ArticlePDF Available

Advanced Design and Synthesis of Composite Photocatalysts for the Remediation of Wastewater: A Review

Authors:

Abstract and Figures

Serious water pollution and the exhausting of fossil resources have become worldwide urgent issues yet to be solved. Solar energy driving photocatalysis processes based on semiconductor catalysts is considered to be the most promising technique for the remediation of wastewater. However, the relatively low photocatalytic efficiency remains a critical limitation for the practical use of the photocatalysts. To solve this problem, numerous strategies have been developed for the preparation of advanced photocatalysts. Particularly, incorporating a semiconductor with various functional components from atoms to individual semiconductors or metals to form a composite catalyst have become a facile approach for the design of high-efficiency catalysts. Herein, the recent progress in the development of novel photocatalysts for wastewater treatment via various methods in the sight of composite techniques are systematically discussed. Moreover, a brief summary of the current challenges and an outlook for the development of composite photocatalysts in the area of wastewater treatment are provided.
Content may be subject to copyright.
catalysts
Review
Advanced Design and Synthesis of Composite
Photocatalysts for the Remediation of Wastewater:
A Review
Jianlong Ge, Yifan Zhang, Young-Jung Heo and Soo-Jin Park *
Department of Chemistry and Chemical Engineering, Inha University, 100 Inharo, Incheon 22212, Korea;
gejianlong1121@126.com (J.G.); zyf910626@inhaian.net (Y.Z.); heoyj1211@inhaian.net (Y.-J.H.)
*Correspondence: sjpark@inha.ac.kr; Tel.: +82-32-860-7234; Fax: +82-32-860-5604
Received: 29 December 2018; Accepted: 28 January 2019; Published: 30 January 2019


Abstract:
Serious water pollution and the exhausting of fossil resources have become worldwide
urgent issues yet to be solved. Solar energy driving photocatalysis processes based on semiconductor
catalysts is considered to be the most promising technique for the remediation of wastewater.
However, the relatively low photocatalytic efficiency remains a critical limitation for the practical
use of the photocatalysts. To solve this problem, numerous strategies have been developed for the
preparation of advanced photocatalysts. Particularly, incorporating a semiconductor with various
functional components from atoms to individual semiconductors or metals to form a composite
catalyst have become a facile approach for the design of high-efficiency catalysts. Herein, the recent
progress in the development of novel photocatalysts for wastewater treatment via various methods
in the sight of composite techniques are systematically discussed. Moreover, a brief summary of the
current challenges and an outlook for the development of composite photocatalysts in the area of
wastewater treatment are provided.
Keywords:
Composite catalysts; photocatalysis; synergy effect; solar energy; wastewater remediation
1. Introduction
In the past several decades, with the booming of industry, the ever-increasing consumption of
natural resources, especially fresh water and fossil resources, have caused alarming damage to the
environment and seriously threaten the sustainability of human society [
1
3
]. As a worldwide concern,
freshwater pollution drives people to seek for an effective approach to repair the polluted water
environment. In general, the contaminants in water are mainly derived from the sewage effluent
of industries (e.g., textile industry, paper industry, the pharmaceutical industry, etc.), and domestic
contaminants (e.g., pharmaceuticals, pesticide, detergent, etc.) [
4
]. Until now, numerous contaminants
have been detected and are classified as inorganic ions, organic chemicals, and pathogens; most
of those contaminants are toxic to organisms [
4
7
]. Up to now, a variety of strategies including
chemical or physical coagulation [
8
], sedimentation [
9
], adsorption [
10
], membrane filtration [
11
],
and biological degradation method [
12
] have been invented to treat wastewater. However, due to the
complex compositions and different physico-chemical properties of the contaminants, there are still
several limitations of these traditional techniques, such as the low efficiency, high energy consumption,
and the risk of secondary pollution [
13
15
]. Consequently, a promoted technique with high efficiency,
low energy consumption, and being environmentally friendly is highly desired for the remediation
of wastewater.
Nowadays, the advanced oxidation processes (AOPs) have been extensively explored to remove
the non-biodegradable and highly stable compounds in water [
16
,
17
]. In fact, the AOPs are chemical
Catalysts 2019,9, 122; doi:10.3390/catal9020122 www.mdpi.com/journal/catalysts
Catalysts 2019,9, 122 2 of 32
processes that can generate highly reactive hydroxyl radicals (
·
OH) in situ. The
·
OH in water exhibits
an extremely strong oxidizing property with a high oxidation potential of 2.80 V/SHE (
·
OH/H
2
O),
such that it can non-selectively oxidize the contaminants and finally convert them to CO
2
, H
2
O, or small
inorganic ions in a short time [
17
,
18
]. In most cases, the
·
OH could be produced with the presence
of one or more primary oxidants, and/or energy sources or catalysts. Therefore, the typical AOPs
could be classified as Fenton reactions, the electrochemical advanced oxidation processes, and the
heterogeneous photocatalysis [
17
]. Compared with the traditional water remediation techniques,
the AOPs exhibit many advantages, which include: (1) the contaminants are directly destroyed or
reduced in the water body, rather than simply coagulated or filtrated from the water, thus the secondary
pollution could be avoided; (2) the AOPs are suitable for a wide range of contaminants including some
inorganics and pathogens because of their robust non-selectively oxidizability; and (3) no hazardous
byproducts will be generated due to the final reduction products of the AOPs being just CO
2
, H
2
O,
or small inorganic ions. With the abovementioned merits, the AOPs have attracted significant attention
from both scientific research and industrial processing [19].
Solar energy is a green, costless, and inexhaustible energy resource. Effective utilization of
solar energy is of vital importance for enhancing the sustainability of industry, reducing pollution,
and retarding global warming. Consequently, solar energy has been widely used in a range of
applications, such as solar heating, photovoltaics, solar thermal energy, solar architecture, artificial
photosynthesis, photocatalysis, etc. [
20
] Among which, photocatalysis is one of the most effective
strategies for the AOPs, which just rely on the light radiation on the photocatalysts to drive the
oxidization reaction at the ambient condition, and during the whole reaction process, no additional
energy is needed and no toxic byproduct will be generated; therefore, it is a green chemical
technique [
21
,
22
]. Actually, the core of photocatalytic AOPs are photocatalysts; semiconductors as the
most employed heterogeneous photocatalysis for the AOPs have attained considerable development
since Fujishima et al. [
23
] carried out the first photo-catalyzed AOP based on the titanium-oxide
(TiO
2
) in 1972. Up to now, a myriad of photocatalytic AOPs have been designed for water treatment
based on various semiconductors. In general, semiconductors are light-sensitive because of their
unique electronic structure with a filled valence band (VB) and an empty conduction band (CB) [
18
,
21
].
Figure 1and Equations (1)–(6) demonstrate the basic reaction process of a semiconductor to generate the
photocatalytic radicals, which could be decomposed in the following steps: (i) photons with a certain
energy are absorbed by the semiconductor; (ii) the absorbed photons with energy greater than the band
gap energy (E
b
) of semiconductors lead to the formation of electrons in the CB and corresponding holes
in the VB; and (iii) the generated electron–hole pairs will migrate to the surface of semiconductors for
redox reactions, and fast recombination in nanoseconds will happen at the same time (it should be
mentioned that this process is negative for the AOPs, which shall be suppressed [21,22]).
Catalysts 2018, 8, x FOR PEER REVIEW 2 of 32
Nowadays, the advanced oxidation processes (AOPs) have been extensively explored to remove
the non-biodegradable and highly stable compounds in water [16,17]. In fact, the AOPs are chemical
processes that can generate highly reactive hydroxyl radicals (·OH) in situ. The ·OH in water exhibits
an extremely strong oxidizing property with a high oxidation potential of 2.80 V/SHE (·OH/H
2
O),
such that it can non-selectively oxidize the contaminants and finally convert them to CO
2
, H
2
O, or
small inorganic ions in a short time [17,18]. In most cases, the ·OH could be produced with the
presence of one or more primary oxidants, and/or energy sources or catalysts. Therefore, the typical
AOPs could be classified as Fenton reactions, the electrochemical advanced oxidation processes, and
the heterogeneous photocatalysis [17]. Compared with the traditional water remediation techniques,
the AOPs exhibit many advantages, which include: (1) the contaminants are directly destroyed or
reduced in the water body, rather than simply coagulated or filtrated from the water, thus the
secondary pollution could be avoided; (2) the AOPs are suitable for a wide range of contaminants
including some inorganics and pathogens because of their robust non-selectively oxidizability; and
(3) no hazardous byproducts will be generated due to the final reduction products of the AOPs being
just CO
2
, H
2
O, or small inorganic ions. With the abovementioned merits, the AOPs have attracted
significant attention from both scientific research and industrial processing [19].
