d) 0.8SCN catalysts; TEM images of 0.6SCN catalyst at (e) 100 nm, (f) 200 nm, (g) 200 nm and
(h) histogram plot of particle size distribution.
the minor changes such as enhanced the particle size and particles micrographs. The internal morphology 0.6SCN perovskite-like catalyst
surface become smooth as shown in Fig. 2(b). In addition, further were also evaluated by TEM micrographs as depicted in Fig. 2(e-g).
increased in copper composition (x = 0.6) in SrCuxNi1-xO3, granules Fig. 2(e) exhibited, nano sized agglomerated particles in the catalyst
particles were agglomerated and formed bigger size particles as depicted were in various shapes e.g., cubic, hexagonal and orthorhombic struc
in Fig. 2(c). Moreover, the addition of high composition of Cu (x = 0.8) ture. These results were well agreement with XRD results exhibited
in catalyst 0.8SCN exhibited the similar morphology like 0.6SCN but it shapes of various phases present (i.e., CuO, NiO, Sr2CuO3 and
was observed that particles density has enhanced. The EDAX spectra of Cu0.2Ni0.8O) in the 0.6SCN catalyst. Fig. 2(f) clearly showed the lattice
SrCuxNi1-xO3 (x = 0, 0.2, 0.4, 0.6, 0.8, 1) catalysts showed intense peaks fringes having the interplaner d-spacing 0.23 nm and 0.21 nm corre
of the elements e.g., Sr, Cu, Ni and O, present in the catalysts as depicted sponding to crystallographic plane d111 = 0.23 nm for monoclinic
in supporting information file (Fig. S2). The weight % of elements pre structure of CuO phase (JCPDS file no. 00–041-0254) [42,43] and plane
sent in catalysts 0.2SCN, 0.4SCN, 0.6SCN and 0.8SCN was almost near d200 = 0.21 nm for cubic crystal structure of NiO phase (JCPDS card no.
the stoichiometric composition indicated in the table in EDAX 01–075-0197) [45]. The interplaner d-spacing of the catalyst shown in
Fig. 3. XPS spectra of fresh and spent 0.6SCN catalyst, survey scan spectra, Cu 2p spectra, Ni 2p spectra, Sr 3d spectra, O1s spectra and thermogravimetric anal
ysis plot.
4
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
Fig. 2(f) reveal that lattice of the catalyst was not destroyed while eV (49.24%) with their respective composition, which could be attrib
introducing the copper composition (x = 0.6) in the 0.6SCN catalyst. uted to Sr2+ as shown in Fig. 3(d) [48][47]. After catalytic oxidation
However, mean size of 0.6SCN catalyst was analyzed by image pro reaction, the composition was slightly changed for both peaks Sr 3d3/2
cessing software (Image J Software). The average mean particle size was (46.27 %) and Sr 3d5/2 (53.73 %). Fig. 3(e), showed the XPS spectra of
estimated 84.61 nm by measuring over 40 particles as depicted in Fig. 2 O1s core levels of fresh and used 0.6SCN catalyst exhibited the three
(h). The average mean particle size was almost close to crystalline size main characteristic peaks at binding energy with their respective
obtained from XRD results. composition ~527.82 eV (44.35 %), ~529.45 eV (29.85 %) and
~530.79 eV (25.80 %), which could be ascribed to lattice oxygen, sur
3.1.3. X-ray photoelectron spectroscopy (XPS) analysis face adsorb oxygen and adsorbed molecular water and carbonates,
In order to explore the qualitative and quantitative information from respectively [35,45,51]. The reduction of lattice oxygen composition
the outermost few atomic layer of the fresh and used catalyst, XPS from 44.35% to 28.72% could be attributed to oxidation of lattice ox
techniques was employed [48,49]. Fig. 3(a), exhibited the various ygen with the reduction of Ni3+/Ni2+ to Cu2+/ Cu+ and also suggested
intense peaks for Sr (3d), Cu (2p), Ni (2p) and O (1 s) in the survey scan that O2− participated in the redox reaction of nickel and copper ions [1].
spectrum of 0.6SCN catalyst which suggested the presence of elements
in the sample. The XPS spectra of Cu 2p in fresh catalyst showed two 3.2. Preliminary experimental study: A comparison of different systems
characteristic peaks at binding energy ~931.53 eV and ~932.87 eV on the degradation of acrylonitrile and acrylamide
could be assign to Cu 2p3/2 with their corresponding composition Cu+
(31.37%) and Cu2+ (35.08%) as depicted in Fig. 3(b) [1,47]. After PMS Some preliminary experiments were conducted to evaluate the ac
activation reaction, the surface composition of Cu 2p significantly tivity of SrCuxNi1-xO3 catalysts for the degradation of acrylonitrile and
changed as Cu+ (16.57 %) and Cu2+ (50.16%) which suggested that acrylamide at room temperature. According to Fig. 4, very low removal
electron transfer occurred from Cu+/Cu2+ during the PMS activation efficiency of acrylonitrile (8%) and acrylamide (5%) were observed over
resulting in enhanced the fraction of Cu2+. The deconvolution of Ni 2p3/ 3 h reaction while 1 mM PMS used alone at natural pH. This signified
2 spectrum displayed two characteristic peaks with binding energies at that acrylonitrile and acrylamide difficult to be degraded by PMS alone.
~853.32 eV and ~855.47 eV, which could be assigned to Ni2+ (35.24%) When 1 g/L catalysts dose with 0SCN, 0.2SCN, 0.4SCN, 0.6SCN, 0.8SCN
and Ni3+ (32.64%) with their respective composition as depicted in and 1SCN was individually testified and obtained the low degradation of
Fig. 3(c) [50]. After PMS activation reaction, large amount of Ni2+ were acrylonitrile (i.e., 7%, 7%, 9%, 13%, 10% and 8%) and acrylamide (i.e.,
oxidized to Ni3+ and the composition of Ni2+ (24.19%) and Ni3+ 4%, 3%, 4%, 6%, 5% and 5%) achieved under identical conditions,
(39.48%) subsequently changed. The XPS spectra of Sr 3d in fresh respectively which was mainly caused of surface adsorption. In addition,
catalyst deconvoluted in two characteristic peaks Sr 3d3/2 and Sr 3d5/2 while PMS dose and catalyst dose simultaneously used maximum
corresponding to binding energies ~131.07 eV (50.74%) and ~132.78 removal of acrylonitrile (i.e., 28%, 31%, 32%, 40%, 34% and 32%) and
Fig. 4. Preliminary experimental study for acrylonitrile and acrylamide degradation. (Reaction conditions: [ACN]0 100 mg/L, [ACM]0 100 mg/L, catalysts dose 1
g/L, PMS dose 1 mM, pH 6.12 and temperature 25 ◦ C.
