Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Environmental Research 203 (2022) 111841

Contents lists available at ScienceDirect

Environmental Research
journal homepage: www.elsevier.com/locate/envres

Mechanochemical synthesis of ternary heterojunctions TiO2(A)/TiO2(R)/


ZnO and TiO2(A)/TiO2(R)/SnO2 for effective charge separation in
semiconductor photocatalysis: A comparative study
M.L. Aruna Kumari a, b, *, L. Gomathi Devi a, Gilberto Maia c, Tse-Wei Chen d, Nabil Al-Zaqri e,
M. Ajmal Ali f
a
Department of Post Graduate Studies in Chemistry, Bangalore University, Bengaluru, 560001, India
b
Department of Chemistry, M. S. Ramaiah College of Arts, Science, and Commerce, Bengaluru, 560054, India
c
Institute of Chemistry, Universidade Federal de Mato Grosso do Sul, Av. Senador Filinto Muller, 1555, Campo Grande, MS, 79074-460, Brazil
d
Department of Materials, Imperial College London, London, SW7 2AZ, United Kingdom
e
Department of Chemistry, College of Science, King Saud University, P.O. Box 2455, Riyadh, 11451, Saudi Arabia
f
Department of Botany and Microbiology, College of Science, King Saud University, Riyadh, 11451, Saudi Arabia

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

Keywords: TiO2, ZnO, and SnO2 metal oxides were synthesized by the sol-gel method and heterojunctions were fabricated
Heterojunctions by combining TiO2 with either ZnO or SnO2 in a 1:1 ratio using mechanochemical ball milling process. The ball
Oxides milling process promotes phase transition of TiO2 from anatase to rutile and yields ternary heterojunction of the
Phase transitions
type TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2 (A-anatase and R-rutile). These ternary heterojunctions
Crystal structure
Defects
were characterized by various analytical techniques and its photocatalytic efficiency is evaluated using 4-Chloro
Phenol as a model compound under UV and solar light. The enhanced catalytic activity of TiO2(A)/TiO2(R)/ZnO
heterojunction is attributed to the formation of Ti3+-Vo defect states which leads to the efficient charge carrier
separation. During the ball milling process severe crystal deformation takes place in TiO2 and ZnO lattices by
creating crystal lattice distortion which leads to the formation of defects due to valency mismatch between Ti4+
and Zn2+. A mechanistic pathway is proposed for the enhanced photocatalytic activity of the ternary
heterojunctions.

1. Introduction et al., 2013; Kim and Choi, 2014). In this regard semiconductors like
TiO2 (Nakata et al., 2012; Schneider et al., 2014), ZnO (Morrison and
The new technologies implementation and improvement of existing Freund, 1967; Gu et al., 2016), SnO2 (Kim et al., 2016; Joudi et al.,
technologies to alters the physicochemical and structural properties has 2019), Fe2O3 (Theerthagiri et al., 2014; Karunakaran and Senthilvelan,
been well focused in recent years. For the abatement of organic pollut­ 2006; Ramu et al., 2020), WO3 (Szilágyi et al., 2012; Do et al., 1994),
ants from the water resources, different types of Advanced Oxidation Nb2O5 (Su et al., 2021; Yan et al., 2014), Bi2O3 (Zhou et al., 2009; Xu
Process (AOP’s) and hybrid AOP’s are implemented (Madhavan et al., et al., 2017) etc., finds extensive application but TiO2 stood as bench­
2019; Yu et al., 2021; Lee et al., 2022). But among those various AOP mark material. TiO2 has three different polymorphs namely anatase,
techniques heterogeneous photocatalysis has gained more interest. The rutile and brookite. Among them, the anatase phase with a higher
fundamental principles of photocatalysis rely on the utilization of pho­ effective mass of electrons with low mobility of charge carriers and its
togenerated charge carriers. To enhance the photocatalytic efficiency high degree of surface hydroxylation exhibits better efficiency compared
the photogenerated charge carriers (electron-hole pairs) should be to other polymorphs. The practical application of anatase and rutile
separated efficiently and to restrain the recombination, thus generated phases is restricted due to their activity only under UV irradiation. To
charge carriers has to move across the surface/interface (Wang et al., increase the photocatalytic efficiency of TiO2 several appropriate mod­
2014a; Pelaez et al., 2012; Daghrir et al., 2013; Dozzi et al., 2013; Park ifications are adopted namely, doping of metal/non-metal ions, noble

* Corresponding author. Department of Post Graduate Studies in Chemistry, Bangalore University, Bengaluru, 560001, India.
E-mail address: dr.arunakumariml@gmail.com (M.L.A. Kumari).

https://doi.org/10.1016/j.envres.2021.111841
Received 7 July 2021; Received in revised form 1 August 2021; Accepted 2 August 2021
Available online 8 August 2021
0013-9351/© 2021 Elsevier Inc. All rights reserved.
M.L.A. Kumari et al. Environmental Research 203 (2022) 111841

Fig. 1. a: Powder X-Ray diffraction pattern of a) TiO2 (A), ZnO and TiO2(A)/TiO2(R)/ZnO samples. b: Powder X-Ray diffraction pattern of TiO2(A), SnO2, and
TiO2(A)/TiO2(R)/SnO2 samples.

