You are on page 1of 9

Chemical Engineering Journal 419 (2021) 129465

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Designing Co3O4/silica catalysts and intensified ultrafiltration


membrane-catalysis process for wastewater treatment
Jia Ding a, *, Gholamreza Vahedi Sarrigani a, Jiangtao Qu b, Amirali Ebrahimi a, Xia Zhong a,
Wen-Che Hou c, Julie M. Cairney b, Jun Huang a, Dianne E. Wiley a, David K. Wang a, *
a
School of Chemical and Biomolecular Engineering, The University of Sydney, New South Wales 2006, Australia
b
School of Aerospace, Mechanical and Mechatronic Engineering, The University of Sydney, New South Wales 2006, Australia
c
Department of Environmental Engineering, National Cheng Kung University, Tainan City 70101, Taiwan, ROC

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

Keywords: An integrated cobalt oxide catalyst supported on silica (Co3O4/SiO2) packed in the lumen of tubular alumina
Wastewater treatment (Al2O3) membrane was designed to realize a Fenton-like oxidation coupling with ultrafiltration water treatment
Dye degradation technology to remove persistent and hazardous pollutants. We employed a selection of SiO2 support materials
Advanced membrane oxidation process
with a range micro/mesostructures to immobilize the Co3O4 catalyst. These were evaluated and screened to
Cobalt catalyst
achieve the highest pollutant decolorization rate and maximum decolorization efficiency. Then, an integrated
Ultrafiltration process intensification
catalyst-membrane was developed as proof-of-concept for advanced water treatment. Batch degradation results
showed acid orange 7 dye (AO7) was effectively removed using Co3O4 deposited on MCM-41 silica, with 97%
decolorization efficiency achieved in 15 mins owing to the large surface area and high porosity of the MCM-41
support. In contrast, humic acid (HA) was relatively difficult to degrade using this catalyst, and was however
found to inhibit AO7 degradation. The integrated catalyst-membrane successfully removed both AO7 and HA
with only a 10% water flux reduction from 25.0 to 22.5 kg m− 2 h− 1 bar− 1 with excellent co-pollutant removal
rate of 99% over 40 h. Furthermore, the integrated catalyst-membrane also removed a range of organic pol­
lutants (neutral red, tetracycline hydrochloride, oxytetracycline) achieving > 90% rejection and 30 kg
m− 2 h− 1 bar− 1. These results show that the integrated catalyst-membrane can effectively purify a binary
pollutant mixture containing both hazardous dye and natural organic matter and other pharmaceutical chemicals
whilst producing process water for recycling and reuse.

1. Introduction hydroxyl radical possesses a high oxidation potential (HO⋅, E = 2.8 V)


[6–9].
Organic contaminants in wastewater pose a serious threat to the Nanomaterials made using transition metals and their oxides such as
environment and human heath [1,2]. The treatment of wastewater has cobalt are highly effective heterogeneous catalysts for the activation of
been an on-going challenge, mainly addressed by technologies based on H2O2 and peroxymonosulfate (PMS), producing radicals and oxide ions
biological and/or chemical degradation. The latter technique has that degrade organic pollutants via a Fenton-like process [10–13].
distinct advantages for the removal of organic contaminants due to the Current research has focused on enhancing the performance of Co3O4
efficient generation of oxidizing free radicals at ambient conditions based catalysts by controlling the morphology, physicochemical prop­
[3–5]. The most widely used methods include Fenton oxidation, ozone erties, and dispersion of the cobalt species [14–18]. In addition, facile
oxidation, chlorine oxidation, persulfate oxidation, electrochemical wet synthesis of catalysts on a variety of ceramic supports has been
catalytic oxidation, and photocatalytic degradation. Among advanced explored as an immobilization strategy for Co3O4 [19–21]. In particular,
oxidation processes (AOPs), Fenton and Fenton-like oxidation with SiO2 supports are attractive materials because of their advantages such
hydrogen peroxide (H2O2) as the oxidant has attracted extensive as abundance of the raw material, flexible textural tunability, and ease
attention for demineralizing refractory organic compounds, because of surface functionalization. Due to these advantages, SiO2 is an
H2O2 is eco-friendly, inexpensive, highly efficient, and the generated extremely versatile material widely investigated as a catalyst support.

* Corresponding authors.
E-mail addresses: jia.ding@sydney.edu.au (J. Ding), david.wang1@sydney.edu.au (D.K. Wang).

https://doi.org/10.1016/j.cej.2021.129465
Received 11 December 2020; Received in revised form 15 March 2021; Accepted 19 March 2021
Available online 29 March 2021
1385-8947/© 2021 Elsevier B.V. All rights reserved.
J. Ding et al. Chemical Engineering Journal 419 (2021) 129465