5
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
acrylamide (i.e., 13%, 14%, 19%, 25%, 21% and 22%) achieved with Further confirmation of reactive species (SO4•− and •OH) generation
corresponding catalysts 0SCN, 0.2SCN, 0.4SCN, 0.6SCN, 0.8SCN and in the 0.6SCN/PMS system, the residual concentration of PMS was
1SCN under identical conditions (catalyst dose 1 g/L, PMS dose 1 mM, detected at various dosage of PMS (0.5, 1, 2, 3, and 5 mM). The con
natural pH 6.12 and room temperature 25 ◦ C). Meanwhile, catalyst sumption of PMS gradually decreased with increased in PMS dosage up
0.6SCN exhibited the highest catalyst activity toward the degradation of to 3 mM, subsequently no significant changed were observed over the
acrylonitrile and acrylamide. Thus, catalyst 0.6SCN was chosen opti excess addition of PMS dosage. However, the degradation of acryloni
mum catalyst and was used in later experiments. trile and acrylamide weaker with low dosage of PMS concentration, it
might be cause of inadequate number of SO4•− radical produced at a
3.3. Effects of PMS dosage fixed dose of catalyst or self-scavenge the SO4•− radical [53]. Despite the
extensive literature on PMS activation, few attempts were made to assess
In order to determine the optimal amount of PMS dosage needed for the efficacy of PMS utilization for mineralization of acrylonitrile and
the degradation of acrylonitrile and acrylamide, various PMS dosage acrylamide. Thus, we evaluated the effectiveness of PMS utilization ef
(0.5–5 mM) were tested as shown in Fig. 5(a). Degradation of acrylo ficiency (η) estimated by following Eq. (3) for different PMS dosage.
nitrile significantly increased from 21% to 57% and acrylamide PMS utilization efficiency defined as the number of mole of pollutant
increased from 16% to 49% with increase in PMS dosage from 0.5 mM to (acrylonitrile/acrylamide) oxidized for each mole of PMS activation.
3 mM, respectively at reaction conditions (1 g/L catalyst dosage, 6.67
natural pH, 3 h reaction time and 25 ◦ C temperature). On the contrary, η=
Δ[Pollutant]degradation
(3)
further increased of PMS dosage from 3 mM to 5 mM, there was no Δ[PMS]decomposition
significant removal of acrylonitrile and acrylamide observed. The
Where Δ[Pollutant] denoted to degradation of pollutants (acryloni
excessive addition of PMS dosage inhibited the degradation of target
trile and acrylamide) in mM and Δ[PMS] denoted to PMS consumed in
pollutants, subsequently self-scavenge the generated ROS species (i.e.,
mM. The corresponding PMS utilization efficiency on various PMS
SO4•− and HO•) and produced the less reactive species perox
dosage 0.5, 1, 2, 3 and 5 mM was obtained 0.65, 0.61, 0.47, 0.42 and
ymonosulfate radicals (SO5•− ) which was weaker oxidizability then
0.26 for acrylonitrile and 0.49, 0.47, 0.40, 0.36 and 0.23 for acrylamide.
SO4•− and •OH as shown in Eq. (1 and 2) [20,22,52]. Consequently, the
PMS utilization efficiency gradually decreased for acrylonitrile and
conversion of active species SO4•− and •OH contributed to suppression
acrylamide degradation while increasing the PMS dosage, which was
of acrylonitrile and acrylamide degradation.
consistent results with previous reported work [20,53,54].
(1)
3.4. Effects of catalyst dosage
(2)
The optimization of catalyst dosage was very necessary for
Fig. 5. Effects of parameters on acrylonitrile and acrylamide removal, (a) PMS dose, (b) catalyst dose, (c) pH and (d) Temperature (Reaction conditions: catalysts
dose 0.950 g/L (a), PMS dose 3 mM (b, c and d), pH 6.12 (a, b, c and d) and temperature 55 ◦ C (d).
6
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
optimization of treatment cost of the process as well as over dosage of and acrylamide 47% to 81% while elevating the reaction temperature
catalyst produced the large quantity of metal leaching in the effluent from 25 ℃ to 55 ℃. Herewith, the consumption of PMS also enhanced
that causes create the secondary pollution. Therefore, effects of various with increase in reaction temperature from 65% (25 ℃) to 95% (55 ℃).
0.6SCN catalyst dosage (0.2, 0.45, 0.7, 0.9 and 1.2 g/L) were assessed While further increase in reaction temperature from 55 ℃ to 65 ℃, no
for degradation of acrylonitrile and acrylamide from aqueous solution. significant changes were observed on removal efficiencies and very
Fig. 5(b), show that the removal of acrylonitrile and acrylamide pro slight consumption of PMS was increased. PMS have asymmetric
gressively enhanced with raise in catalyst dosage from 0.2 to 0.9 g/L structure and generally stable at room temperature. While increasing the
thereafter on further addition of catalyst dosage over 0.9 g/L, no sig reaction temperature reactive species SO4•− and •OH progressively
nificant changes were observed at reaction conditions (i.e., PMS dosage generated [55]. PMS is unstable at high temperature, above greater than
3 mM, natural pH 6.12, temperature 25 ◦ C and reaction time 3 h). Upon 55 ℃ temperature fission of bond O − O in PMS structure easily broken
increasing of catalyst dosage enhances the active site and produce the and consequently accelerate the side reaction Eqs. (6)–(9) and scavenge
more free radical (i.e., SO4•− and HO•) by rapid decomposition of PMS, the generated reactive species SO4•− and •OH which was responsible for
which effectively enhanced the degradation of acrylonitrile and acryl degradation [56,59]. Furthermore, PMS utilization efficiencies at
amide [20,55]. The consumption of PMS progressively increased and various temperatures i.e., 25, 35, 45, 55 and 65 ℃ were estimated 2.58,
generating the more free radical species (i.e., SO4•− and HO•) which 1.34, 0.68, 0.50 and 0.29 for acrylonitrile and 2.18, 1.04, 0.57, 0.42 and
leads to enhanced the degradation of acrylonitrile and acrylamide. The 0.25 for acrylamide degradation.
maximum consumption of PMS 65% was observed with catalyst dosage
(6)
0.9 g/L and further increase of catalyst dosage upto 1.2 g/L, PMS con
sumption increased very slightly. Moreover, the PMS utilization effi (7)
ciencies on different catalyst dosage 0.2, 0.45, 0.7, 0.9 and 1.2 g/L was
obtained 1.09, 0.90, 0.63, 0.43 and 0.25 for acrylonitrile and 0.98, 0.77, (8)
0.54, 0.36 and 0.21 for acrylamide degradation. The maximum utiliza
tion efficiency was observed at low catalyst dosage and gradually (9)
decreased with increase the catalyst dosage. Many authors have been
reported the similar decreasing trend for utilizing efficiency [20,53].