in increasing the efficiency under visible wavelength region. The tuning


Table 1 of electronic properties takes place in heterojunction structures due to
Percentage phase composition and band gap values of prepared samples.
the formation of heterojunctions between three components. The charge
Sample % of Anatasae % of Rutile % of % of Bandgap migration mechanism across the heterojunctions will produce a redshift
TiO2 TiO2 ZnO SnO2 (Eg)
in the optical absorption spectrum. Heterojunctions containing the rutile
TiO2 100 – – – 3.2 eV phase can reduce the rate of photo-generated charge carrier recombi­
ZnO – – 100 – 3.2 eV nation since its conduction band edge position is 0.2 eV below the
SnO2 – – – 100 3.6 eV
TiO2A/ 22.7 18.5 58.7 – 3.0 eV
anatase phase thereby facilitating smooth migration of charges at the
TiO2R/ heterojunction interface (Wang et al., 2013; Dahl et al., 2014). Besides,
ZnO the heterojunction between the two crystalline phases enhances the
TiO2A/ 27.5 24.5 – 47.9 3.0 eV quantum efficiency of the photocatalyst due to the vectorial transfer of
TiO2R/
charge carriers. At the interface, lattice space difference between the
SnO2
semiconductors induces lattice mismatch and create several defects al­
ters the electronic band structure, and induces lattice strain, which
metal deposition, surface sensitization, and formation of heterojunctions hinders the recombination of photogenerated charge carriers by trap­
by combining with other semiconducting materials of different bandgap ping (Wang et al., 2013). Recent studies have revealed that ternary
(Dozzi et al., 2013; Park et al., 2013). Researchers have paid more heterojunction exhibiting better catalytic property compared to their
attention to the electronic coupling of two different bandgap semi­ individual counterpart and binary heterojunctions (Varaprasad et al.,
conductors to form heterojunction because of their unique properties 2021; Mohanta and Ahmaruzzaman, 2021a, 2021b).
arising from the interfacial interaction that is not present in the pure In the present research work preparation of TiO2(A)/TiO2(R)/ZnO
counterparts (Theerthagiri et al., 2015, 2017, 2018, 2019; Fateh et al., and TiO2(A)/TiO2(R)/SnO2 heterojunctions (A-anatase and R-rutile) is
2014; Xiao et al., 2014; Naik et al., 2021). attempted by combining anatase TiO2 with either ZnO or SnO2 via high-
It is widely accepted that the solid interfacial structure in the het­ energy ball milling technique. The influence of phase composition,
erojunctions plays a key role in improving the charge transfer process electronic band structure, defects and reaction conditions on, photo­
between different semiconductors. This improvement is addressed to catalytic activity were investigated in detail. In the present study, weight
build in electric field at junction interface, band bending and vectorial proportional ratio of two coupled photocatalysts was kept at 1:1. The
transfer of electrons and holes across the heterojunction (Kumar and study pertaining to the effect of mixing of proportional ratios of these
Rao, 2015; Zhang and YatesJr, 2012; Wang et al., 2013). It is well known coupled photocatalysts on the photocatalytic performance is still under
that an optimal ratio of anatase-rutile mixture leads to better photo­ investigation.
catalytic activity compared to their pure phases. The mixed-phase TiO2
like Degussa P25 exhibits higher photocatalytic activity compared to
pure anatase or rutile phase. Apart from usage of mixed-phase TiO2,
heterojunction between TiO2 and non-TiO2 materials is quite promising

2
M.L.A. Kumari et al. Environmental Research 203 (2022) 111841

anatase TiO2 powder along with ZnO or SnO2 were loaded in 1:1 ratio in
the ball-milling reactor followed by addition of 10 ml of water. Ball
milling process was carried out for about 3 hours in 600 rpm speed, the
obtained powder is further dried at a temperature of 110 ◦ C in an oven.
The mechanical force in ball milling process promotes the TiO2 phase
transition from anatase to rutile yielding a ternary heterojunctions of the
type TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2.

2.3. Characterization

Phase compositions of prepared samples were determined using


powder X-ray diffraction (PXRD) using Brucker-D8 Advance diffrac­
tometer. The diffraction patterns were recorded using Cu Kα radiation
with Ni filter in the specific 2θ range from 20◦ to 80◦ at a scan rate of 2◦
per min. The UV–vis diffused reflectance spectra were recorded using
Shimadzu-UV 3101 PC UV-VIS-NIR spectrophotometer, in which BaSO4
was used as the background in range of 200–800 nm. The X-ray
photoelectron spectroscopy (XPS) measurements were carried out using
AXIS ULTRA from AXIS 165, integrated with Kratos patented magnetic
immersion lens along with the charge neutralization system and
spherical mirror analyzer. All the binding energies of the elements were
calibrated with respect to C 1s (284.6 eV). The morphological features
were studied using TECNAIF-30 Transmission Electron Microscopy
(TEM) and High-Resolution Transmission Electron Microscopy
(HRTEM) techniques. Image J software was used for particle size
analysis.

2.4. Photocatalytic activity

Photocatalytic activity was evaluated by using 4-CP as a model


organic pollutant compound under both UV/Solar light illumination
(Experimental setup details are given in S2). In a typical test, 100 mg/L
of the photocatalyst was dispersed in 10 mg/L of 4-CP solution under the
specified conditions and the solution was stirred vigorously using
magnetic stirrer. 5 ml aliquot samples were collected from the reaction
suspension at definite time interval and subjected to centrifugation. The
solution was then filtered through 0.45 μm Millipore filter to remove the
catalyst particles and the residual concentration is measured at char­
Fig. 2. a: UV–visible diffuse reflectance spectra of all the photocatalysts. b: The acteristic band of 4-CP (225 nm) using UV–visible spectroscopic
plots of [F (R∞) hv ] ½ v/s photon energy (hv) for TiO2(A)/TiO2(R)/ZnO and technique.
TiO2(A)/TiO2(R)/SnO2 composites.
3. Results and discussion
2. Experimental section
3.1. Phase characterization
2.1. Materials
X-ray diffraction patterns of TiO2, ZnO, SnO2, TiO2(A)/TiO2(R)/ZnO
Titanium (IV) chloride (TiCl4 ≥ 99.9 %), zinc nitrate hexahydrate and TiO2(A)/TiO2(R)/SnO2 samples are shown in Fig. 1. The search
(Zn(NO3)2.6H2O ≥ 99 %), tin (IV) chloride (SnCl4.5H2O) were obtained match data base analysis confirms that, prepared photocatalyst powders
from Merck Chemicals Limited, ammonia and 4-chlorophenol have well indexed phase composition of anatase TiO2 phase (JCPDS #
(C6H5OHCl) were obtained from SD Fine Chemicals. Sulphuric acid 21–1272), hexagonal wurtzite ZnO phase (JCPDF # 36–1451) and
(H2SO4) and ethanol were from Sisco-chemical industries. All the re­ tetragonal rutile SnO2 phase (JCPDF # 41–1445). In the case of TiO2(A)/
agents used were of analytical grade. Double distilled water was used TiO2(R)/ZnO (Fig. 1a) heterojunction diffraction pattern shows a multi-
throughout the experiment. phased pattern consists of both anatase and rutile TiO2 (JCPDF #
21–1276) along with hexagonal wurtzite phase of ZnO. TiO2(A)/
TiO2(R)/SnO2 (Fig. 1b) also shows multi phased pattern of TiO2 similar
2.2. Catalyst preparation to the previous case along with tetragonal rutile phase of SnO2. The
rutile phase formation in the heterojunction samples suggests that high
TiO2 (anatase), ZnO and SnO2 metal oxides were prepared by sol-gel mechanical energy produced during ball-milling process accelerates the
method through hydrolysis of TiCl4, Zn(NO3)2.6H2O and SnCl4 respec­ anatase to rutile phase transformation.
tively. The detailed experimental process was presented in the supple­ It is reported that, phase transformation of the anatase TiO2 to
mental file (S1) (Audaithai and Kutty, 1987; Zhang et al., 2013; Gu et al., thermodynamically stable rutile is possible during the ball milling pro­
2004). TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2 hetero­ cesses. leads the undergo lattice deformations due to the stress and strain
structure preparation were carried out in the ball miller (DDR-GM 9458, induced. The stress and strains and leads to lattice deformation and
Fritsch Pulverisette-7, Germany). The ball milling reactor had two steel create many defects possessing high lattice and surface energies which
balls with a diameter of 10 mm and the mixing was carried out at reduces the energy activation for elements diffusion at room tempera­
ambient conditions (room temperature 25 ± 2 ◦ C, 1 atm pressure). Pure ture. The collision between balls of the ball miller and grains of the

