Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Chemosphere 287 (2022) 132278

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

Modified TiO₂ nanotubes-zeolite composite photocatalyst: Characteristics,


microstructure and applicability for degrading triclocarban
Siti Nor Hidayah Arifin a, Radin Maya Saphira Radin Mohamed a, **, Adel Ali Al-Gheethi a, *,
Lai Chin Wei b, G. Yashni c, Nurina Fitriani d, Mu. Naushad e, Ahmad B. Albadarin f
a
Micropollutant Research Centre (MPRC), Faculty of Civil Engineering and Built Environment, Universiti Tun Hussein Onn Malaysia (UTHM), 86400, Parit Raja, Batu
Pahat, Johor, Malaysia
b
Nanotechnology & Catalysis Research Centre (NANOCAT), Institute of Postgraduate Studies (IPS), University of Malaya, 3rd Floor, Block A, 50603, Kuala Lumpur,
Malaysia
c
School of Applied Science, Faculty of Engineering, Science and Technology, Nilai University, 71800, Nilai, Negeri Sembilan, Malaysia
d
Biology Department, Faculty of Science and Technology, Universitas Airlangga, Kampus C Jl.Mulyorejo, Surabaya, 60115, Indonesia
e
Department of Chemistry, College of Science, King Saud University, P.O. Box 2455, Riyadh-11451, Saudi Arabia
f
Department of Chemical Sciences, Bernal Institute, University of Limerick, Limerick, Ireland

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• The characteristics of modified TiO2


nanotubes with zeolite composite was
described.
• MTNZC exhibited 81.2% of the triclo­
carban degradation within 50 min.
• The reusability of MTNZC exhibited be­
tween 94.2 and 77.4% of TCC
degradation.

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

Handling Editor: Derek Muir The study explored the characteristics and effectiveness of modified TiO2 nanotubes with zeolite as a composite
photocatalyst (MTNZC) for the degradation of triclocarban (TCC) from the aqueous solution. MTNZC samples
Keywords: have been produced via electrochemical anodisation (ECA) followed by electrophoretic deposition (EPD). Three
Electrochemical anodisation independent factors selected include MTNZC size (0.5–1 cm2), pH (3–10), and irradiation time (10–60 min). The
Electrophoretic deposition
observation revealed that the surface of Ti substrate by the 40 V of anodisation and 3 h of calcination was
Photocatalytic degradation
covered with the array ordered, smooth and optimum elongated nanotubes with average tube length was
Triclocarban
Zeolite approximately 5.1 μm. EDS analysis proved the presence of Si, Mg, Al, and Na on MTNZC due to the chemical
composition present in the zeolite. The average crystallite size of TiO₂ nanotubes increased from 2.07 to 3.95 nm
by increasing anodisation voltage (10, 40, and 60 V) followed by 450 ◦ C of calcination for 1, 3, and 6 h,
respectively. The optimisation by RSM shows the F-value (36.12), the p-value of all responses were less than
0.0001, and the 95% confidence level of the model by all the responses indicated the model was significant. The
R2 in the range of 0.9433–0.9906 showed the suitability of the model to represent the actual relationship among

* Corresponding author.
** Corresponding author.
E-mail addresses: maya@uthm.edu.my (R.M.S. Radin Mohamed), adel@uthm.edu.my (A.A. Al-Gheethi).

https://doi.org/10.1016/j.chemosphere.2021.132278
Received 4 July 2021; Received in revised form 2 September 2021; Accepted 16 September 2021
Available online 18 September 2021
0045-6535/© 2021 Elsevier Ltd. All rights reserved.
S.N.H. Arifin et al. Chemosphere 287 (2022) 132278

the parameters. The photocatalytic degradation rate of TCC from the first and the fifth cycles were 94.2 and
77.4%, indicating the applicability of MTNZC to be used for several cycles.

1. Introduction Zhang et al., 2018). Coating TiO₂ with zeolite could have significant
oxidation; either lower or higher oxidation states improve the electrical
The continuous persistence of triclocarban (TCC) in rivers and conductivity. The lower oxidation state assists in achieving
streams is a major issue to the environment and humankind due to the electro-neutrality by removing electrons from the valence band (Prakash
toxicity of the substance and the hazardous effect on the ecosystem et al., 2019).
(Rochester et al., 2017). TCC is an antimicrobial agent that has been Zeolite seems like an attractive candidate that could improve the
used in several personal care products, such as soap and toothpaste. Still, performance of TiO₂ nanotubes photocatalyst for photocatalytic degra­
it can end up in the food chain by discharging bathroom greywater dation due to its advantages (Prakash et al., 2019; Koohsaryan and
directly to the stream, neither regulated nor monitored. The European Anbia, 2016). The characteristic of zeolite has proven to have a unique
Chemical Agency and British Environment Agency proposed triclo­ relationship with titania structures, and its modified morphologies are
carban as an organic compound of potential concern because the TCC in promising for better photocatalytic activities. Zeolite has proven to have
surface water was found to be at a risk level that needs to be controlled the ability to accommodate a variety of ions which is cations including
(Vimalkumar et al., 2018). TCC potentially can impact aquatic life with K⁺, Na⁺, Mg2⁺, Ca2⁺ and more (Koohsaryan and Anbia, 2016;
the level of 3 μg L− 1 to 10 μg L− 1 of concentration (Barros et al., 2017; Pérez-Ramírez et al., 2008). Freshly, zeolite involving positive transition
Musee, 2018). TCC is known to cause methemoglobinemia in humans, metal ions have potential in many research areas, especially to enhance
cause cancer and baby blue syndrome, a condition of decreasing photocatalysis processes due to its ability to exchange within the process
oxygen-carrying capacity of haemoglobin in babies leading to death for others in reaction solution (Wang et al., 2015).
(Bomar et al., 2017). Among the group of TiO2-modified photocatalysts, ions like zeolite
Electrochemical anodisation (ECA) is an electrolytic process that shows a great potential with the light irradiation (Wang et al., 2015).
produces an oxidised layer over a metallic surface such as Ti foil (Sahrin Therefore, the study intends to propose modification on the photo­
et al., 2020; Lai and Sreekantan, 2011). Synthesis of TiO₂ nanotubes via catalyst, which requires metal cation (zeolite) as a precursor to be coated
electrochemical anodisation method using appropriate electrolyte and on TiO₂ nanotubes to enhance the photocatalyst activity—in this case,
processing parameters could produce uniformly array ordered and the photocatalytic degradation of TCC in water given that zeolite can
straight-up aligned TiO₂ nanotubes (Montakhab et al., 2020; Yoo et al., separate the recombination of electron-hole (Prakash et al., 2019;
2018). The significant processing parameters of ECA are anodisation Priyanka et al., 2020). The study aims to modify TiO₂ nanotubes pho­
voltage, processing time, type of electrolyte, countered electrodes tocatalyst by coating it with zeolite to efficiently degrade the TCC
(anode and cathode), calcination temperature and time, and ramping through photocatalytic degradation to enhance the knowledge on pho­
rate (Mazierski et al., 2016). By using Ti foil as substrate, TiO₂ nanotubes tocatalysis degradation of wastewater systems. This approach attracts
are formed on Ti foil as a result of a process that begins with the much attention from researchers and scientists globally because the
development of the first layer before the formation of a rather benefits of an environmental-friendly approach have always been at the
well-defined nanotubes structure that could be formed after a complete forefront of sustainable development.
process of electrochemical anodisation (Yoo et al., 2018; Mazierski The current work aimed to enhance the photocatalyst performance
et al., 2016; Macak et al., 2007) Generally, vertically uniform array for photocatalytic degradation of TCC by synthesising a modified TiO₂
nanotubes promise a large specific surface area, which has tube struc­ nanotubes-zeolite composite photocatalyst (MTNZC) and investigate its
tures with a circular nanotubular opening that could anchor characteristics, microstructure and applicability of degrading triclo­
light-harvesting assemblies and provide a more active site of photo­ carban in an aqueous solution. The degradation process was optimised
catalyst (Sahrin et al., 2020; Mazierski et al., 2016; Zakir et al., 2020; using response surface methodology (RSM), while the reusability of
Sun et al., 2010). The diameter of the nanotubular opening and the MTNZC was investigated for 5 cycles.
length of the tube are depending on the processing parameter. The base
part of the nanotubes is called the barrier layer, which is the nanotubes 2. Experimental procedures
attached to the Ti foil with the typical shape of hexagonal or pentagonal
(Mohan et al., 2020; Mor et al., 2011; Paulose et al., 2007). 2.1. Synthesis of MTNZC
TiO₂ is one of the general photocatalysts applied in photocatalytic
degradation that plays the main role in the process (Hunge et al., 2021). The synthesis of TiO₂ Nanotubes was conducted as described in the
However, the TiO₂ photocatalyst alone is inadequate to degrade recal­ previous study (Arifin et al., 2021). In brief, the titanium substrate was
citrant compounds, especially TCC, because TiO₂ photocatalyst has high anodised in a two electrodes electrochemical setup using a platinum rod
recombination of electron-hole that could decrease the photocatalytic (Nischk et al., 2016). The ethylene glycol was used as the electrolyte of
performance activities (Nguyen et al., 2020). Consequently, unmodified the electrochemical anodisation process with the mixture of fluoride
TiO₂ usually needs a solution to undertake practical applications of in­ ions by 98% v/v of ethylene glycol, 2% v/v of H₂O, and 0.09 M of NH₄F.
dustrial and environmental interest (Daghrir et al., 2013). This phe­ The Ti foil substrate underwent an electrochemical anodisation process
nomenon could lead to the loss of some operation effects. The mentioned for 1 h at 10, 40 and 60 V of the applied voltage using a DC power
issues could be prevented by selecting a significant cation-based mate­ supply. The TiO₂ nanotubes were rinsed using acetone, followed by
rial for the TiO₂ nanotubes modification, such as zeolite. Titanium di­ deionised water, dried at 80 ◦ C for 24 h and calcined at 450 ◦ C with 2
oxide coated with cation—for example, zeolite—could upgrade the ◦
C/min of heating rate for 6 h (Li et al., 2018) (Fig. 1a). The MTNZC was
exhibition of photocatalysis (Azizi-Lalabadi et al., 2019; Prakash et al., produced using pure TiO₂ nanotubes modified with transition metal
2019). Azizi-Lalabadi et al. (2019), Mao et al. (2017), and Zhang et al. nanoparticles (zeolite) by electrophoresis deposition (EPD) (Fig. 1b).
(2018) studied the modification of TiO₂ nanotubes to improve the A fixed volume (100 mL) solution was prepared by the mixing of 98%
photocatalytic activities. The TiO₂ nanotubes coated with zeolite v/v of ethanol, 0.2 g of zeolite powder, 2% v/v of deionised water and
hypothesised could widen the range of light absorption, improve the 0.04 g of magnesium nitrate (Mg(NO₃)₂) as dispersing agent. The pre­
redox potential of hydroxyl radical, and prevent the recombination of pared solution was ultrasonically cleaned and stirred by a magnetic
electron-hole on the conduction and valence band (Mao et al., 2017; stirrer until all the mixture were concentrated. The EPD experimental

