You are on page 1of 11

Journal of Environmental Management 296 (2021) 113305

Contents lists available at ScienceDirect

Journal of Environmental Management


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

Polymer-based immobilized Fe2O3–TiO2/PVP catalyst preparation method


and the degradation of triclosan in treated greywater effluent by
solar photocatalysis
Sarath Chandra Pragada *, Arun Kumar Thalla
Department of Civil Engineering, National Institute of Technology Karnataka, Surathkal, 575025, India

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

Keywords: The present study involves a novel protocol to develop a ternary composite catalyst for an effective post-
Greywater treatment technique for greywater. The ternary film of Fe2O3–TiO2/polyvinyl pyrrolidine (PVP) is coated on a
Triclosan glass tube using spray coating with annealing at 320 ◦ C. The structure, thermal, microstructure, and surface
Immobilized Fe2O3–TiO2/PVP catalyst
properties of the coated film are characterized by X-ray diffraction (XRD), Fourier Transform Infrared Spec­
Spray coating
Solar photocatalysis
troscopy (FTIR), Field Emission Scanning Electron Microscopy (FESEM), and Thermo Gravimetric Analysis
(TGA). The scratch hardness of photocatalysts at different Fe2O3/TiO2 compositions is investigated based on the
width measurement of scratch using FESEM analysis. Results show that at an optimum coating of 5% of Fe2O3/
TiO2 composition catalytic film, the maximum scratch hardness (7.984 GPa) is obtained. Also, the photocatalyst
has the highest cohesive bond strength and wearing resistance. The degradation of triclosan (TCS) in treated
greywater, discharged from the anaerobic-aerobic treatment system, is investigated at a lab-scale using a solar
photocatalytic reactor. The response surface analysis has been performed from the different sets of experimental
trials for various optimal parameters. It is observed that the TCS degradation efficiency of 83.27% has resulted
under optimum conditions.

1. Introduction techniques (Ansari and Shrikhande, 2019). The term PPCPs refers to a
broad range of uncontrolled products, including anti-depresent, anti-­
Water inadequacy has pushed researchers into ecological, economic, inflammatory drugs, antibiotics, and antimicrobial agents, which are
and low-energy alternative aquatic resources, out of which greywater utilized by human health care, cosmetic reasons, and livestock produc­
has also been considered (Abu Ghunmi et al., 2010; Boano et al., 2020). tivity (Sharma et al., 2020). For instance, triclosan (TCS) is one such
Greywater is the wastewater collected from domestic sources, such as common PPCPs.
bathtubs, laundry, hand basins, showers, air condition outlets, etc., TCS (C12H7Cl3O2) is a broad-spectrum synthetic antimicrobial agent,
without input from the toilet (Laaffat et al., 2019; Samayamanthula under the category of emerging contaminants, widely used in PPCPs,
et al., 2019; Vuppaladadiyam et al., 2019). Significant constituents in soaps, toothpaste, shower gels, body lotions, deodorants, cleansers, and
greywater are wide-ranging and depend on lifestyle patterns, customs, hand sanitizers (Jagini et al., 2019). Liquid hand soaps generally contain
and climatic conditions. It generally contains high concentrations of TCS in concentrations of 0.1%–0.45% (% w/v) (Harrow et al., 2011).
nutrients, xenobiotic organic compounds (XOCs), and biological mi­ Due to its extensive use as a prophylactic, the high concentrations at
crobes (Shaikh and Ahammed, 2020). Greywater disposal with a high which it is employed, and its inherent stability, TCS accumulates to high
amount of XOCs can reduce the surface tension of water and create levels in the environment (Orhon et al., 2017). Moreover, TCS reacts
undesired environmental conditions for aquatic organisms. However, with free chlorine and can form chloroform, categorized as a carcinogen
recent studies have found many pharmaceuticals and personal care (Lépesová et al., 2019). Besides, several studies have confirmed that
products (PPCPs), pesticides, pigments, and toxic heavy metals in conventional wastewater treatment systems cannot eliminate TCS
appreciable concentrations in greywater (Czech et al., 2020). Hence, it is complexes due to their least decomposability (Pycke et al., 2010).
essential to treat greywater with low-cost physicochemical or biological Consequently, it is necessary to obtain an effective process that allows its

* Corresponding author.
E-mail addresses: sarathpragada@gmail.com (S. Chandra Pragada), arunkumar@nitk.edu.in (A.K. Thalla).

https://doi.org/10.1016/j.jenvman.2021.113305
Received 31 January 2021; Received in revised form 11 July 2021; Accepted 14 July 2021
Available online 26 July 2021
0301-4797/© 2021 Elsevier Ltd. All rights reserved.
S. Chandra Pragada and A.K. Thalla Journal of Environmental Management 296 (2021) 113305

degradation to non-toxic substances. immobilized by an ultrasonic-assisted spray coating technique on a glass