3
M.L.A. Kumari et al. Environmental Research 203 (2022) 111841

Fig. 3. a: Zn 2p XPS spectra of ZnO and TiO2(A)/TiO2(R)/ZnO. b: Sn 3d XPS spectra of SnO2 and TiO2(A)/TiO2(R)/SnO2. c: Ti 2p XPS spectra of TiO2, TiO2(A)/
TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2. d: O1s XPS spectra of TiO2, ZnO, SnO2, TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2.

sample powder produces a rise in interface temperature leads a strong


coupling reaction and leads to the formation of heterojunction (Shifu (1 − ​ R∞ )2
F(R∞ ) = (2)
et al., 2006a; Yin et al., 2003). The XRD results confirmed the formation 2 ​ R∞
of triphasic heterojunctions of the type TiO2(A)/TiO2(R)/ZnO and The TiO2, ZnO and SnO2 exhibits strong absorption below the
TiO2(A)/TiO2(R)/SnO2. Besides the qualitative analysis, a quantitative wavelength of 380 nm (Fig S3), which is attributed to intrinsic band gap
estimation of phase content in heterojunctions were estimated using absorption due to electronic transition between valence band (VB) to
semi quantitative formula (Liao et al., 2010). conduction band (CB). In contrast, TiO2(A)/TiO2(R)/ZnO and TiO2(A)/
Hi TiO2(R)/SnO2 heterojunction shows noticeable red shift in absorption
Ci = ∑ × 100% (1) edge towards 400–430 nm range, which can be attributed to the char­
Hi
acteristic rutile phase of TiO2 (Eg = 3.0 eV) formed during the ball-
where Ci is percentage of phase i, Hi denotes the strongest characteristic milling process, which is also confirmed by PXRD studies. The calcu­

peak intensity of phase i, and Hi represents sum of the strongest lated values of band gap energy are given in Table 1. The change in the
characteristic peak intensities of all the phases. The results are listed in optical band gap of TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2
Table 1. heterojunctions can be attributed to the i) mixing of energy band levels
of the heterojunction semiconductors, ii) interfacial coupling effect at
junctions between different phases like TiO2(A) and TiO2(R), TiO2(R)
3.2. Optical properties
and ZnO, TiO2(A) and ZnO. In the case of TiO2(A)/TiO2(R)/SnO2 the
effect is between TiO2(A) and TiO2(R), TiO2(R) and SnO2, TiO2(A) and
The optical absorbance from diffused reflectance spectra of TiO2,
SnO2 junctions, iii) different band edge positions of all the pristine
ZnO, SnO2, TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2 hetero­
semiconductors (Xu et al., 2014; Yan et al., 2012a).
junctions are given in Fig. 2a. The band gap energy (Eg) was calculated
The TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2 hetero­
using Kubelka Munk function. The Eg values were obtained by extrap­
junctions have lower energy band gap values compared to their indi­
olation of linear portion of the plot of [F (R∞) hv ] ½ v/s photon energy
vidual metal oxides with an added advantage of response in the visible
(hv) as shown in Fig. 2b. The Kubelka Munk function F (R∞) and photon
region. Interestingly, the TiO2(A)/TiO2(R)/ZnO heterojunction exhibits
energy (hv) can be calculated by following equation. Where R∞ is
a wide absorption peak in the range 420–580 nm (blue, green, and
reflection coefficient of the sample (Wan et al., 2007).

4
M.L.A. Kumari et al. Environmental Research 203 (2022) 111841

Fig. 3. (continued).