2
S.N.H. Arifin et al. Chemosphere 287 (2022) 132278

setup was conducted at room temperature within two counter elec­ voltage was applied in the range of 5 kV at a working distance of 6 mm.
trodes, which is TiO₂ nanotubes as a cathode (− ) and Pt rod as an anode The surface elemental analysis focused on five different areas of sam­
(+). Both electrodes were dipped into the 100 mL prepared solution with pling aiming at composition and characteristic determination of the
a 1–1.5 cm gap between TiO₂ nanotubes and Pt rod. The samples were samples. The chemical composition of the analysed materials was ob­
electrophoresis deposited at 40 V of applied voltage for 60 s by using DC tained at a different energy level of excitation of the TiO₂ nanotubes
power. The obtained samples were dried before cleaning to strengthen using the EDS spectrum. X-ray diffractometry (XRD) was used to
the deposition of zeolite particles into the surface of the TiO₂ nanotubes. determine the chemical structure and the crystallinity of the MTNZC.
After the samples were completely dried, distilled water followed by The XRD patterns are recorded using a PanAnalytical X-ray diffrac­
acetone were used to rinse the samples and dried again at 80 ◦ C for 24 h tometer with copper Kα target (40 kV, 30 mA, λ = 1.5404 Å). The
using a drying oven. Finally, the samples were calcined at 450 ◦ C for 3 h scanning range was applied in a range of 2θ = 20–80◦ at a scan step of
within 5 ◦ C min− 1 of ramping time. 0.01◦ . Furthermore, the estimated crystal size was calculated according
to Sherrer’s formula (Ref).

2.2. Characterisation of MTNZC


2.3. Photocatalytic degradation setup for optimisation of TCC
The morphology structure of MTNZC was described using Field degradation
Emission-Scanning Electron Microscopy (FESEM), while the chemical
compositions were determined using energy-dispersive X-ray spectros­ The TCC degradation was optimised using a factorised central
copy (EDS) (Zhang et al., 2014; Yan and Zhou, 2011; Bessekhouad et al., composite design (FCCD) based on response surface methodology (RSM)
2006). The micrographs from the scanned image were obtained to via Design-Expert 11. Table 1 shows experimental variables and levels in
analyse the microstructure of the TiO₂ nanotubes arrays. FESEM images the factorised central composite design. Three independent factors were
were obtained via JEOL/JAPAN microscope combines with Oxford selected and included MTNZC size (0.5–1 cm2), pH (3–10) and irradia­
instrument/USA energy dispersive X-ray analyser. The accelerating tion time (10–60 min), according to Farouq et al. (2018). A total of 20

Fig. 1. a: Electrochemical anodisation process for TiO₂ Nanotubes. Fig. 1b: Electrophoresis deposition for MTNZC. Fig. 1c: Sample extraction procedures. Fig. 1d:
Schematic diagram of High-Performance Liquid Chromatography (HPLC).

3
S.N.H. Arifin et al. Chemosphere 287 (2022) 132278

Table 1 sunlight irradiation for 60 min. The samples were analysed every 10 min
Experiment variables and level in the factorised central composite design. of time interval to identify the performance ability of MTNZC. In addi­
Symbols Variables Coded Level tion, before MTNZC was reused for the repeated experiment, it was
washed with acetone and rinsed with ultra-pure water. After that, the
− 1 0 +1
MTNZC catalyst was dried in a drying oven at 80 ◦ C within 24 h. The
A pH value 3 6.5 10 reused MTNZC were characterised by using FESEM and EDS.
B Irradiation Time (minutes) 10 35 60
C TNTs/Zeo Catalyst Size (cm2) 0.5 0.75 1.0
3. Results and discussion

experimental runs were performed; for each experimental run, a fixed 3.1. Morphological structure of MTNZC
volume (100 mL) of TCC solution was prepared by mixing 5 mL of TCC
stock solution and 95 mL of distilled water. The photocatalytic experi­ The diameter length and wall thickness of nanotube dimensions as
ments were conducted in triplicates to obtain accurate results for the determined using FESEM images are illustrated in Fig. 2 and Table 1.
degradation performance. The MTNZC with different concentrations The samples were set under 10, 40, and 60 V electrochemical anodisa­
was mixed with the TCC solution; pH was adjusted to the desired value tion within 1 h with ethylene glycol as electrolyte solution and calcined
using 0.1 N of HCl and 0.1 N of NaOH. The mixture was placed under the
natural sunlight radiation between 11 a.m. and 4 p.m., the mixture was
aerated throughout the photocatalysis process. A fixed volume (10 mL)
of the treated samples was collected after 10, 20, 30, 40, 50 and 60 min
of the photocatalysis process to determine the TCC degradation.