3.7. Effect of co-existing ions
3.5. Effects of pH In the real wastewater various kind of inorganic ions present viz.,
chloride (Cl− ), carbonate (CO32–), and nitrate (NO3− ) which affect the
Various researcher has been demonstrated that pH value of the so performance of pollutants degradation in sulfate radical based AOPs
lution strongly influences on the generation of ROS species in hetero through multitudinous ways including capturing the in-situ generated
geneous PMS activation based process [20,56,57]. Accordingly, a series ROS species and influencing the decomposition of PMS molecule [60-
of experiments conducted with 0.6SCN catalyst at various adjusted pHs 62]. Therefore, the investigation on interferences of acrylonitrile and
2, 4, 6.12, 8 and 10 to explore the degradation of acrylonitrile and acrylamide degradation over the 0.6SCN/PMS system was very neces
acrylamide. As shown in Fig. 5(c), the maximum degradation of acry sary. The effects of typical inorganic ions were evaluated on acrylonitrile
lonitrile (56%) and acrylamide (47%) were observed with maximum and acrylamide degradation by adding various ion concentrations (0, 5
PMS consumption (65%) at natural pH (6.12). While adjusting the so and 50 mM) separately added into 0.6SCN/PMS system as showed re
lution pH in alkaline or acidic medium, removal efficiencies of acrylo sults in Table S1 (in Supporting Information). As shown in Fig. 6(a-b),
nitrile and acrylamide suppressed and simultaneously decreased the the addition of Cl− ions in 0.6SCN/PMS system at optimum operating
PMS consumption. In the strong acidic medium, dominant species H+ conditions had a greatly positive effects on acrylonitrile and acrylamide
ions quench the generated reactive oxygen species (SO4•-) and (•OH) degradation. The acrylonitrile degradation enhanced from 85% to 99%
according to reaction Eqs. (4 & 5). The another reason for reduction of and acrylamide degradation 64% to 94% with the Cl− dosage increased
removal efficiency in highly acidic medium was high metal leaching (Cu from 0 to 50 mM. The complete degradation of acrylonitrile was ach
and Ni) [58]. ieved within 1 h at 50 mM of Cl− ions dosage. However, the rate constant
(4) i.e., 7 × 10− 7, 9.4 × 10− 7 and 51.2 × 10− 7 M− 1s− 1 for acrylonitrile and i.
e., 2 × 10− 7, 13 × 10− 7 and 28 × 10− 7 M− 1s− 1 for acrylamide was
OH + H+ + e- → H2O (5) obtained by second order degradation kinetic model which was
•
7
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
Fig. 6. Effects of (a-b) Cl– ions concentration, (c-d) HCO3– ions concentration and (e-f) NO3– ions concentration, on acrylonitrile and acrylamide removal. Reaction
conditions: PMS dose 3 mM, catalysts dose 0.950 g/L, pH 6.12, and temperature 55 ◦ C.
removal efficiencies of acrylonitrile and acrylamide as depicted in Fig. 6 Moreover, another negative impact of HCO3− ions on acrylonitrile
(c-d). The rate of degradation suppressed with rate constant i.e., 7 × and acrylamide degradation was raised the pH of the solution while
10− 7, 6 × 10− 7 and 5 × 10− 7 M− 1s− 1 for acrylonitrile and i.e., 2.5 × adding the HCO3− ions, which leads to suppressed the degradation ef
10− 7, 2.3 × 10− 7 and 2.1 × 10− 7 M− 1s− 1 for acrylamide obtained by ficiencies [60]. Furthermore, the effects of NO3− ions on acrylonitrile
second order degradation kinetic model. The inhibitory effects of acry and acrylamide degradation were also evaluated as shown in Fig. 6(e-f).
lonitrile and acrylamide degradation might be cause of scavenging of The degradation efficiency inhibited from 85% to 81% for acrylonitrile
SO4•− and •OH radicals by HCO3− as via Eqs. (15) and (16) [63,64]. and 64% to 61% for acrylamide in the existing of 5 mM NO3− dosage and
further dropped to 72% and 54% for acrylonitrile and acrylamide while
(15)
increasing the NO3− dosage upto 50 mM. However, the rate constant i.e.,
7 × 10− 7, 6 × 10− 7 and 4 × 10− 7 M− 1s− 1 for acrylonitrile and i.e., 2.5 ×
HCO3− + OH → CO3− + H2O (16)
•
10− 7, 2.4 × 10− 7 and 2.2 × 10− 7 M− 1s− 1 for acrylamide was obtained as
8
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
depicted in Fig. 6(f). The rate of acrylonitrile and acrylamide degrada optimum operating conditions e.g., 3 mM PMS, 950 mg/L 0.6SCN dose,
tion was significantly dropped due to formation of lower potential 6.12 pH, 55 ⁰C reaction temperature and 3-hour reaction time, as results
species i.e., NO3• and NO2• via Eqs. (17)–(19), which have lower stan shown in Table S2 (Supporting Information). The above results sug
dard redox potential E0(NO3•/ NO3− ) = 2.30 V and E0(NO2•/ NO3− ) = gested that EtOH showed much more significant inhibition effects on
1.03 V. acrylonitrile and acrylamide degradation compare than both LH and
TBA. Therefore, influences of high dosage of EtOH (10–200 mM) on
(17)
acrylonitrile and acrylamide were also explored at the same optimum
(18) operating condition. The acrylonitrile degradation inhibited from 66%
(EtOH, 10 mM) to 35% (EtOH, 200 mM) and acrylamide degradation
(19) inhibited from 58% (EtOH, 10 mM) to 35% (EtOH, 200 mM). Authors
Yan et al., reported that sulfamethoxazole degradation inhibited more
than 52% in the presence of 500 mM EtOH and more than 72% sup
3.8. Radical scavenger study pressed at 500 mM TBA and similarly authors Wu et al., reported the
carbamazepine degradation suppressed to 84% and 15% while using the
Generally, in heterogeneous catalysis with PMS activation reaction 100 mM TBA and 100 mM EtOH dosage, respectively [60,62]. It has
various types of reactive oxygen species in-situ generated viz., sulfate
been demonstrated from literature, EtOH rapidly react with either SO4•–
radicals (SO4•–) and hydroxyl radical (•OH), and singlet oxygen (1O2). or •OH radical having rate constant k = 1.6–7.7 × 107 M− 1s− 1 for SO4•–
Accordingly, to determining the main ROS species in 0.6SCN/PMS sys
and k = 1.2–2.8 × 109 M− 1s− 1 for •OH while TBA more rapidly react
tem a series of quenching experiments were carried out at optimum about 1000-fold faster rate constant (k = 3.8–7.6 × 108 M− 1s− 1) with
operating condition with their respective scavengers e.g., ethanol •
OH radical than SO4•– radical (k = 4–9.1 × 105 M− 1s− 1) [60,62,63,66].
(EtOH) for both SO4•– and •OH radical quencher, TBA for only •OH In the present study, degradation kinetics with rate constants k = 340 ×
radical quencher, and L-histidine for 1O2 [62,65,66]. As shown in Fig. 7
108 M− 1s− 1 for acrylonitrile and k = 110 × 108 M− 1s− 1 for acrylamide
(a-b), 96% of acrylonitrile and 81% of acrylamide degradation were were estimated when no scavenger was used. While, rate constants k =
observed while no scavenger dosage was introduced. The degradation of
16 × 108 M− 1s− 1, 10 × 108 M− 1s− 1 and 7.8 × 108 M− 1s− 1 were esti
acrylonitrile suppressed from 96% to 85%, 69%, and 66%; and acryl mated for acrylonitrile degradation and k = 30 × 108 M− 1s− 1, 10 × 108
amide suppressed from 81% to 74%, 63%, and 58% when 10 mM of each
M− 1s− 1 and 10 × 108 M− 1s− 1 were estimated for acrylamide
LH, TBA and EtOH were introduced separately in 0.6SCN/PMS system at
Fig. 7. Effect of quenching scavengers on, (a) acrylonitrile and (b) acrylamide degradation; and (c) EPR spectra of DMPO-OH•– and DMPO-SO4•–. Reaction con
ditions: PMS dose 3 mM, catalysts dose 0.950 g/L, pH 6.12, and temperature 55 ◦ C.