yellow region). This mainly arises due to the Ti3+ associated oxygen is well fitting to the standard reference value of Zn2+ elemental state
vacancies (Ti3+- Vo) which is alternatively referred to as F+ centers in (Xiao et al., 2014). The positive shift in the BE of TiO2(A)/TiO2(R)/ZnO
anatase TiO2 (Santara et al., 2013). One of the reasons for the formation indicated the strong interaction between ZnO and TiO2 at the interface
of these defects in TiO2(A)/TiO2(R)/ZnO heterojunction materials could and can be attributed to the different chemical environment of Zn at the
probably be attributed to high energy ball milling process. In the process interface (Zn2+-O-Ti4+ instead of Zn2+-O-Zn2+) (Chen et al., 2012).
of ball milling, crystal lattices of TiO2 and ZnO undergo severe plastic However ionic state of Zn will not change which is confirmed by XPS.
deformation, producing stresses and strains. This deformation creates High resolution Sn 3d XPS spectra of SnO2 and TiO2(A)/TiO2(R)/
crystal lattice distortion and leads to the formation of several defects due SnO2 are given in Fig. 3b. In SnO2, Sn 3d region exhibits spin-coupled
to mismatch of ionic valencies (Ti4+ and Zn2+) (Shifu et al., 2008). The doublet peaks with BE’s of 486.6 and 495.0 eV ascribed to Sn 3d5/2
absence of such F+ centers in TiO2(A)/TiO2(R)/SnO2 heterojunction can and Sn 3d3/2 respectively. But in the case of TiO2(A)/TiO2(R)/SnO2 the
be attributed to similar valencies of Ti4+ and Sn4+. characteristic peaks are observed at 485.1 and 493.5 eV with slight
negative shift (− 1.5 eV). Both Sn 3d peaks were highly symmetric and
the spin orbital splitting energy of Sn 3d peaks was found to be 8.4 eV,
3.3. X-ray photoelectron spectroscopic (XPS) studies which is in agreement with the literature. This splitting energy value
confirms the ionic state of tin as Sn4+ in both SnO2 and TiO2(A)/
Fig. 3a displays high resolution XPS spectra of Zn 2p region of bare TiO2(R)/SnO2 samples (Yang et al., 2012; Liu et al., 2014). The observed
ZnO and TiO2(A)/TiO2(R)/ZnO samples. In pure ZnO, Zn 2p region negative shift (− 1.5 eV) is due to the process of electron transfer from Sn
showed a spin-coupled doublet (2p3/2 and 2p1/2) with peaks centered at to Ti at the interface (Sn–O–Ti) and it is also consistent with the greater
1021.1 and 1044.2 eV. But in the case of TiO2(A)/TiO2(R)/ZnO the basicity of SnO2 relative to TiO2. This leads to decrease in electron cloud
characteristic peaks were observed at 1022.3 and 1045.4 eV. Slight density on Sn atom which in turn decreases the BE of Sn 3d (Xu et al.,
positive shift (+1.2 eV) in binding energy (BE) was observed in the case 2012).
of TiO2(A)/TiO2(R)/ZnO compared to ZnO. Spin-orbital splitting energy The high-resolution Ti 2p region of TiO2, TiO2(A)/TiO2(R)/ZnO and
for both ZnO and TiO2(A)/TiO2(R)/ZnO was found to be 23.1 eV, which

5
M.L.A. Kumari et al. Environmental Research 203 (2022) 111841

Fig. 3. (continued).

TiO2(A)/TiO2(R)/SnO2 samples exhibits spin coupled peaks pertaining 2013; Kang et al., 2014). But in the case of TiO2(A)/TiO2(R)/ZnO and
to Ti 2p3/2 and Ti 2p1/2 as shown in Fig. 3c. The spin-orbital splitting TiO2(A)/TiO2(R)/SnO2 heterojunctions a broadened Ti 2p peak is
energy between Ti 2p3/2 and Ti 2p1/2 was found to be 5.7 eV. This value observed due to the heterogeneous environment caused by the mixing of
confirms the presence of titanium in Ti4+oxidation state (Uddin et al., anatase and rutile TiO2 phases. Compared to anatase TiO2, a negative

6
M.L.A. Kumari et al. Environmental Research 203 (2022) 111841

Fig. 4. TEM image and plot of frequency distribution versus particle size of a) TiO2(A)/TiO2(R)/ZnO b) TiO2(A)/TiO2(R)/SnO2 composite.

shift in Ti 2p peak is observed for TiO2(A)/TiO2(R)/ZnO and 532.4 eV arises due to the oxygen found in the surface adsorbed hy­
TiO2(A)/TiO2(R)/SnO2 heterojunctions by ~0.3 eV, and ~0.8 eV droxyl group and the peak at 534.0 eV can be attributed to oxygen atom
respectively. The observed negative shift in heterojunctions can be in vicinity of Ti3+ associated oxygen vacancies (Ti3+- Vo- O). The O 1s
attributed to formation of heterogeneous environment of two different XPS spectra of TiO2(A)/TiO2(R)/SnO2 heterojunction shows two BE
TiO2 phases. A small BE peak observed at 461.4 eV in the case of TiO2 peaks. The peak around 529.0 eV is due to the lattice oxygen and the
(A)/TiO2(R)/ZnO indicates the presence of small fraction of titanium in peak at 531.2 eV arises due to the oxygen of surface bound hydroxyl
Ti3+ oxidation state at the interface adjacent to Zn2+ ions (Kang et al., groups (Young et al., 2011; Yan et al., 2012b).
2014).
Fig. 3d shows the O1s XPS spectra pertaining to bare TiO2, ZnO,
SnO2, TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2 samples. The O 3.4. TEM analysis of TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/
1s spectra of bare TiO2 show a peak at 533.0 eV can be assigned to lattice SnO2 heterojunctions
oxygen. However, ZnO and SnO2 exhibits two BE peaks at 530.6 eV and
530.5 eV due to lattice oxygen present in the respective metal oxides TEM images and particle size distribution of as prepared TiO2(A)/
(O–Zn–O, O–Sn–O). Hydroxyl group bound on surface of ZnO and SnO2 TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2 is shown in Fig. 4. The low
shows BE peaks at 532.2 eV and 532.6 eV respectively. magnification image shows microstructures formed by the random
A wide and asymmetric peak of O1s spectrum observed in the case of attachment of the nanoparticles of sizes ranging from 20 to 80 nm and
TiO2(A)/TiO2(R)/ZnO heterojunction exhibits three BE components, 20–110 nm for TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2 het­
which implies the presence of oxygen in more than one chemical state. erojunctions respectively. Average particle size is found to be 22.3 nm
The BE peak at 530.9 eV can be attributed to lattice oxygen the peak at and 24.8 nm for TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2
respectively.

7
M.L.A. Kumari et al. Environmental Research 203 (2022) 111841

Fig. 5. HRTEM image of a) TiO2(A)/TiO2(R)/ZnO composite b) TiO2(A)/TiO2(R)/SnO2 composite.