Solar energy is a green, costless, and inexhaustible energy resource. Effective utilization of solar
energy is of vital importance for enhancing the sustainability of industry, reducing pollution, and
retarding global warming. Consequently, solar energy has been widely used in a range of
applications, such as solar heating, photovoltaics, solar thermal energy, solar architecture, artificial
photosynthesis, photocatalysis, etc. [20] Among which, photocatalysis is one of the most effective
strategies for the AOPs, which just rely on the light radiation on the photocatalysts to drive the
oxidization reaction at the ambient condition, and during the whole reaction process, no additional
energy is needed and no toxic byproduct will be generated; therefore, it is a green chemical technique
[21,22]. Actually, the core of photocatalytic AOPs are photocatalysts; semiconductors as the most
employed heterogeneous photocatalysis for the AOPs have attained considerable development since
Fujishima et al. [23] carried out the first photo-catalyzed AOP based on the titanium-oxide (TiO
2
)
in
1972. Up to now, a myriad of photocatalytic AOPs have been designed for water treatment based on
various semiconductors. In general, semiconductors are light-sensitive because of their unique
electronic structure with a filled valence band (VB) and an empty conduction band (CB) [18,21].
Figure 1 and Equations (1)–(6) demonstrate the basic reaction process of a semiconductor to generate
the photocatalytic radicals, which could be decomposed in the following steps: i) photons with a
certain energy are absorbed by the semiconductor; ii) the absorbed photons with energy greater than
the band gap energy (E
b
) of semiconductors lead to the formation of electrons in the CB and
corresponding holes in the VB; and iii) the generated electron–hole pairs will migrate to the surface
of semiconductors for redox reactions, and fast recombination in nanoseconds will happen at the
same time (it should be mentioned that this process is negative for the AOPs, which shall be
suppressed [21,22]).
Figure 1.
Schematic illustration of the photocatalytic reaction process of a semiconductor. Adapted with
permission from Reference [18]. Copyright (2012) Elsevier.
Catalysts 2019,9, 122 3 of 32
Excitation: Photon (hv) + Semiconductor eCB + h+VB (1)
Recombination: e+ h+energy (2)
Oxidation of H2O: H2O+h+VB → •OH + H+(3)
Reduction of adsorbed O2: O2+ eO2(4)
Reaction with H+: O2+ H+→ •OOH (5)
Electrochemical reduction: OOH + OOH H2O2+ O2(6)
However, it remains a significant challenge to fabricate a high-efficiency visible light photocatalyst
solely based on an individual semiconductor photocatalyst. For example, the TiO
2
, as the most used
photocatalyst, possesses various advantages with excellent chemical stability, large surface area, non-toxicity,
and low cost [
24
]; however, its wide energy band gap (3.0–3.2 eV) means it can only be excited by the
UV light (
λ
< 400 nm), such that less than 5% of the irradiated solar energy can be effectively used [
25
].
Moreover, the fast recombination speed of electron–hole pairs seriously limits the further improvement of
its photocatalytic activity [
18
,
22
]. On the other hand, although the recently developed visible light response
semiconductors have a lower energy band gap (<3 eV), such as BiOX (X = I or Br) [
26
], they still suffer
from serious photo-corrosion problems in aqueous media via redox reactions and the fast recombination of
electron–hole pairs during the reaction process. Therefore, it is highly urgent to find an effect strategy to
further improve the performance of semiconductor photocatalysts.
From ancient times, people have recognized that the incorporation of two or more constituent
materials could obtain various composite materials with intriguing properties superior to the individual
components. Nowadays, a myriad of functional composite materials have been developed for different
applications [27,28]. Actually, the enhanced performance of a composite material is mainly attributed to
the synergistic effect of its individual constituent materials; meanwhile, this principle is also appropriate
to the design of semiconductor photocatalysts. Up to now, there have been numerous pioneering studies
reporting the design and fabrication of composite semiconductor photocatalysts via various methods,
such as doping heteroatoms or constructing heterojunctions via directly compositing with individual
semiconductors or carbonaceous nanomaterials, among others. Therefore, as shown in Scheme 1, in this
review, we aim to provide a systematic appraisal of the recent development in the design and fabrication
of various composite photocatalysts for the application of wastewater treatment. Meanwhile, some
representative photocatalysts with composite structures and morphologies from the atomic scale to
macroscopic scale are reviewed. Finally, the current developing status, challenges, and evolution trend of
the composite semiconductor photocatalysts for wastewater remediation are briefly proposed.
Catalysts 2018, 8, x FOR PEER REVIEW 3 of 32
Figure 1. Schematic illustration of the photocatalytic reaction process of a semiconductor. Adapted
with permission from Reference [18]. Copyright (2012) Elsevier.
Excitation: Photon (hv) + Semiconductor e
CB
+ h
+VB
(1)
Recombination: e
+ h
+
energy (2)
Oxidation of H
2
O: H
2
O + h
+VB
•OH + H
+
(3)
Reduction of adsorbed O
2
: O
2
+ e
O
2
(4)
Reaction with H
+
: O
2
+ H
+
•OOH (5)
Electrochemical reduction: •OOH + •OOH H
2
O
2
+ O
2
(6)
However, it remains a significant challenge to fabricate a high-efficiency visible light
photocatalyst solely based on an individual semiconductor photocatalyst. For example, the TiO
2
, as
the most used photocatalyst, possesses various advantages with excellent chemical stability, large
surface area, non-toxicity, and low cost [24]; however, its wide energy band gap (3.0–3.2 eV) means
it can only be excited by the UV light (λ < 400 nm), such that less than 5% of the irradiated solar
energy can be effectively used [25]. Moreover, the fast recombination speed of electron–hole pairs
seriously limits the further improvement of its photocatalytic activity [18,22]. On the other hand,
although the recently developed visible light response semiconductors have a lower energy band gap
(<3 eV), such as BiOX (X = I or Br) [26], they still suffer from serious photo-corrosion problems in
aqueous media via redox reactions and the fast recombination of electron–hole pairs during the
reaction process. Therefore, it is highly urgent to find an effect strategy to further improve the
performance of semiconductor photocatalysts.
From ancient times, people have recognized that the incorporation of two or more constituent
materials could obtain various composite materials with intriguing properties superior to the
individual components. Nowadays, a myriad of functional composite materials have been developed
for different applications [27,28]. Actually, the enhanced performance of a composite material is
mainly attributed to the synergistic effect of its individual constituent materials; meanwhile, this
principle is also appropriate to the design of semiconductor photocatalysts. Up to now, there have
been numerous pioneering studies reporting the design and fabrication of composite semiconductor
photocatalysts via various methods, such as doping heteroatoms or constructing heterojunctions via
directly compositing with individual semiconductors or carbonaceous nanomaterials, among others.
Therefore, as shown in Scheme 1, in this review, we aim to provide a systematic appraisal of the
recent development in the design and fabrication of various composite photocatalysts for the
application of wastewater treatment. Meanwhile, some representative photocatalysts with composite
structures and morphologies from the atomic scale to macroscopic scale are reviewed. Finally, the
current developing status, challenges, and evolution trend of the composite semiconductor
photocatalysts for wastewater remediation are briefly proposed.
Scheme 1.
The schematic illustration demonstrating the design and synthesis strategies for
composite photocatalysts.
Catalysts 2019,9, 122 4 of 32
2. Principle of the Semiconductor Photocatalysts for Wastewater Remediation
As mentioned above, the trace contaminants (e.g., phenol, chlorophenol, oxalic acid) derived
from the dyeing industry, petrochemical industry, and the agricultural chemicals are quite difficult to
remove from the water due to the low concentration and complex compositions [
4
]. A photocatalytic
degradation method is considered as the most promising strategy to deal with this problem.