9
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
degradation with 10 mM of each LH, TBA and EtOH, respectively. reduction reaction and contribute the role in acrylonitrile and acryl
Moreover, degradation rate suppressed while enhancing the EtOH amide degradation. All these observations were well accord with the
dosage from k = 7.8 × 108 M− 1s− 1 (10 mM) to k = 3 × 108 M− 1s− 1 (200 results of quenching experiments.
mM) for acrylonitrile and k = 10 × 108 M− 1s− 1 (10 mM) to k = 3 × 108
M− 1s− 1 (200 mM) for acrylamide, respectively. Hence, the above find
ings concluded that both reactive species SO4•– and •OH contributed role
in the degradation of acrylonitrile and acrylamide.
Table 1
The contributions of ROS species (SO4•– and •OH) were further Two-step degradation kinetics for acrylonitrile and acrylamide at various
validated by EPR study for the degradation of acrylonitrile and acryl temperatures.
amide, EPR study was conducted with EtOH (10 mM and 100 mM) and
Acrylonitrile degradation kinetics
10 mM of 5,5-dimethyl-1-pyrollidone N-oxide (DMPO) use as a spin
trapping agent at optimum operating conditions e.g., 3 mM PMS, 950 Temperature Fast Step Slow Step
(K)
mg/L 0.6SCN dose, 6.12 pH, 55 ⁰C reaction temperature. As shown in
k1 R2 Ea (kJ k2 R2 Ea (kJ
Fig. 7(c), very high intense signal of DMPO/SO4•– and DMPO/•OH (min− 1) mol− 1) (min− 1) mol− 1)
adduct were observed when low dosage of EtOH 10 mM was introduced
298 9.5 × 0.989 19.78 2.8 × 0.930 32.61
in aqueous solution. The hyperfine splitting constants of aN = 13.7 G, aH 10-3 10-3
= 10.3 G, aH = 1.48 G, and aH = 0.76 G are accordance with DMPO/ 308 11.7 × 0.987 5.3 × 0.851
SO4•– adducts, while the hyperfine splitting constant of aN = 14.8 G, aH 10-3 10-3
= 15.2 G are consistent with DMPO/•OH adduct. High intense signal of 318 14.9 × 0.993 7.5 × 0.983
10-3 10-3
DMPO/•OH adducts indicated that generated SO4•– radicals rapidly
328 19.4 × 0.995 11.6 × 0.975
transformed in to •OH radical via Eqs. (20) and (21). 10-3 10-3
338 24.1 × 0.993 13.7 × 0.991
10-3 10-3
(20) Acrylamide degradation kinetics
298 7.8 × 0.987 11.59 1.2 × 0.984 28.43
(21) 10-3 10-3
308 8.9 × 0.993 1.8 × 0.982
Various authors have been reported the similar type of results [67-
10-3 10-3
73]. Furthermore, weak signal of DMPO/SO4•– and DMPO/•OH adduct 318 10.6 × 0.994 2.4 × 0.979
were observed while high dosage of EtOH (100 mM) used in reaction 10-3 10-3
solution. It might be caused of quenching of SO4•– and •OH radicals by 328 12.2 × 0.987 3.8 × 0.812
10-3 10-3
using high dosage of EtOH (100 mM) and some of radicals consumed by
338 13.3 × 0.981 4.5 × 0.790
target pollutant [20,74]. This phenomenon suggested that both radicals 10-3 10-3
SO4•– and •OH simultaneously generated during the oxidation/
Fig. 8. Two-step degradation kinetics at various temperature with their activation energy for (a) acrylonitrile and (b) acrylamide; (c) reusability study of 0.6SCN
catalyst and (d) effect of pH with Cu and Ni leaching at optimum operating conditions.
10
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
The degradation kinetics of acrylonitrile and acrylamide were well The reusability of catalyst is of great significance for practical
fitted by first-order kinetics model with two-stage degradation at opti application. The reusability potential and stability of 0.6SCN catalyst
mum operating conditions with various temperature (e.g., 298, 308, were assessed over the four consecutive cycles under optimum operating
318, 328 and 338 K). First order reaction kinetics model can be conditions (i.e., 0.6SCN catalyst dosage 950 mg/L, PMS dosage 3 mM,
expressed by Eq. (22). reaction temperature 55 ◦ C, pH 6.12 and time 3 h) in 0.6SCN/PMS
system as results depicted in Fig. 8(c) and Table S3 (Supporting Infor
ln(C0 /Ct ) = kt (22)
mation). The maximum degradation of acrylamide 81% and acryloni
Where, k denotes to rate constant, C0 and Ct denote to initial and trile 95% were observed in the first cycle. At the end of each cycle, spent
final concentration of acrylonitrile and acrylamide. As shown in Fig. 8(a- catalyst was subjected to centrifugal separation and subsequently gently
b), acrylonitrile and acrylamide degradation displayed bilinear plots rinsed several times in Milli pore water and then dried at 120 ◦ C for 12 h.