In this context, various Advanced Oxidation Processes (AOPs) have tube. Its usage for the photocatalytic degradation of TCS is employed
been recognized and employed as efficient methods to remove recalci­ under solar light irradiation. Moreover, the present work aims:
trant environmental pollutants like TCS from wastewater. As one of the
AOPs, the heterogeneous semiconductor photocatalytic process, utiliz­ i) To study the degradation capability of TCS in greywater effluent,
ing catalysts (TiO2, ZnO, Fe2O3, CdS, GaP, and ZnS), provides an inno­ using the continuous photocatalytic reactor coated with ternary
vative and effective, low-cost, environmental friendly, and sustainable composite.
treatment technology for the degradation of persistent organic com­ ii) To describe the effect of Fe2O3–TiO2/PVP on TCS degradation
pounds (POPs) for the specified wastewater (Manu and Thalla, 2017; using the Box Behnken model and to optimize the significance of
Poulopoulos et al., 2019). However, low photo-quantum efficiency and the independent parameters.
inactivity under visible light are significant problems with the use of iii) To identify the intermediate products formed in the photo­
TiO2 as a catalyst. It also further occurs from the rapid recombination of catalytic oxidation of TCS through Liquid Chromatography and
photogenerated electrons and holes. In order to boost its response and Mass Spectroscopy (LC-MS) analysis.
minimize the recombination of electrons and holes by immobilization,
dopants, such as transitional metals, have been added to the TiO2 2. Materials and methods
catalyst (Nawawi et al., 2016). Similar transition metals, such as Fe2O3,
can be easily integrated into the crystal lattice of TiO2 owing to their 2.1. Preparation and characterization of synthetic greywater
related ionic radii. Fe2O3, with a lower bandgap energy of 2.3 eV, has
emerged as a suitable compound for the formulation of TiO2 composites, Synthetic greywater was prepared by mixing urea (85.4 mg/L), sol­
with enhanced optical and morphological properties (Abbas et al., 2016; uble starch (275 mg/L), milk powder (70 mg/L), potassium dihydrogen
Banisharif et al., 2015). phosphate (20 mg/L), detergent powder (50 mg/L), and liquid hand
In suspended photocatalysis processes, the separation of photo­ washes (50 mg/L) in tap water. The resultant synthetic wastewater was
catalysts from treated effluent is essential. It prevents the accumulation used in the pilot-scale integrated anaerobic-aerobic sequential batch
of catalytic powder in the environment and minimizes water treatment reactor (fabricated at NITK Surathkal, Environmental engineering lab),
costs (Behpour et al., 2019). However, it is not easy to separate and reuse coupled with photocatalytic treatment. The effluent from the integrated
the photocatalyst. Such a photocatalyst in powder form cannot be uti­ anaerobic-aerobic sequential batch reactor was fed into an immobilized
lized, particularly for drinking water purification applications. So, it is photocatalytic reactor set up to study the TCS removal efficiency. The
advantageous to use immobilized photocatalysts, which can be easily characteristics of greywater effluent are summarized in Table 1.
isolated and repeatedly used (Belapurkar et al., 2006). The effective
coating of these photocatalytic powders on any surface is challenging to 2.2. Preparation of ternary composite (Fe2O3–TiO2/PVP) solution and
implement due to the poor film-forming ability of the metal oxide. With catalytic film coating
polymers, film formation exhibits very high mechanical stability,
excellent glass adhesion, and photocatalytic activity. Therefore, this A ternary composite (Fe2O3–TiO2/PVP) solution was prepared by
catalyst film formability can be improved using a suitable polymeric mixing different atomic weight % (1%, 3%, 5%, 7%, and 9%) of Fe2O3/
binder. The catalyst immobilization with polymers has been recognized TiO2 with pure commercial anatase and 3.5% (w/v) of PVP in 2-methoxy
to improve photoactivity due to increased specific surface area (Nawawi ethanol. The resulted solution was mixed using a magnetic stirrer at
et al., 2016). There is a plethora of polymers used for photocatalysis 350–450 rpm for 120–180 min. It was then homogenized with an ul­
purpose. Because of its excellent surface passivation and adsorption trasonic bath for 30 min. A catalytic film was prepared by spraying the
ability, polyvinyl pyrrolidine (PVP) with film-forming characteristics is colloidal solution on a hollow cylindrical borosilicate glass pipe at a
the most suitable capturing agent of metal oxide particles (Yashni et al., constant pressure of 60 kPa. The wet coated film was allowed to dry in a
2020). hot air oven, fixed around 60–80 ◦ C, for 2–3 h. The glass pipe was then
Numerous studies have been carried out on the immobilization of placed in a furnace chamber, which was initially maintained at 4 ◦ C/min
TiO2 in photocatalysis (Shan et al., 2010). Several attempts have been until a constant temperature of 320 ◦ C was achieved. The specimen was
made to utilize visible light instead of UV light due to the challenges allowed to maintain this constant temperature for more than 2 h for the
associated with UV light (Mukherjee et al., 2014). Qin et al. (2014) proper immobilization process. Finally, the coated material was allowed
studied the modification of TiO2 photocatalyst with a suitable to cool down at the same rate (i.e., 4 ◦ C/min) to room temperature, in
co-catalyst to provide excellent photocatalytic activity to degrade order to avoid cracks on coated layers and reactions with impurities in
organic contaminants. Liu et al. (2011) studied the bandgap of the atmosphere; the temperature was set to come down from 320 ◦ C up
Fe2O3–TiO2 composite, which exhibited not only stronger absorption in to 50 ◦ C, to maintain the rate of 4 ◦ C/min.
the ultraviolet region but also adequate and strong absorption in the
visible light region.
Table 1
A few research studies have been only focused on the removal of TCS Physical and chemical characteristics of synthetic greywater effluent
by photocatalyst via an immobilized catalyst. Fidelis et al. (2019) characteristics.
investigated the photocatalytic degradation of TCS in both ultra-pure
Parameter Synthetic greywater effluent from the integrated system
and Cl⁻ containing water using an immobilized Fe/Nb2O5 catalyst.