According to the previous studies [
18
,
29
,
30
], as shown in Figure 2, the basic mechanism of the
photocatalytic degradation process of a contaminant could be characterized as the following steps:
(i) the target contaminants transfer from the water body to the surface of the photocatalysts, in which
the migration rate of corresponding contaminants may be influenced by the morphology and surface
properties of the catalysts (e.g., surface area, porosity, and surface charges); (ii) the contaminants
are adsorbed on the surface of catalysts with photon excited reaction sites, therefore a high surface
area of the catalysts can provide more active sites for the reaction; (iii) the redox reactions of the
photon activated sites with the adsorbed contaminants and the degraded intermediates are produced,
which are finally degraded to CO
2
and H
2
O; (iv) part of the generated intermediates and the resultant
mineralization products (CO
2
and H
2
O) desorb from the surface of catalysts to expose the active sites
for the subsequent reactions; and (v) the desorbed intermediates transfer from the interface of catalysts
and water to the bulk liquid, and part of the intermediates will repeat the procedure i–v until they are
completely degraded to CO
2
and H
2
O. Based on the abovementioned principles of the semiconductor
photocatalysts for water contaminants degradation, five main criteria for the design of an effective
photocatalyst could be proposed as follow: (1) a semiconductor with a lower E
g
is preferred so that the
electron–hole pair could be excited easier; (2) the photon absorption capacity of the catalysts shall be
as high as possible to generate more electron–hole pairs; (3) the recombination process of electron–hole
pairs must be prevented as much as possible to enhance the quantum efficiency of the photo-generated
electron–hole pairs; (4) the surface area of the catalysts shall be large to provide more reaction sites;
and (5) the chemical and physical structures of photocatalysts must be stable and be beneficial for the
mass transfer in water. To meet the abovementioned requirements, a variety of strategies have been
developed for the design, some of the most-used strategies will be summarized in this review.
Catalysts 2018, 8, x FOR PEER REVIEW 4 of 32
Scheme 1. The schematic illustration demonstrating the design and synthesis strategies for composite
photocatalysts.
2. Principle of the Semiconductor Photocatalysts for Wastewater Remediation
As mentioned above, the trace contaminants (e.g., phenol, chlorophenol, oxalic acid) derived
from the dyeing industry, petrochemical industry, and the agricultural chemicals are quite difficult
to remove from the water due to the low concentration and complex compositions [4]. A
photocatalytic degradation method is considered as the most promising strategy to deal with this
problem. According to the previous studies [18,29,30], as shown in Figure 2, the basic mechanism of
the photocatalytic degradation process of a contaminant could be characterized as the following steps:
i) the target contaminants transfer from the water body to the surface of the photocatalysts, in which
the migration rate of corresponding contaminants may be influenced by the morphology and surface
properties of the catalysts (e.g., surface area, porosity, and surface charges); ii) the contaminants are
adsorbed on the surface of catalysts with photon excited reaction sites, therefore a high surface area
of the catalysts can provide more active sites for the reaction; iii) the redox reactions of the photon
activated sites with the adsorbed contaminants and the degraded intermediates are produced, which
are finally degraded to CO
2
and H
2
O; iv) part of the generated intermediates and the resultant
mineralization products (CO
2
and H
2
O) desorb from the surface of catalysts to expose the active sites
for the subsequent reactions; and v) the desorbed intermediates transfer from the interface of catalysts
and water to the bulk liquid, and part of the intermediates will repeat the procedure i–v until they
are completely degraded to CO
2
and H
2
O. Based on the abovementioned principles of the
semiconductor photocatalysts for water contaminants degradation, five main criteria for the design
of an effective photocatalyst could be proposed as follow: 1) a semiconductor with a lower E
g
is
preferred so that the electron–hole pair could be excited easier; 2) the photon absorption capacity of
the catalysts shall be as high as possible to generate more electron–hole pairs; 3) the recombination
process of electron–hole pairs must be prevented as much as possible to enhance the quantum
efficiency of the photo-generated electron–hole pairs; 4) the surface area of the catalysts shall be large
to provide more reaction sites; and 5) the chemical and physical structures of photocatalysts must be
stable and be beneficial for the mass transfer in water. To meet the abovementioned requirements, a
variety of strategies have been developed for the design, some of the most-used strategies will be
summarized in this review.
Figure 2. Schematic diagram demonstrating the removal of contaminants in water with the presence
of photocatalysts [18,29,30].
Figure 2.
Schematic diagram demonstrating the removal of contaminants in water with the presence of
photocatalysts [18,29,30].
Catalysts 2019,9, 122 5 of 32
3. Heteroatoms Doping
Recently, the strategy of introducing heteroatoms into the lattice of corresponding semiconductors
has been widely employed to regulate the band gap of the semiconductor photocatalysts so as to
improve their absorption capacity for visible lights, which takes up almost 45% in the solar light
spectrum [
31
]. In general, the most commonly used dopants in semiconductors (e.g., TiO
2
) could be
classified as the metal cations and the non-metallic elements [32,33].
3.1. Metal Cations Doping
The most-used metal cation dopants for semiconductors mainly involve transition metal ions, such
as Fe
3+
, Co
3+
, Mo
5+
, Ru
3+
, Ag
+
, Cu
2+
, Rb
+
, Cr
3+
, V
4+
, etc. [
32
,
34
37
]. In most cases, the redox energy
states of those employed metal cations lie within the band gap states of corresponding semiconductors
(e.g., TiO
2
); therefore, the introduction of those metal ions will result in an intraband state near
the CB or VB edge of a semiconductor. Consequently, the red shift in band gap absorption of a
metal-cation-doped semiconductor is mainly contributed by the charge migration between the d
electrons of the doped cations and the CB (or VB) of the corresponding semiconductors. In addition,
the doped metal cations could act as an electron–hole trap, regulating the charge carrier equilibrium
concentration [
38
40
]. Although some transition metal cations could provide new energy levels
as electron donors or acceptors, and virtually improved the visible light absorption capacity of
corresponding semiconductors, this approach is also known to suffer from many disadvantages,
such as bad thermal stability, significant increase in the carrier-recombination centers, and the high
cost for an expensive facility, which are critical limitations for the generalization of this strategy.
3.2. Non-Metallic Anions Doping
Alternatively, doping the semiconductors with appropriate non-metallic anions has been
proven to be a facile way to regulate the intrinsic electronic structure of semiconductors and could
construct various heteroatomic surface structures such that the resultant non-metallic-anion-doped
semiconductors exhibit improved photocatalytic performances under solar light irradiation [
33
,
41
].
In general, the chemical states and locations are key factors for the regulation of the electronic state of
the dopant and the corresponding heteroatomic surface structures of the composite semiconductor
catalysts. According to a previous study [
18
], three requirements needed to be satisfied for the doping
of a semiconductor: (i) the doping process should construct states in the band gap of corresponding
semiconductors with an enhanced visible light absorption capacity, (ii) the CB minimum including
the doped states should be equal to that of the semiconductor’s or higher than that of the H
2
/H
2
O
level such that the photoreduction can be conducted, and (iii) the states in the gap should sufficiently
overlap with the band states of semiconductors to ensure that the photoexcited carriers could migrate
to the surface of catalysts within their lifetime. Based on the abovementioned principles, various
elements, including C, N, F, P, and S, were employed to substitute for the O in TiO
2
[
42
], and the results
showed that N was the most effective dopant for the improvement of visible-light photocatalysis of
TiO
2
because the pstates of N can narrow the band gap of N-doped TiO
2
via mixing with the O 2p
states [
43
]. Moreover, owing to the comparable atomic size with oxygen, small ionization energy,
and high stability, the nitrogen has been one of the most promising elements for promoting the
photocatalysis performance of the semiconductors. In general, the doped N in the TiO
2
could be
classified as the substitutional type and interstitial type (Figure 3), the substitutional type N-doped
TiO2is attributed to the oxygen replacement, while the interstitial type is attributed to the additional
N element in the lattice of TiO
2
[
41
]. Up to now, the N-doping of semiconductors can be realized via
several strategies, and the most-used techniques with certain industrial application prospects could be
mentioned as the magnetron sputtering, ion implantation, chemical vapor deposition, atomic layer
deposition, and sol-gel and combustion method, which will be discussed as follows.