which indicated that more than one process controlling the oxidation Regenerated catalysts were further used in next cycle and remarkable
reaction. In the fast stage, target pollutants oxidized into intermediates results were obtained. The maximum removal of acrylamide i.e., 76%,
and intermediates were mineralized into CO2 and H2O [75,76]. As 74% and 70%; and acrylonitrile i.e., 94%, 91% and 89% were observed
shown in Table 1., rate of acrylonitrile degradation progressively during consecutively cycles 2nd, 3rd and 4th. Removal efficiencies
enhanced from k1 = 9.5 × 10-3 min− 1 to k1 = 24.1 × 10-3 min− 1 in fast gradually decreased in each cycle might be causes of several reasons: (a)
step and k2 = 2.8 × 10-3 min− 1 to k2 = 13.7 × 10-3 min− 1 in slow step leaching of active metal ions copper and nickel which leads to sup
while reaction temperature raises from 298 K to 338 K. Similarly, pressed the active sites from the catalysts surface, and (b) blockage of
acrylamide degradation rate progressively enhanced from k1 7.8 × 10-3 active sites on the catalyst surface due to carbon deposition (carbon
min− 1 to 13.3 × 10-3 min− 1 in fast step and k2 1.2 × 10-3 min− 1 to 4.5 × present in the form of intermediate products) [22]. Considering the
10-3 min− 1 in slow step while elevation of reaction temperature from previous facts, metals leaching was evaluated by ICP-MS technique in
298 K to 338 K. However, the rate of degradation of the fast stage (k1) each cycle of experiments. The leaching of Cu and Ni in first cycle was
was much higher than that of the slow stage (k2). The determined values observed 1.89 mg/L and 1.24 mg/L subsequently leaching of Cu ions
of the rate constants k1 and k2 were then subsequently used to activation dropped to e.g., 6%, 33% and 41%; and Ni leaching dropped to e.g.,
energies assessment through Arrhenius equation as follows Eq. (23). 20%, 52% and 72% in second, third and fourth cycle consecutively. The
present values of metal ions leaching followed the discharge permissible
Ea 1
− lnk = − lnA (23) limit standard of Central Pollution Control Board, India [81]. Various
R T
authors have been observed very low metal ions leaching in perovskite-
Where, Ea, A and k represent the activation energy, pre-exponential like catalysts during the degradation experiments and also reported very
factor and rate constant at temperature T, respectively. The activation high stability which does not lost their catalytic activity during multiple
energy for acrylonitrile degradation was estimated 19.78 kJ/mol in fast degradation cycles [35,82]. The chemical stability of catalyst was also
stage and 32.61 kJ/mol in slow stage; and similarly for acrylamide strongly affected by reaction pH [83,84]. Meanwhile, the chemical
removal activation energies obtained 11.59 kJ/mol in fast stage and stability of 0.6SCN catalyst was also assessed at the same optimum
28.43 kJ/mol in slow stage. Similar values of activation energies be conditions and different pHs e.g., 2, 4, 6.12, 8 and 10. As shown in Fig. 8
tween 10.36 and 34.48 kJ/mol have been reported in our preceding (d), very high leaching of Ni and Cu were observed in strong acidic
research for acrylonitrile removal from aqueous solution by heteroge medium and correspondingly very low removal of acrylonitrile and
neous catalysis with PMS activation process and catalytic wet peroxi acrylamide were observed. With increased in reaction pH from 2 to 6.12
dation process [39,77]. Likewise, many authors reported the similar leaching of Cu and Ni gradually decreased and correspondingly activity
values of activation energies i.e., 41.29 kJ/mol for pyrrole, 29.4 kJ/mol of catalyst was enhanced and maximum removal of acrylonitrile (96%)
for quinoline and19.10 kJ/mol for pyridine degradation from aqueous and acrylamide (81%) were observed. Authors Krol et al., also reported
solution, respectively [78-80]. very high leaching of Cu and Ni in strong acidic medium [84]. Authors
Qin et al., reported that leaching of iron ions gradually decreased e.g.,
2.12%, 1.88%, 1.65%, and 1.28% with increase in corresponding pHs i.
e., 2.1, 3.8, 5.8 and 10.2, respectively catalysis with Fe/ACP for benzoic
Fig. 9. Proposed reaction pathways of acrylonitrile and acrylamide degradation with PMS induced SO4•– and •OH radical mechanism.
11
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
acid degradation [83]. Authors Lei et al., reported the similar findings as
very high leaching of Cu (~35 mg/L) observed in strong acidic medium
(at pH 2.5). Cu leaching gradually suppressed with raise in pH from 2.5
to 10 subsequently no leaching was observed at pHs 11 and 12 [85].
(30)
Thus, above results elucidated that catalytic activity and stability of
(31)
Based on the EPR test, quenching experiments and XPS analysis, a (33)
possible PMS activation mechanism has been proposed for mineraliza
tion of acrylonitrile and acrylamide in 0.6SCN/PMS system as shown in
(34)
Fig. 9. Initially, when the catalyst dosage dispersed into the solution,
H2O molecule physically adsorbed on the active sites offered by 0.6SCN
catalyst to create ≡ Cu+ /Cu2+ − OH and ≡ Ni2+ /Ni3+ − OH on the In addition, the degradation pathways of acrylonitrile and acryl
catalyst surface. Thereafter, adding the PMS dosage, formed amide elucidated by intermediates identified by GC/MS in 0.6SCN/PMS
≡ Cu+ /Cu2+ − OH and ≡ Ni2+ /Ni3+ − OH species further react with PMS system and literature studies [13]. Various types of intermediate prod
(HSO5–) and produced the SO4•– radical species along with SO5•– via ucts were detected during GC/MS analysis i.e., propionic acid, d-
Eqs. (24)–(26) [66]. Besides, •OH radical also produced through the alanine, propenoic acid, acetamide, dimethylamine, and acetic acid etc.,
reaction between ≡ Cu+ /Ni3+ and PMS on the catalysts surface via Eq. as shown in Figs. S3 (in Supplementary Information). As shown in Fig. 8,
(27) [70]. Furthermore, water adsorbed active species react with HSO5– generated reactive species SO4•– and •OH react with acrylonitrile to
to produce the sulfate complexes ≡ Cu+ /Cu2+ − (OH)OSO−3 and produce intermediates dimethylamine, propionic acid and alanine
≡ Ni2+ /Ni3+ − (OH)OSO−3 which could also produce the SO4•– radical which was further oxidized into acetic acid. In another pathway, acry
species via Eqs. (29)–(32) [66,86,87]. Meanwhile, the oxidation and lonitrile decomposes into acetamide which was then oxidized into acetic
acid by deamination and subsequently mineralized into CO2 and H2O.
reduction cycle of ≡ Cu+ /Cu2+ and ≡ Ni2+ /Ni3+ simultaneously carried
Moreover, reactive species SO4•– and •OH react with acrylonitrile
out via Eq. (33) and continuously produce the reactive species [1].
(propanamide) to produce intermediates propenoic acid and acetic acid.
Generated reactive species oxidized the compounds acrylonitrile and
Intermediate product like acetic acid was finally mineralized into CO2
acrylamide into intermediate products which was again oxidized in to
and H2O in all the reaction pathways [88].
CO2 and H2O via Eq. (34).
(24)
(25)
(29)
12
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
13
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
[15] R.F. de Sena, R.d.F.P.M. Moreira, H.J. José, Assessment of polyacrylamide contaminant of emergent concern-cotinine- in aqueous medium using the magnetic
degradation using advanced oxidation processes and ferrate(Vi) oxidation, Chem. double perovskite oxide Sr2FeCuO6 as a highly stable catalayst: Degradation
Eng. Commun. 200 (2) (2013) 235–252. kinetics and oxidation products, Appl. Catal. B Environ. 240 (2019) 201–214.
[16] Y.G. Maksimova, D.M. Vasil’ev, A.S. Zorina, G. V Ovechkina, A.Y. Maksimov, [41] Wen-Da Oh, Shun-Kuang Lua, Zhili Dong, Teik-Thye Lim, High surface area DPA-
Acrylamide and acrylic acid biodegradation by Alcaligenes faecalis 2 planktonic hematite for efficient detoxification of bisphenol A via peroxymonosulfate
cells and biofilms, Appl. Biochem. Microbiol. 54 (2018) 158–164. activation, J. Mater. Chem. A. 2 (38) (2014) 15836–15845.