For example, Zhou et al. [11] chose diatomite as a cobalt catalyst sup­ experiments.
port, and demonstrated a bicarbonate-activated H2O2 as a novel oxidant
source for the degradation of dyes and phenol. Liu et al. [12] employed 2.1.2. Preparation of Co3O4/SiO2 catalyst
sol–gel chemistry with a rapid thermal process to prepare Co3O4-silica The Co3O4/SiO2 catalysts were prepared by an impregnation
nanocomposites and demonstrated superior performance for acid or­ method. Typically, 2.45 g of Co(NO3)2⋅6H2O was dissolved in 4 mL
ange 7 (AO7) degradation with a Fenton-like HCO–3/H2O2 catalytic ethanol while stirring to form a homogenous solution. Subsequently, 2 g
system. of silica support (silica xerogel, silicalite–1, MCM–41 or fumed silica)
Apart from organic dyes in wastewater, natural organic matters was added into the cobalt solution. The mixture was then ultrasonically
(NOMs) such as humic acids (HA) are recognized as persistent con­ treated for 1 h (ultrasonic power: 180 W) before drying in an oven at 80
taminants in the industrial effluents that need to be removed before ℃ for 12 h. The Co3O4/SiO2 catalysts were obtained by calcining the
discharge because of aesthetics and health concerns [22]. Many studies mixture in a pre-heated furnace at 600 ℃ for 1 h.
have shown that traditional Fenton processes are not adequate for
treating HA due to their large molecular size [23–25]. More importantly, 2.2. Characterization
they are known to generate carcinogenic substances during post disin­
fection treatment due to chlor(am)ination [26]. Furthermore, HA in The X-ray diffraction (XRD) patterns of the samples were recorded on
wastewater can inhibit the degradation of organic dyes and pollutants a PANalytical X’Pert3 powder diffractometer. The N2 adsorption/
by blocking the catalyst active sites or quenching the radicals through desorption isotherms were collected at 77 K on a Quantachrome
carboxyl and hydroxyl scavenging groups [27,28]. Therefore, it is Autosorb-iQ gas adsorption analyzer. The total specific surface area was
extremely important to demonstrate the effectiveness of organic calculated from the N2 adsorption branch in the range of relative pres­
pollutant removal by including HA in the reaction mixture. sure from 0.05 to 0.30 by the Brunauer–Emmett–Teller (BET) method.
Membrane technologies have proven to be very effective in removing The nonlocal density functional theory (NLDFT) and Bar­
large molecules, algae, and particles from water or wastewater [29–31]. rett–Joyner–Halenda (BJH) methods in the Quantachrome ASiQwin
By combining the advantages of AOP and membrane filtration, several software were used for estimation of micropore and mesopore size dis­
studies have designed membrane reactors and catalytic membranes to tributions from the adsorption isotherms. Scanning electron microscopy
enhance wastewater treatment and purification efficiency for the (SEM) images were obtained using a Phenom XL G2 Desktop. The FEI
removal of different organic pollutants [32–35]. Themis–Z aberration corrected scanning/transmission electron micro­
In this paper, a series of Co3O4/SiO2 nanocatalysts were prepared scopy (S/TEM) equipped with Super X energy dispersive X-ray spec­
using silica xerogel, silicalite-1, fumed silica and MCM-41 supports for troscopy (EDS) detector was used for characterization, and the images
AO7 degradation via a Fenton-like process. The properties of the SiO2 were obtained at 300 kV acceleration voltage. The Fourier transform
supports and Co3O4/SiO2 catalysts were systematically characterized to infrared-attenuated total reflectance (FTIR–ATR) spectra were collected
reveal the structure and reactivity relationship between the nanoporous on a Nicolet 6700 Fourier Transform infrared spectrophotometer with
support materials and the catalyst. The best Co3O4/MCM-41 nano­ an ATR accessory. The UV–Vis diffuse reflectance spectra (DRS) were
catalyst was then packed in the lumen of an Al2O3 ultrafiltration collected by a Shimadzu UV-3600 UV–VIS-NIR Spectrophotometer and
membrane tube to develop an intensified ultrafiltration-catalysis process BaSO4 was used as the reference material. The AO7 and HA concen­
for treating both AO7 and HA over an extended 40 h period. trations were analyzed by measuring the absorbance of the solution at
λmax = 484 and 410 nm respectibely, using an Agilent Cary 60 UV–Vis
2. Experimental spectrophotometer. The total organic carbon (TOC) was tested by using
the Shimadzu TOC-L Analyzer.
2.1. Sample preparation

2.1.1. Materials 2.3. Wastewater treatment performance


Four different SiO2 materials (silica xerogel, silicalite–1, MCM–41
and fumed silica; Fig. 1) were used as the support for the synthesis of 2.3.1. Batch test of Fenton-like degradation
Co3O4/SiO2 composites. The preparation procedures of the silica xero­ AO7 was selected as a model dye pollutant to investigate the catalytic
gel, zeolitic silicalite-1 and mesoporous MCM-41 supports are described activity of the Co3O4/SiO2 catalysts. All degradation experiments were
in the Supporting Information. The fumed silica (99.8%) was purchased performed using 50 ppm AO7 aqueous solution. Firstly, a pre-
from Sigma-Aldrich. Cobalt nitrate hexahydrate (Co(NO3)2⋅6H2O, A.R., determined amount of Co3O4/SiO2 catalyst (0.2 g L-1) and NaHCO3
Sigma–Aldrich) and ethanol (A.R., Sigma–Aldrich) were used as the (3.6 mM) were added into 250 mL AO7 solution at room temperature,
cobalt precursor and solvent for the preparation of Co3O4/SiO2 com­ followed by stirring for 30 min to reach adsorption equilibrium in dark
posites, respectively. Acid orange 7 (AO7, Sigma–Aldrich), humic acid conditions before adding H2O2 (11–100 mM) to initiate the reaction.
(HA, technical, Sigma–Aldrich), neutral red (≥90%, Sigma–Aldrich), Solutions were periodically withdrawn, filtered through a 0.2 μm Mil­
tetracycline hydrochloride (≥98%, Cayman) and oxytetracycline lipore syringe filter, and analyzed immediately to avoid further degra­
(≥95%, Cayman) were used as the model pollutants for degradation dation. As the concentration of the AO7 and the apparent kinetic
constant (kapp) follow a pseudo-first-order relationship, the

Fig. 1. Schematic of the different silica supports.