Kosera et al. (2017) presented the utilization of ZnO immobilized in Range Average

biopolymer as a photocatalyst for TCS degradation in ultra-pure water. pH 6.68 ̶ 7.32 7


Besides, Köwitsch and Mehring (2021) also studied the photocatalytic Temperature, ◦ C 26 –28 27
Turbidity, NTU 3.9 –7.2 5.8
degradation of TCS using magnetic composites of iron oxide (α-Fe2O3
Chemical oxygen 39.5 ̶ 69.97 54.75
and Fe3O4) and carbon nitride materials (CN) synthesized via a demand, mg/L
microwave-assisted hydrothermal. So far, TCS degradation studies have Biological oxygen 8.55 ̶ 34.25 21.4
mostly been analyzed on the distilled water matrices. However, the ef­ demand, mg/L
ficiency of photocatalysis for TCS degradation on greywater and the TCS, mg/L 2.80 ̶ 4.68 3.87
Total Phosphorous, mg/L 0.63 ̶ 1.91 1.27
effect of operational parameters have not been widely reported, which
Total Nitrogen, mg/L 4.62 ̶ 11.02 7.82
still is a research gap in this study field.
In the present study, the polymer-based Fe2O3–TiO2/PVP catalyst is

2
S. Chandra Pragada and A.K. Thalla Journal of Environmental Management 296 (2021) 113305

2.3. Experimental setup for photocatalytic treatment the wear performance of the coated sample, sliding wear tests were
conducted on the pin on disc tribometer (TR-20LE-PHM 400-CHM 600,
The detailed representation of the solar photocatalytic reactor setup Ducom Instruments Pvt. Ltd, India) against EN 32 steel counter body.
is shown in Fig. 1. The Parabolic Trough Concentrator (PTC) was made Besides, a 3D surface profilometer (NANOVEA ST400, California) was
up of a 2 mm thick mirror-polished stainless-steel sheet, along with used to examine the wear mechanism. Furthermore, the adhesion
stands, to hold the coated glass pipe. Moreover, the PTC orientation was strength of the catalytic film was analyzed by a Pull-off adhesion tester
also made to rotate around the complete reactor setup to adjust the di­ (POsiTest AT-A, USA).
rection of solar irradiation. The coated-hollow cylindrical glass tube X-ray diffraction (XRD) (JEOL-JPX 8, Japan) scan, with a scanning
consisted of a 30 mm outer diameter and was covered with a 50 mm rate of 2◦ /min, was conducted using an X-ray source of Cu Kα radiation
outer diameter hollow circular glass made case allocated on the (λ = 1.5418 A◦ ) to determine the composition of the ternary composite
concentrated parabolic reactor. Hence, about 12.5 mm, horizontal space film. The sample scans were carried out with a diffraction angle range of
for the water and two taps were fixed onto the outer glass tube to feed 2θ = 10◦ –90◦ . Furthermore, Fourier Transform Infrared Spectroscopy
and discharge the water sample. (FTIR) (Bruker Alpha 400, Germany) with an attenuated total reflection
(ATR) detector was used to examine the chemical functional groups
2.4. Operational procedure present in the catalytic film. Finally, surface roughness analysis was
performed on the coated film using a surface roughness meter (surf test
The photocatalytic activity of the immobilized Fe2O3–TiO2/PVP SJ-301, Mitutoyo, Japan) to determine the average thickness and sur­
ternary composite film was evaluated using a continuous mode PTC face roughness.
reactor under solar irradiation. Three parameters, namely pH, H2O2, and
contact time, were considered during this experimental work. The 2.5.2. Kinetics
treated greywater effluent from the photocatalytic reactor was with­ The observed rate constant (kobs) has been employed as the basic
drawn at each pre-determined interval (30 min irradiation time). The kinetic parameter to evaluate the performance of different photo­
sample was further analyzed for TCS removal efficiency, and all the solar catalysts, i.e., ln [C0/C] = kobs. C0 indicates the initial concentration of
irradiation experiments were conducted in triplicates. The washing TCS, and C is the concentration of TCS remaining in the greywater
procedure was carried out to clean the immobilized Fe2O3–TiO2/PVP effluent at irradiation time (t) (Kelly et al., 2021; Ye et al., 2021; Zhou
glass pipe of all undesirable contaminants. This process was done by et al., 2019).
irradiating the immobilized glass pipe with deionized water for 30 min
to attain zero contamination. 2.5.3. Physicochemical analysis
A standard physicochemical analysis method was used to characterize
the treated effluent from the photocatalytic reactor for selected parame­
2.5. Analytical and computational methods
ters. Accordingly, TCS was analyzed using a UV–visible spectroscopy
technique. In this experimentation, 1.4 mL of 0.020 mol L− 1 sodium nitrite
2.5.1. Characterization of ternary composite (Fe2O3–TiO2/PVP) film
solutions; 1 mL of 0.25% p-sulfanilic acid solution containing 0.09 mol L− 1
Thermogravimetric Analysis (TGA) (EXSTAR TG-DTA 6300, Seiko
hydrochloric acid; 3.5 mL of TCS standard or sample solution; and 2.2 mL
Instruments, Inc., Japan) was used to determine the optimal annealing
of glycine buffer solution were gradually shifted into a 10 mL volumetric
temperature for immobilized film formation on the glass pipe. Scratch
flask. Finally, the solution was diluted, with deionized water, to a volu­
hardness testing of the samples was performed by scratch hardness
metric flask of 10.00 mL before detection at 452 nm wavelength of
testers (ERICHSEN, LINEARTESTER 249, Germany). In this method, the
extreme absorption (adopted from Lu et al., 2009).
coated sample was drawn against a stylus tip composed of a tungsten
During the degradation of TCS in greywater effluent, the intermediate
carbide insert with a tip diameter of 1 mm at a constant load of 20 N.
products were tentatively identified, using Liquid Chromatograph Mass
Using the Field Emission Scanning Electron Microscope (FESEM) (ZEISS
Spectrometer (LCMS-2020, Shimadzu, USA); a C18 reversed-phase liquid
GeminiSEM 300, Australia), the morphological study, coating thickness,
chromatography column; single quadrupole mass spectrometer; coupled
and width measurement of the scratches on the samples were carried
with Electron Spray Ionisation (ESI) operating in positive mode. The
out. In addition, an Energy Dispersive X-ray (EDX) attached to the
sample injection volume was 20 L, and the UV–Vis detector wavelength
FESEM was used to analyze the elemental composition and uniform
was set to 280 nm. The other conditions were oven temperature of 40 ◦ C,
dispersion of each element in the obtained film. In order to investigate