[17] M. Lakshmikandan, K. Sivaraman, S. Elaiya Raja, P. Vasantha kumar, R.P. Rajesh, [42] Nina Perkas, Poernomo Gunawan, Galina Amirian, Zhan Wang, Ziyi Zhong,
K. Sowparthani, S.E. Joshua Jebasingh, Biodegradation of acrylamide by Aharon Gedanken, The sonochemical approach improves the CuO-ZnO/TiO2
acrylamidase from Stenotrophomonas acidaminiphila MSU12 and analysis of catalyst for WGS reaction, Phys. Chem. Chem. Phys. 16 (16) (2014) 7521–7530.
degradation products by MALDI-TOF and HPLC, Int. Biodeterior. Biodegrad. 94 [43] Chengjun Dong, Xuechun Xiao, Gang Chen, Hongtao Guan, Yude Wang, Applied
(2014) 214–221. surface science morphology control of porous CuO by surfactant using combustion
[18] Young Ho Shin, Nae Chul Shin, Bambang Veriansyah, Jaehoon Kim, Youn- method, Appl. Surf. Sci. 349 (2015) 844–848.
Woo Lee, Supercritical water oxidation of wastewater from acrylonitrile [44] Ashish Aphale, Md Aman Uddin, Boxun Hu, Su Jeong Heo, Junsung Hong,
manufacturing plant, J. Hazard. Mater. 163 (2-3) (2009) 1142–1147. Prabhakar Singh, Synthesis and stability of SrxNiyOz chromium getter for solid
[19] Bo Lai, Yuexi Zhou, Ping Yang, Jinghui Yang, Juling Wang, Degradation of 3,3’- oxide fuel cells, J. Electrochem. Soc. 165 (9) (2018) F635–F640.
iminobis-propanenitrile in aqueous solution by Fe0/GAC micro-electrolysis system, [45] A. Kumar, B. Prasad, Catalytic peroxidation of acrylonitrile aqueous solution by Ni-
Chemosphere 90 (4) (2013) 1470–1477. doped CeO2 catalysts: characterization, kinetics and thermodynamics, Int. J.
[20] S. Zhu, Y. Xu, Z. Zhu, Z. Liu, W. Wang, Activation of peroxymonosulfate by Environ. Sci. Technol. 17 (2020) 1809–1824.
magnetic Co-Fe/SiO2 layered catalyst derived from iron sludge for ciprofloxacin [46] Huihuang Chen, Julius Motuzas, Wayde Martens, João C. Diniz da Costa,
degradation, Chem. Eng. J. 384 (2020), 123298. Degradation of azo dye Orange II under dark ambient conditions by calcium
[21] Farshid Ghanbari, Mahsa Moradi, Application of peroxymonosulfate and its strontium copper perovskite, Appl. Catal. B Environ. 221 (2018) 691–700.
activation methods for degradation of environmental organic pollutants: Review, [47] H. Chen, Engineering SrCuxO composition to tailor the degradation activity toward
Chem. Eng. J. 310 (2017) 41–62. organic pollutant under dark ambient conditions, (2019) 16449–16456.
[22] R. Zhang, Y. Wan, J. Peng, G. Yao, Y. Zhang, B. Lai, Efficient degradation of [48] Jun Li, Yi Ren, Fangzhou Ji, Bo Lai, Heterogeneous catalytic oxidation for the
atrazine by LaCoO3/Al2O3 catalyzed peroxymonosulfate: Performance, degradation of p-nitrophenol in aqueous solution by persulfate activated with
degradation intermediates and mechanism, Chem. Eng. J. 372 (2019) 796–808. CuFe2O4 magnetic nano-particles, Chem. Eng. J. 324 (2017) 63–73.
[23] J. Sharma, I.M. Mishra, D.D. Dionysiou, V. Kumar, Oxidative removal of bisphenol [49] Jerzy A. Mielczarski, Gonzalo Montes Atenas, Ela Mielczarski, Role of iron surface
a by UV-C/peroxymonosulfate (PMS): Kinetics, influence of co-existing chemicals oxidation layers in decomposition of azo-dye water pollutants in weak acidic
and degradation pathway, Chem. Eng. J. 276 (2015) 193–204. solutions, Appl. Catal. B Environ. 56 (4) (2005) 289–303.
[24] Songshan Jiang, Huiping Zhang, Ying Yan, Xinya Zhang, Stability and deactivation [50] H. Chen, J. Motuzas, W. Martens, J.C. Diniz, Separation and Puri fi cation
of Fe-ZSM-5 zeolite catalyst for catalytic wet peroxide oxidation of phenol in a Technology Degradation of orange II dye under dark ambient conditions by
membrane reactor, RSC Adv. 5 (51) (2015) 41269–41277. MeSrCuO (Me = Mg and Ce) metal oxides, 205 (2018) 293–301.
[25] Fassi Soumia, Christian Petrier, Effect of potassium monopersulfate (oxone) and [51] Chao Su, Xiaoguang Duan, Jie Miao, Yijun Zhong, Wei Zhou, Shaobin Wang,
operating parameters on sonochemical degradation of cationic dye in an aqueous Zongping Shao, Mixed conducting perovskite materials as superior catalysts for fast
solution, Ultrason. Sonochem. 32 (2016) 343–347. aqueous-phase advanced oxidation: A mechanistic study, ACS Catal. 7 (1) (2017)
[26] Shaoxia Yang, Michèle Besson, Claude Descorme, Catalytic wet air oxidation of 388–397.
formic acid over Pt/CexZr1− xO2 catalysts at low temperature and atmospheric [52] Chaoqun Tan, Naiyun Gao, Yang Deng, Jing Deng, Shiqing Zhou, Jun Li,
pressure, Appl. Catal. B Environ. 100 (1-2) (2010) 282–288. Xiaoyan Xin, Radical induced degradation of acetaminophen with Fe3O4 magnetic
[27] P. Hu, H. Su, Z. Chen, C. Yu, Q. Li, B. Zhou, P.J.J. Alvarez, M. Long, Selective nanoparticles as heterogeneous activator of peroxymonosulfate, J. Hazard. Mater.
degradation of organic pollutants using an efficient metal-free catalyst derived 276 (2014) 452–460.
from carbonized polypyrrole via peroxymonosulfate activation, Environ. Sci. [53] Yong Feng, Po-Heng Lee, Deli Wu, Kaimin Shih, Surface-bound sulfate radical-
Technol. 51 (2017) 43. dominated degradation of 1,4-dioxane by alumina-supported palladium (Pd/
[28] Xiaodi Duan, Shanshan Yang, Stanisław Wacławek, Guodong Fang, Ruiyang Xiao, Al2O3) catalyzed peroxymonosulfate, Water Res. 120 (2017) 12–21.
Dionysios D. Dionysiou, Limitations and prospects of sulfate-radical based [54] S. Wang, J. Wang, A novel strategy of successive non-radical and radical process for
advanced oxidation processes, J. Environ. Chem. Eng. 8 (4) (2020) 103849, enhancing the utilization efficiency of persulfate, Chemosphere 245 (2020),
https://doi.org/10.1016/j.jece:2020.103849. 125555.