2
J. Ding et al. Chemical Engineering Journal 419 (2021) 129465

Langmuir–Hinshelwood (L–H) model was used to quantify the AO7 solutions over the bare Al2O3 membrane were also conducted.
decolorization rate [12]. The kapp for the different catalysts were
calculated using the equation: ln(C/C0) = − kapp*t, where C/C0 is the 3. Results and discussions
concentration ratio of AO7 at time, t, over the initial concentration, C0.
For comparison, the degradation of HA was also conducted. The HA 3.1. Catalyst structure and morphology
stock solution was prepared by dissolving 200 mg of HA in 80 mL of
0.01 M NaOH solution, followed by stirring at room temperature for 24 h The FTIR-ATR spectra of the Co3O4/silica composites in Fig. 3a show
and adjusting to neutral pH with 1 mol L-1 HCl. Finally, a certain amount the characteristic Si-O-Si symmetric/asymmetric stretching vibrations
of deionized water was added to the above solution to obtain a HA with bands near 800 cm− 1 and 1000–1250 cm− 1 in all four catalysts
concentration of 2 g L-1. The HA stock solution was stored in the [21,36]. These are also evident in the spectra of the pure silica supports
refrigerator. The degradation of HA (15 ppm) was performed using the (Fig. S1). For the Co3O4/silica composites, two new peaks at 568 and
same procedure as the AO7 degradation experiment. 660 cm− 1 were observed after cobalt impregnation and calcination, due
to the formation of Co3O4 species [21,37,38]. As compared to other
2.3.2. Integrated Co3O4/MCM-41@Al2O3 membrane for AO7 degradation samples, the Co3O4/silicalite-1 possesses additional bands at 550 and
and HA filtration 1228 cm− 1, due to the five-membered ring of the zeolitic structure in an
Tubular Al2O3 membranes (6 mm i.d., 10 mm o.d., 40 mm length) asymmetric stretching mode [39].
were heat-treated at 1200 ℃ for 3 h at a heating rate of 5 ℃ min− 1. The The XRD patterns in Fig. 3b show that all Co3O4/SiO2 samples dis­
powdered Co3O4/MCM–41 was pressed, crushed and sieved into ~ 50 played strong diffraction lines at 19.0◦ , 31.3◦ , 36.8◦ , 44.8◦ and 59.4◦ , in
mesh granules. Then, 800 mg of Co3O4/MCM–41 granules were packed accordance with the patterns of standard Co3O4 (PDF no. 43–1003)
inside the lumen of the Al2O3 membrane to fabricate the integrated [40]. The additional peaks at 7.9◦ , 8.9◦ and 23-25◦ in the spectrum of
Co3O4/MCM–41@Al2O3 membrane for the wastewater treatment the Co3O4/silicalite-1 are due to the distinctive zeolitic structure of the
(Fig. 2). One end of the Co3O4/MCM–41@Al2O3 membrane was blocked silicalite-1 support [39]. In addition, there are no other CoO diffraction
before immersing the packed membrane into a model wastewater con­ peaks that can be observed for all samples in the XRD patterns [41].
taining NaHCO3 (3.6 mM) and H2O2 (100 mM) and a model pollutant It can be seen in Fig. 3b that the intensity and width of the XRD peak
(50 ppm AO7, pH = 8.5; or 15 ppm HA, pH = 8.2; or a solution mixture at 36.8◦ associated with the Co3O4 are dependent on the SiO2 support
of 50 ppm AO7/15 ppm HA, pH = 8.4). The feed solution was then properties. The peak intensity decreased as Co3O4/silica xerogel >
pressurized to 0.5 bar to commence the filtration process, which allowed Co3O4/silicalite-1 > Co3O4/fumed silica > Co3O4/MCM-41, which is
the solution to come in contact with the Co3O4/MCM–41 packed catalyst related to the crystallinity degradation. The width of the XRD peak is an
in the lumen. The weight of the treated solution (the solution that indication of the particle size of the tricobalt tetroxides [42]. By using
permeated through the membrane) was recorded by an on-line digital the Scherrer formula, the average size of Co3O4 nanoparticles was
balance. The quality of the treated solution was analyzed using the calculated based on the width of the (3 1 1) planes at 36.8◦ . As suggested
UV–Vis spectrophotometer. To understand the role of the integrated by the results summarized in Table 1, the average Co3O4 crystallite sizes
Co3O4/MCM–41@Al2O3 membrane, filtration tests for HA and AO7 calculated by Scherrer’s equation are in the order of Co3O4/silica

Fig. 2. Schematic of the tubular membrane set up for wastewater treatment.

3
J. Ding et al. Chemical Engineering Journal 419 (2021) 129465

Fig. 3. (a) FTIR-ATR spectra, (b) XRD patterns, and (c) UV–Vis diffuse reflectance spectra of the Co3O4/SiO2 catalysts.

region of 300–450 nm is ascribed to the presence of ligand to metal


Table 1
charge transfer transition of O2–→Co2+ and 1A1g→1T2g. The second band
Textural properties of different SiO2 materials, the particle size of supported
observed at around 600–800 nm is indicative of O2–→Co3+ charge
Co3O4 and apparent kinetic constant of the Co3O4/SiO2 catalysts.
transfer transition and 1A1g→1T1g transition [44]. In addition, there are
Catalyst SiO2 support Co3O4 crystallite kappd
no absorption bands at 460, 510 and 550 nm, indicating that no CoO
sizec (nm) (min− 1)
SBET a
Vporeb phase is present in these samples [45,46], which is in accordance with
(m2 g− 1) 3 − 1
(cm g ) the XRD results.
Co3O4/silica 9 0.035 56 0.0011 The N2 adsorption–desorption isotherms of the SiO2 supports are
xerogel displayed in Fig. 4a. The physisorption isotherm of the silica xerogel
Co3O4/ 262 0.89 40 0.0037
indicates that this support is dense compared to the other SiO2 materials,
silicalite-1
Co3O4/fumed 336 2.3 24 0.013 with a very small surface area of 9 m2 g− 1 and a pore volume of 0.035
silica cm3 g− 1. The silicalite-1 sample shows a typical microporous zeolite
Co3O4/MCM- 1079 0.84 16 0.019 isotherm with a steep increase at very low relative pressure (P/P0 < 0.1).
41 In addition, a small type-H3 hysteresis loop can be seen upon reaching
a
Surface area determined by the Brunauer-Emmett-Teller (BET) method. P/P0 > 0.9, due to the open mesopores structure of the silicalite–1
b
Total pore volume at P/P0 = 0.995. nanocrystals [47]. The pore size distribution curves in Fig. 4b,c indicate
c
Average crystallite size of Co3O4 supported catalysts calculated by the that the silicalite–1 sample contains pore volume made up of micropores
Scherrer formula based on the width of the (3 1 1) planes. with a size of < 1 nm, mesopores with sizes 20–50 nm and macropores
d
Apparent kinetic constant for AO7 decolorization on the Co3O4 supported with sizes > 50 nm. The fumed silica sample exhibits a very large type-
catalysts. H3 hysteresis loop, where a gradually increased adsorption appears at
intermediate and high pressures (0.4 < P/P0 < 1), in agreement with the
xerogel > Co3O4/silicalite-1 > Co3O4/fumed silica > Co3O4/MCM-41. open pore structure of commercial fumed silica materials. The BJH pore
The Co3O4/MCM-41 catalyst possesses the smallest average Co3O4 size distribution curves in Fig. 4c show that the fumed silica has many
crystallite sizes of ~ 16 nm. more mesopores (>5 nm) than any of the other supports. In contrast, the
The UV–Vis diffuse reflectance spectra of the Co3O4/SiO2 catalysts N2 adsorption–desorption isotherm and BJH pore size distribution curve
with different SiO2 supports are shown in Fig. 3c. All samples exhibit of the MCM-41 sample show that typical MCM–41 zeolite with mesopore
two broad reflection bands around 400 and 710 nm, which are char­ size of 2.7 nm was successfully synthesized. After loading of Co3O4, the
acteristic of the Co3O4 [37,43,44]. The first broad band located in the obtained Co3O4/MCM–41 catalyst shows a decrease in the BET surface

Fig. 4. (a) N2 adsorption and desorption isotherms, (b) non-local density functional theory (NLDFT) micropore size distribution, and (c) Barrett-Joyner-Halenda
(BJH) meso-/macro-pore size distribution of the SiO2 supports.