Fig. 1. a) The experimental setup for photocatalytic degradation of TCS in greywater effluent. ((a) Inlet chamber, (b) Peristaltic pump, (c) Hollow cylindrical glass
casing, (d) Film-coated glass pipe, (e) Stainless steel parabolic concentrator, and (f) Outlet). b) Synthesis coated sample and solar photocatalysis reactor setup.

3
S. Chandra Pragada and A.K. Thalla Journal of Environmental Management 296 (2021) 113305

N2 as a carrier gas, and pump pressure of 79 bars (Wang and Wang, 2019). the proper coating film at the required temperature. Fig. 2 (a) displays
Eventually, the amount of titanium (Ti) and iron (Fe) elements in grey­ the TGA thermogram of the ternary composite solution. For ’A’ (at
water effluent were determined using Inductively Coupled Plasma-Optical 60–80 ◦ C) and ’B’ (at 320–350 ◦ C), the two significant weight losses are
Emission Spectroscopy (ICP-OES) (5100 series; Agilent Technologies, observed. The corresponding Differential Thermal Analysis (DTA) re­
USA) to study the leaching (Cunha et al., 2018). sponses (A′ and B′ ; Fig. 2 (b)) suggests that the loss of weight around
60–80 ◦ C is due to solvent evaporation; similarly, the drastic loss of
2.5.4. Experimental design and optimization weight at 320–350 ◦ C may be attributed to the decomposition of PVP.
Many independent and influencing parameters were effective in the The decrease in the decomposition temperature of PVP may be due to
treatment of greywater effluent using photocatalysis. Box Behnken the presence of TiO2 and Fe2O3.
Design (BBD) is one of the standard designs in the Response Surface A single mechanical test is not sufficient to prove the overall me­
Methodology (RSM) used to determine the independent variables. chanical stability of the coated layer since it is affected by various pa­
Accordingly, the total number of experiments obtained using the BBD, as rameters, viz. stress, vibration, impact, etc. Therefore, it is required to
given by Eq. (1). conduct different tests such as scratch hardness, wear resistance, and
adhesion strength. The Fe2O3–TiO2/PVP coated substrate was drawn
Number of experiments = 2k + 2k + n (1)
against the stylus tip during the scratch hardness test, resulting in the
shearing of binding atoms. When the cohesive bond strength value is
Where k = no. of studied parameters; n = no. of central points.
high, the scratch hardness is expected to be maximized. Scratch hard­
ness, therefore, is an indirect measure of the cohesive bond strength of
3. Results and discussions
the catalytic film. The values of the scratch hardness (Hs) are obtained
from the below Eq. (2) (Augustin et al., 2016).
3.1. Structures and surface morphologies of immobilized ternary
composite (Fe2O3–TiO2/PVP) coated film Hs = 8FN/πb2 (2)

Thermal analysis of the precursor solution was carried out in order to Where FN is the normal force used, and b is the average scratch width.
determine the temperature at which the composite solution alters into Fig. 2 (c) shows the top view of the scratches imprinted on the coated

Fig. 2. (a) TGA thermogram (b) DTA trace for Fe2O3–TiO2/PVP composite solution and (c) FESEM micrographs of scratches on coated samples with scratch width
dimensions at different ratios of Fe2O3/TiO2 compositions of (i) 1 atomic wt.%, (ii) 3 atomic wt.%, (iii) 5 atomic wt.%, (iv) 7 atomic wt.% and (v) 9 atomic wt.%.