[29] Adarlêne M. Silva, Rosa M.F. Lima, Versiane A. Leão, Mine water treatment with [55] Feng Jiang, Bin Qiu, Dezhi Sun, Advanced degradation of refractory pollutants in
limestone for sulfate removal, J. Hazard. Mater. 221-222 (2012) 45–55. incineration leachate by UV/Peroxymonosulfate, Chem. Eng. J. 349 (2018)
[30] M. Mori, Y. Itagaki, Y. Sadaoka, Effect of VOC on ozone detection using 338–346.
semiconducting sensor with SmFe1-xCoxO3 perovskite-type oxides, Sensors [56] J. Cao, L. Lai, B. Lai, G. Yao, X. Chen, L. Song, Degradation of tetracycline by
Actuators, B Chem. 163 (2012) 44–50. peroxymonosulfate activated with zero- valent iron: Performance, intermediates,
[31] Arvind Kumar, Basheswar Prasad, Krishan Kishor Garg, Catalytic peroxidation of toxicity and mechanism, Chem. Eng. J. 364 (2019) 45–56.
acrylic acid from aqueous solution incorporated with highly active La0.5Sr0.5BO3 [57] A. Kumar, B. Prasad, V.K. Sandhwar, K.K. Garg, Mechanistic Insight into
(B=Cu, Fe and Ni) perovskite-like catalysts, J. Environ. Heal. Sci. Eng. 18 (2) Heterogeneous Fenton-like Catalysis with M-Al2O3/SiO2 (M = Fe, Co and Ni) for
(2020) 897–913. Acrylonitrile Mineralization from Real ABS Resin Wastewater: Optimization and
[32] A. Weidenkaff, Preparation and Application of Nanostructured Perovskite Phases, toxicity assessment, J. Environ. Chem. Eng. 9 (2021), 105177.
Adv. Eng. Mater. 6 (9) (2004) 709–714. [58] Z. Yang, Y. Li, X. Zhang, X. Cui, S. He, H. Liang, A. Ding, Sludge activated carbon-
[33] Ji-Jing Xu, Dan Xu, Zhong-Li Wang, Heng-Guo Wang, Lei-Lei Zhang, Xin-Bo Zhang, based CoFe2O4-SAC nanocomposites used as heterogeneous catalysts for degrading
Synthesis of perovskite-based porous La0.75Sr0.25MnO3 Nanotubes as a highly antibiotic norfloxacin through activating peroxymonosulfate, Chem. Eng. J. 384
efficient electrocatalyst for rechargeable lithium-oxygen batteries, Angew. Chemie (2020), 123319.
Int. Ed. 52 (14) (2013) 3887–3890. [59] Jianlong Wang, Shizong Wang, Activation of persulfate (PS) and
[34] J. Miao, J. Sunarso, X. Duan, W. Zhou, S. Wang, Z. Shao, Nanostructured Co-Mn peroxymonosulfate (PMS) and application for the degradation of emerging
containing perovskites for degradation of pollutants: Insight into the activity and contaminants, Chem. Eng. J. 334 (2018) 1502–1517.
stability, J. Hazard. Mater. 349 (2018) 177–185. [60] Z. Wu, Y. Wang, Z. Xiong, Z. Ao, S. Pu, G. Yao, B. Lai, Core-shell magnetic Fe3O4@
[35] Oxana P. Taran, Artemiy B. Ayusheev, Olga L. Ogorodnikova, Igor P. Prosvirin, Zn/Co-ZIFs to activate peroxymonosulfate for highly efficient degradation of
Lyubov A. Isupova, Valentin N. Parmon, Perovskite-like catalysts LaBO3 (B = Cu, carbamazepine, Appl. Catal. B Environ. 277 (2020), 119136.
Fe, Mn Co, Ni) for wet peroxide oxidation of phenol, Appl. Catal. B Environ. 180 [61] A. Jawad, J. Lang, Z. Liao, A. Khan, J. Ifthikar, Z. Lv, S. Long, Z. Chen, Z. Chen,
(2016) 86–93. Activation of persulfate by CuOx@Co-LDH: A novel heterogeneous system for
[36] Samia Ben Hammouda, Feiping Zhao, Zahra Safaei, Varsha Srivastava, contaminant degradation with broad pH window and controlled leaching, Chem.
Deepika Lakshmi Ramasamy, Sidra Iftekhar, Simo kalliola, Mika Sillanpää, Eng. J. 335 (2018) 548–559.
Degradation and mineralization of phenol in aqueous medium by heterogeneous [62] J. Yan, J. Li, J. Peng, H. Zhang, Y. Zhang, B. Lai, Efficient degradation of
monopersulfate activation on nanostructured cobalt based-perovskite catalysts sulfamethoxazole by the CuO@Al2O3 (EPC) coupled PMS system: Optimization,
ACoO3 (A = La, Ba, Sr and Ce): Characterization, kinetics and mechanism study, degradation pathways and toxicity evaluation, Chem. Eng. J. 359 (2019)
Appl. Catal. B Environ. 215 (2017) 60–73. 1097–1110.
[37] Kun-Yi Andrew Lin, Yu-Chien Chen, Yi-Feng Lin, LaMO3 perovskites (M=Co, Cu, Fe [63] J. Peng, X. Lu, X. Jiang, Y. Zhang, Q. Chen, B. Lai, G. Yao, Degradation of atrazine
and Ni) as heterogeneous catalysts for activating peroxymonosulfate in water, by persulfate activation with copper sulfide (CuS): Kinetics study, degradation
Chem. Eng. Sci. 160 (2017) 96–105. pathways and mechanism, Chem. Eng. J. 354 (2018) 740–752.
[38] Rafael R. Solís, F. Javier Rivas, Olga Gimeno, Removal of aqueous metazachlor, [64] Z. Shen, H. Zhou, Z. Pan, Y. Guo, Y. Yuan, G. Yao, B. Lai, Degradation of atrazine by
tembotrione, tritosulfuron and ethofumesate by heterogeneous monopersulfate Bi2MoO6 activated peroxymonosulfate under visible light irradiation, J. Hazard.
decomposition on lanthanum-cobalt perovskites, Appl. Catal. B Environ. 200 Mater. 400 (2020), 123187.
(2017) 83–92. [65] J. You, W. Sun, S. Su, Z. Ao, C. Liu, G. Yao, B. Lai, Degradation of bisphenol A by
[39] A. Kumar, B. Prasad, K.K. Garg, Efficiently degradation of acrylonitrile from peroxymonosulfate activated with oxygen vacancy modified nano-NiO-ZnO
aqueous solution by La0.5Ce0.5MO3 (M = Fe, Cu and Co) perovskite-like catalyst: composite oxides: A typical surface-bound radical system, Chem. Eng. J. 400
Optimization and reaction pathways, J. Water Process Eng. 36 (2020) 162–180. (2020), 125915.