4
J. Ding et al. Chemical Engineering Journal 419 (2021) 129465

area and total pore volume from 1079 to 701.4 m2 g− 1, and 0.84 to 0.68 catalysts exhibited very different dye degradation performance. The
cm3 g− 1 (Fig. S2a), respectively. The mesopore structure of MCM-41 Co3O4/silica xerogel delivered the slowest AO7 decolorization profile,
support remains relatively unchanged after impregnation and calcina­ only reaching C/C0 = 0.94 at 60 min. In contrast, the Co3O4/MCM–41
tion (Fig. S2b). sample displayed the best catalytic performance, with 70% of the AO7
The bright-field TEM images, HAADF images and the EDS maps of being degraded after 60 min (C/C0 = 0.30). As can be seen by the kapp
the Co3O4/silica catalysts are shown in Fig. 5. The Co3O4/SiO2 xerogel values listed in Table 1, the catalyst reactivity follows the order of
contained spherical Co3O4 particles in the range of 20 to 80 nm on the Co3O4/silica xerogel < Co3O4/silicalite-1 < Co3O4/fumed silica <
silica xerogel support (Fig. 5a). In comparison, the Co3O4 aggregates Co3O4/MCM-41.
were wrapped by small silicalite-1 particles as shown in Fig. 5b. Inter­ To explain this trend, it is very important to consider the catalyst
estingly, the open-pore structure of the fumed silica resulted in flower- structures, accessibility and availability. Fig. 6c shows the average
like Co3O4 aggregates of approximately 150–500 nm in diameter Co3O4 crystallite size decreases as the surface area of the SiO2 support
(Fig. 5c), which are locally dispersed throughout the matrix. In contrast increases, indicating that a large SiO2 surface area facilitated the for­
to the cobalt species in the Co3O4/SiO2 xerogel, Co3O4/silicalite-1, and mation of smaller Co3O4 crystallite size and the better dispersion of
Co3O4/fumed silica catalysts, the Co3O4/MCM-41 catalyst displayed Co3O4 (Fig. 5, TEM images). As a result, these properties (catalyst size
better dispersion (Fig. 5d). An abundance of Co3O4 nanoparticles with and dispersion) are well correlated with higher kapp value (faster
an average size of ~ 20 nm can be observed within the support matrix decolorization rate). Furthermore, it appears that the apparent rate
(in the inset; Fig. 5d1). There are also a large number of small particles constant scales linearly and inversely with respect to the catalyst crys­
(~3 nm) incorporated into the mesopores of MCM-41 (Fig. S3). The tallite size below 40 nm. This can be rationalized by the increased
HAADF–STEM image and the corresponding elemental maps of Co and number of cobalt sites within and on the SiO2 support that are available
Si elements (Fig. 5d2-4) indicate that the cobalt species in the Co3O4/ to react with H2O2 for the same molar fraction per volume, thus,
MCM–41 catalyst are dispersed throughout the MCM–41 matrix. increasing the rate of reaction in the heterogeneous system.
The Fenton-like AO7 degradation process in this bicarbonate/H2O2
3.2. Catalytic degradation performance system is well–known to follow the APMC mechanism, in which the
contact of Co3O4/SiO2 catalysts with the H2O2 initiates the formation of
The catalytic performance of the as-synthesized Co3O4/SiO2 was hydroxyl and carbonate radicals [12,48]. Then, superoxide radicals
firstly evaluated for the degradation of AO7 as a model dye pollutant, via (perhydroxyl radical, superoxide ion, and singlet oxygen) are generated
the activated peroxymonocarbonate (APMC) reaction mechanism, and contribute to the destruction of organic compounds [49,50]. Hence,
which is well-described in the Fenton-like process literature [11,12]. the availability and accessibility of Co3O4 within and on the SiO2 sup­
The AO7 concentration during catalytic degradation was followed by port play a critical role in determining the degree of H2O2 contacting
UV–Vis spectroscopy (absorbance profiles are shown in the Supporting with the catalytic Co species. This directly affects the production of the
Information). Fig. 6 and Fig. S4 shows that the four Co3O4 supported radicals and their concentrations available for dye degradation.

Fig. 5. The bright-field TEM images, HAADF images and the corresponding EDS maps of Co and Si of (a1-a4) Co3O4/silica xerogel, (b1-b4) Co3O4/silicalite-1, (c1-c4)
Co3O4/fumed silica and (d1-d4) Co3O4/MCM-41 catalysts.

5
J. Ding et al. Chemical Engineering Journal 419 (2021) 129465

concentration increased from 11 to 25 mM, the corresponding decol­


orization accelerated significantly, with a 12.5 times increase in the kapp
value (Fig. 7b). As the concentration of H2O2 increased from 25 to 100
mM, the decolorization of AO7 was further enhanced, but with less
enhancement between 50 and 100 mM. The 100 mM H2O2 in the solu­
tion gave the highest kapp value of 0.23 min− 1, and achieved 97% of AO7
decolorization after 15 min. In stark contrast, the chemical oxidation of
AO7 by 100 mM H2O2 alone (in the absence of catalyst) is very slow
achieving only 3% at 15 min and gradually reaching 5% of AO7
decolorization at the end of 30 min. This highlights the critical role that
the catalyst played in the heterogeneous catalytic system. In addition,
the catalytic results showed that the NaHCO3 concentration played an
important role in the AO7 degradation (Fig. S5), indicating it indeed
followed the APMC reaction mechanism.
In addition to the organic dye pollutants, the ubiquitous natural
organic matters (NOMs, such as humic acid, fulvic acid, tannic acid, etc.)
and their by-products in wastewater pose a serious risk to the ecological
and human health because they are carcinogenic and/or mutagenic
toxic substances [22,33,51,52]. In this work, we used humic acid (HA)
as a representative NOM due to its well-studied ecological impact and
recently reported inhibitory effect on organic pollutant degradation
[53–56]. Fig. 8a and Fig. S6 show the HA decolorization performance of
Co3O4/MCM–41 catalyst in the Fenton-like process. When the reaction
was commenced by adding the H2O2, the absorbance intensity gradually
decreased, indicating HA was degraded by the Co3O4/MCM–41 catalyst.
However, an overall HA decolorization rate of ~ 50% was obtained after
120 min, as compared to 97% of AO7 at 15 min time under the same
conditions (Fig. 8b). In addition, Fig. 8b shows the decolorization effi­
ciency of AO7 by the same catalyst was marginally inhibited in the first
15 min (23% decrease in kapp) in the presence of HA but reached about
the same AO7 decolorization percentage of 99% at 25 min compared to
the control experiment (no HA addition). Therefore, these interactions
effects between different organic pollutants during degradation were
further investigated using an integrated, compact catalyst-membrane
process to achieve a continuous throughput of water recovery.