2.4. Sample extraction

The treated samples were extracted using disperse liquid-liquid


microextraction (DLLME). In this experiment, 10 mL of the sample
was transferred into a 60 mL QuEChErs tube and mixed with 10 mL of
sodium chloride (NaCl). A mixture of acetonitrile (4 mL) as extraction
solvent and hexane (1 mL) as dispersive solvent were rapidly injected
into the sample. The mixture was concentrated by gradually shaking the
sample tube and left the mixture to form the sedimentation phase. The
supernatant was transferred by micropipette into a clean QuEChErs tube
and mixed with calcium chloride (CaCl) as a drying agent to absorb the
moisture in the sample. The precipitate was transferred into a 2 mL vial,
sealed with aluminium foil and parafilm and stored below − 4 ◦ C for
preservation before high-performance liquid chromatography (HPLC)
analysis. The procedures of the DLLME of the sample were illustrated in
Fig. 1c.

2.5. Analytical procedure

TCC concentration was measured using HPLC analysis. The chro­


matographic system used for HPLC was Agilent 1260 Infinity DAD HPLC
system, 1260 Quat Pump VL, 1260 ALS Auto-Injector, and ZORBAX
Eclipse Plus C18 column was used with the rapid resolution 4.6 × 100
mm, 35 μ. The mobile phase (flow rate 1 mL min⁻1) applied contained
80% acetonitrile and 20% deionised water by 25 min of stop time, 5 min
post-run, 25 ± 1 ◦ C column temperature, and 20 μL of sample injection
volume. HPLC analysis was conducted by 265 nm (10) of absorbance
wavelength, 360 nm (100) of absorbance reference wavelength and
190–400 nm spectrum (Fig. 1d). The degradation efficiency was calcu­
lated according to Eq. (1).
A0 − A
Degradation (%) = × 100 (1)
A0

Where;
A0 represents the TCC initial concentration (before the adsorption),
while A is the final concentration.

2.6. Reusability of MTNZC

The reusability assessment of the MTNZC was determined by


repeating TCC photocatalytic degradation experimental set up as
described in Section 2.4 several times. The photocatalytic degradation of
TCC was conducted by using the same MTNZC five times with the same Fig. 2. The influence of anodisation voltage on the morphology of the resulting
MTNZC in 5 ppm of TCC initial concentration and pH 11 under natural thin TiO₂ films: (a) 10V, (b) 40V, (c) 60V

4
S.N.H. Arifin et al. Chemosphere 287 (2022) 132278

at 450 ◦ C for 1, 3, and 6 h with 2 ◦ C min− 1 of ramping rate. The obtained


pure TiO₂ nanotubes by the 10V anodisation have a very small nano­
tubes surface with the tube’s diameter of approximately 50 nm (Fig. 2a).
The smaller TiO₂ nanotubes formed due to less electric field dissolution
occurs at low voltage as 10 V, and the TiO₂ layer becomes compact with
not the smooth nanotubular structure can be observed as shown in
Fig. 2a. On the contrary, a well-defined TiO₂ nanotubes structure could
be seen when the voltage is high (Ozkan et al., 2017). However, it could
be noticed that the obtained nanotubes from 10 V anodisation are not in
uniform array order to be compared with 40 V and 60 V of TiO₂ nano­
tubes. The TiO₂ nanotubes obtained by 40 V anodisation were in uniform
array ordered and smooth structure due to the effect of electrolyte type
presented in Fig. 2a. It was shown that the morphology structure of the
TiO₂ nanotubes layer is influenced by the electrolyte solution composi­
tion, such as fluoride ions concentration and water content (Montakhab
et al., 2020). As reported by Macak et al. (2007), the anodic growth of
the oxidised layers on metal surfaces and the formation of tubes are
created by the reaction between anodisation and chemical dissolution
toward the metallic surface. The balance between water content and
fluoride concentration in electrolyte solution influenced the formation
rate of the nanotubes (Qin et al., 2021).
The differences of the TiO₂ nanotubes morphology structure affected
by calcination time within 450 ◦ C of calcination temperature by 40 V of
anodisation is shown in Fig. 3. The nanotubular structure of 3 h calci­
nation time has a smooth arrangement compared to 6 h of calcination.
The nanotubes formed in 40 V anodisation are much longer than those
prepared in 10 V anodisation, as shown in Fig. 4, and all dimensions’
details have presented in Table 1. The differences in tube length of TiO₂
nanotubes by 10–60 V of anodisation reached about 0.56 μm–20.64 μm,
respectively. Hence, the length of TiO₂ nanotubes is proportional to the

Fig. 4. The influence of anodisation voltage on the morphology of the resulting


on TiO₂ nanotubes length: (a) 10V, (b) 40V, (c) 60V

anodisation voltage. Furthermore, studies showed that the range of


voltage forming TiO₂ nanotubes strongly relates to an electrolyte solu­
tion. The anodisation potential exhibits a significant influence on the
growth of TiO₂ nanotube arrays, which exhibited a slight influence in
non-aqueous electrolytes such as ethylene glycol electrolyte solution.
The voltage value significantly reduces non-aqueous electrolytes, which
is attributed to a large extent to the low conductivity of organic elec­
trolytes (Fu and Mo, 2018). However, Mor et al. (2006) mentioned that
with the longer length of the nanotubes, the base could no longer hold
the nanotubes root, and it ended up broken and formed another layer of
nanotubes. The higher anodisation voltage grows longer the length of
the nanotubes, and 60 V of anodisation gave substantial modification of
nanotubes base.
Fig. 3. The influence of calcination time on the surface morphology of TiO₂ The FESEM image of MTNZC is presented in Fig. 5 (a, b). The
nanotubes at 40 V anodisation: (a) 3 h calcination, (b) 6 h calcination. observation of FESEM images revealed that the TiO₂ nanotubes were