Catalysts 2019,9, 122 6 of 32
Catalysts 2018, 8, x FOR PEER REVIEW 6 of 32
Figure 3. Schematic diagram demonstrating the N doping in the lattice of TiO
2
.
Adapted with
permission from Reference [41]. Copyright (2011) Royal Society of Chemistry.
3.2.1. Magnetron Sputtering Method
The magnetron sputtering method is widely used for the preparation of various hybrid
semiconductors. For example, Kitano et al. [44] fabricated nitrogen-substituted TiO
2
thin films by
using a radio frequency magnetron sputtering (RF-MS) method. The N
2
/Ar gas mixtures with
different concentration of N
2
was used as the sputtering gas. They systematically investigated the
influence of nitrogen content on the properties of the obtained N-TiO
2
thin films via regulating the
concentration of N
2
in the sputtering gases. Meanwhile, they proved that the extent of substitution of
oxygen positions with N in the lattice of TiO
2
as well as the surface morphologies of TiO
2
could be
controlled well. As a result, the visible light absorption capacity of the obtained N-TiO
2
was obviously
enhanced with bands up to 550 nm, and it was found that the band red shift extent was closely related
to the content of the substituted N element in the TiO
2
lattice. Moreover, they found that the as-
prepared N-TiO
2
photocatalyst exhibited an optimized photocatalysis reactivity with the N content
of 6%. This result was because of the excessive substituted N, which causes the formation of
undesirable Ti
3+
species and acts as the recombination centers to decrease the photocatalytic activity
[44]. Apart from the TiO
2
, some other N-doped semiconductors could also be prepared based on the
RF-MS method. Recently, Salah et al. [45] fabricated a series of N-doped ZnO nanoparticles films by
employing the RF-MS method. As shown in Figure 4, the obtained N-doped ZnO films exhibited an
improved response to the visible light, and possessed significantly enhanced
degradation/mineralization performance for 2-chlorophenol (2-CP), 4-chlorophenol (4-CP), and 2,4-
dichlorophenoxyaceticacid (2,4-D) solely under the drive of natural sunlight.
Figure 3.
Schematic diagram demonstrating the N doping in the lattice of TiO
2
. Adapted with
permission from Reference [41]. Copyright (2011) Royal Society of Chemistry.
3.2.1. Magnetron Sputtering Method
The magnetron sputtering method is widely used for the preparation of various hybrid
semiconductors. For example, Kitano et al. [
44
] fabricated nitrogen-substituted TiO
2
thin films by using
a radio frequency magnetron sputtering (RF-MS) method. The N
2
/Ar gas mixtures with different
concentration of N
2
was used as the sputtering gas. They systematically investigated the influence of
nitrogen content on the properties of the obtained N-TiO
2
thin films via regulating the concentration of
N
2
in the sputtering gases. Meanwhile, they proved that the extent of substitution of oxygen positions
with N in the lattice of TiO
2
as well as the surface morphologies of TiO
2
could be controlled well.
As a result, the visible light absorption capacity of the obtained N-TiO
2
was obviously enhanced with
bands up to 550 nm, and it was found that the band red shift extent was closely related to the content
of the substituted N element in the TiO
2
lattice. Moreover, they found that the as-prepared N-TiO
2
photocatalyst exhibited an optimized photocatalysis reactivity with the N content of 6%. This result
was because of the excessive substituted N, which causes the formation of undesirable Ti
3+
species
and acts as the recombination centers to decrease the photocatalytic activity [
44
]. Apart from the TiO
2
,
some other N-doped semiconductors could also be prepared based on the RF-MS method. Recently,
Salah et al. [
45
] fabricated a series of N-doped ZnO nanoparticles films by employing the RF-MS
method. As shown in Figure 4, the obtained N-doped ZnO films exhibited an improved response to
the visible light, and possessed significantly enhanced degradation/mineralization performance for
2-chlorophenol (2-CP), 4-chlorophenol (4-CP), and 2,4-dichlorophenoxyaceticacid (2,4-D) solely under
the drive of natural sunlight.
Catalysts 2019,9, 122 7 of 32
Catalysts 2018, 8, x FOR PEER REVIEW 7 of 32
Figure 4. (a) The degradation and (b) mineralization of N-doped ZnO films for 2-CP, 4-CP, and 2,4-
D. (c) The stability of pristine ZnO and N-doped ZnO film. (d) The stability and reusability of an N-
doped ZnO film for the degradation of 2-CP. Adapted with permission from Reference [45].
Copyright (2016) Elsevier.
3.2.2. Ion Implantation Method
The ion implantation method as a typical materials engineering strategy that can effectively
regulate the physical, chemical, and electronic properties of semiconductors, and the operation
process does not involve any other elements except the selected element, which ensures the purity of
the dopant [46]. Moreover, owing to the controllable parameters of ion beam implantation, such as
ion element, ion energy, ion density, uniformity of ion beam, and the doping efficiency, ion beam
implantation is a powerful approach for the heteroatom doping of semiconductors. For example,
Tang et al. [47] fabricated an N-doped TiO
2
layer with macrospores on a titanium substrate by using
the plasma-based ion implantation method. The fabrication process involves four steps: i) a helium
plasma was employed to generate He bubbles in the substrate, ii) an oxygen plasma treatment and a
followed annealing in air were used to obtain rutile and anatase phases of TiO
2
, iii) an Ar ion
sputtering method was used to exposure the He bubbles on the surface; and iv) the pre-treated
samples were doped by nitrogen though the nitrogen beam ion implantation method. Moreover, co-
doping of two or more non-metallic anions into a semiconductor photocatalyst (e.g., TiO
2
) could also
be realized using the ion implantation method. For example, Song et al. [48] prepared C/N-implanted
single-crystalline rutile TiO
2
nanowire arrays by using carbon and nitrogen ions beam to treat the as-
prepared TiO
2
nanowire arrays. After an annealing treatment, the obtained C/N-doped TiO
2
nanowire arrays exhibited a superior visible light response activity, which was attributed to the
synergistic effect between the doped C and N atoms. Their work proved that the co-doped C and N
in the lattice of TiO
2
not only greatly improves the visible light absorption capability, but also
enhances the separating and transferring property of photo-generated electron–hole pairs (Figure 5).
Figure 4.
(
a
) The degradation and (
b
) mineralization of N-doped ZnO films for 2-CP, 4-CP, and 2,4-D.
(
c
) The stability of pristine ZnO and N-doped ZnO film. (
d
) The stability and reusability of an
N-doped ZnO film for the degradation of 2-CP. Adapted with permission from Reference [
45
].
Copyright (2016) Elsevier.
3.2.2. Ion Implantation Method
The ion implantation method as a typical materials engineering strategy that can effectively
regulate the physical, chemical, and electronic properties of semiconductors, and the operation process
does not involve any other elements except the selected element, which ensures the purity of the
dopant [
46
]. Moreover, owing to the controllable parameters of ion beam implantation, such as
ion element, ion energy, ion density, uniformity of ion beam, and the doping efficiency, ion beam
implantation is a powerful approach for the heteroatom doping of semiconductors. For example,
Tang et al. [47]
fabricated an N-doped TiO
2
layer with macrospores on a titanium substrate by using the
plasma-based ion implantation method. The fabrication process involves four steps: (i) a helium plasma
was employed to generate He bubbles in the substrate, (ii) an oxygen plasma treatment and a followed
annealing in air were used to obtain rutile and anatase phases of TiO
2
, (iii) an Ar ion sputtering method
was used to exposure the He bubbles on the surface; and (iv) the pre-treated samples were doped by
nitrogen though the nitrogen beam ion implantation method. Moreover, co-doping of two or more
non-metallic anions into a semiconductor photocatalyst (e.g., TiO
2
) could also be realized using the ion
implantation method. For example, Song et al. [48] prepared C/N-implanted single-crystalline rutile
TiO
2
nanowire arrays by using carbon and nitrogen ions beam to treat the as-prepared TiO
2
nanowire
arrays. After an annealing treatment, the obtained C/N-doped TiO
2
nanowire arrays exhibited a
superior visible light response activity, which was attributed to the synergistic effect between the doped
C and N atoms. Their work proved that the co-doped C and N in the lattice of TiO
2
not only greatly
improves the visible light absorption capability, but also enhances the separating and transferring
property of photo-generated electron–hole pairs (Figure 5).