4
S. Chandra Pragada and A.K. Thalla Journal of Environmental Management 296 (2021) 113305

film with different Fe2O3/TiO2 compositions. It has been observed that


the 5% Fe2O3/TiO2 composition catalytic film has the optimum coating
with higher scratch hardness 7.984 GPa when compared to other Fe2O3/
TiO2 composition catalysts (Table 2).
At higher cohesive bond strength, the wear resistance of the coated
film is also enhanced. In order to understand the wear mechanism, 3D
Profilometer studies and sliding wear operations were performed. The
3D surface profiles and the corresponding profile curves of the coatings
are shown in Fig. S1. From 3D profilometer studies, maximum wear
depth at 5N and 10N are observed as 7.28 and 12.56 μm. According to
the sliding wear test, the wear of the coating is noted as 5 μm at 5N load.
As increasing load to 10N, it is observed that the wear increased to 11
μm. The stable friction coefficients of the coatings under 5N and 10N are
noted as 0.155 and 0.216, respectively (Fig. S2). The delamination of
coating observed at some locations is attributed to the formation of
micro-cracks during sliding wear operation. These cracks progress and
reach critical length as load increases. Furthermore, the mechanical
stability can also be analyzed by a pull-off test. In this experiment, the
adhesion strength of the coating film is obtained to be 0.92 MPa/S.
Eventually, the mechanical stability of the coated film is adequate in
order to utilize it as a photocatalyst.
The XRD patterns of PVP, Fe2O3, TiO2, and the ternary composites
(Fe2O3–TiO2/PVP) are shown in Fig. 3 (a). The XRD analysis for the
composite samples was carried out at an angle ranging from 10◦ to 90◦ .
Diffraction peaks are identified as too broad and shallow for PVP. This
broad curve and low intensity indicate that the PVP is amorphous or
very poor crystalline in nature. The compounds that the distinct peaks
have identified are shown in Table S1, indicating the presence of TiO2
and Fe2O3 in the composite samples. The XRD peaks of Fe2O3 (JCPDS
file No. 33–0664) in Fe2O3–TiO2/PVP (high temp at 320 ◦ C) are found at
35.71◦ , 57.39◦ , and 82.92◦ whereas, in Fe2O3–TiO2/PVP (low temp at
60 ◦ C), no peak that is related to Fe2O3 due to its low amount (only 5
atomic wt.%); its amorphous phase (small size of crystallites). Hence,
the high-temperature treatments improve the crystallinity of the coated
film (Banisharif et al., 2015).
In the FTIR spectrum, the C– – O groups of PVP exhibit a peak at
1659.4 cm‒1. This characteristic peak was investigated to explore the
interaction between PVP and metal ions (Mn+). The observed peaks in Fig. 3. (a) XRD patterns, (b) FTIR spectra of coated samples.
the ternary composite, a PVP-titanium-iron complex, are at 1403, 1287,
and 1630 cm− 1, corresponding to C–N–C, N–C, and C– – O, respectively
(K V et al., 2014). Due to both O–atom of C– – O group and N-atom of the
Fe2O3 are in a spherical-like structure with a smooth surface. It is also
heterocyclic ring can form coordinate bonds with Mn + ions, as in Fig. 3 seen that the homogeneity in morphology has been well observed. This
(b); it is evident that there is a shift in C–N–C, N–C, and C– – O group
homogeneity helps in the uniform coating of this ternary composite on
stretching frequencies in the samples. There is a slight shift in the peak the surface of the glass pipe. The obtained film thickness of the
position of the C–– O and N–C functional groups. The change in the peaks
Fe2O3–TiO2/PVP ternary composite film is in the range of 13.8–16.4 μm
towards the lower wavenumber region shows that the C–N–C bond has (Fig. 4 b) for the ternary composite obtained after calcination at 320 ◦ C.
weakened. There is an interaction between titanium ions, ferric ions, It has also been observed that the film obtained is more uniform and has
and PVP through the oxygen of the C– – O group of the polymer. This
low thickness. The immobilized catalytic film thickness varies in a
characterization represents the formation of C– – O, C–N–C, and –N–C
range, from 6.9 to 34.1 μm, and is practically suitable for photocatalysis
bonds from oxidized PVP, which act as an electron donor in enhancing (Han and Bai, 2011). The R-profile is determined by a surface roughness
photoactivity. meter for the coated film of Fe2O3–TiO2/PVP, whose thickness is found
The surface morphology of Fe2O3–TiO2/PVP composites was exam­ to be 15.84 μm (shown in Fig. S3). Furthermore, EDAX spectrum has
ined by FESEM, as shown in Fig. 4 (a). It is observed that the TiO2 and confirmed the Fe2O3–TiO2/PVP ternary composite elements (Fig. 4 c), as
it exhibits the characteristic peaks of O, C, N, Ti, and Fe. The uniform
distribution of elements on the surface of the ternary composites is
observed on the surface elemental mapping (see Fig. 4 d-h).
Table 2
Scratch hardness at different ratio Fe2O3–TiO2 coated samples.
Fe2O3/TiO2 Scratch width (μm) Scratch 3.2. Immobilized photocatalytic degradation of TCS
composition hardness
Position1 Position Position Average
(GPa)
2 3 3.2.1. Kinetics of photodegradation
1% 108.3 107.7 114.8 110.27 4.186 Fig. S4 (a) plots the variation in ln [C0/C] as a function of reaction
3% 108.3 91.8 141.6 113.9 3.924 time. The Fe2O3–TiO2/PVP catalyst confirms that the photocatalysis
5% 79.3 74.38 85.86 79.85 7.984 process involves k1 of 0.105 min–1 and k2 of 0.015 min–1 under sunlight.
7% 102.3 100.6 101.2 101.37 4.954 The Langmuir-Hinshelwood kinetic model exhibits as the best suitable
9% 109.9 106.6 109.9 108.8 4.3
for photodegradation of TCS since the R2 value is 0.95 (shown in Fig. S4

5
S. Chandra Pragada and A.K. Thalla Journal of Environmental Management 296 (2021) 113305

Fig. 4. (a) FESEM image, (b) Thickness, (c) EDAX spectrum, (d) O mapping, (e) C mapping, (f) N mapping, (g) Ti mapping, (h) Fe mapping for immobilized
photocatalyst.

b). The best fit model is the pseudo-first-order kinetic model from this the process, using 1–10 mg/L TCS, demonstrates that the observed rate
analysis, as can be seen when comparing it with the pseudo-second- is constant (kTCS) and ranged from 0.3405 to 0.0687 min− 1 under UV
order kinetic model (R2 = 0.88) (shown in Fig. S5). Further, kTCS has light. Hence, the kTCS (i.e., k1 = 0.105 min− 1) for this present study has
been compared with different photocatalytic treatment systems of TCS been observed in the adequate range, under sunlight, with the immo­
in the present study. Through a pseudo-first-order model, the kinetics of bilized Fe2O3–TiO2/PVP catalysts (Azarpira et al., 2019).

6
S. Chandra Pragada and A.K. Thalla Journal of Environmental Management 296 (2021) 113305

3.2.2. Response Surface Methodology 4.15) has shown an indirect relationship between the initial concen­
The effect of operational parameters on the TCS degradation output tration of H2O2 and TCS degradation, which decreases as an increase in
has been optimized using a BBD. The BBD has allowed the most signif­ H2O2 concentrations. When the concentration of H2O2 is higher, the
icant parameters and their interactions to be studied, specifically, the excess H2O2 entraps the OH• radicals to form weaker oxidant H2O•
contact time, the initial concentration of H2O2, and the pH. Table 3 radicals (Reza et al., 2017; Tseng et al., 2012). This undesirable reaction
summarizes the efficacy of 17 sets of BBD-based tests. has retarded the TCS removal efficiency.