The selected area electron diffraction (SAED) pattern of TiO2(A)/ both UV and solar illumination. This suggests that ternary hetero­
TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2 heterojunctions are provide in junction system showed better activity when compared to pristine in­
supplementary material. Strong Debye–Scherrer rings in the electron dividual photocatalysts (Fig. 6). The enhanced activity of
diffraction pattern indicate excellent crystallinity and it shows a super heterojunctions may be attributed to the higher specific surface area. It
imposition of three different sets of diffraction patterns. TiO2(A)/ is well known that the ball milling process enhances the specific surface
TiO2(R)/ZnO heterojunction shows the super imposition of anatase area which will lead to the higher extent of adsorption of pollutant
TiO2, rutile TiO2 and wurtzite ZnO. The dominant diffraction spot rings substrate molecules thereby enhances the photocatalytic activity (Shifu
could be indexed to the crystal planes of ZnO [(100), (002), (101), (102), et al., 2005a, 2005b, 2006b). The rate constant values are presented in
(110), (112)], anatase TiO2 [(101), (200)] and rutile TiO2 [(110), (101), Table 2 and inference from the results implicates higher activity of
(111), (211)] (Wang et al., 2014b). Superimposition of diffraction pat­ TiO2(A)/TiO2(R)/ZnO catalyst under both UV and solar light
terns corresponding to anatase TiO2, rutile TiO2 and rutile SnO2 is irradiation.
observed for TiO2(A)/TiO2(R)/SnO2. The diffraction spot rings can be The interfaces in the heterojunction play a definite role in aiding the
indexed to the crystal planes of anatase TiO2 [(101), (200)], rutile SnO2 overall photocatalytic activity. The band gap values of TiO2 (A),
[(101), (111), (211), (002), (202), (112)] and rutile TiO2 [(101), (111), TiO2(R), ZnO and SnO2 are 3.2 eV, 3.0 eV, 3.37 eV and 3.6 eV respec­
(211), (002), (202), (112)] (Lee et al., 2012). tively as shown in Fig. S5. These values suggest that each of them to have
Diffraction patterns of both the heterojunctions show multi-phased different electron affinity and band configuration. The CB edge of
structures of anatase and rutile TiO2 components, which further anatase TiO2 is in between the CB edge positions of SnO2 and ZnO.
confirm the formation of rutile phase during the ball milling process. Further CB edge of rutile TiO2 is 0.2 eV lower than the anatase CB edge
HRTEM analysis of TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/ leading to the “staggered” type II heterojunctions at the interface.
SnO2 heterojunctions were taken, precisely focusing at heterojunction Therefore, a tentative mechanism for the band configuration at the
interface which displays three types of clear lattice fringes (Fig. 5). In contacted interface of TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/
case of TiO2(A)/TiO2(R)/ZnO shows d-spacing of 0.35 nm, 0.32 nm and SnO2 is proposed. Figs. 7 and 8 shows the schematic diagram of the
0.24 nm could be ascribed to crystal plane (101) of anatase TiO2, (110) energy band gap and the charge transfer mechanism in the hetero­
of rutile TiO2 and (101) of ZnO respectively (Johra and Jung, 2015). junction semiconductor system.
TiO2(A)/TiO2(R)/SnO2 heterojunction also shows three distinctive lat­
tice fringes 0.35 nm, 0.32 nm, and 0.33 nm which could be ascribed to 3.5.1. UV light driven photocatalytic activity
the crystal plane (101) of anatase TiO2, (110) of rutile TiO2, and (110) of Under UV light irradiation TiO2(A)/TiO2(R)/ZnO and TiO2(A)/
rutile SnO2 respectively, which are in consistent with the XRD analysis. TiO2(R)/SnO2 heterojunctions exhibit better photocatalytic activity
Figure also indicates the overlapping of mixed fringe lattices, suggesting compared to bare TiO2, ZnO and SnO2 (as shown in Fig. 6). The higher
the presence of well-crystallized mixed phases within the photocatalytic activity of heterojunctions can be attributed to its band
heterojunction. configuration. In the case of TiO2(A)/TiO2(R)/ZnO, anatase TiO2 and
ZnO has same band-gap energy of 3.2 eV, but the CB of ZnO is situated at
higher energy than that of the anatase., Under UV light illumination
3.5. Photocatalytic degradation studies
band gap excitation takes plays only in ZnO and anatase TiO2. The
photogenerated electron in the CB of ZnO may jump to either the CB of
The photocatalytic activity of the prepared samples is evaluated by
anatase or CB of rutile TiO2, since its CB is 0.2 eV lower than the anatase
the degradation of 4-CP under the illuminations of UV and solar light. A
TiO2 (as shown in Fig. 7a). Thus, rutile serves as a passive electron sink,
blank experiment containing only 4-CP in the absence of photocatalyst
hindering the recombination in the anatase phase. Likewise, the posi­
but under UV/solar illumination was performed in order to determine
tions of VB energy of anatase TiO2 is lower than ZnO leading to the hole
the contribution of direct photolysis. The experimental results
transfer from anatase TiO2 to ZnO (Pei and Leung, 2013). Among these
confirmed the absence of contribution from direct photolysis and there
three semiconductors, the photoinduced electron–hole pairs are effec­
was no significant degradation of the 4-CP in presence of photocatalysts
tively separated due to the suitable band energy positions and the life­
under the dark conditions suggesting that the 4-CP degradation arises
time of charge carriers is increased which intern improves
mainly from the photocatalytic activity of excited semiconductor. The
photocatalytic reaction.
decreasing order of photocatalytic activities is observed to be TiO2(A)/
In case of TiO2 (A)/TiO2(R)/SnO2 heterojunction under UV light
TiO2(R)/ZnO > TiO2(A)/TiO2(R)/SnO2 > TiO2 > ZnO > SnO2 under

8
M.L.A. Kumari et al. Environmental Research 203 (2022) 111841

Fig. 6. The plot of C/Co versus time for the degradation of 4-CP under a) UV illumination b) Solar irradiation.