Catalysts 2019,9, 122 8 of 32
Catalysts 2018, 8, x FOR PEER REVIEW 8 of 32
Figure 5. (a) UV–vis absorption spectra of TiO2 and various doped TiO2. (b) Linear sweep
voltammograms of C/N-TiO2 and TiO2. (c) Photo-response of TiO2 and the various doped TiO2
samples under visible light. (d) Incident photon-to-current conversion efficiency spectra of TiO2 and
various doped TiO2. Adapted with permission from Reference [48]. Copyright (2018) Wiley.
3.2.3. Chemical Vapor Deposition Method
Chemical vapor deposition (CVD) is a low-cost and scalable technique, which can directly grow
a solid-phase material from a gas phase containing specific precursors. The CVD method has been
widely used for the fabrication of semiconductors and the corresponding composite of oxides,
sulfides, nitrides, and other mixed anion materials [49]. For example, Lee et al. [50] prepared TiO2
composite materials doped by C (TiOC) and N (TiON) with the titanium tetraisopropoxide (TTIP),
oxygen, and NH3 as the precursors via combing the CVD method with a fluidized bed. The results
demonstrated that the visible light photocatalysis performance of the composite TiO2 (e.g., TiON)
was significantly improved compared to the commercial TiO2 catalyst (P25, Degussa). Similarly,
Kafizas et al. [51] employed a combinatorial atmospheric pressure chemical vapor deposition
(cAPCVD) method to prepare an anatase TiO2 film with a gradating N content. The obtained TiO2
film exhibited a gradating substitutional (Ns) and interstitial (Ni) nitrogen concentration, and the
transition process from predominantly Ns-doped TiO2 to Ns/Ni mixtures, and finally to purely Ni-
doped TiO2 was precisely characterized. In addition, the UV and visible light photocatalytic activities
of the obtained N-doped TiO2 were evaluated. As a result, this work demonstrated that Ni-doped
anatase TiO2 results in a better visible light photocatalytic activity than that of predominantly Ns-
doping. They proved that the different influences of substitutional and interstitial nitrogen doping
on the photocatalytic activity of TiO2 were due to that the greater stability of electron–holes in Ni-
doped TiO2 compare with that of Ns-doped TiO2, while the propensity of the Ns-doped TiO2 for
recombination is greater. This result indicated that the doped structures is well-deigned to improve
the photocatalytic activity of a semiconductor. Additionally, the CVD could also be combined with
other materials synthesis strategy; for example, as shown in Figure 6, Youssef et al. [52] prepared the
Figure 5.
(
a
) UV–vis absorption spectra of TiO
2
and various doped TiO
2
. (
b
) Linear sweep
voltammograms of C/N-TiO
2
and TiO
2
. (
c
) Photo-response of TiO
2
and the various doped TiO
2
samples under visible light. (
d
) Incident photon-to-current conversion efficiency spectra of TiO
2
and
various doped TiO2. Adapted with permission from Reference [48]. Copyright (2018) Wiley.
3.2.3. Chemical Vapor Deposition Method
Chemical vapor deposition (CVD) is a low-cost and scalable technique, which can directly
grow a solid-phase material from a gas phase containing specific precursors. The CVD method
has been widely used for the fabrication of semiconductors and the corresponding composite of
oxides, sulfides, nitrides, and other mixed anion materials [
49
]. For example, Lee et al. [
50
] prepared
TiO
2
composite materials doped by C (TiOC) and N (TiON) with the titanium tetraisopropoxide
(TTIP), oxygen, and NH
3
as the precursors via combing the CVD method with a fluidized bed.
The results demonstrated that the visible light photocatalysis performance of the composite TiO
2
(e.g., TiON) was significantly improved compared to the commercial TiO
2
catalyst (P25, Degussa).
Similarly, Kafizas et al. [
51
] employed a combinatorial atmospheric pressure chemical vapor deposition
(cAPCVD) method to prepare an anatase TiO
2
film with a gradating N content. The obtained TiO
2
film
exhibited a gradating substitutional (N
s
) and interstitial (N
i
) nitrogen concentration, and the transition
process from predominantly N
s
-doped TiO
2
to N
s
/N
i
mixtures, and finally to purely N
i
-doped TiO
2
was precisely characterized. In addition, the UV and visible light photocatalytic activities of the
obtained N-doped TiO
2
were evaluated. As a result, this work demonstrated that N
i
-doped anatase
TiO
2
results in a better visible light photocatalytic activity than that of predominantly N
s
-doping.
They proved that the different influences of substitutional and interstitial nitrogen doping on the
photocatalytic activity of TiO
2
were due to that the greater stability of electron–holes in N
i
-doped TiO
2
compare with that of N
s
-doped TiO
2
, while the propensity of the N
s
-doped TiO
2
for recombination is
greater. This result indicated that the doped structures is well-deigned to improve the photocatalytic
activity of a semiconductor. Additionally, the CVD could also be combined with other materials
Catalysts 2019,9, 122 9 of 32
synthesis strategy; for example, as shown in Figure 6, Youssef et al. [
52
] prepared the N-doped
anatase films via a one-step low-frequency plasma enhanced chemical vapor deposition (PECVD)
process. Furthermore, they demonstrated that this method did not need the subsequential annealing
step or post-incorporation of the doping agent, and the as–prepared N-TiO
2
film exhibited good
visible-light-induced photocatalytic performance.
Catalysts 2018, 8, x FOR PEER REVIEW 9 of 32
N-doped anatase films via a one-step low-frequency plasma enhanced chemical vapor deposition
(PECVD) process. Furthermore, they demonstrated that this method did not need the subsequential
annealing step or post-incorporation of the doping agent, and the as–prepared N-TiO
2
film exhibited
good visible-light-induced photocatalytic performance.
Figure 6. Schematic view of the capacitively-coupled low frequency PECVD reactor. Adapted with
permission from Reference [52]. Copyright (2017) Elsevier.
3.2.4. Atomic Layer Deposition
The atomic layer deposition (ALD) method is a recently developed and facile strategy for the
element doping of semiconductors. Actually, ALD is a gas-phase deposition process based on
alternate surface reactions of the substrates, and the ALD method possesses several advantages, such
as good reproducibility, considerable conformality, and excellent uniformity [53]. Consequently, the
ALD method is considered as a promising strategy for the preparation of doped and composite
photocatalysts [54]. For example, Pore et al. [55] prepared a series of N-TiO
2
films via employing the
ALD processes. In this study, TiCl
4
was used as the titanium precursor and there were two ALD
cycles during the fabrication process: i) a thin layer of TiN was grown from the TiCl
4
and NH
3
; and
ii) TiO
2
was deposited on the surface of TiN layer from TiCl
4
and H
2
O, meanwhile the as-prepared
TiN layer was part-oxidized to TiO
2
, thus resulting in the TiO
2x
N
x
. Moreover, the nitrogen
concentration of the obtained TiO
2x
N
x
could be well controlled via changing the ratio of TiN and
TiO
2
deposition cycles. Similarly, Lee et al. [56] reported a facile and effective vapor-phase synthesis
strategy to prepare a conformal N-TiO
2
thin film based on the ALD process. As shown in Figure 7,
the fabrication process of the corresponding N-TiO
2
film involved four main steps: (i) pulse the TiCl
4
vapor on the surfaces of a substrate to produce a monolayer of chemisorbed TiCl
x
species; (ii) remove
the remaining unreacted TiCl
4
and corresponding HCl byproducts using nitrogen gas; (iii) NH
4
OH
as the nitrogen source was subsequently pulsed to generate a mixture of gaseous H
2
O and NH
3
,
which react with the as-prepared TiCl
x
species to obtain the N-TiO
2
; and (iv) remove the unreacted
precursors and HCl byproducts again. This cycle could be repeated to achieve the N-TiO
2
film with
the desired thickness. The as-prepared N-TiO
2
exhibited significantly enhanced photocatalytic
degradation performance for organic pollutants solely driven by the solar irradiation.
Figure 6.