3.2.2.1. RSM results. The model equation for the TCS degradation 3.2.2.2. Model validation. The analysis of variance (ANOVA) is used to
yield, attained using statistical tests after 17 experiments and clearance verify the adequacy of the second-order model, showing the importance
the negligible effects, is as follows (Eq. (3)): of the effect of the individual parameter. The results are shown in
Table S2. The F-value in the model of photocatalysis is 93.55, which
Y = 43.33 + 17.33 x1 − 4.15 x2 + 17.43 x3 − 3.01 x1 x3 + 6.86 x12 (3)
indicates that the model is significant. As the P-value is below 0.05, the
The statistical model has made it possible to predict the contact time model parameters are vital in predicting the TCS removal efficiency. The
(x1), the initial concentration of H2O2 (x2), and the pH (x3) with good predicted R2 of 0.9346 is in reasonable agreement with the Adjusted R2
accuracy (the correlation coefficient is found to be R2 = 0.99). of 0.9812; i.e., the difference is less than 0.2. The closer the R2 to 1, the
An explicit agreement between the values observed and those pre­ better the relationship between the independent variables is expressed,
dicted by the statistical model is shown by comparing experimental and and the response becomes more precise as well. The adjusted R2 = 0.98
calculated responses (see Fig. 5a). The residual plot of random distri­ is very close to the R2 value for photocatalysis, indicating no additional
bution (see Fig. 5 b) shows a lack of trend, indicating that the mathe­ terms in the model. Pred-R2 for the process is equal to 0.93, which shows
matical model is adequate. There is no inconsistency between the the ability of the model to predict TCS removal. The COV value of the
experimental and measured response values. According to the equation presented model in the photocatalysis process is 5.59%, which is less
of regression (Eq. (3)), the contact time (x1) and pH solution (x3) has the than 10%; the model proposed is reliable and accurate (Hassanshahi and
greatest impact on the response (b1 = +17.33 and b3 = +17.43). The Karimi-Jashni, 2018).
positive effect of these variables specifies that TCS degradation has
improved for increasing the contact time and pH solution. 3.2.2.3. Optimization. In order to determine the optimal value of the
TCS has a pKa of 7.9–8.1 (a weak base) and readily ionizes at envi­ independent parameters, the pH is initially maximized to be similar to
ronmental pH (Quan et al., 2019). Based on the pH effect study, the the pH of raw greywater for the highest removal efficiency of TCS in the
removal efficiency of TCS at pH > pKa is about 79.21%, and it decreases photocatalytic process; to minimize the cost of chemicals required for
marginally (to ~ 72.6%) because pH is neutral. Fe2O3–TiO2/PVP has a pH adjustment. In addition, the use of metal oxide concentration is
point of zero charges at pH = 6.8 (Fig. S6), indicating that its surface has minimized to be economical in the immobilized coating technique.
a positive charge at pH lower than 6.8 and a deprotonation negative Finally, to reduce the use of the UV lamp, the reaction time is moderate
charge at pH higher than 6.8. The substantial removal of TCS from a with solar light absorption. After performing trials of the various oper­
strongly basic solution is, therefore, due to other adsorption processes, ating parameters by BBD, the response surface and contour plot have
such as hydrogen bonding and precipitation. Dissociation of TCS occurs been drawn using Design-Expert 11 software (see Fig. 5c and d). This
at pH < pKa, resulting in a positive charge. Thus, between 6.8 (PZC of plot analysis allows the following optimal conditions to be deduced for
Fe2O3–TiO2/PVP) and 7.9–8.1 (pKa of TCS), electrostatic attraction the TCS degradation at pH of 10, a contact time of 300 min, and an
plays an important role in attachment TCS to the positively charged initial H2O2 concentration of 1 mM, resulting in TCS degradation of
Fe2O3–TiO2/PVP surface. Fe2O3–TiO2/PVP become positively charged 83.27%. The comparison of the present study with other photocatalytic
at pH < 6.8, resulting in a decrease in TCS removal (~51%) as elec­ treatment of TCS in an aqueous medium is shown in Table 4.
trostatic repulsion (between the TCS positively charged and the catalyst
surface positively charged) is dominant. However, the reduction of TCS 3.2.3. Plausible degradation pathway of TCS
is significant between pH 7 and 10 (increasing from ~72.6 to 79.21%), According to preliminary studies, TCS is not completely degraded
and the target TCS is more efficient of photocatalytic degradation in during the photocatalysis process. It is also made possible that the
strongly alkaline conditions. The other significant parameter is the photocatalytic process could result in some intermediates derived from
initial concentration of H2O2. The negative sign of this coefficient (b2 = - the parent compound. The final irradiated sample was analyzed using
LC-MS to identify the intermediate compounds formed during photo­
catalysis. Since the photocatalytic degradation of TCS is carried out on
Table 3 the synthetic greywater effluent, there is a possibility of different noise
Box Behnken Design matrix in the photocatalyst process. peaks (as shown in Fig. S7) and interference of other constituents pre­
Irradiation time H2O2 pH Actual response (TCS removal sent in greywater.
(min) (mM) efficiency, %) Majorly, TCS has undergone three different reaction mechanisms to
180 25.5 7 45.81
produce different intermediates through hydroxylation, OH radical
60 1 7 31.28 attack on the aromatic ring, and de-chlorination reactions. These in­
180 50 10 54.35 termediates subsequently degraded to other smaller organic in­
180 1 10 61.65 termediates, which can further mineralize to CO2 and water. Based on
300 50 7 62.69
these LC-MS data and comparison with the available literature, the
300 25.5 4 51.03
300 1 7 72.59 plausible degradation pathway of TCS is elucidated. Eleven intermediate
180 25.5 7 41.22 products have been identified, and their molecular structures are
180 25.5 7 46.25 inferred at m/z 318, m/z 306, m/z 255, m/z 248, m/z 235, m/z 199, m/
300 25.5 10 79.21 z 163, m/z 147, m/z 142, m/z 113, m/z 109 and m/z 94. Fig. 6 displays
60 25.5 10 51.46
180 1 4 29.69
the photocatalytic degradation pathway of TCS. Firstly, the attack of the
60 25.5 4 11.26 h+ and •OH radicals, produced from the solar irradiation of the immo­
180 25.5 7 42.13 bilized Fe2O3–TiO2/PVP catalyst, has led to forming the chlorophenol
60 50 7 29.69 intermediates I1 (m/z 306) by hydroxylation reaction and dispropor­
180 50 4 15.29
tionation mechanism. The chlorophenol intermediates are unstable.
180 25.5 7 41.25

7
S. Chandra Pragada and A.K. Thalla Journal of Environmental Management 296 (2021) 113305

Fig. 5. Response surface analysis (a) experimental and theoretical responses, (b) predictable model’s residual analysis, (c) Response surfaces, and (d) contour plots
showing the effect of the contact time and the pH solution on the TCS degradation.

Table 4
Comparison of the present study with other photocatalytic treatments of TCS from the aqueous medium.
S.No. PPCPs Catalysts combination Light source Optimal conditions efficiency (%) References

1 TCS TiO2 nanotubes UV light Time >30 min 67.2 Liu et al. (2013)
2 TCS TiO2 films on porous Tezontle stone Solar simulator pH = 5 65.9 Martínez et al. (2014)
pH = 7 58.5
pH = 9 74.7
3 TCS ZnO powder UV light Time = 90 min 90 Kosera et al. (2017)
4 TCS immobilized catalyst Fe/Nb2O5 UV light pH = 8.5, Time = 80 min 91 Fidelis et al. (2019)
5 TCS Ag/BiVO4/rGO Simulated sunlight Time = 120 min 59.7 Li et al. (2019)
6 TCS TiO2 nano tubes UV light pH = 5.8, Time>30 min 62.2 Arifin et al. (2021)
7 TCS Fe2O3–TiO2/PVP Sun light pH = 10, Time = 300 min 83.72 Present study

Therefore, they are further oxidized to quinone, I2 (m/z 318), and intermediate I4 is reduced to another intermediate (I5), identifies as
finally mineralized into insignificant molecular inorganic acids (Li et al., chlorobenzene (m/z 113), through de-chlorination reaction (Xie et al.,
2019). The intermediate I2 sequentially undergoes C–O bond cleavage 2019). Through hydroxylation reaction, the intermediate I4 is further
due to the •OH radical attack and another product at m/z 142 (I3). transformed into another product with m/z 163 (I6). And it is also
Secondly, •OH radicals and h+ attack the ether bond of TCS; another reduced to another compound, I7, through successive de-chlorination
intermediate I4; identified as m-Dichlorobenzene (m/z 147), which is reactions, which are identified as phenol (m/z 94).
generated by OH radical attack and aromatic ring cleavage. Further, On the other hand, the intermediate, (I8) 4-chlororesorcinol (m/z

8
S. Chandra Pragada and A.K. Thalla Journal of Environmental Management 296 (2021) 113305

Fig. 6. Plausible degradation pathway of TCS by immobilized ternary composite (Fe2O3–TiO2/PVP) coating catalyst.