9
M.L.A. Kumari et al. Environmental Research 203 (2022) 111841

Table 2
Kinetic studies under UV/Solar light.
Photocatalysts UV light Solar light
− 2 − 1 2 1
Rate constant(k) in 10 min Percentage degradation in 70 min Rate constant(k) in 10− min− Percentage degradation in 150 min

TiO2 1.37 64.2 0.08 12.5


ZnO 1.19 53.9 0.10 14.2
SnO2 1.02 36.7 0.05 7.5
TiO2A/TiO2R/ZnO 3.71 100 1.52 92.7
TiO2A/TiO2R/SnO2 2.34 82.2 1.02 72.3

illumination, the photogenerated electrons are transferred from the CB neighboring Ti atoms around the oxygen vacancy site, and form shallow
of anatase TiO2 to CB’s of SnO2 and rutile TiO2. The photogenerated donor states below the conduction band originating from Ti 3d orbitals.
holes transfer from the VB of SnO2 to VB’s of anatase or rutile TiO2 (as Thus, it acts as electron traps not as hole traps and inhibits the electron-
shown in Fig. 7b) (Zhang et al., 2014; Abdel-Messih et al., 2013). The hole recombination (Fig. 8a) (Nakamura et al., 2000). Further, the
accumulated electrons in the CB can be transferred to molecular oxygen as-formed Ti3+ defect states can take part in a new photoexcitation
reducing it to super oxide radical anion and photogenerated holes can process by creating a shallow donor level just below the CB, which will
oxidize the 4-CP molecules by the production of hydroxyl radicals. also contribute to the visible light response (Pan et al., 2013; Serpone,
Among these two heterojunctions TiO2(A)/TiO2(R)/ZnO shows 2006). Even though TiO2(A)/TiO2(R)/SnO2 has higher amount of rutile
higher activity compared to TiO2(A)/TiO2(R)/SnO2, which can be TiO2 content than TiO2(A)/TiO2(R)/ZnO it exhibits poor visible light
attributed to the extent of generation of superoxide radicals on the activity due to the absence of Ti3+ induced defect states (Fig. 8b).
surface of the catalyst. This is because SnO2 can only trap the electrons
from anatase CB. But further movement of this electron to the molecular 4. Conclusion
oxygen is inhibited due to the mismatch of redox potential of O2/O2•-
with the CB edge position of SnO2 (O2+ e− →O2•- E = − 0.28 eV) (de In summary, the ternary heterojunction systems have been success­
Mendonça et al., 2014). fully synthesized by mechano-chemical process and its activity was
Further, the proportion of rutile content is high compared to the studied using 4-CP as the model compound. The material composition,
photoactive anatase phase in TiO2(A)/TiO2(R)/SnO2 heterojunction, optical properties, as well as chemical nature and morphology have been
since SnO2 can favorably promote the rutile phase formation in the ball verified by various characterization techniques. Ball milling process
milling conditions. This will also decrease extent of hydroxyl radical induces the phase transition of TiO2 from anatase to rutile. This process
generation in TiO2(A)/TiO2(R)/SnO2 heterojunction which intern de­ leads to the formation of F+ centers (Ti3+-Vo) in TiO2(A)/TiO2(R)/ZnO
creases the degradation rate. heterojunction due to mismatch of valencies of titanium and zinc at the
Further the higher photocatalytic activity of TiO2(A)/TiO2(R)/ZnO interface as confirmed by UV–visible diffused reflectance and XPS
can be accounted to the formation of Ti3+-Vo state below the CB of techniques. Heterojunction systems showed better photocatalytic per­
anatase TiO2. Formation of Ti3+-Vo state is confirmed by UV–visible formance under both UV/Solar light irradiation by effectively inhibiting
diffused reflectance and XPS spectroscopic techniques. This defect en­ electron− hole recombination at the interfaces. The results show that the
ergy level will serve as shallow trap for the CB electrons, where an photocatalytic degradation ability of the TiO2(A)/TiO2(R)/ZnO is
electron can be trapped momentarily and further it can be detrapped to significantly higher than TiO2(A)/TiO2(R)/SnO2 under both UV/Solar
the molecular oxygen quite easily. This process hinders the photo­ light. This can be attributed to the formation of Ti3+ induced defect
generated charge carrier recombination effectively and promote the states, which will serve as shallow traps for the CB electrons and this
oxidation of 4-CP (Liqiang et al., 2004). state can act as an energy level to which photoexcitation can take place
under solar light irradiation. The lower activity of TiO2(A)/TiO2(R)/
3.5.2. Solar light driven photocatalytic activity SnO2 compared to TiO2(A)/TiO2(R)/ZnO can be accounted to unsuitable
Under solar light illumination TiO2, ZnO and SnO2 samples did not CB band edge position of SnO2 for reducing molecular oxygen and this
show significant activity compared to heterojunctions due to their results in poor generation of superoxide radicals. The ternary hetero-
higher band gaps (as shown in Fig. 6b). However, both the hetero­ heterojunctions could provide a new path for further developments in
junctions show higher activity under solar light due to the mixing up of the photocatalytic applications and can be a suitable candidate for
energy levels. Further, the interfacial mixed polymorph structure con­ various eco-friendly energy and environmental applications.
tains certain amount of tetrahedral Ti4+ sites in rutile phase, which are
more active than octahedral coordinated Ti4+ sites present in anatase Author contributions
phase under solar light illumination. These tetrahedral Ti4+ sites could
serve as catalytic hot spots at the heterojunction interface and thus avail M. L. Aruna Kumari: Data curation, Writing – original draft prepa­
the mixed polymorph nanocrystals into an effective photocatalytic relay ration, Methodology, Review and revision. L. Gomathi Devi: Investiga­
for solar energy utilization. Under solar light illumination TiO2(A)/ tion, concept of study, critical review and Supervision. Gilberto Maia:
TiO2(R)/ZnO exhibits higher photocatalytic activity than TiO2(A)/ Review, revision and Data curation. Tse-Wei Chen: Data curation. Nabil
TiO2(R)/SnO2. The band gap of rutile is favorable for visible light Al-Zaqri: Data curation. M. Ajmal Ali: Data curation.
excitation as the CB edge of rutile lies 0.2 eV below the CB edge of
anatase.
In the case of TiO2(A)/TiO2(R)/ZnO the photogenerated electron in Declaration of competing interest
the CB of rutile TiO2 may jump to the defect energy level originating
from oxygen vacancies located around 0.75–1.18 eV below the CB of The authors declare that they have no known competing financial
TiO2. The oxygen vacancies formed by removal of neutral oxygen atoms interests or personal relationships that could have appeared to influence
cause the redistribution of the excess electrons among the nearest the work reported in this paper.