[40] S. Ben Hammouda, C. Salazar, F. Zhao, D.L. Ramasamy, E. Laklova, S. Iftekhar, [66] M. Xu, H. Zhou, Z. Wu, N. Li, Z. Xiong, G. Yao, B. Lai, Efficient degradation of
I. Babu, M. Sillanpää, Efficient heterogeneous electro-Fenton incineration of a sulfamethoxazole by NiCo2O4 modified expanded graphite activated
14
A. Kumar and B. Prasad Separation and Purification Technology 279 (2021) 119677
peroxymonosulfate: Characterization, mechanism and degradation intermediates, [78] S. Garg, V.C. Srivastava, S. Singh, T.K. Mandal, Catalytic Degradation of Pyrrole in
J. Hazard. Mater. 399 (2020). Aqueous Solution by Cu/SBA-15, Int. J. Chem. React. Eng. 13 (2015) 437–445.
[67] L. Liu, H. Mi, M. Zhang, F. Sun, R. Zhan, H. Zhao, S. He, L. Zhou, Efficient [79] V. Gosu, A. Dhakar, P. Sikarwar, U.K.A. Kumar, V. Subbaramaiah, T.C. Zhang, Wet
moxifloxacin degradation by CoFe2O4 magnetic nanoparticles activated peroxidation of resorcinol catalyzed by copper impregnated granular activated
peroxymonosulfate: Kinetics, pathways and mechanisms, Chem. Eng. J. 127201 carbon, J. Environ. Manage. 223 (2018) 825–833.
(2020). [80] Nilam J. Pachupate, Prakash D. Vaidya, Catalytic wet oxidation of quinoline over
[68] Y. Wang, Z. Chi, C. Chen, C. Su, D. Liu, Y. Liu, X. Duan, S. Wang, Facet- and defect- Ru/C catalyst, J. Environ. Chem. Eng. 6 (1) (2018) 883–889.
dependent activity of perovskites in catalytic evolution of sulfate radicals, Appl. [81] Central Pollution Control Board (CPCB), Pollution Control Acts, and notification
Catal. B Environ. 272 (2020), 118972. under CPCB, India (2001) https://www.cpcb.nic.
[69] W. Qin, G. Fang, Y. Wang, D. Zhou, Mechanistic understanding of polychlorinated [82] Huiyuan Zhu, Pengfei Zhang, Sheng Dai, Recent Advances of lanthanum-based
biphenyls degradation by peroxymonosulfate activated with CuFe2O4 perovskite oxides for catalysis, ACS Catal. 5 (11) (2015) 6370–6385.
nanoparticles: Key role of superoxide radicals, Chem. Eng. J. 348 (2018) 526–534. [83] H. Qin, R. Xiao, J. Chen, Catalytic wet peroxide oxidation of benzoic acid over Fe/
[70] J. Peng, H. Zhou, W. Liu, Z. Ao, H. Ji, Y. Liu, S. Su, G. Yao, B. Lai, Insights into AC catalysts: Effect of nitrogen and sulfur co-doped activated carbon, Sci. Total
heterogeneous catalytic activation of peroxymonosulfate by natural chalcopyrite: Environ. 626 (2018) 1414–1420.
pH-dependent radical generation, degradation pathway and mechanism, Chem. [84] Anna Król, Kamila Mizerna, Marta Bożym, An assessment of pH-dependent release
Eng. J. 397 (2020), 125387. and mobility of heavy metals from metallurgical slag, J. Hazard. Mater. 384 (2020)
[71] Wei Xiang, Mingjie Huang, Yifan Wang, Xiaohui Wu, Fugang Zhang, Dan Li, 121502, https://doi.org/10.1016/j.jhazmat.2019.121502.
Tao Zhou, New insight in the O2 activation by nano Fe/Cu bimetals: The synergistic [85] Y. Lei, C.S. Chen, Y.J. Tu, Y.H. Huang, H. Zhang, Heterogeneous degradation of
role of Cu(0) and Fe(II), Chinese Chem. Lett. 31 (10) (2020) 2831–2834. organic pollutants by persulfate activated by CuO-Fe3O4: Mechanism, stability, and
[72] M. Huang, W. Xiang, C. Wang, T. Zhou, J. Mao, X. Wu, F. Zhang, D. Li, X. Lu, effects of pH and bicarbonate ions, Environ. Sci. Technol. 49 (2015) 6838–6845.
Ultrafast O2 activation by copper oxide for 2,4-dichlorophenol degradation: The [86] Lingcong Li, Ningqiang Zhang, Rui Wu, Liyun Song, Guizhen Zhang, Hong He,
size-dependent surface reactivity, Chinese Chem. Lett. 31 (2020) 2769–2773. Comparative study of moisture-treated Pd@CeO2/Al2O3 and Pd/CeO2/Al2O3
[73] Wei Xiang, Tao Zhou, Yifan Wang, Mingjie Huang, Xiaohui Wu, Juan Mao, catalysts for automobile exhaust emission reactions: Effect of core-shell interface,
Xiejuan Lu, Beiping Zhang, Catalytic oxidation of diclofenac by hydroxylamine- ACS Appl. Mater. Interfaces. 12 (9) (2020) 10350–10358.
enhanced Cu nanoparticles and the efficient neutral heterogeneous-homogeneous [87] Guoying Wang, Cheng Cheng, Jianchao Zhu, Lijun Wang, Shengwang Gao,
reactive copper cycle, Water Res. 153 (2019) 274–283. Xunfeng Xia, Enhanced degradation of atrazine by nanoscale LaFe1-xCuxO3-δ
[74] Y. Wang, S. Chen, X. Yang, X. Huang, Y. Yang, E. kai He, S. Wang, R. Qiu, perovskite activated peroxymonosulfate: Performance and mechanism, Sci. Total
Degradation of 2,2′′ ,4,4′′ -tetrabromodiphenyl ether (BDE-47) by a nano zerovalent Environ. 673 (2019) 565–575.
iron-activated persulfate process: The effect of metal ions, Chem. Eng. J. 317 [88] Rohit Chauhan, Vimal Chandra Srivastava, Ajay Devidas Hiwarkar,
(2017) 613–622. Electrochemical mineralization of chlorophenol by ruthenium oxide coated
[75] Seema Singh, Shang-Lien Lo, Catalytic performance of hierarchical metal oxides for titanium electrode, J. Taiwan Inst. Chem. Eng. 69 (2016) 106–117.
per-oxidative degradation of pyridine in aqueous solution, Chem. Eng. J. 309 [89] T. González, J.R. Dominguez, E.M. Cuerda-Correa, S.E. Correia, G. Donoso,
(2017) 753–765. Selecting and improving activated homogeneous catalytic processes for pollutant
[76] V. Subbaramaiah, Vimal Chandra Srivastava, Indra Deo Mall, Catalytic Activity of removal. Kinetics, mineralization and optimization, J. Environ. Manage. 256
Cu/SBA-15 for Peroxidation of Pyridine Bearing Wastewater at Atmospheric (2020) 109972, https://doi.org/10.1016/j.jenvman.2019.109972.
Condition, AIChE J. 59 (7) (2013) 2577–2586. [90] Vijayalakshmi Gosu, Prerana Sikarwar, V. Subbaramaiah, Mineralization of
[77] Arvind Kumar, Basheswar Prasad, Krishan Kishor Garg, Efficiently degradation of pyridine by CWPO process using nFe0/GAC catalyst, J. Environ. Chem. Eng. 6 (1)
acrylonitrile from aqueous solution by La0.5Ce0.5MO3 (M = Fe, Cu and Co) (2018) 1000–1007.
perovskite-like catalyst: Optimization and reaction pathways, J. Water Process
Eng. 36 (2020) 101314, https://doi.org/10.1016/j.jwpe.2020.101314.
15