143) is derived from TCS due to C–O cleavage (Chen et al., 2015); compounds by a ring-opening mechanism.
further, it is reduced to another intermediate, (I9); the de-chlorination
reaction can also form resorcinol (m/z 109) from 4-chlororesorcinol. 3.2.4. Toxicity studies
Similarly, in another reaction mechanism, 5-chloro-2-(2chlorophenoxy) The developed Fe2O3–TiO2/PVP catalyst has applications in tertiary
phenol, I10 (m/z 255) is generated from TCS by a de-chlorination re­ treatments such as photocatalysis that are exposed to aquatic environ­
action since the C–Cl bond cleavage formation in the effluent treatment. ments. So, it is important to investigate the stability and toxicity of any
The same compound has also been identified in the literature of other metals leached out of the coated film into the aquatic environment.
works. Further, I10 undergoes Cl substitution by OH and results in Leaching of metals from the Fe2O3–TiO2/PVP catalyst causes metals
another intermediate I11, identified as 5-(4-Chlorobenzyl)-2-furoic acid, bioaccumulation and biomagnification, which may have short and long-
with the m/z of 235 and is released of HCl. Further, the intermediate I11 term effects on aquatic life and human health.
can result in two more transformation products with hydroxylation and As shown in Table S3, the leaching analysis of the catalyst with ICP/
de-chlorination reactions. First, the intermediate I12 with m/z 248 has OES shows that the Ti and Fe concentrations in the immersion solution
resulted from hydroxylation reaction on the side chain of the interme­ are less than the detection limit of WHO standards (i.e., 0.1 mg/L for Ti
diate I10. Second, the same intermediate I10 endures a de-chlorination and 0.3 mg/L for Fe) at all conditions. The amount of Ti released is
reaction and produces another intermediate I13 at m/z 199. All the almost negligible, which is caused by the addition of Fe2O3 to the
formed intermediates are possibly mineralized to simpler inorganic catalyst (Wahyuningsih et al., 2018). So, the catalytic film not only

9
S. Chandra Pragada and A.K. Thalla Journal of Environmental Management 296 (2021) 113305

provides a stable surface that prevents the peel out of metals from the to pursue his research studies at NITK Surathkal. Also, the Authors thank
coating layer but also delivers the highest TCS degradation rate. Dept. of materials & metallurgy (NITK), Dept. of Chemistry & Physics
(NITK), Dept. of Chemical engineering (NITK), Dept. of Mechanical
3.3. Practical applications and future research prospects engineering (NITK), and Central Research Facility (NITK), for extending
their help and support in instrumental analysis.
In developing countries, as rate of generation of greywater rises due
to population growth, the emerging contaminants have become an Appendix A. Supplementary data
increasing area of concern due to the adverse effects of these compounds
on human health and environment. So, greywater has been identified as Supplementary data to this article can be found online at https://doi.
an important waste that must be treated with appropriate treatment org/10.1016/j.jenvman.2021.113305.
technology. Many developed countries, however, have implemented
conventional treatment techniques that are not only ineffective at Credit author statement
eliminating emerging contamination, but also expensive and energy-
intensive. To address these challenges, the present study is conducted Sarath Chandra Pragada: Conceptualization; Formal analysis;
with immobilized Fe2O3–TiO2/PVP catalyst under natural sunlight are Investigation; Methodology; Validation; Visualization; Roles/Writing -
proven to be one of the effective techniques for the greywater treatment. original draft. Arun Kumar Thalla: Conceptualization; Investigation;
Although applications of immobilized photocatalysts on various Funding acquisition; Supervision; Writing - review & editing.
supports have been reported in the literature, they are still limited to the
large-scale applications. Most of the studies were on simulated waste­ References
water, and now significant focus on actual field water is essential. The
design of the system is also an essential aspect in the technology Abbas, N., Shao, G.N., Haider, M.S., Imran, S.M., Park, S.S., Kim, H.T., 2016. Sol–gel
implementation to the real field. synthesis of TiO2-Fe2O3 systems: effects of Fe2O3 content and their photocatalytic
properties. J. Ind. Eng. Chem. 39, 112–120. https://doi.org/10.1016/j.
Prevention of pollution is one of the practical approaches that re­ jiec.2016.05.015.
duces, eliminates, or prevents waste generation at its source. Novel Abu Ghunmi, L., Zeeman, G., Fayyad, M., Lier, J.B. van, 2010. Grey water treatment in a
composite catalyst developed from this study used to treat greywater series anaerobic - aerobic system for irrigation. Bioresour. Technol. 101, 41–50.
https://doi.org/10.1016/j.biortech.2009.07.056.
effluent that ultimately minimizes ecological impacts effectively.
Ansari, K., Shrikhande, A.N., 2019. Feasibility on grey water treatment by
Similar results were observed in the past (Priyanka et al., 2019). electrocoagulation process: a review. Int. J. Emerg. Technol. 10, 85–92.
Reclaiming greywater is considered as a sustainable approach to Arifin, S., nor, H., Mohamed, R., Al-Gheethi, A., Lai, C.W., Yashni, G., 2021.
wastewater reuse and a feasible option to decrease domestic demand for Heterogeneous photocatalysis of triclocarban and triclosan in greywater: a
systematic and bibliometric review analysis. Int. J. Environ. Anal. Chem. 1–19.
freshwater in accordance with the circular bio-economy approach. https://doi.org/10.1080/03067319.2020.1863391.
Furthermore, the regeneration and stability of the photocatalyst can be Augustin, A., Huilgol, P., Udupa, K.R., K, U.B., 2016. Effect of current density during
viewed as influential factors in providing catalyst recovery benefits in electrodeposition on microstructure and hardness of textured Cu coating in the
application of antimicrobial Al touch surface. J. Mech. Behav. Biomed. Mater. 63,
treating greywater. 352–360. https://doi.org/10.1016/j.jmbbm.2016.07.013.
Azarpira, H., Sadani, M., Abtahi, M., Vaezi, N., Rezaei, S., Atafar, Z., Mohseni, S.M.,
4. Conclusions Sarkhosh, M., Ghaderpoori, M., Keramati, H., Hosseini Pouya, R., Akbari, A.,
fanai, V., 2019. Photo-catalytic degradation of triclosan with UV/iodide/ZnO
process: performance, kinetic, degradation pathway, energy consumption and
The photocatalytic degradation of TCS, with immobilized toxicology. J. Photochem. Photobiol. Chem. 371, 423–432. https://doi.org/
Fe2O3–TiO2/PVP on a cylindrical glass reactor, by a spray coating 10.1016/j.jphotochem.2018.10.041.
Banisharif, A., Khodadadi, A.A., Mortazavi, Y., Anaraki Firooz, A., Beheshtian, J.,
technique, is found to be a useful method. This immobilization strategy Agah, S., Menbari, S., 2015. Highly active Fe2O3-doped TiO2 photocatalyst for
shows a few advantages: rapid and effective surface layer, no require­ degradation of trichloroethylene in air under UV and visible light irradiation:
ment of high temperature (>400 ◦ C) process. The coating structure has experimental and computational studies. Appl. Catal. B Environ. 165, 209–221.
https://doi.org/10.1016/j.apcatb.2014.10.023.
resulted in higher scratch hardness (7.984 GPa), more homogeneous,
Behpour, M., Shirazi, P., Rahbar, M., 2019. Immobilization of the Fe2O3/TiO2
and uniform surface morphology for 5% Fe2O3/TiO2 composite catalyst. photocatalyst on carbon fiber cloth for the degradation of a textile dye under visible
Meanwhile, the obtained film thickness of the ternary composite is light irradiation. React. Kinet. Mech. Catal. 127, 1073–1085. https://doi.org/
approximately 15.84 μm, which is suitable for the photocatalytic pro­ 10.1007/s11144-019-01581-1.
Belapurkar, A.D., Sherkhane, P., Kale, S.P., 2006. Disinfection of drinking water using
cess. The pseudo-first-order kinetic model is well fitted (R2 = 0.95) than photocatalytic technique. Curr. Sci. 91, 73–76.
the pseudo-second-order kinetic model for the photodegradation of TCS Boano, F., Caruso, A., Costamagna, E., Ridolfi, L., Fiore, S., Demichelis, F., Galvão, A.,
under solar irradiation. Surface response analysis and contour plots have Pisoeiro, J., Rizzo, A., Masi, F., 2020. A review of nature-based solutions for
greywater treatment: applications, hydraulic design, and environmental benefits.
exploited the optimal conditions for degradation of TCS, resulting in Sci. Total Environ. 711, 134731. https://doi.org/10.1016/j.scitotenv.2019.134731.
83.27% of TCS removal in greywater effluent. The plausible TCS Chen, X., Casas, M.E., Nielsen, J.L., Wimmer, R., Bester, K., 2015. Identification of
degradation pathway has been proposed based on LC-MS results. Triclosan-O-Sulfate and other transformation products of Triclosan formed by
activated sludge. Sci. Total Environ. 505, 39–46. https://doi.org/10.1016/j.
scitotenv.2014.09.077.
Declaration of competing interest Cunha, D.L., Kuznetsov, A., Achete, C.A., Eduardo, A., Marques, M., 2018. Immobilized
TiO2 on Glass Spheres Applied to Heterogeneous Photocatalysis : Photoactivity,
Leaching and Regeneration Process, pp. 1–19. https://doi.org/10.7717/peerj.4464.
The authors declare that they have no known competing financial Czech, B., Zygmunt, P., Kadirova, Z.C., Yubuta, K., Hojamberdiev, M., 2020. Effective
interests or personal relationships that could have appeared to influence photocatalytic removal of selected pharmaceuticals and personal care products by
the work reported in this paper. elsmoreite/tungsten oxide@ ZnS photocatalyst. J. Environ. Manag. 270 https://doi.
org/10.1016/j.jenvman.2020.110870.
Fidelis, M.Z., Abreu, E., Dos Santos, O.A.A., Chaves, E.S., Brackmann, R., Dias, D.T.,
Acknowledgment Lenzi, G.G., 2019. Experimental design and optimization of triclosan and 2.8-
diclorodibenzeno-p-dioxina degradation by the Fe/Nb2O5/UV system. Catalysts 9,
1–18. https://doi.org/10.3390/catal9040343.
All photocatalysis experiments conducted in the present study have
Han, H., Bai, R., 2011. Effect of thickness of photocatalyst film immobilized on a buoyant
been carried out as a part of the research project titled "Small scale and substrate on the degradation of methyl orange dye in aqueous solutions under
sustainable household greywater recycling (S3HWR)". Grant no. different light irradiations. Ind. Eng. Chem. Res. 50, 11922–11929. https://doi.org/
IMPRINT 5670 is funded by the MHRD & MOUD Government of India. 10.1021/ie200787j.
Harrow, D.I., Felker, J.M., Baker, K.H., 2011. Impacts of triclosan in greywater on soil
The Authors thank the Ministry of Human Resources Development, microorganisms. Appl. Environ. Soil Sci. 1–8. https://doi.org/10.1155/2011/
Govt. of India, for providing fellowship to Mr. Sarath Chandra Pragada 646750, 2011.