3.3. Integrated Co3O4/MCM-41@Al2O3 membrane for wastewater


treatment

This study focused on designing a Co3O4/MCM–41 catalyst packed


bed within the lumen of a tubular Al2O3 membrane (Fig. S7) for the
removal of AO7 and HA pollutants, and recovery of water for reuse.
Water flux and pollutant removal ability of the Al2O3 membrane and
Co3O4/MCM–41@Al2O3 membrane were independently investigated
using three different pollutant water, viz. AO7 (50 ppm), HA (15 ppm)
and AO7/HA mixture (50/15 ppm). As shown in Fig. 9, the water fluxes
of all pollutant waters treated using the Co3O4/MCM–41@Al2O3 mem­
brane were lower than the bare Al2O3 membrane, especially for the HA
and the AO7/HA mixture. For the AO7 solution, the integrated Co3O4/
MCM–41@Al2O3 membrane displayed a water flux of 28.1 ± 3.0 kg
m− 2 h− 1 bar− 1, which decreased to 24.4 ± 1.5 kg m− 2 h− 1 bar− 1 for HA
solution. The mixed pollutant solution containing AO7 and HA led to an
even lower flux of 22.6 ± 3.1 kg m− 2 h− 1 bar− 1.
Fig. 6. (a) AO7 decolorization performance, (b) apparent kinetic constant The permeated water solutions under the different feed conditions
(kapp), and (c) the relationship between the Co3O4 crystallite size, silica surface were analyzed by UV–Vis spectrophotometry to evaluate the decolor­
area and kapp of the Co3O4/SiO2 catalysts. Reaction conditions: 50 ppm AO7, ization efficiency. Fig. 9 and Fig. S8 show that only ~ 15% of the AO7
0.2 g L-1 catalyst, 3.6 mM NaHCO3, 11 mM H2O2. was removed using the bare Al2O3 membrane within 2 h. The result
indicates that the bare membrane is not able to remove AO7 by itself.
Another important parameter in the Fenton-like process that can There are three possible reasons for the 15% AO7 removal, i.e., partial
enhance radical formation and dye degradation is the initial concen­ rejection by the bare Al2O3 membrane; adsorption by both the Al2O3
tration of H2O2. Fig. 7a shows the effect of H2O2 concentration on the membrane surface and the membrane pore wall; slow chemical oxida­
decolorization performance of the Co3O4/MCM–41 catalyst. When no tion by the H2O2 as shown in Fig. 7a. Meanwhile, the bare Al2O3
H2O2 was added to the reaction mixture, the AO7 concentration membrane shows a very high rejection for the HA (96%), which is most
remained relatively unchanged over 30 min, as there was no apparent probably purely attributable to filtration rather than chemical oxidation
production of hydroxyl, hydroperoxyl and other radicals. As the H2O2 because HA was observed to be very stable in H2O2/NaHCO3 solution
(Fig. 8a).

6
J. Ding et al. Chemical Engineering Journal 419 (2021) 129465

Fig. 7. (a) AO7 decolorization performance and (b) the apparent kinetic constant of the Co3O4/MCM–41 catalyst for different H2O2 concentrations. Reaction
conditions: 50 ppm AO7, 0.2 g L-1 catalyst, 3.6 mM NaHCO3.

Fig. 8. (a) Comparison of HA decolorization performance with and without the Co3O4/MCM–41 catalyst, and (b) effect of HA addition on the AO7 decolorization
performance. Reaction conditions: 50 ppm AO7 and/or 15 ppm HA, 0.2 g L-1 catalyst, 3.6 mM NaHCO3, 100 mM H2O2.

Fig. 9. (a) The water flux and pollutant decolorization using the bare Al2O3 membrane and Co3O4/MCM–41@Al2O3 membrane with different wastewater feed for
the first 2 h, and (b) long-term wastewater (AO7/HA) treatment performance of the Co3O4/MCM–41@Al2O3 membrane. Test conditions: 50 ppm AO7 and/or 15 ppm
HA, 3.6 mM NaHCO3, 100 mM H2O2, 0.5 bar.

When the AO7 and HA solutions were treated using the integrated respectively, by the bare Al2O3 membrane. These results demonstrate
Co3O4/MCM–41@Al2O3 membrane, the removal performance is > 96% that a vast amount of the AO7 removal by the integrated membrane is
for the AO7 and > 98% for the HA, as compared to only 15% and 96% due to degradation by the radicals formed from H2O2/NaHCO3 in the