Schematic view of the capacitively-coupled low frequency PECVD reactor. Adapted with
permission from Reference [52]. Copyright (2017) Elsevier.
3.2.4. Atomic Layer Deposition
The atomic layer deposition (ALD) method is a recently developed and facile strategy for the
element doping of semiconductors. Actually, ALD is a gas-phase deposition process based on
alternate surface reactions of the substrates, and the ALD method possesses several advantages,
such as good reproducibility, considerable conformality, and excellent uniformity [
53
]. Consequently,
the ALD method is considered as a promising strategy for the preparation of doped and composite
photocatalysts [
54
]. For example, Pore et al. [
55
] prepared a series of N-TiO
2
films via employing the
ALD processes. In this study, TiCl
4
was used as the titanium precursor and there were two ALD cycles
during the fabrication process: (i) a thin layer of TiN was grown from the TiCl
4
and NH
3
; and (ii) TiO
2
was deposited on the surface of TiN layer from TiCl
4
and H
2
O, meanwhile the as-prepared TiN layer
was part-oxidized to TiO
2
, thus resulting in the TiO
2x
N
x
. Moreover, the nitrogen concentration of
the obtained TiO
2x
N
x
could be well controlled via changing the ratio of TiN and TiO
2
deposition
cycles. Similarly, Lee et al. [
56
] reported a facile and effective vapor-phase synthesis strategy to
prepare a conformal N-TiO
2
thin film based on the ALD process. As shown in Figure 7, the fabrication
process of the corresponding N-TiO
2
film involved four main steps: (i) pulse the TiCl
4
vapor on the
surfaces of a substrate to produce a monolayer of chemisorbed TiCl
x
species; (ii) remove the remaining
unreacted TiCl
4
and corresponding HCl byproducts using nitrogen gas; (iii) NH
4
OH as the nitrogen
source was subsequently pulsed to generate a mixture of gaseous H
2
O and NH
3
, which react with the
as-prepared TiCl
x
species to obtain the N-TiO
2
; and (iv) remove the unreacted precursors and HCl
byproducts again. This cycle could be repeated to achieve the N-TiO
2
film with the desired thickness.
The as-prepared N-TiO
2
exhibited significantly enhanced photocatalytic degradation performance for
organic pollutants solely driven by the solar irradiation.
Catalysts 2019,9, 122 10 of 32
Catalysts 2018, 8, x FOR PEER REVIEW 10 of 32
Figure 7. Schematic illustration demonstrating the synthesis process of the N-doped TiO
2
conformal
film via the ALD method. Adapted with permission from Reference [56]. Copyright (2017) Wiley.
3.2.5. Sol-Gel and Combustion Method
Compared with the abovementioned synthesis approaches, the sol-gel and combustion method
is a facile and low-cost strategy for the preparation of various semiconductors and the corresponding
hybrid semiconductors. With the merits of simplicity and the possibility of controlling the synthesis
conditions, the sol-gel methods have been well developed and several extended sol-gel techniques
have been invented to fabricate new types of semiconductor photocatalysts. For example, Albrbar et
al. [57] reported the synthesis of a series of mesoporous anatase TiO
2
powders doped by N, and S, as
well as the N,S co-doped anatase TiO
2
powder using a non-hydrolytic sol-gel process. During the gel
synthesis process, titaniumtetrachloride and titaniumisopropoxide were used as the precursor of Ti,
dimethylsulfoxide (DMSO) was used as the precursor of S, and NH
3
was used as the precursor of N.
For the preparation of S-doped TiO
2
, the obtained gel derived from the solvent of DMSO was calcined
in air, while N and S co-doped TiO
2
was obtained when the gel was annealed in the atmosphere of
NH
3
. In addition, the pristine TiO
2
and corresponding N-doped TiO
2
was further obtained via
calcining the gel derived from the solvent of cyclohexane in air and NH
3
, respectively. In their studies,
the photocatalytic activities of the samples were evaluated via the degradation of dye C.I. Reactive
Orange16 in water under the irradiation of visible light. The obtained results showed that the N-
doped TiO
2
exhibited better visible-light photocatalytic activity compared with the pristine TiO
2
and
S-doped TiO
2
. Similarly, the sol-gel method is also versatile enough to be combined with other
materials synthesis techniques. Most recently, Rajoriya et al. [58] successfully fabricated a samarium
(Sm) and nitrogen (N) co-doped TiO
2
photocatalyst through an ultrasound-assisted sol-gel process
(Figure 8), where they found that after doping TiO
2
with Sm and N, the photocatalytic degradation
performance of the TiO
2
for 4-acetamidophenol was greatly improved owing to the significantly
improved separation efficiency of the photo-generated electron–hole pair.
Figure 7.
Schematic illustration demonstrating the synthesis process of the N-doped TiO
2
conformal
film via the ALD method. Adapted with permission from Reference [56]. Copyright (2017) Wiley.
3.2.5. Sol-Gel and Combustion Method
Compared with the abovementioned synthesis approaches, the sol-gel and combustion method is
a facile and low-cost strategy for the preparation of various semiconductors and the corresponding
hybrid semiconductors. With the merits of simplicity and the possibility of controlling the synthesis
conditions, the sol-gel methods have been well developed and several extended sol-gel techniques have
been invented to fabricate new types of semiconductor photocatalysts. For example,
Albrbar et al. [57]
reported the synthesis of a series of mesoporous anatase TiO
2
powders doped by N, and S, as well
as the N,S co-doped anatase TiO
2
powder using a non-hydrolytic sol-gel process. During the gel
synthesis process, titaniumtetrachloride and titaniumisopropoxide were used as the precursor of Ti,
dimethylsulfoxide (DMSO) was used as the precursor of S, and NH
3
was used as the precursor of N.
For the preparation of S-doped TiO
2
, the obtained gel derived from the solvent of DMSO was calcined
in air, while N and S co-doped TiO
2
was obtained when the gel was annealed in the atmosphere
of NH
3
. In addition, the pristine TiO
2
and corresponding N-doped TiO
2
was further obtained via
calcining the gel derived from the solvent of cyclohexane in air and NH
3
, respectively. In their
studies, the photocatalytic activities of the samples were evaluated via the degradation of dye C.I.
Reactive Orange16 in water under the irradiation of visible light. The obtained results showed that the
N-doped TiO
2
exhibited better visible-light photocatalytic activity compared with the pristine TiO
2
and S-doped TiO
2
. Similarly, the sol-gel method is also versatile enough to be combined with other
materials synthesis techniques. Most recently, Rajoriya et al. [
58
] successfully fabricated a samarium
(Sm) and nitrogen (N) co-doped TiO
2
photocatalyst through an ultrasound-assisted sol-gel process
(Figure 8), where they found that after doping TiO
2
with Sm and N, the photocatalytic degradation
performance of the TiO
2
for 4-acetamidophenol was greatly improved owing to the significantly
improved separation efficiency of the photo-generated electron–hole pair.
Catalysts 2019,9, 122 11 of 32
Catalysts 2018, 8, x FOR PEER REVIEW 11 of 32
Figure 8. Schematic illustrating the ultrasound assisted sol-gel synthesis process of the Sm/N doped
TiO2. Adapted with permission from Reference [58]. Copyright (2019) Elsevier.
4. Heterojunctions Construction
Besides the abovementioned heteroatoms doping strategy, constructing heterojunctions in
photocatalysts is also considered as one of the most promising approaches for improving the
photocatalysis performance of semiconductors due to its feasibility and effectiveness for the spatial
separation of electron–hole pairs. More specifically, the heterojunction is defined as the formed
interface between two semiconductors with the unequal band structure, which can form band
alignments [59,60]. In fact, there have been several types of heterojunction structures, which could be
considered as the conventional heterojunction structures, and the new generation of heterojunction
structures.