5
S.N.H. Arifin et al. Chemosphere 287 (2022) 132278

by the supporting material applied in the EDS analysis. The relative


amounts of Ti and O regarding P and C content, as estimated by EDS
analysis in Table 2. It was noted that only 40 V anodisation voltage of
TiO₂ shows the presence of Si, which might be related to the unwanted
element; however, it does not affect the characterisation of TiO₂. The
amount of Ti weight percentages was decreased by the increase of the
anodisation voltage. Table 2 shows the differences of Ti loading on the
sample, which decreases with the amount from 52.36 of weight per­
centage (10 V) to 49.19 of weight percentage (60 V). Its atomic per­
centage pattern also has a similar pattern to the weight percentage.
However, the weight of O remains almost similar with only a bit dif­
ference of amount that can be concluded that anodisation voltage is not
the only factor that can affect the loading of O in TiO₂ nanotubes.
The deposition of zeolite on the MTNZC was confirmed by EDS
analysis by the performed modifiers as 30s_40V, 60s_40V, 30s_60V,
60s_60V. Table 3 shows the elements of chemical composition. It was
noted that the presence of Si, Mg, Al, and Na might be due to the ele­
ments present in the zeolite. The amount of zeolite deposition must be
combined with the entire element together. However, the amount of the
deposited zeolite on samples at the 30 s was much lower than the 60 s
samples with 4.11% and 5.11% for the 30 s zeolite deposition times and
17.16% and 18.02% of the 60 s deposition time. The atomic ratios on
samples at the 30 s were much lower than the 60 s samples within 3.68%
and 4.4% for the 30 s zeolite deposition times and 12.84% and 13.34%
of the 60 s deposition time. The deposition of zeolite into TiO₂ nanotubes
was depended on the deposition time and voltage applied. Deposition
time significantly influenced the zeolite deposition where only 30 s gap
has given a big difference in weight percentage and atomic percentage of
the element. However, the applied voltage also strongly impacted this
deposition technique, proven by the amount differences between 40 V
and 60 V applied voltages. The 60 V applied has a bit higher weight and
atomic amount on the samples, as shown in Table 3.
Fig. 5. Homogeneous distribution of zeolite on the surface of TiO₂ nanotubes: The crystalline composition of MTNZC is depicted in Fig. 6. The re­
(a) NT_40V_30s, (b) NT_60V_30s. sults illustrate the outcomes of XRD measurement analysis for the ob­
tained MTNZC by applying ECA in ethylene glycol as electrolyte solution
uniformly array of ordered, elongated nanotubes formed, smooth at 10, 40 and 60 V of applied anodisation voltage calcined at 450 ◦ C
nanotube walls, and covered zeolite for the modification. The length, within 1, 3 and 6 h. The pattern of the TiO₂ nanotubes are also revealed
inner diameter, and wall thickness of the tubes were approximately 5.1 in Fig. 8 (a, b, c). The form of anatase and rutile of the TiO₂ nanotubes
μm, 120.3 nm, and 14.5 nm, respectively. The electrophoresis deposi­ has been confirmed by the XRD pattern analysis consisting of the re­
tion process for zeolite deposition has directed to the size-changing of flexes to both mentioned forms. In Fig. 8c, the rutile form could be seen
the tubes for fewer nanometres due to the presence of metal nano­ at the 2Ɵ = 25◦ where the presence of a bump in the pattern from
particles in zeolite by not greater than 1.2 nm in additional (Macak et al., samples anodised at 10 V and 6 h of calcination time. The event might be
2007; Ge et al., 2016). The electrophoresis voltage applied on the due to the changing formation at the TiO₂ state form anatase to rutile
charged particles (zeolite) is the product of electric field strength and the between 430 and 450 ◦ C of calcination temperature range (Yoo et al.,
surface charge of particles. Therefore, the deposition rate generally 2018).
increased with the increased applied deposition voltage and The increasing of anodisation voltage and calcination time (Fig. 7 a.
time-related (Zaidi et al., 2020). By the intense observation, the zeolite b.c) increases the intensity of the anatase reflexes. It is caused by the
particles possibly deposited pointedly on the wall of TiO₂ nanotubes at gaining thickness of the nanotubes’ walls. 10 V anodisation voltage of
60 V within 30 s of the electrophoresis process. samples shows the synthesis not reached to TiO₂ and the Ti shows all
The EDS analysis of MTNZC is presented in Table 2, which revealed over the pattern. After 6 h of the calcination time, the rutile reflex was
the presence of Ti and O elements. The presence of P and C were caused located at the beginning of 2Ɵ. All patterns show more than 45% crys­
tallinity, from 49.9% to 75.3%, which is predicted as an anatase state.
The crystallisation occurs during the calcination process and it is
Table 2
Dimension measurement of obtained TiO₂ nanotubes.
Sample Nanotubes’ Approximated Approximated wall Table 3
Label diameter, d (nm) nanotubes’ length, L, thickness, b (nm) Chemical composition elements detected on TiO₂ nanotubes by EDS analysis.
(μm)
Element 10 V 40 V 60 V
NT_10V_1hr 30.3 ± 0.4 0.76 ± 0.05 11.1 ± 0.11
Weight Atomic Weight Atomic Weight Atomic
NT_40V_1hr 93.9 ± 1.66 3.31 ± 0.01 12.8 ± 0.32
(%) (%) (%) (%) (%) (%)
NT_60V_1hr 116.5 ± 0.49 17.5 ± 0.34 13.8 ± 0.22
NT_10V_3hr 25.6 ± 0.52 0.64 ± 0.04 11.3 ± 0.2 Ti 52.36 26.93 50.07 25.22 49.19 24.55
NT_40V_3hr 120.3 ± 0.49 5.1 ± 0.12 14.5 ± 0.11 O 44.58 68.65 44.58 67.23 46.86 70.02
NT_60V_3hr 132.2 ± 1.92 19.3 ± 0.23 19.1 ± 0.33 C 1.57 3.23 2.20 4.42 1.96 3.90
NT_10V_6hr 26.1 ± 0.27 0.56 ± 638.7 10.5 ± 0.6 P 1.49 1.19 1.40 1.09 1.98 1.53
NT_40V_6hr 104.1 ± 0.58 4.5 ± 0.01 14 ± 0.3 Si ND ND 0.45 0.39 ND ND
NT_60V_6hr 121.2 ± 3.7 20.64 ± 0.09 18.9 ± 0.3 F ND ND 1.30 1.65 ND ND

6
S.N.H. Arifin et al. Chemosphere 287 (2022) 132278

Fig. 7. XRD pattern of TiO₂ Nanotubes modified with zeolite with different
deposition voltage and deposition time: (a) 40 V and 60 V at 30s, (b) 40 V and
60 V at 60s.

TiO₂ nanotubes which shows 40 V anodisation and 3 h of calcination


time is the most significant to be compared with other samples. Based on
the literature data, the structure crystalline is energetically relying on
the specific electrochemical anodisation parameters involving applied
voltage, anodisation time, or ramping rate of the calcination. The low
anodisation voltage applied in the range below 20 V has been reported
to produce amorphous TiO₂ nanotubes on Ti foil. In order to form a
crystalline structure, a higher anodisation voltage should be applied.
The crystal structure of the TiO₂ nanotubes was revealed to be anatase,
the mixture of anatase and rutile or rutile influenced by ECA parameters
condition (Sahrin et al., 2020; Zakir et al., 2020; Nischk et al., 2016).
The average crystallite size of MTNZC increased from 2.07 nm to
3.95 nm by the increasing of anodisation voltage (10 V, 40 V and 60 V)
followed by 450 ◦ C of calcination for 1, 3, and 6 h, respectively. It was
Fig. 6. XRD patterns for the samples obtained with the use of ethylene glycol- proven that the samples calcined for 3 h give significantly larger crys­
based electrolyte and anodisation voltage in different conditions: (a) 1 h, (b) 3
tallite sizes compared with 6 h of calcination. The 3 h of calcination time
h, (c) 6 h.
showed the largest crystallite size with 40 V of anodisation from 2.57 nm
(1 h) up to 3.53 nm (3 h) and went down to 2.07 ± 5 nm (6 h). According
dependent on the calcination temperature and time (Sahrin et al., 2020). to Lai and Sreekantan (Yoo et al., 2018), the crystallinity of the samples
It can be seen that calcination influenced the intensity, the intensity prepared by ECA depends on the explicit anodisation factors, such as the
ratio and the width of the (1 0 1) and (0 0 4) reflexes of anatase (Nischk anodisation voltage, anodisation time, calcination temperature, calci­
et al., 2016). It verified that the nanotubes were significantly in uniform nation time, and ramping time. Due to the factors, it has been reported
array ordered and the growth of nanotubes direction is at (1 0 1) di­ that the structure of the TiO₂ can be amorphous at a low voltage as lower
rection. The best crystallinity percentage up to 68.6% has been revealed than 10 V. Meanwhile, the higher anodisation voltage could affect the
by XRD pattern analysis from 40 V of anodisation voltage and 450 ◦ C of crystallisation of the TiO₂ nanotubes. Regarding to the electrochemical
calcination temperature within 3 h of TiO₂ nanotubes. The effect of the anodisation condition, the crystalline structure could be predicted as
calcination time on the structure of the nanotubes has been demon­ anatase with the mixture of rutile (Lai et al., 2014).
strated in Fig. 7 (a, b, c). It can be observed in Table 4 that calcination The XRD patterns of modified TiO₂ nanotubes with zeolite (MTNZC)
time and anodisation voltage influence the intensity and crystallinity of