7
J. Ding et al. Chemical Engineering Journal 419 (2021) 129465

presence of the packed Co3O4/MCM–41 catalyst. The HA removal for generating a highly stable effluent output for possible industrial
the integrated membrane was marginally higher than that of the bare wastewater applications.
Al2O3 membrane, and is most likely attributable to a modest HA
degradation capacity by the catalyst in conjunction with membrane CRediT authorship contribution statement
filtration removal.
To further demonstrate the performance of this advanced oxida­ Jia Ding: Conceptualization, Methodology, Formal analysis, Inves­
tion–separation process, a longer term experiment was conducted using tigation, Writing - original draft. Gholamreza Vahedi Sarrigani:
the mixed AO7/HA solution as HA is well-known to block the catalytic Investigation. Jiangtao Qu: Investigation. Amirali Ebrahimi: Concep­
sites and scavenge the formed radicals as shown by Fig. 8b and these tualization. Xia Zhong: Investigation. Wen-Che Hou: Conceptualiza­
references [27,28]. Our integrated Co3O4/MCM-41@Al2O3 membrane, tion, Writing - review & editing. Julie M. Cairney: Resources. Jun
however, can eliminate those adverse effects by HA filtration prior to Huang: Resources. Dianne E. Wiley: Writing - review & editing, Su­
dye degradation. Fig. 9b shows the water flux of the integrated catalyst pervision. David K. Wang: Conceptualization, Writing - review &
membrane gradually decreased from 25.0 to 22.5 kg m− 2 h− 1 bar− 1 editing, Supervision.
(~10%) over the testing time of 40 h, which can be understood by the
fouling effect of HA on/in the Al2O3 membrane. Meanwhile, the overall
decolorization rate gradually increased to more than 99% at the end of Declaration of Competing Interest
40 h. The results demonstrate that the integrated Co3O4/MCM-
41@Al2O3 membrane has good catalytic degradation and stability for The authors declare that they have no known competing financial
potentially treating dye/HA-contaminated wastewater solutions. The interests or personal relationships that could have appeared to influence
TOC values of the feed and permeate solutions are 34.1 mg/L and 17.6 the work reported in this paper.
mg/L, corresponding to a TOC removal rate of 48%, which is lower than
the AO7 removal rate as determined by UV–Vis because of the inter­ Acknowledgments
mediate compounds in the permeate solution. According to the litera­
ture, the active species in the APMC process included the perhydroxyl The authors gratefully acknowledge the financial support provided
radicals (⋅OH), superoxide ions (O−2 ⋅), and singlet oxygen (O12) by the Australian Research Council (DP190101734), and acknowledge
[11,57,58]. In the presence of HCO−3 , the azo dye AO7 was firstly co­ the facilities from Australian Centre for Microscopy and Microanalysis
ordinated with Co2+ ions of the Co3O4/silica catalyst [13]. Due to the (ACMM).
existence of H2O2, “crypto-⋅OH” radicals were then produced, which
was tightly associated with the cobalt complex. At the same time, Co(III)
Appendix A. Supplementary data
species were reduced to Co(II) by another H2O2 molecule with the for­
mation of O−2 ⋅ superoxide radicals. Subsequently, these radicals attack
Supplementary data to this article can be found online at https://doi.
the N atom of AO7, generating AO7⋅+ via electron transfer, which was
org/10.1016/j.cej.2021.129465.
then decomposed to p-phenolsulfonic acid and 1,2-naphthaquinone
through the contemporaneous break of two C–N bonds. After that, the
References
radicals further reacted with 1,2-naphthaquinone to form hydroxylated
or polyhydroxylated derivatives, and finally produced carboxylic acids [1] P. Hu, M. Long, Cobalt-catalyzed sulfate radical-based advanced oxidation: A
by ring-opening. Since the intermediates are also degradable by the review on heterogeneous catalysts and applications, Appl. Catal. B 181 (2016)
APMC process, the TOC removal rate can be further improved in prac­ 103–117.
[2] M. Pera-Titus, V. Garcı ́a-Molina, M.A. Baños, J. Giménez, S. Esplugas, Degradation
tical applications, such as recycling the permeate solution. In addition,
of chlorophenols by means of advanced oxidation processes: a general review,
the degradation performance of the Co3O4/MCM–41@Al2O3 membrane Appl. Catal. B 47 (4) (2004) 219–256.
for neutral red, tetracycline hydrochloride and oxytetracycline were also [3] H. Wu, X. Xu, L. Shi, Y.u. Yin, L.-C. Zhang, Z. Wu, X. Duan, S. Wang, H. Sun,
Manganese oxide integrated catalytic ceramic membrane for degradation of
tested. As shown in Fig. S9, the Co3O4/MCM–41@Al2O3 membrane can
organic pollutants using sulfate radicals, Water Res. 167 (2019) 115110.
effectively remove the model dyes and antibiotics with a rate of > 90% [4] Y.u. Chen, G. Zhang, H. Liu, J. Qu, Confining free radicals in close vicinity to
and showed similar water flux of ~ 30 kg m− 2 h− 1 bar− 1, indicating that contaminants enables ultrafast Fenton-like processes in the interspacing of MoS2
the Co3O4/MCM–41@Al2O3 membrane has good universal applicability membranes, Angew. Chem. Int. Ed. 58 (24) (2019) 8134–8138.
[5] M.M. Khin, A.S. Nair, V.J. Babu, R. Murugan, S. Ramakrishna, A review on
for the different organic pollutants. nanomaterials for environmental remediation, Energy Environ. Sci. 5 (8) (2012)
8075.
4. Conclusions [6] Y. Yin, L. Shi, W. Li, X. Li, H. Wu, Z. Ao, W. Tian, S. Liu, S. Wang, H. Sun, Boosting
Fenton-like reactions via single atom Fe catalysis, Environ. Sci. Technol. 53 (2019)
11391–11400.
In this paper, four silica-based supports with different micro-/meso- [7] Q. Jin, J. Kang, Q. Chen, J. Shen, F. Guo, Z. Chen, Efficiently enhanced Fenton-like
structures and textures were employed for the preparation of Co3O4/ reaction via Fe complex immobilized on silica particles for catalytic hydrogen
peroxide degradation of 2,4-dichlorophenol, Appl. Catal. B 268 (2020), 118453.
SiO2 catalysts. It was found that the texture properties of SiO2 greatly [8] C. Ma, S. Feng, J. Zhou, R. Chen, Y. Wei, H. Liu, S. Wang, Enhancement of H2O2
influenced the size of Co3O4 nanoparticles and degradation capacity. decomposition efficiency by the co-catalytic effect of iron phosphide on the Fenton
The MCM–41 with the largest surface area produced the smallest and reaction for the degradation of methylene blue, Appl. Catal. B 259 (2019), 118015.
[9] Y. Sun, Z. Yang, P. Tian, Y. Sheng, J. Xu, Y.-F. Han, Oxidative degradation of
highly dispersed Co3O4 particles, which thus achieved the best reactivity nitrobenzene by a Fenton-like reaction with Fe-Cu bimetallic catalysts, Appl. Catal.
for AO7 decolorization. Furthermore, by increasing the H2O2 concen­ B 244 (2019) 1–10.
tration, the catalytic performance of Co3O4/MCM–41 was significantly [10] X. Long, Z. Yang, H. Wang, M. Chen, K. Peng, Q. Zeng, A. Xu, Selective Degradation
Of Orange II with the cobalt(II)–bicarbonate–hydrogen peroxide system, Ind. Eng.
enhanced. However, the HA was difficult to be degraded under the same Chem. Res. 51 (2012) 11998–12003.
reaction conditions owing to its persistent and non-degradable nature, [11] L. Zhou, W. Song, Z. Chen, G. Yin, Degradation of organic pollutants in wastewater
which also led to adverse effects on AO7 decolorization in the HA/AO7 by bicarbonate-activated hydrogen peroxide with a supported cobalt catalyst,
Environ. Sci. Technol. 47 (2013) 3833–3839.
mixture. Process intensification was enabled by designing an integrated
[12] L. Liu, J. Ding, G.V. Sarrigani, P. Fitzgerald, Z.M. Aljunid Merican, J.-W. Lim, H.-
Co3O4/MCM–41@Al2O3 membrane, which achieved a very high H. Tseng, F. Xie, B. Zhang, D.K. Wang, Enhanced catalyst dispersion and structural
removal rate of > 96% for both AO7 and HA at a stable water flux over control of Co3O4-silica nanocomposites by rapid thermal processing, Appl. Catal. B
40 h. This advanced oxidation membrane separation process is prom­ 262 (2020), 118246.
[13] M. Luo, L. Lv, G. Deng, W. Yao, Y. Ruan, X. Li, A. Xu, The mechanism of bound
ising to effectively remove HA and ameliorate the interference of HA in a hydroxyl radical formation and degradation pathway of acid orange II in Fenton-
heterogenous AOP system for dye degradation and removal whilst like Co2+-HCO-3 system, Appl. Catal. A 469 (2014) 198–205.