4.1. Conventional Heterojunctions
In general, the conventional heterojunctions can be classified as three types depending on the
different band gaps of the composite semiconductors, which are type I with a straddling gap, type II
with a staggered gap, and type III with a broken gap (Figure 9) [59]. As for the type I heterojunction,
the VB and CB of semiconductor A are lower and higher than the corresponding VB and CB of
semiconductor B, respectively. As a result, the photo-generated electrons and holes transfer to the CB
and VB of semiconductor B, which is negative for the separation of electron–hole pairs. Moreover,
the redox reaction of the composite semiconductors with a type I heterojunction will conduct on the
surface of semiconductor B with a lower redox potential, therefore the redox ability of the whole
photocatalyst may be suppressed. Meanwhile, in the composite semiconductor system with type II
heterojunctions, the VB and CB of semiconductor A are higher than that of semiconductor B, thus the
photo-generated electrons will migrate from the CB of semiconductor A to that of semiconductor B
with a lower reduction potential, and the corresponding holes in the VB of semiconductor B will
migrate to semiconductor A with a lower oxidation potential, thus a spatial separation of electron–
hole pairs will be completed. However, the band gap of the two semiconductors will not overlap in
the type III heterojunctions, and as a result, there is no transmission or separation of electrons and
holes between semiconductor A and semiconductor B. Consequently, the type II heterojunction is the
most effective structure for improving the photocatalysis performance of semiconductors, and has
received a great deal of research attention.
Figure 8.
Schematic illustrating the ultrasound assisted sol-gel synthesis process of the Sm/N doped
TiO2. Adapted with permission from Reference [58]. Copyright (2019) Elsevier.
4. Heterojunctions Construction
Besides the abovementioned heteroatoms doping strategy, constructing heterojunctions in
photocatalysts is also considered as one of the most promising approaches for improving the
photocatalysis performance of semiconductors due to its feasibility and effectiveness for the spatial
separation of electron–hole pairs. More specifically, the heterojunction is defined as the formed interface
between two semiconductors with the unequal band structure, which can form band alignments [
59
,
60
].
In fact, there have been several types of heterojunction structures, which could be considered as the
conventional heterojunction structures, and the new generation of heterojunction structures.
4.1. Conventional Heterojunctions
In general, the conventional heterojunctions can be classified as three types depending on the
different band gaps of the composite semiconductors, which are type I with a straddling gap, type II
with a staggered gap, and type III with a broken gap (Figure 9) [
59
]. As for the type I heterojunction,
the VB and CB of semiconductor A are lower and higher than the corresponding VB and CB of
semiconductor B, respectively. As a result, the photo-generated electrons and holes transfer to the CB
and VB of semiconductor B, which is negative for the separation of electron–hole pairs. Moreover,
the redox reaction of the composite semiconductors with a type I heterojunction will conduct on the
surface of semiconductor B with a lower redox potential, therefore the redox ability of the whole
photocatalyst may be suppressed. Meanwhile, in the composite semiconductor system with type II
heterojunctions, the VB and CB of semiconductor A are higher than that of semiconductor B, thus the
photo-generated electrons will migrate from the CB of semiconductor A to that of semiconductor
B with a lower reduction potential, and the corresponding holes in the VB of semiconductor B will
migrate to semiconductor A with a lower oxidation potential, thus a spatial separation of electron–hole
pairs will be completed. However, the band gap of the two semiconductors will not overlap in the
type III heterojunctions, and as a result, there is no transmission or separation of electrons and holes
between semiconductor A and semiconductor B. Consequently, the type II heterojunction is the most
effective structure for improving the photocatalysis performance of semiconductors, and has received
a great deal of research attention.
Catalysts 2019,9, 122 12 of 32
Catalysts 2018, 8, x FOR PEER REVIEW 12 of 32
Figure 9. Schematic illustrating the photocatalysis mechanism of the three different types of
heterojunction photocatalysts: (a) type-I, (b) type-II, and (c) type-III. Adapted with permission from
Reference [59]. Copyright (2017) Wiley.
Up to now, several type-II heterojunction photocatalysts have been developed by creating two
different phases in the same semiconductor, or directly compositing different semiconductors
together [60,61]. For example, Yu et al. [62] once created the anatase-brookite dual-phase in a TiO
2
photocatalyst to form a type-II heterojunction via hydrolyzing the titanium tetraisopropoxide in
water and an ethanol-H
2
O mixture solution. They found that the co-presence of brookite and anatase
phases in the TiO
2
significantly enhanced the photocatalysis performance. After that, Uddin et al. [63]
successfully fabricated the mesoporous SnO
2
-ZnO heterojunction photocatalysts using a two-step
synthesis strategy. Furthermore, they had carefully examined the band alignment, the results showed
that the obtained SnO
2
-ZnO heterojunction photocatalyst possessed a type-II band alignment and
exhibited higher photocatalytic activity for the degradation of methyl blue in water than that of the
individual SnO
2
and ZnO nanocatalysts (Figure 10). Apart from the inorganic semiconductors,
organic semiconductors could also be incorporated with the semiconductors to form the type-II
heterojunction. For example, Shirmardi et al. [64] used polyaniline (PANI) as the organic
semiconductor combined with ZnSe nanoparticles via a simple and cost-effective co-precipitation
method in the ambient conditions. The obtained ZnSe/PANI nanocomposites exhibited obvious
enhancement in the photocatalytic performance compared to that of the pristine ZnSe nanoparticles.
Figure 9.
Schematic illustrating the photocatalysis mechanism of the three different types of
heterojunction photocatalysts: (
a
) type-I, (
b
) type-II, and (
c
) type-III. Adapted with permission from
Reference [59]. Copyright (2017) Wiley.
Up to now, several type-II heterojunction photocatalysts have been developed by creating
two different phases in the same semiconductor, or directly compositing different semiconductors
together [
60
,
61
]. For example, Yu et al. [
62
] once created the anatase-brookite dual-phase in a TiO
2
photocatalyst to form a type-II heterojunction via hydrolyzing the titanium tetraisopropoxide in water
and an ethanol-H
2
O mixture solution. They found that the co-presence of brookite and anatase
phases in the TiO
2
significantly enhanced the photocatalysis performance. After that, Uddin et al. [
63
]
successfully fabricated the mesoporous SnO
2
-ZnO heterojunction photocatalysts using a two-step
synthesis strategy. Furthermore, they had carefully examined the band alignment, the results showed
that the obtained SnO
2
-ZnO heterojunction photocatalyst possessed a type-II band alignment and
exhibited higher photocatalytic activity for the degradation of methyl blue in water than that of the
individual SnO
2
and ZnO nanocatalysts (Figure 10). Apart from the inorganic semiconductors, organic
semiconductors could also be incorporated with the semiconductors to form the type-II heterojunction.
For example, Shirmardi et al. [
64
] used polyaniline (PANI) as the organic semiconductor combined with
ZnSe nanoparticles via a simple and cost-effective co-precipitation method in the ambient conditions.
The obtained ZnSe/PANI nanocomposites exhibited obvious enhancement in the photocatalytic
performance compared to that of the pristine ZnSe nanoparticles.
Catalysts 2019,9, 122 13 of 32
Catalysts 2018, 8, x FOR PEER REVIEW 13 of 32
Figure 10. (a) Nanostructures of SnO
2
ZnO composite photocatalysts. (b) The corresponding
photocatalytic performances of SnO
2
–ZnO (red line with square dots), SnO
2
(green line with triangle
dots), and ZnO (blue line with circle dots). Adapted with permission from Reference [63]. Copyright
(2012) American Chemical Society.
4.2. New Generation of Heterojunctions
Although the conventional type-II heterojunctions are capable of spatially separating the photo-
generated electron–hole pairs, there remain several critical limitations, such as the relatively weak
redox ability due to the lower reduction and oxidation potentials, and the suppressed migration of
electrons and holes due to the electrostatic repulsion [59]. Recently, in order to overcome the
abovementioned limitations, a new generation of heterojunctions have been developed, including the
p-n heterojunctions, the surface heterojunctions, the Z-scheme heterojunctions, and the
semiconductor/carbon heterojunctions. Here we will give a brief introduction of each kind of these
newly developed heterojunctions.