7
S.N.H. Arifin et al. Chemosphere 287 (2022) 132278

3.2. Optimisation of TCC degradation using MTNZC

The screening for TCC degradation using MTNZC is illustrated in


Table 6. The maximum TCC removal (92%) was recorded at pH 6.5, after
35 min and with 0.75 cm2 of MTNZC, while the minimum degradation
efficiency (62.79%) was obtained at pH 3, after 10 min and with 0.5 cm2
of MTNZC. The ANOVA analysis indicated that the photocatalytic
degradation process of TCC was most suitably described with a
quadratic model (p < 0.05). The Lack of Fit F-value of all responses
implied the non-significant Lack of Fit relative to the pure error. The
model showed a non-significant lack of fit that indicated the significance
of the model and the model is applicable. The F-value (36.12) of the
model for all responses presented the model was significant. The p-value
of all responses was less than 0.0001, and the model was statistically
significant at the 95% confidence level. A quite high coefficient of
determination, with R2 in the range of 0.9433–0.9906, showed the
suitability of the model to represent the actual relationship among the
parameters (Chaibakhsh et al., 2016) (see Table 7).
The actual and coded equations for all responses’ models are pre­
sented in Eqs. (2) and (3).

yactual = 90.379 − 0.804x1 + 3.814x2 + 4.428x3 − 1.53625x1 x2 − 0.98625x1 x3


+ 0.19125x1 x3 − 10.9264x12 − 3.25636x22 − 2.30636x23
(2)

ypredicted = − 0.262283 + 12.8255x1 + 0.60844x2 + 79.32015x3


− 0.01756x1 x2 − 1.12714x1 x2 + 0.0306x1 x3 − 0.891948x12 − 0.00521x22
− 36.90182x32
(3)
( ( ))
x1 (pH), x2 (time, min), x3 MTNZC size cm2 , y (TCC regredation %)

The three-dimensional surface plots of interaction between pH value


versus irradiation time are graphical representations of the regression
equation for the optimisation of reaction conditions and are the most
useful approaches in revealing the conditions of the reaction system.
Fig. 8 shows the eight plots of 3-D response surface that illustrate the
interaction between pH value and irradiation time affecting TCC
degradation rate. The relationship of pH and time is depicted in Fig. 8a.
It was noted that the interaction effects of pH value and irradiation time
on the reduction of TCC depicts a canopy-like surface. With the
increasing pH value and irradiation time to the optimum points, the
reduction rate of TCC approaches the maximum level. In the literature,
the authors claimed that irradiation time has the greatest impact on
photocatalytic activity, followed by oxidant concentration and photo­
catalyst amount (Dostanić et al., 2013).
The interaction between pH and MTNZC loading on the removal rate
of TSS show a half-cylindrical response shape (Fig. 8b). The pH and
MTNZC loads exhibited a synergistic interaction that significantly
improved the TSS degradation rate (p < 0.05). Chaibakhsh et al. (2016)
indicated that the molecule is broken into smaller pieces and de­
composes completely at higher pH due to the hydroxyl radicals. The pH
value of the solution could influence photocatalytic degradation due to
Fig. 8. Response surface methodology of TCC degradation using MTNZC. the effect of hydroxyl radical formation (Zulfiqar et al., 2019). The
significant TCC degradation due to the higher pH can be identified by
are shown in Fig. 9. Ti foil contained primarily pure titanium with a the characteristic of the photocatalyst surface charge. By increasing the
small amount of crystallinity. According to Grimes (Mazierski et al., pH values, the number of negative charge sites gradually increase
2016), the increasing nanotubes length and wall thickness could in­ (Farrokhi-Rad et al., 2018). The positive effect of MTNZC catalyst size in
crease the growth rate of the crystallites of TiO₂ nanotubes anatase alkaline solution may reduce the recombination of electrons hole-pair
formed. The crystallites size of MTNZC were determined by using on catalyst surface by taking the electrons from the conduction band.
Scherrer’s formula (Macak et al., 2007; Robin et al., 2014). Based on the Hence, both pH values of the alkaline region resulted in an increase in
data presented in Table 5, 38.1–40.1 nm of anatase crystallites size was TCC photocatalytic degradation rate.
in the range for all obtained MTNZC, indicating that the chosen oper­ The 3-D surface plot of photo-degradation of TCC as an interaction of
ating factors for electrophoresis deposition of zeolite has significantly catalyst loading size MTNZC and irradiation time is shown in Fig. 8c.
affected the crystallinity properties of the samples (Awad et al., 2018). The elliptical shape of the contour plot indicated that the interaction of

8
S.N.H. Arifin et al. Chemosphere 287 (2022) 132278

Table 4
Chemical composition elements detected on TiO₂ Nanotubes-Zeolite Composite.
Element 30s_40V 60s_40V 30s_60V 60s_60V

Weight (%) Atomic (%) Weight (%) Atomic (%) Weight (%) Atomic (%) Weight (%) Atomic (%)

Ti 47.82 23.91 24.51 10.55 44.24 21.49 22.33 9.56


O 45.97 68.82 55.08 71.02 48.25 70.19 57.34 73.50
C 1.28 2.55 3.25 5.58 1.39 2.69 1.75 2.99
F 0.81 1.02 ND ND 1.01 1.23 0.56 0.61
Na ND ND 0.01 0.01 ND ND ND ND
Mg 1.31 1.29 2.05 1.74 1.28 1.23 1.80 1.51
Si 2.80 2.39 15.07 11.07 3.83 3.17 16.00 11.69
Al ND ND 0.04 0.03 ND ND ND ND
P ND ND − 0.01 − 0.01 ND ND 0.22 0.14

Table 5
Crystallinity properties of obtained TiO₂ nanotubes.
Sample Label Crystallinity (%) Average Crystallite Size (nm)

NT_10V_1hr 49.90 3.95


NT_40V_1hr 54.30 2.57
NT_60V_1hr 75.30 3.53
NT_10V_3hr 63.10 2.83
NT_40V_3hr 68.60 3.53
NT_60V_3hr 63.40 2.64
NT_10V_6hr 57.60 3.86
NT_40V_6hr 68.50 2.07
NT_60V_6hr 69.10 2.36

Table 6
Crystallinity properties of obtained TiO₂ Nanotubes-Zeolite Composite.
Sample Deposition Deposition Crystallite Crystallinity
Label Voltage (V) Time (s) size (nm) (%)

NT_40V_30s 40 30 38.1 71
NT_40V_60s 40 60 39.5 56.3
NT_60V_30s 60 30 40.1 60
NT_60V_60s 60 60 36.5 69.8

Table 7
The RSM FCCD responses to the TCC photocatalytic degradation.
Run x1 x2 x3 y

1 6.5 35 0.5 85.63


2 6.5 35 1 91.06
3 10 10 0.5 68.34
4 6.5 35 0.75 91.19
5 6.5 35 0.75 90.6
6 10 35 0.75 78.43
7 6.5 10 0.75 82.69
8 3 35 0.75 81.02
9 3 60 0.5 74.53
10 10 10 1 73.83
11 6.5 35 0.75 89.04
12 3 10 0.5 62.79
Fig. 9. TCC photocatalytic degradation versus irradiation time by the different 13 3 60 1 84.73
size of MTNZC loading (A); MTNZC applicability performance in TCC photo­ 14 10 60 1 80.19
15 6.5 35 0.75 91.85
catalytic degradation and its reusability (B).
16 6.5 60 0.75 92.1
17 3 10 1 75.96
MTNZC size and irradiation time was effective towards photo- 18 10 60 0.5 70.2
degradation of TCC, a significant interaction (p < 0.05) was recorded 19 6.5 35 0.75 86.06
20 6.5 35 0.75 92.44
between MTNZC size and time, which enhanced the photocatalytic
degradation of TCC. RSM has been used to optimise the photocatalytic x1 (pH), x2 (time, min), x3 (MTNZC size (cm2)), y (TCC regredation %).
performance of TiO₂ in the degradation of pollutants under simulated
sunlight (Dostanić et al., 2013). Based on the RSM analysis, the size of catalyst loading size could enhance the TCC degradation rate because
the MTNZC catalyst has great interaction with the TCC molecule and more active sites will be available to treat the TCC molecules. The au­
hydroxyl radical, where the adequate size of the photocatalyst active thors claimed that irradiation time has the greatest impact on photo­
sites enhance the TCC photocatalytic degradation rate along with catalytic activity, followed by photocatalyst size loading. This
reduction of electron-hole recombination. Hence, increasing the MTNZC phenomenon could be explained as the longer irradiation time and the