10
S. Chandra Pragada and A.K. Thalla Journal of Environmental Management 296 (2021) 113305

Hassanshahi, N., Karimi-Jashni, A., 2018. Comparison of photo-Fenton, O3/H2O2/UV Priyanka, K., Remya, N., Behera, M., 2019. Comparison of titanium dioxide based
and photocatalytic processes for the treatment of gray water. Ecotoxicol. Environ. catalysts preparation methods in the mineralization and nutrients removal from
Saf. 161, 683–690. https://doi.org/10.1016/j.ecoenv.2018.06.039. greywater by solar photocatalysis. J. Clean. Prod. 235, 1–10. https://doi.org/
Jagini, S., Konda, S., Bhagawan, D., Himabindu, V., 2019. Emerging contaminant 10.1016/j.jclepro.2019.06.314.
(triclosan) identification and its treatment: a review. SN Appl. Sci. 1, 1–15. https:// Pycke, B.F.G., Crabbé, A., Verstraete, W., Leys, N., 2010. Characterization of triclosan-
doi.org/10.1007/s42452-019-0634-x. resistant mutants reveals multiple antimicrobial resistance mechanisms in
K V, A., Veeraiah, M.K., P, H., M, M., 2014. Synthesis and characterisation of poly rhodospirillum rubrum S1H. Appl. Environ. Microbiol. 76, 3116–3123. https://doi.
(vinylpyrrolidone) – nickel (II) complexes. IOSR J. Appl. Chem. 7, 61–66. https:// org/10.1128/AEM.02757-09.
doi.org/10.9790/5736-07816166. Qin, L., Pan, X., Wang, L., Sun, X., Zhang, G., Guo, X., 2014. Facile preparation of
Kelly, J., McDonnell, C., Skillen, N., Manesiotis, P., Robertson, P.K.J., 2021. Enhanced mesoporous TiO2(B) nanowires with well-dispersed Fe2O3 nanoparticles and their
photocatalytic degradation of 2-methyl-4-chlorophenoxyacetic acid (MCPA) by the photochemical catalytic behavior. Appl. Catal. B Environ. 150– 151, 544–553.
addition of H2O2. Chemosphere 275, 130082. https://doi.org/10.1016/j. https://doi.org/10.1016/j.apcatb.2013.12.055.
chemosphere.2021.130082. Quan, B., Li, X., Zhang, H., Zhang, C., Ming, Y., Huang, Y., Xi, Y., Weihua, X., Yunguo, L.,
Kosera, V.S., Cruz, T.M., Chaves, E.S., Tiburtius, E.R.L., 2017. Triclosan degradation by Tang, Y., 2019. Technology and principle of removing triclosan from aqueous media:
heterogeneous photocatalysis using ZnO immobilized in biopolymer as catalyst. a review. Chem. Eng. J. 378, 122185. https://doi.org/10.1016/j.cej.2019.122185.
J. Photochem. Photobiol. Chem. 344, 184–191. https://doi.org/10.1016/j. Reza, K.M., Kurny, A., Gulshan, F., 2017. Parameters affecting the photocatalytic
jphotochem.2017.05.014. degradation of dyes using TiO2: a review. Appl. Water Sci. 7, 1569–1578. https://
Köwitsch, I., Mehring, M., 2021. Coatings of magnetic composites of iron oxide and doi.org/10.1007/s13201-015-0367-y.
carbon nitride for hotocatalytic water purification. RSC Adv. 11, 14053–14062. Samayamanthula, D.R., Sabarathinam, C., Bhandary, H., 2019. Treatment and effective
https://doi.org/10.1039/d1ra00790d. utilization of greywater. Appl. Water Sci. 9, 1–12. https://doi.org/10.1007/s13201-
Laaffat, J., Aziz, F., Ouazzani, N., Mandi, L., 2019. Biotechnological approach of 019-0966-0.
greywater treatment and reuse for landscape irrigation in small communities. Saudi Shaikh, I.N., Ahammed, M.M., 2020. Quantity and quality characteristics of greywater: a
J. Biol. Sci. 26, 83–90. https://doi.org/10.1016/j.sjbs.2017.01.006. review. J. Environ. Manag. 261, 110266. https://doi.org/10.1016/j.
Lépesová, K., Krahulcová, M., Mackuľak, T., Bírošová, L., 2019. Sewage sludge as a jenvman.2020.110266.
source of triclosan-resistant bacteria. Acta Chim. Slovaca 12, 34–40. https://doi.org/ Shan, A.Y., Ghazi, T.I.M., Rashid, S.A., 2010. Immobilisation of titanium dioxide onto
10.2478/acs-2019-0006. supporting materials in heterogeneous photocatalysis: a review. Appl. Catal. Gen.
Li, M., Xu, G., Guan, Z., Wang, Y., Yu, H., Yu, Y., 2019. Synthesis of Ag/BiVO4/rGO 389, 1–8. https://doi.org/10.1016/j.apcata.2010.08.053.
composite with enhanced photocatalytic degradation of triclosan. Sci. Total Environ. Sharma, K., Kaushik, G., Thotakura, N., Raza, K., Sharma, N., Nimesh, S., 2020.
664, 230–239. https://doi.org/10.1016/j.scitotenv.2019.02.027. Enhancement effects of process optimization technique while elucidating the
Liu, H., Cao, X., Liu, G., Wang, Y., Zhang, N., Li, T., Tough, R., 2013. degradation pathways of drugs present in pharmaceutical industry wastewater using
Photoelectrocatalytic degradation of triclosan on TiO2 nanotube arrays and toxicity Micrococcus yunnanensis. Chemosphere 238, 124689. https://doi.org/10.1016/j.
change. Chemosphere 93, 160–165. https://doi.org/10.1016/j. chemosphere.2019.124689.
chemosphere.2013.05.018. Tseng, D.H., Juang, L.C., Huang, H.H., 2012. Effect of oxygen and hydrogen peroxide on
Liu, H., Shon, H.K., Sun, X., Vigneswaran, S., Nan, H., 2011. Preparation and the photocatalytic degradation of monochlorobenzene in TiO2 aqueous suspension.
characterization of visible light responsive Fe2O3-TiO2 composites. Appl. Surf. Sci. Int. J. Photoenergy. https://doi.org/10.1155/2012/328526.
257, 5813–5819. https://doi.org/10.1016/j.apsusc.2011.01.110. Vuppaladadiyam, A.K., Merayo, N., Prinsen, P., Luque, R., Blanco, A., Zhao, M., 2019.
Lu, H., Ma, H., Tao, G., 2009. Spectrophotometric determination of triclosan in personal A review on greywater reuse: quality, risks, barriers and global scenarios. Rev.
care products. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 73, 854–857. Environ. Sci. Biotechnol. 18, 77–99. https://doi.org/10.1007/s11157-018-9487-9.
https://doi.org/10.1016/j.saa.2009.04.007. Wahyuningsih, S., Ramelan, A.H., Prasetyawati, L., Saputri, L.N.M.Z., Ichsan, S.,
Manu, D.S., Thalla, A.K., 2017. Influence of Various Operating Conditions on Wastewater Kristiawan, Y.R., 2018. The influence of Fe2O3 addition on the TiO2 structure and
Treatment in an AS-biofilm Reactor and Post-treatment Using TiO2 Based Solar/UV photoactivity properties. IOP Conf. Ser. Mater. Sci. Eng. 333 https://doi.org/
Photocatalysis 3330. https://doi.org/10.1080/09593330.2017.1420697. 10.1088/1757-899X/333/1/012033.
Martínez, S., Morales-Mejía, J.C., Hernández, P.P., Santiago, L., Almanza, R., 2014. Solar Wang, S., Wang, J., 2019. Activation of peroxymonosulfate by sludge-derived biochar for
photocatalytic oxidation of Triclosan with TiO2 immobilized on volcanic porous the degradation of triclosan in water and wastewater. Chem. Eng. J. 356, 350–358.
stones on a CPC pilot scale reactor. Energy Procedia 57, 3014–3020. https://doi.org/ https://doi.org/10.1016/j.cej.2018.09.062.
10.1016/j.egypro.2014.10.337. Xie, X., Chen, C., Wang, X., Li, J., Naraginti, S., 2019. Efficient detoxification of triclosan
Mukherjee, D., Barghi, S., Ray, A.K., 2014. Preparation and characterization of the TiO2 by a S-Ag/TiO2@g-C3N4 hybrid photocatalyst: process optimization and bio-toxicity
immobilized polymeric photocatalyst for degradation of aspirin under UV and solar assessment. RSC Adv. 9, 20439–20449. https://doi.org/10.1039/c9ra03279g.
light. Processes 2, 12–23. https://doi.org/10.3390/pr2010012. Yashni, G., Al-Gheethi, A., Mohamed, R., Hossain, M.S., Kamil, A.F., Abirama
Nawawi, W., Zaharudin, R., Ishak, M., Ismail, K., Zuliahani, A., 2016. The preparation Shanmugan, V., 2020. Photocatalysis of xenobiotic organic compounds in greywater
and characterization of immobilized TiO2/PEG by using DSAT as a support binder. using zinc oxide nanoparticles: a critical review. Water Environ. J. 0– 1 https://doi.
Appl. Sci. 7, 24. https://doi.org/10.3390/app7010024. org/10.1111/wej.12619.
Orhon, K.B., Orhon, A.K., Dilek, F.B., Yetis, U., 2017. Triclosan removal from surface Ye, X., Wang, M., Wang, P., Li, R., Guo, M., Yu, P., Liu, H., Shi, F., Wang, Y., Du, J., 2021.
water by ozonation - kinetics and by-products formation. J. Environ. Manag. 204, Nanoscale localization of the near-surface nitrogen vacancy center assisted by a
327–336. https://doi.org/10.1016/j.jenvman.2017.09.025. silicon atomic force microscopy probe. JPhys Photonics 3. https://doi.org/10.1088/
Poulopoulos, S.G., Yerkinova, A., Ulykbanova, G., Inglezakis, V.J., 2019. Photocatalytic 2515-7647/abcd86.
treatment of organic pollutants in a synthetic wastewater using UV light and Zhou, G., Zhao, T., Qian, R., Xia, X., Dai, S., Alsaedi, A., Hayat, T., Pan, J.H., 2019.
combinations of TiO2, H2O2 and Fe(III). PloS One 14, 1–20. https://doi.org/ Decorating (0 0 1) dominant anatase TiO2 nanoflakes array with uniform WO3
10.1371/journal.pone.0216745. clusters for enhanced photoelectrochemical water decontamination. Catal. Today
335, 365–371. https://doi.org/10.1016/j.cattod.2018.12.004.

11

You might also like