10
M.L.A. Kumari et al. Environmental Research 203 (2022) 111841

Fig. 7. Schematic representation of a model depicting photocatalytic reaction mechanism of a) TiO2(A)/TiO2(R)/ZnO b) TiO2(A)/TiO2(R)/SnO2 under UV light.

11
M.L.A. Kumari et al. Environmental Research 203 (2022) 111841

Fig. 8. Schematic representation of a model depicting photocatalytic reaction mechanism of a) TiO2(A)/TiO2(R)/ZnO b)TiO2(A)/TiO2(R)/SnO2 under solar light.

Acknowledgment Science and Technology (DST), Government of India. The authors also
extend their appreciation to the Researchers Supporting Project number
L. Gomathi Devi and M. L. Aruna Kumari acknowledge the financial (RSP-2021/396), King Saud University, Riyadh, Saudi Arabia.
assistance from University Grants Commission (UGC) and Department of

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.envres.2021.111841.

12
M.L.A. Kumari et al. Environmental Research 203 (2022) 111841

References Schneider, J., Matsuoka, M., Takeuchi, M., Zhang, J., Horiuchi, Y., Anpo, M.,
Bahnemann, D.W., 2014. Chem. Rev. 114, 9919–9986.
Serpone, N., 2006. J. Phys. Chem. B 110, 24287–24293.
Abdel-Messih, M.F., Ahmed, M.A., El-Sayed, A.S., 2013. J. Photochem. Photobiol., A 260,
Shifu, C., Lei, C., Shen, G., Gengyu, C., 2005a. Powder Technol. 160, 198–202.
1–8.
Shifu, C., Lei, C., Shen, G., Gengyu, C., 2005b. Chem. Phys. Lett. 413, 404–409.
Audaithai, M., Kutty, T.R.N., 1987. Mater. Res. Bull. 22, 641–650.
Shifu, C., Lei, C., Shen, G., Gengyu, C., 2006a. Mater. Chem. Phys. 98, 116–120.
Chen, Y.-W., Lee, D.-S., Chen, H.-J., 2012. Int. J. hydrogen energ. 37, 15140–15155.
Shifu, C., Lei, C., Shen, G., Gengyu, C., 2006b. Mater. Chem. Phys. 98, 116–120.
Daghrir, R., Drogui, P., Robert, D., 2013. Ind. Eng. Chem. Res. 52, 3581–3599.
Shifu, C., Wei, Z., Wei, L., Sujuan, Z., 2008. Appl. Surf. Sci. 255, 2478–2484.
Dahl, M., Liu, Y., Yin, Y., 2014. Chem. Rev. 114, 9853–9889.
Su, K., Liu, H., Gao, Z., Fornasiero, P., Wang, F., 2021. Adv. Sci. 8, 2003156.
de Mendonça, V.R., Lopes, O.F., Fregonesi, R.P., Giraldi, T.R., Ribeiro, C., 2014. Appl.
Szilágyi, I.M., Fórizs, B., Rosseler, O., Szegedi, Á., Németh, P., Király, P., Tárkányi, G.,
Surf. Sci. 298, 182–191.
Vajna, B., Varga-Josepovits, K., László, K., Tóth, A.L., 2012. J. Catal. 294, 119–127.
Do, Y.R., Lee, W., Dwight, K., Wold, A., 1994. J. Solid State Chem. 108, 198–201.
Theerthagiri, J., Senthil, R.A., Priya, A., Madhavan, J., Michael, R.J.V., Ashokkumar, M.,
Dozzi, M.V., Selli, E., Photochem, J., Photobiol, C., 2013. Photochem. Rev. 14, 13–28.
2014. RSC Adv. 4, 38222–38229.
Fateh, R., Dillert, R., Bahnemann, D., 2014. Appl. Mater. Inter. 6, 2270–2278.
Theerthagiri, J., Senthil, R.A., Priya, A., Madhavan, J., Ashokkumar, M., 2015. New J.
Gu, F., Wang, S.F., Lü, M.K., Zhou, G.J., Xu, D., Yuan, D.R., 2004. J. Phys. Chem. B 108,
Chem. 39, 1367–1374.
8119–8123.
Theerthagiri, J., Senthil, R.A., Senthilkumar, B., Polu, A.R., Madhavan, J.,
Gu, X., Li, C., Yuan, S., Ma, M., Qiang, Y., Zhu, J., 2016. Nanotechnology 27, 402001.
Ashokkumar, M., 2017. J. Solid State Chem. 252, 43–71.
Johra, F.T., Jung, W.-G., 2015. Appl. Catal. Gen. 491, 52–57.
Theerthagiri, J., Chandrasekaran, S., Salla, S., Elakkiya, V., Senthil, R.A.,
Joudi, F., Naceur, J.B., Ouertani, R., Chtourou, R., 2019. J. Mater. Sci. Mater. Electron.
Nithyadharseni, P., Maiyalagan, T., Micheal, K., Ayeshamariam, A., Arasu, M.V., Al-
30, 167–179.
Dhabi, N.A., Kim, H.-S., 2018. J. Solid State Chem. 267, 35–52.
Kang, S.H., Song, K., Jung, J., Jo, M.R., Kang, Y.-M., 2014. J. Mater. Chem. 2,
Theerthagiri, J., Salla, S., Senthil, R.A., Nithyadharseni, P., Madankumar, A.,
19660–19664.
Arunachalam, P., Maiyalagan, T., Kim, H.-S., 2019. Nanotechnology 30, 392001.
Karunakaran, C., Senthilvelan, S., 2006. Electrochem. Commun. 8, 95–101.
Uddin, M.T., Nicolas, Y., Olivier, C., Toupance, T., Muller, M.M., Kleebe, H.-J.,