9
S.N.H. Arifin et al. Chemosphere 287 (2022) 132278

greater the MTNZC size devoted to the reaction, the more exposure the RSM analysis regarding optimisation of TCC photocatalytic degradation,
MTNZC surface to visible light. the ANOVA analysis with a quadratic model (p < 0.05) and the coeffi­
The best operating parameters for TCC degradation was recorded at cient rate (R2) of the model operation is in the range of 0.9433–0.9906,
pH 10.11, after 50 min using 1 cm2 of the MTNZC size. An additional indicating that the optimisation of the process was significant. The best
experiment was conducted within the suggested optimum conditions to operating parameters for TCC degradation was recorded at pH 10.11,
compare the predicted and experimental values. The TCC degradation after 50 min and with 1 cm2 of the MTNZC size, where 81.2 vs. 79.29%
was 81.2 vs 79.29% of the actual and predicted results, respectively. The of the actual and predicted degradation were obtained. The photo­
error values of the actual and predicted value were below 5% indicates catalytic degradation rate of TCC from the first and the fifth cycles was
that the experimental values were closely agreed to the predicted values. 94.2% and 77.4%, indicating the applicability of MTNZC for several
cycles.
3.3. Performance of MTNZC in TCC photocatalytic degradation
Credit author statement
The effectiveness of a catalyst defines the stability of the catalyst and
effectiveness of its photocatalytic performance. The trend of TCC pho­ Siti Nor Hidayah Arifin: Methodology Investigation, Writing draft
tocatalytic degradation as a response for the MTNZC size and time is manuscript, Radin Maya Saphira Radin Mohamed; Conceptualization,
presented in Fig. 9a. The results revealed that 0.75 cm2 and 1 cm2 of Supervision, Writing - review & editing, Adel Ali Al-Gheethi; Concep­
MTNZC size exhibited a similar effect toward the TCC photocatalytic tualization, Supervision, Writing - review & editing, Lai Chin Wei;
degradation. The TCC degradation rate has slightly different, which is Conceptualization, Supervision, Writing - review & editing, G. Yashni:
92.0% for 1 cm2 and 94.2% for 0.75 cm2 of MTNZC. However, the software and data analysis, Nurina Fitriani: writing—review and edit­
remained concentration of TCC after photocatalytic degradation treat­ ing. Mu. Naushad: writing—review and editing. Ahmad B. Albadarin;
ment by the 0.75 cm2 was 0.29 ppm compared to 1 cm2, which is the writing—review and editing.
remaining TCC concentration was 0.36 ppm. It proves that the larger
size of MTNZC does not affect the TCC photocatalytic degradation Declaration of competing interest
process. The obtained results indicate that the performance of photo­
catalytic degradation of TCC improved in the alkaline condition of so­ The authors declare that they have no known competing financial
lution regarding to the highest degradation rate of TCC by 95.2% with interests or personal relationships that could have appeared to influence
1.0 cm2 in size of MTNZC at pH 11 in 5 ppm of TCC initial concentration the work reported in this paper.
within 60 min of irradiation with the remain TCC concentration 0.24
ppm respectively. In alkaline solutions, the reactivity of the amino Acknowledgement
groups is induced due to the protonation that occurred because the pH
level was above pKa of the functional group (Ding et al., 2015). It also This research was supported by the Ministry of Higher Education,
can be hypothesised the improvement of photocatalytic degradation of Malaysia (MOHE) through the Fundamental Research Grant Scheme:
TCC in alkaline conditions was due to the structural orientation of FRGS/1/2018/WAB05/UTHM/02/2 (An Insight of Xenobiotic Organic
substrate is significant for the attack of the reactive species under this Compounds (XOCs) Degradation in Greywater via Photocatalytic
condition alongside protonation (Ji et al., 2013). Enhancement for Water Resources Protection) vot K090 and Universiti
In order to verify the reusability of MTNZC on TCC photocatalytic, Tun Hussein Onn Malaysia for GPPS grant, vot U777 for the financial
the same experiments were repeated five times with the same MTNZC in support to this project. Mu. Naushad is grateful to the Researchers
5 ppm of TCC initial concentration at pH 11 under natural sunlight Supporting Project number (RSP-2021/8), King Saud University,
irradiation for 60 min. Before MTNZC was reused for the repeated Riyadh, Saudi Arabia, for the support.
experiment, it was washed with acetone and rinsed using ultra-pure
water. After that, the MTNZC was dried in a drying oven at 80 ◦ C Appendix A. Supplementary data
within 24 h. The maintained high reactivity of MTNZC size after the fifth
cycle is presented in Fig. 9b. The photocatalytic degradation rate of TCC Supplementary data to this article can be found online at https://doi.
from the fifth cycle was 77.4%, where it still can be considered a sig­ org/10.1016/j.chemosphere.2021.132278.
nificant degradation rate (Wang et al., 2016). The result demonstrated
the MTNZC has a stable and effective performance for TCC photo­ References
catalytic degradation and exhibited good recycle ability.
Arifin, S. nor H., Mohamed, R., Al-Gheethi, A., Lai, C.W., Yashni, G., 2021.
Heterogeneous photocatalysis of triclocarban and triclosan in greywater: a
4. Conclusion systematic and bibliometric review analysis. Int. J. Environ. Anal. Chem. (December)
https://doi.org/10.1080/03067319.2020.1863391.
The preparation of the MTNZC depended on anodisation voltage and Awad, N.K., Edwards, S.L., Morsi, Y.S., 2018. A review of TiO₂ NTs on Ti metal:
electrochemical synthesis, functionalization and potential use as bone implants.
calcination time. The TiO₂ nanotubes prepared by 40 V of anodisation Mater. Sci. Eng. C 76, 1401–1412. https://doi.org/10.1016/j.msec.2017.02.150.
and 3 h of calcination process were the most significant catalyst due to Azizi-Lalabadi, M., Ehsani, A., Divband, B., Alizadeh-Sani, M., 2019. Antimicrobial
its smooth and uniformly ordered nanotubes with average tube’s length activity of Titanium dioxide and Zinc oxide nanoparticles supported in 4A zeolite
and evaluation the morphological characteristic. Sci. Rep. 9 (1), 1–10. https://doi.
was approximately 5.1 μm. The inner diameter and wall thickness were org/10.1038/s41598-019-54025-0.
120.3 and 14.5 nm, respectively and the zeolite particles possibly Barros, S., Montes, R., Quintana, J.B., Rodil, R., Oliveira, J.M.A., Santos, M.M.,
deposited pointedly on the wall of TiO₂ nanotubes. The presence of el­ Neuparth, T., 2017. Chronic effects of triclocarban in the amphipod Gammarus
locusta: behavioural and biochemical impairment. Ecotoxicol. Environ. Saf. 135
ements Al, Mg, Si and Na has been confirmed by EDS analysis, indicating (September 2016), 276–283. https://doi.org/10.1016/j.ecoenv.2016.10.013.
that zeolite’s deposition into TiO₂ nanotubes has significantly occurred. Bessekhouad, Y., Chaoui, N., Trzpit, M., Ghazzal, N., Robert, D., Weber, J.V., 2006. UV-
TiO₂ nanotubes obtained from 3 h of calcination time shows the highest vis versus visible degradation of Acid Orange II in a coupled CdS/TiO2
semiconductors suspension. J. Photochem. Photobiol. Chem. 183 (1–2), 218–224.
crystallite size with 40 V of anodisation from 2.57 nm (1 h) up to 3.53
https://doi.org/10.1016/j.jphotochem.2006.03.025.
nm (3 h) and went down to 2.07 ± 5 nm (6 h). 38.1–40.1 nm of anatase Bomar, E., Owens, G., Murray, G., 2017. Nitrate ion selective electrode based on ion
crystallites size was recorded in obtained by 40 and 60 V of electro­ imprinted poly(N-methylpyrrole). Chemosensors 5 (1), 2.
phoresis voltage of MTNZC within 30 s, indicating that the chosen Chaibakhsh, N., Ahmadi, N., Zanjanchi, M.A., 2016. Optimization of photocatalytic
degradation of neutral red dye using TiO2 nanocatalyst via Box-Behnken design.
operating factors for electrophoresis deposition of zeolite has signifi­ Desalination and Water Treatment 57 (20), 9296–9306. https://doi.org/10.1080/
cantly affected the crystallinity properties of the samples. Based on the 19443994.2015.1030705.