8
J. Ding et al. Chemical Engineering Journal 419 (2021) 129465

[14] X. Dong, X. Duan, Z. Sun, X. Zhang, C. Li, S. Yang, B. Ren, S. Zheng, D.D. Dionysiou, [37] G.A.M. Ali, O.A. Fouad, S.A. Makhlouf, Structural, optical and electrical properties
Natural illite-based ultrafine cobalt oxide with abundant oxygen-vacancies for of sol–gel prepared mesoporous Co3O4/SiO2 nanocomposites, J. Alloys Compd. 579
highly efficient Fenton-like catalysis, Appl. Catal. B 261 (2020), 118214. (2013) 606–611.
[15] L. Hu, F. Yang, W. Lu, Y. Hao, H. Yuan, Heterogeneous activation of oxone with [38] W. Zhang, J. Díez-Ramírez, P. Anguita, C. Descorme, J.L. Valverde, A. Giroir-
CoMg/SBA-15 for the degradation of dye Rhodamine B in aqueous solution, Appl. Fendler, Nanocrystalline Co3O4 catalysts for toluene and propane oxidation: Effect
Catal. B 134–135 (2013) 7–18. of the precipitation agent, Appl. Catal. B 273 (2020) 118894.
[16] S. Lin, G. Su, M. Zheng, D. Ji, M. Jia, Y. Liu, Synthesis of flower-like Co3O4–CeO2 [39] J. Ding, S. Fan, P. Chen, T. Deng, Y.e. Liu, Y. Lu, Vapor-phase transport synthesis of
composite oxide and its application to catalytic degradation of 1,2,4- microfibrous-structured SS-fiber@ZSM-5 catalyst with improved selectivity and
trichlorobenzene, Appl. Catal. B 123–124 (2012) 440–447. stability for methanol-to-propylene, Catal. Sci. Technol. 7 (10) (2017) 2087–2100.
[17] H. Xu, N. Jiang, D. Wang, L. Wang, Y. Song, Z. Chen, J. Ma, T. Zhang, Improving [40] L. Chen, J. Hu, R. Richards, S. Prikhodko, S. Kodambaka, Synthesis and surface
PMS oxidation of organic pollutants by single cobalt atom catalyst through hybrid activity of single-crystalline Co3O4 (111) holey nanosheets, Nanoscale 2 (9) (2010)
radical and non-radical pathways, Appl. Catal. B 263 (2020), 118350. 1657.
[18] Z. Jiang, J. Zhao, C. Li, Q. Liao, R. Xiao, W. Yang, Strong synergistic effect of Co3O4 [41] D. Barreca, C. Massignan, S. Daolio, M. Fabrizio, C. Piccirillo, L. Armelao,
encapsulated in nitrogen-doped carbon nanotubes on the nonradical-dominated E. Tondello, Composition and microstructure of cobalt oxide thin films obtained
persulfate activation, Carbon 158 (2020) 172–183. from a novel cobalt(II) precursor by chemical vapor deposition, Chem. Mater. 13
[19] W. Zhang, H.L. Tay, S.S. Lim, Y. Wang, Z. Zhong, R. Xu, Supported cobalt oxide on (2001) 588–593.
MgO: Highly efficient catalysts for degradation of organic dyes in dilute solutions, [42] A.L. Patterson, The Scherrer formula for X-Ray particle size determination, Phys.
Appl. Catal. B 95 (2010) 93–99. Rev. 56 (1939) 978–982.
[20] C. Cai, S. Kang, X. Xie, C. Liao, Ultrasound-assisted heterogeneous [43] J. Gamonchuang, P. Poosimma, K. Saito, N. Khaorapapong, M. Ogawa, The effect
peroxymonosulfate activation with Co/SBA-15 for the efficient degradation of of alcohol type on the thickness of silica layer of Co3O4@SiO2 core-shell particle,
organic contaminant in water, J. Hazard. Mater. 385 (2020), 121519. Colloids Surf. A 511 (2016) 39–46.
[21] P. Shukla, H. Sun, S. Wang, H.M. Ang, M.O. Tadé, Nanosized Co3O4/SiO2 for [44] L.G.A. van de Water, G.L. Bezemer, J.A. Bergwerff, M. Versluijs-Helder, B.
heterogeneous oxidation of phenolic contaminants in waste water, Sep. Purif. M. Weckhuysen, K.P. de Jong, Spatially resolved UV–vis microspectroscopy on the
Technol. 77 (2011) 230–236. preparation of alumina-supported Co Fischer-Tropsch catalysts: Linking activity to
[22] A. Matilainen, M. Sillanpää, Removal of natural organic matter from drinking Co distribution and speciation, J. Catal. 242 (2006) 287–298.
water by advanced oxidation processes, Chemosphere 80 (2010) 351–365. [45] C. de Julián Fernández, G. Mattei, C. Sada, C. Battaglin, P. Mazzoldi,
[23] Y. Wu, S. Zhou, X. Ye, R. Zhao, D. Chen, Oxidation and coagulation removal of Nanostructural and optical properties of cobalt and nickel–oxide/silica
humic acid using Fenton process, Colloids Surf. A 379 (2011) 151–156. nanocomposites, Mater. Sci. Eng. C 26 (2006) 987–991.
[24] M.J. Farré, X. Doménech, J. Peral, Combined photo-Fenton and biological [46] N. Koshizaki, K. Yasumoto, T. Sasaki, Mechanism of optical transmittance change
treatment for Diuron and Linuron removal from water containing humic acid, by NOx in CoO/SiO2 nanocomposites films, Sens. Actuators B 66 (2000) 122–124.
J. Hazard. Mater. 147 (2007) 167–174. [47] H. Zhang, Y. Ma, K. Song, Y. Zhang, Y. Tang, Nano-crystallite oriented self-
[25] Y. Wu, S. Zhou, F. Qin, K. Zheng, X. Ye, Modeling the oxidation kinetics of Fenton’s assembled ZSM-5 zeolite and its LDPE cracking properties: Effects of accessibility
process on the degradation of humic acid, J. Hazard. Mater. 179 (2010) 533–539. and strength of acid sites, J. Catal. 302 (2013) 115–125.