Kim, G. Zhang G., Choi, W., 2014. Energy Environ. Sci. 7, 954–966.
Rachut, K., Ziegler, J., Klein, A., Jaegermann, W., 2013. J. Phys. Chem. C 117,
Kim, S.P., Choi, M.Y., Choi, H.C., 2016. Mater. Res. Bull. 74, 85–89.
22098–22110.
Kumar, S.G., Rao, K.S.R.K., 2015. RSC Adv. 5, 3306–3351.
Varaprasad, H.S., Sridevi, P.V., Anuradha, M.S., 2021. Adv. Powder Technol. 32,
Lee, S.S., Bai, H., Liu, Z., Sun, D.D., 2012. Int. J. Hydrogen Energy 37, 10575–10584.
1472–1480.
Lee, S.J., Theerthagiri, J., Choi, M.Y., 2022. Chem. Eng. J. 471, 130970.
Wan, L., Li, J.F., Feng, J.Y., Sun, W., Mao, Z.Q., 2007. Mater. Sci. Eng. B 139, 216–220.
Liao, Y., Li, H., Liu, Y., Zou, Z., Zeng, D., Xie, C., 2010. J. Comb. Chem. 12, 883–889.
Wang, Y., Wang, Q., Zhan, X., Wang, F., Safdar, M., He, J., 2013. Nanoscale 5,
Liqiang, J., Xiaojun, S., Baifu, X., Baiqi, W., Weimin, C., Honggang, F., 2004. J. Solid
8326–8339.
State Chem. 177, 3375–3382.
Wang, H., Zhang, L., Chen, Z., Hu, J., Li, Shijie, Wang, Zhaohui, Liu, Jianshe, 2014a.
Liu, S., Huang, G., Yu, J., Ng, T.W., Yip, H.Y., Wong, P.K., 2014. Appl. Mater. Inter. 6,
Xinchen Wang Chem. Soc. Rev. 43, 5234–5244.
2407–2414.
Wang, Y., Zhu, S., Chen, X., Tang, Y., Jiang, Y., Peng, Z., Wang, H., 2014b. Appl. Surf.
Madhavan, J., Theerthagiri, J., Balaji, D., Sunitha, S., Choi, M.Y., Ashokkumar, M., 2019.
Sci. 307, 263–271.
Molecules 24, 3341.
Xiao, S., Zhao, L., Leng, X., Lang, X., Lian, J., 2014. Appl. Surf. Sci. 299, 97–104.
Mohanta, D., Ahmaruzzaman, Md, 2021a. Chemosphere 285, 131395.
Xu, L., Steinmiller, E.M.P., Skrabalak, S.E., 2012. J. Phys. Chem. C 116, 871–877.
Mohanta, D., Ahmaruzzaman, Md, 2021b. Environ. Res. 201, 111586.
Xu, H., Liu, W., Cao, L., Su, G., Duan, R., 2014. Appl. Surf. Sci. 301, 508–514.
Morrison, S.R., Freund, T., 1967. J. Chem. Phys. 47, 1543–1551.
Xu, D., Hai, Y., Zhang, X., Zhang, S., He, R., 2017. Appl. Sur. Sci. 400, 530–536.
Naik, S.S., Lee, S.J., Theerthagiri, J., Yu, Y., Choi, M.Y., 2021. J. Hazard Mater. 418,
Yan, X., Zou, C., Gao, X., Gao, W., 2012a. J. Mater. Chem. 22, 5629–5640.
126269.
Yan, X., Zou, C., Gao, X., Gao, W., 2012b. J. Mater. Chem. 22, 5629–5640.
Nakamura, I., Negishi, N., Kutsuna, S., Ihara, T., Sugihara, S., Takeuchi, K., 2000. J. Mol.
Yan, J., Wu, G., Guan, N., Li, L., 2014. Appl. Catal. B Environ. 152, 280–288.
Catal. Chem. 161, 205–212.
Yang, G., Yan, Z., Xiao, T., 2012. Appl. Surf. Sci. 258, 8704–8712.
Nakata, K., Fujishima, A., Photochem, J., Photobiol, C., 2012. Photochem. Rev. 13,
Yin, S., Yamaki, H., Komatsu, M., Zhang, Q., Wang, J., Tang, Q., Saito, F., Sato, T., 2003.
169–189.
J. Mater. Chem. 13, 2996–3001.
Pan, X., Yang, M.-Q., Fu, X., Zhang, N., Xuv, Y.-J., 2013. Nanoscale 5, 3601–3614.
Young, Ku, Huang, Y.-H., Chou, Y.-C., 2011. J. Mol. Catal. Chem. 342–343, 18–22.
Park, H., Park, Y., Kim, W., Choi, W., Photochem, J., Photobiol, C., 2013. Photochem.
Yu, Y., Naik, S.S., Oh, Y., Theerthagiri, J., Lee, S.J., Choi, M.Y., 2021. J. Hazard Mater.
Rev. 15, 1–20.
420, 126585.
Pei, C.C., Leung, W.W.-F., 2013. Catal. Commun. 37, 100–104.
Zhang, Z., YatesJr, J.T., 2012. Chem. Rev. 112, 5520–5551.
Pelaez, M., Nolan, N.T., Pillai, S.C., Seery, M.K., Falaras, P., Kontos, A.G., Dunlop, P.S.M.,
Zhang, D., Gong, J.Y., Ma, J.J., Han, G.Q., Tong, Z.W., 2013. Dalton Trans. 42,
Hamilton, J.W.J., Byrne, J.A., OShea, K., Entezari, M.H., Dionysiou, D.D., 2012.
16556–16561.
Appl. Catal. B Environ. 125, 331–349.
Zhang, L., Li, Y., Zhang, Q., Wang, H., 2014. Appl. Surf. Sci. 319, 21–28.
Ramu, A.G., Telmenbayar, L., Theerthagiri, J., Yang, D., Song, M., Choi, D., 2020. New J.
Zhou, L., Wang, W., Xu, H., H, Sun, S., Shang, M., 2009. Chem. Eur J. 15, 1776–1782.
Chem. 44, 18633–18645.
Santara, B., Giri, P.K., Imakita, K., Fujii, M., 2013. J. Phys. Chem. C 117, 23402–23411.

13

You might also like