10
S.N.H. Arifin et al. Chemosphere 287 (2022) 132278

Daghrir, R., Drogui, P., Robert, D., 2013. Modified TiO2 for environmental photocatalytic Nischk, M., Mazierski, P., Wei, Z., Siuzdak, K., 2016. Applied Surface Science Enhanced
applications: a review. Ind.Chem.Res 52, 3581–3599. https://doi.org/10.1021/ photocatalytic, electrochemical and photoelectrochemical properties of TiO₂
ie303468t. nanotubes arrays modified with Cu, AgCu and Bi nanoparticles obtained via
Ding, S.L., Wang, X.K., Jiang, W.Q., Zhao, R.S., Shen, T.T., Wang, C., Wang, X., 2015. radiolytic reduction. Appl. Sur. Sci. 387, 89–102. https://doi.org/10.1016/j.
Influence of pH, inorganic anions, and dissolved organic matter on the photolysis of apsusc.2016.06.066.
antimicrobial triclocarban in aqueous systems under simulated sunlight irradiation. Ozkan, S., Nguyen, N.T., Mazare, A., Hahn, R., Cerri, I., Schmuki, P., 2017. Fast growth
Environ. Sci. Pollut. Control Ser. 22 (7), 5204–5211. https://doi.org/10.1007/ of TiO2 nanotube arrays with controlled tube spacing based on a self-ordering
s11356-014-3686-x. process at two different scales. Electrochem. Commun. 77, 98–102. https://doi.org/
Dostanić, J., Lončarević, D., Rožić, L., Petrović, S., Mijin, D., Jovanović, D.M., 2013. 10.1016/j.elecom.2017.03.007.
Photocatalytic degradation of azo pyridone dye: optimization using response surface Paulose, M., Prakasam, H.E., Varghese, O.K., Peng, L., Popat, K.C., Mor, G.K., Grimes, C.
methodology. Desalination and Water Treatment 51 (13–15), 2802–2812. https:// A., 2007. TiO₂ nanotube arrays of 1000 μm length by anodization of titanium foil:
doi.org/10.1080/19443994.2012.750699. phenol red diffusion. J. Phys. Chem. C 111 (41), 14992–14997. https://doi.org/
Farouq, R., Abd-Elfatah, M., Ossman, M.E., 2018. Response surface methodology for 10.1021/jp075258r.
optimization of photocatalytic degradation of aqueous ammonia. J. Water Supply Pérez-Ramírez, J., Christensen, C.H., Egeblad, K., Christensen, C.H., Groen, J.C., 2008.
Res. Technol. - Aqua 67 (2), 162–175. https://doi.org/10.2166/aqua.2018.121. Hierarchical zeolites: enhanced utilisation of microporous crystals in catalysis by
Farrokhi-Rad, M., Mohammadalipour, M., Shahrabi, T., 2018. Electrophoretic deposition advances in materials design. Chem. Soc. Rev. 37 (11), 2530–2542. https://doi.org/
of titania nanostructured coatings for photodegradation of methylene blue. Ceram. 10.1039/b809030k.
Int. 44 (9), 10716–10725. https://doi.org/10.1016/j.ceramint.2018.03.106. Prakash, K., Karuthapandian, S., Senthilkumar, S., 2019. Zeolite nanorods decorated g-
Fu, Y., Mo, A., 2018. A review on the electrochemically self-organized titania nanotube C3N4 nanosheets: a novel platform for the photodegradation of hazardous water
Arrays: synthesis, modifications, and biomedical applications. Nanoscale Research contaminants. Mater. Chem. Phys. 221, 34–46. https://doi.org/10.1016/j.
Letters 13. https://doi.org/10.1186/s11671-018-2597-z. matchemphys.2018.09.026.
Ge, M.Z., Cao, C.Y., Huang, J.Y., Li, S.H., Zhang, S.N., Deng, S., Li, Q.S., Zhang, K.Q., Priyanka, R.N., Joseph, S., Abraham, T., Plathanam, N.J., Mathew, B., 2020. Novel La
Lai, Y.K., 2016. Synthesis, modification, and photo/photoelectrocatalytic (OH)3-integrated sGO-Ag3VO4/Ag nanocomposite as a heterogeneous photocatalyst
degradation applications of TiO₂ nanotube arrays: a review. Nanotechnol. Rev. 5 (1), for fast degradation of agricultural and industrial pollutants. Catalysis Science and
75–112. https://doi.org/10.1515/ntrev-2015-0049. Technology 10 (9), 2916–2930. https://doi.org/10.1039/d0cy00104j.
Hunge, Y.M., Yadav, A.A., Khan, S., Takagi, K., Suzuki, N., Teshima, K., Fujishima, A., Qin, J., Cao, Z., Li, H., Su, Z., 2021. Formation of anodic TiO2 nanotube arrays with
2021. Photocatalytic degradation of bisphenol A using titanium dioxide@ ultra-small pore size. Surf. Coating. Technol. 405 (November 2020), 126661.
nanodiamond composites under UV light illumination. J. Colloid Interface Sci. 582, https://doi.org/10.1016/j.surfcoat.2020.126661.
1058–1066. https://doi.org/10.1016/j.jcis.2020.08.102. Robin, A., Ribeiro, M. B. a, Rosa, J.L., Nakazato, R.Z., Paulo, U.D.S., Lorena, E. D. E. De,
Ji, Y., Zhou, L., Ferronato, C., Yang, X., Salvador, A., Zeng, C., Chovelon, J.M., 2013. Brazil, L., 2014. Formation of TiO₂ nanotubes produced by anodization in ethylene
Photocatalytic degradation of atenolol in aqueous titanium dioxide suspensions: glycol-H₂O electrolyte Experimental procedure. J. Surf. Eng. Mater. Adv. Technol. 4,
kinetics, intermediates and degradation pathways. J. Photochem. Photobiol. Chem. 123–130. https://doi.org/10.4236/jsemat.2014.43016.
254, 35–44. https://doi.org/10.1016/j.jphotochem.2013.01.003. Rochester, J.R., Bolden, A.L., Pelch, K.E., Kwiatkowski, C.F., 2017. Potential
Koohsaryan, E., Anbia, M., 2016. Nanosized and hierarchical zeolites: a short review. developmental and reproductive impacts of triclocarban: a scoping review.
Cuihua Xuebao/Chinese Journal of Catalysis 37 (4), 447–467. https://doi.org/ J. Toxicol. 2017, 1–15. https://doi.org/10.1155/2017/9679738.
10.1016/S1872-2067(15)61038-5. Sahrin, N.T., Nawaz, R., Kait, C.F., Lee, S.L., Wirzal, M.D.H., 2020. Visible light
Lai, C.W., Sreekantan, S., 2011. Effect of applied potential on the formation of self- photodegradation of formaldehyde over TiO2 nanotubes synthesized via
organized TiO₂ nanotube arrays and its photoelectrochemical response. electrochemical anodization of titanium foil. Nanomaterials 10 (1). https://doi.org/
J. Nanomater. 2011 https://doi.org/10.1155/2011/142463. 10.3390/nano10010128.
Lai, C.W., Juan, J.C., Ko, W.B., Bee Abd Hamid, S., 2014. An overview: recent Sun, L., Zhang, S., Sun, X., He, X., 2010. Effect of the geometry of the anodized titania