[26] I.A. Ike, Y. Lee, J. Hur, Impacts of advanced oxidation processes on disinfection [48] A. Xu, X. Li, S. Ye, G. Yin, Q. Zeng, Catalyzed oxidative degradation of methylene
byproducts from dissolved organic matter upon post-chlor(am)ination: A critical blue by in situ generated cobalt (II)-bicarbonate complexes with hydrogen
review, Chem. Eng. J. 375 (2019), 121929. peroxide, Appl. Catal. B 102 (2011) 37–43.
[27] L. Kong, G. Fang, X. Xi, Y. Wen, Y. Chen, M. Xie, F. Zhu, D. Zhou, J. Zhan, A novel [49] X. Xu, D. Tang, J. Cai, B. Xi, Y. Zhang, L. Pi, X. Mao, Heterogeneous activation of
peroxymonosulfate activation process by periclase for efficient singlet oxygen- peroxymonocarbonate by chalcopyrite (CuFeS2) for efficient degradation of 2,4-
mediated degradation of organic pollutants, Chem. Eng. J. 403 (2021), 126445. dichlorophenol in simulated groundwater, Appl. Catal. B 251 (2019) 273–282.
[28] Z.-H. Diao, F.-X. Dong, L. Yan, Z.-L. Chen, W. Qian, L.-J. Kong, Z.-W. Zhang, [50] E.V. Bakhmutova-Albert, H. Yao, D.E. Denevan, D.E. Richardson, Kinetics and
T. Zhang, X.-Q. Tao, J.-J. Du, D. Jiang, W. Chu, Synergistic oxidation of Bisphenol Mechanism of Peroxymonocarbonate Formation, Inorg. Chem. 49 (2010)
A in a heterogeneous ultrasound-enhanced sludge biochar catalyst/persulfate 11287–11296.
process: Reactivity and mechanism, J. Hazard. Mater. 384 (2020) 121385. [51] Z. Domany, I. Galambos, G. Vatai, E. Bekassy-Molnar, Humic substances removal
[29] S.N. Abd Jalil, D.K. Wang, C. Yacou, J. Motuzas, S. Smart, J.C. Diniz da Costa, from drinking water by membrane filtration, Desalination 145 (2002) 333–337.
Vacuum-assisted tailoring of pore structures of phenolic resin derived carbon [52] Q.V. Ly, C. Matindi, A.T. Kuvarega, H.H. Ngo, Q.V. Le, V.H. Nam, J. Li, Exploring
membranes, J. Membr. Sci. 525 (2017) 240–248. the novel PES/malachite mixed matrix membrane to remove organic matter for
[30] D.K. Wang, M. Elma, J. Motuzas, W.C. Hou, F. Xie, X. Zhang, Rational design and water purification, Chem. Eng. Res. Des. 160 (2020) 63–73.
synthesis of molecular-sieving, photocatalytic, hollow fiber membranes for [53] X. Ding, L. Gutierrez, J.-P. Croue, M. Li, L. Wang, Y. Wang, Hydroxyl and sulfate
advanced water treatment applications, J. Membr. Sci. 524 (2017) 163–173. radical-based oxidation of RhB dye in UV/H2O2 and UV/persulfate systems:
[31] J. Wang, A. Cahyadi, B. Wu, W. Pee, A.G. Fane, J.W. Chew, The roles of particles in Kinetics, mechanisms, and comparison, Chemosphere 253 (2020), 126655.
enhancing membrane filtration: A review, J. Membr. Sci. 595 (2020) 117570. [54] W. Li, S. Li, Y. Tang, X. Yang, W. Zhang, X. Zhang, H. Chai, Y. Huang, Highly
[32] Y. Fan, Y. Zhou, Y. Feng, P. Wang, X. Li, K. Shih, Fabrication of reactive flat-sheet efficient activation of peroxymonosulfate by cobalt sulfide hollow nanospheres for
ceramic membranes for oxidative degradation of ofloxacin by peroxymonosulfate, fast ciprofloxacin degradation, J. Hazard. Mater. 389 (2020), 121856.
J. Membr. Sci. 611 (2020) 118302. [55] F. Chen, Q. Yang, D. Wang, F. Yao, Y. Ma, X. Li, J. Wang, L. Jiang, L. Wang, H. Yu,
[33] H. Lin, Q. Fang, W. Wang, G. Li, J. Guan, Y.i. Shen, J. Ye, F.u. Liu, Prussian blue/ Highly-efficient degradation of amiloride by sulfate radicals-based photocatalytic
PVDF catalytic membrane with exceptional and stable Fenton oxidation processes: Reactive kinetics, degradation products and mechanism, Chem. Eng. J.
performance for organic pollutants removal, Appl. Catal. B 273 (2020) 119047. 354 (2018) 983–994.
[34] X. Wang, Y. Li, H. Yu, F. Yang, C.Y. Tang, X. Quan, Y. Dong, High-flux robust [56] F. Rehman, M. Sayed, J.A. Khan, N.S. Shah, H.M. Khan, D.D. Dionysiou, Oxidative
ceramic membranes functionally decorated with nano-catalyst for emerging micro- removal of brilliant green by UV/S2O2- -
8 , UV/HSO5 and UV/H2O2 processes in
pollutant removal from water, J. Membr. Sci. 611 (2020) 118281. aqueous media: A comparative study, J. Hazard. Mater. 357 (2018) 506–514.
[35] Z. Zhu, L. Zhong, Z. Zhang, H. Li, W. Shi, F. Cui, W. Wang, Gravity driven ultrafast [57] J.M. Lin, M. Liu, Chemiluminescence from the decomposition of
removal of organic contaminants across catalytic superwetting membranes, peroxymonocarbonate catalyzed by gold nanoparticles, J. Phys. Chem. B 112
J. Mater. Chem. A 5 (48) (2017) 25266–25275. (2008) 7850–7855.
[36] X.J. Yin, K. Peng, A.P. Hu, L.P. Zhou, J.H. Chen, Y.W. Du, Preparation and [58] M. Liu, L. Zhao, J.-M. Lin, Chemiluminescence energy transfer reaction for the on-
characterization of core–shell structured Co/SiO2 nanosphere, J. Alloys Compd. line preparation of peroxymonocarbonate and Eu(II)− Dipicolinate complex,
479 (1-2) (2009) 372–375. J. Phys. Chem. A 110 (2006) 7509–7514.

You might also like