development of titanium oxide nanotubes as photocatalyst for dye degradation. Int. nanotube Array on the performance of dye-sensitized solar cells. J. Nanosci.
J. Photoenergy 1–14. https://doi.org/10.1155/2014/524135. Nanotechnol. 10 (7), 4551–4561. https://doi.org/10.1166/jnn.2010.1695.
Li, C., Jin, F., Snyder, S.A., 2018. Recent advancements and future trends in analysis of Vimalkumar, K., Arun, E., Krishna-kumar, S., Poopal, R.K., Nikhil, N.P., Subramanian, A.,
nonylphenol ethoxylates and their degradation product nonylphenol in food and Babu-rajendran, R., 2018. Science of the Total Environment Occurrence of
environment. Trac. Trends Anal. Chem. 107, 78–90. https://doi.org/10.1016/j. triclocarban and benzotriazole ultraviolet stabilizers in water, sediment, and fish
trac.2018.07.021. from Indian rivers. Sci. Total Environ. 625, 1351–1360. https://doi.org/10.1016/j.
Macak, J.M., Tsuchiya, H., Ghicov, A., Yasuda, K., Hahn, R., Bauer, S., Schmuki, P., 2007. scitotenv.2018.01.042.
TiO₂ nanotubes: self-organized electrochemical formation, properties and Wang, C., Li, Y., Shi, H., Huang, J., 2015. Preparation and characterization of natural
applications. Curr. Opin. Solid State Mater. Sci. 11, 3–18. https://doi.org/10.1016/j. zeolite supported nano TiO2 photocatalysts by a modified electrostatic self-assembly
cossms.2007.08.004. method. Surf. Interface Anal. 47 (1), 142–147. https://doi.org/10.1002/sia.5686.
Mao, C., Wang, Y., Jiao, W., Chen, X., Lin, Q., Deng, M., Feng, P., 2017. Integrating Wang, J., Sun, Y., Feng, J., Xin, L., Ma, J., 2016. Degradation of triclocarban in water by
zeolite-type chalcogenide with titanium dioxide nanowires for enhanced dielectric barrier discharge plasma combined with TiO2/activated carbon fibers:
photoelectrochemical activity. Langmuir 33 (47), 13634–13639. https://doi.org/ effect of operating parameters and byproducts identification. Chem. Eng. J. 300,
10.1021/acs.langmuir.7b02403. 36–46. https://doi.org/10.1016/j.cej.2016.04.041.
Mazierski, P., Nischk, M., Gołkowska, M., Lisowski, W., Gazda, M., Winiarski, M.J., Yan, J., Zhou, F., 2011. TiO₂ nanotubes: structure optimization for solar cells. J. Mater.
Zaleska-Medynska, A., 2016. Photocatalytic activity of nitrogen doped TiO 2 Chem. 21 (26), 9406–9418. https://doi.org/10.1039/c1jm10274e.
nanotubes prepared by anodic oxidation: the effect of applied voltage, anodization Yoo, H., Kim, M., Kim, Y.-T., Lee, K., Choi, J., 2018. Catalyst-Doped anodic TiO2
time and amount of nitrogen dopant. Appl. Catal. B Environ. 196, 77–88. https:// nanotubes: binder-free electrodes for (Photo)Electrochemical reactions. Catalysts 8
doi.org/10.1016/j.apcatb.2016.05.006. (11), 555. https://doi.org/10.3390/catal8110555.
Mohan, L., Dennis, C., Padmapriya, N., Anandan, C., Rajendran, N., 2020. Effect of Zaidi, S.Z.J., Harito, C., Bavykin, D.V., Martins, A.S., Yuliarto, B., Walsh, F.C., Ponce de
electrolyte temperature and anodization time on formation of TiO2 nanotubes for León, C., 2020. Photocatalytic degradation of methylene blue dye on reticulated
biomedical applications. Materials Today Communications 23 (March), 101103. vitreous carbon decorated with electrophoretically deposited TiO2 nanotubes. Diam.
https://doi.org/10.1016/j.mtcomm.2020.101103. Relat. Mater. 109 (April), 108001. https://doi.org/10.1016/j.
Montakhab, E., Rashchi, F., Sheibani, S., 2020. Modification and photocatalytic activity diamond.2020.108001.
of open channel TiO₂ nanotubes array synthesized by anodization process. Appl. Zakir, O., Idouhli, R., Elyaagoubi, M., Khadiri, M., Aityoub, A., Koumya, Y.,
Surf. Sci. 534 (August), 147581. https://doi.org/10.1016/j.apsusc.2020.147581. Outzourhit, A., 2020. Fabrication of TiO2Nanotube by electrochemical anodization:
Mor, G.K., Varghese, O.K., Paulose, M., Shankar, K., Grimes, C.A., 2006. A review on toward photocatalytic application. J. Nanomater. 2020 https://doi.org/10.1155/
highly ordered, vertically oriented TiO₂ nanotube arrays: fabrication, material 2020/4745726.
properties, and solar energy applications. Solar Energy Mat. Sol. Cell. 90 (14), Zhang, Q., Zhu, J., Wang, Y., Feng, J., Yan, W., Xu, H., 2014. Electrochemical assisted
2011–2075. https://doi.org/10.1016/j.solmat.2006.04.007. photocatalytic degradation of salicylic acid with highly ordered TiO₂ nanotube
Mor, G.K., Varghese, O.K., Paulose, M., Shankar, K., Ã, C.A.G., 2011. A Review on Highly electrodes. Appl. Sur. Sci. 308, 161–169. https://doi.org/10.1016/j.
Ordered, Vertically Oriented TiO₂ Nanotube Arrays: Fabrication, Material Properties, apsusc.2014.04.125.
and Solar Energy Applications, vol. 90, pp. 2011–2075. https://doi.org/10.1016/j. Zhang, G., Song, A., Duan, Y., Zheng, S., 2018. Enhanced photocatalytic activity of TiO2/
solmat.2006.04.007, 2006. zeolite composite for abatement of pollutants. Microporous Mesoporous Mater. 255,
Musee, N., 2018. Environmental risk assessment of triclosan and triclocarban from 61–68. https://doi.org/10.1016/j.micromeso.2017.07.028.
personal care products in South Africa. Environ. Pollut. 242, 827–838. https://doi. Zulfiqar, M., Samsudin, M.F.R., Sufian, S., 2019. Modelling and optimisation of
org/10.1016/j.envpol.2018.06.106. photocatalytic degradation of phenol via TiO2 nanoparticles: an insight into
Nguyen, H.H., Gyawali, G., Martinez-Oviedo, A., Kshetri, Y.K., Lee, S.W., 2020. response surface methodology and artificial neural network. J. Photochem.
Physicochemical properties of Cr-doped TiO2 nanotubes and their application in Photobiol. Chem. 384 (May), 112039. https://doi.org/10.1016/j.
dye-sensitized solar cells. J. Photochem. Photobiol. Chem. 397 (March), 112514. jphotochem.2019.112039.
https://doi.org/10.1016/j.jphotochem.2020.112514.

11

You might also like