4.2.1. p–n Heterojunctions
The p-n heterojunctions could be obtained by incorporating a p-type semiconductor with an n-
type semiconductor, and it has been proved that the formation of p-n heterojunctions are effective
for improving the photocatalytic performance of composite catalysts [65,66]. In general, before the
irradiation of light, there is an internal electric field in the region closed to the p-n interface due to the
electron–hole diffusion tendency of the composite semiconductors system with unequal Fermi levels
[59,67]. Alternatively, when the composite semiconductors are irradiated by a light, and the energy
state of the photon is beyond the band gaps of both p-type and n-type semiconductors, electron–hole
pairs will be generated in the corresponding semiconductors. However, due to the presence of an
internal electric field, the photo-generated electrons and holes will transfer to the CB of the n-type
semiconductor and the VB of p-type semiconductor, respectively. Furthermore, it has been proved
that this spatial separation of the photo-generated electron–hole pairs is much more efficient
compared with that of conventional type-II heterojunction because of the synergy of the internal
electric field and band alignment [59,68]. As a result, a variety of composite semiconductors with the
p-n heterojunctions have been created for the application of photocatalysis. For example, Wen et al.
[69] reported the fabrication of a BiOI/CeO
2
p-n junction using a facile in situ chemical bath method.
The result demonstrated that the BiOI/CeO
2
composite with a mole ratio of 1:1 exhibited a superior
photocatalytic performance for the decomposition of bisphenol A (BPA) and methylene orange under
visible light irradiation. Most recently, as shown in Figure 11, our group reported a facile method for
the preparation of SnS
2
/MoO
3
hollow nanotubes based on the hydrothermal method [70]. The
obtained SnS
2
/MoO
3
hollow nanotubes exhibit a typical p-n heterojunction structure, and a
synergistic effect between MoO
3
and SnS
2
was proven to yield an optimal hydrogen peroxide
production performance.
Figure 10.
(
a
) Nanostructures of SnO
2
ZnO composite photocatalysts. (
b
) The corresponding
photocatalytic performances of SnO
2
–ZnO (red line with square dots), SnO
2
(green line with
triangle dots), and ZnO (blue line with circle dots). Adapted with permission from Reference [
63
].
Copyright (2012) American Chemical Society.
4.2. New Generation of Heterojunctions
Although the conventional type-II heterojunctions are capable of spatially separating the
photo-generated electron–hole pairs, there remain several critical limitations, such as the relatively
weak redox ability due to the lower reduction and oxidation potentials, and the suppressed
migration of electrons and holes due to the electrostatic repulsion [
59
]. Recently, in order to
overcome the abovementioned limitations, a new generation of heterojunctions have been developed,
including the p-n heterojunctions, the surface heterojunctions, the Z-scheme heterojunctions, and the
semiconductor/carbon heterojunctions. Here we will give a brief introduction of each kind of these
newly developed heterojunctions.
4.2.1. p–n Heterojunctions
The p-n heterojunctions could be obtained by incorporating a p-type semiconductor with an
n-type semiconductor, and it has been proved that the formation of p-n heterojunctions are effective
for improving the photocatalytic performance of composite catalysts [
65
,
66
]. In general, before
the irradiation of light, there is an internal electric field in the region closed to the p-n interface
due to the electron–hole diffusion tendency of the composite semiconductors system with unequal
Fermi levels [59,67].
Alternatively, when the composite semiconductors are irradiated by a light,
and the energy state of the photon is beyond the band gaps of both p-type and n-type semiconductors,
electron–hole pairs will be generated in the corresponding semiconductors. However, due to the
presence of an internal electric field, the photo-generated electrons and holes will transfer to the CB of
the n-type semiconductor and the VB of p-type semiconductor, respectively. Furthermore, it has been
proved that this spatial separation of the photo-generated electron–hole pairs is much more efficient
compared with that of conventional type-II heterojunction because of the synergy of the internal electric
field and band alignment [
59
,
68
]. As a result, a variety of composite semiconductors with the p-n
heterojunctions have been created for the application of photocatalysis. For example,
Wen et al. [69]
reported the fabrication of a BiOI/CeO
2
p-n junction using a facile in situ chemical bath method.
The result demonstrated that the BiOI/CeO
2
composite with a mole ratio of 1:1 exhibited a superior
photocatalytic performance for the decomposition of bisphenol A (BPA) and methylene orange under
visible light irradiation. Most recently, as shown in Figure 11, our group reported a facile method for the
preparation of SnS
2
/MoO
3
hollow nanotubes based on the hydrothermal method [
70
]. The obtained
SnS
2
/MoO
3
hollow nanotubes exhibit a typical p-n heterojunction structure, and a synergistic effect
between MoO
3
and SnS
2
was proven to yield an optimal hydrogen peroxide production performance.
Catalysts 2019,9, 122 14 of 32
Catalysts 2018, 8, x FOR PEER REVIEW 14 of 32
Figure 11. Schematic illustration of the SnS
2
/MoO
3
hollow nanotubes and its photocatalysis
mechanism with a two-channel pathway. Adapted with permission from Reference [70]. Copyright
(2018) Royal Society of Chemistry.
4.2.2. Surface Heterojunctions
As reported before, a surface heterojunction can be created between two crystal facets of a single
semiconductor [59,71]. For example, Yu et al. [72] proved that the formation of a heterojunction
between the (001) and (101) facets in TiO
2
contribute significantly toward the enhancement of
photocatalytic activity. This method enables the construction of a heterojunction on the surface of a
single semiconductor, which is less costly because only one semiconductor is used. They also
demonstrate that there is an optimal ratio for the (001) and (101) facets in the anatase TiO
2
for the
improvement of its photocatalysis performance. Subsequently, Gao et al. [73] found that the surface
heterojunction of TiO
2
could be self-adjusted, and its photocatalytic activity could be significantly
improved via combining a proper surface heterojunction with the Schottky junction. Apart from the
TiO
2
, Bi-based semiconductors could also be employed for the design of photocatalysts with surface
heterojunctions. Most recently, as shown in Figure 12, Lu et al. [74] synthesized a tetragonal BiOI
photocatalyst by regulating the amount of water in the hydrolysis process at room temperature. The
as-prepared photocatalyst possessed a typical surface heterojunction structure between (001) facets
and (110) facets, and exhibited a promoted photocatalytic performance for the degradation of organic
contaminants in water under visible light.
Figure 11.
Schematic illustration of the SnS
2
/MoO
3
hollow nanotubes and its photocatalysis
mechanism with a two-channel pathway. Adapted with permission from Reference [
70
].
Copyright (2018) Royal Society of Chemistry.
4.2.2. Surface Heterojunctions
As reported before, a surface heterojunction can be created between two crystal facets of a single
semiconductor [
59
,
71
]. For example, Yu et al. [
72
] proved that the formation of a heterojunction
between the (001) and (101) facets in TiO
2
contribute significantly toward the enhancement of
photocatalytic activity. This method enables the construction of a heterojunction on the surface
of a single semiconductor, which is less costly because only one semiconductor is used. They also
demonstrate that there is an optimal ratio for the (001) and (101) facets in the anatase TiO
2
for the
improvement of its photocatalysis performance. Subsequently, Gao et al. [
73
] found that the surface
heterojunction of TiO
2
could be self-adjusted, and its photocatalytic activity could be significantly
improved via combining a proper surface heterojunction with the Schottky junction. Apart from
the TiO
2
, Bi-based semiconductors could also be employed for the design of photocatalysts with
surface heterojunctions. Most recently, as shown in Figure 12, Lu et al. [
74
] synthesized a tetragonal
BiOI photocatalyst by regulating the amount of water in the hydrolysis process at room temperature.
The as-prepared photocatalyst possessed a typical surface heterojunction structure between (001) facets
and (110) facets, and exhibited a promoted photocatalytic performance for the degradation of organic
contaminants in water under visible light.
Catalysts 2019,9, 122 15 of 32
Catalysts 2018, 8, x FOR PEER REVIEW 15 of 32
Figure 12. (a,b) Schematic illustrating the growth of TiO2 nanosheets at different conditions. (c) UV-
vis images of the related samples. (d) Photocatalytic degradation efficiency of different catalysts for
methyl orange. (e) Schematic demonstrating the migration of electrons and holes in the surface
heterojunction. Adapted with permission from Reference [74]. Copyright (2018) Elsevier.
4.2.3. Z-Scheme Heterojunctions
Z-scheme heterojunctions were constructed to overcome the limitation of the lower redox
potential of the heterojunction systems. [59,75] In general, the Z-scheme heterojunction is composed
of two different semiconductors and an electron acceptor/donor pair. During the photocatalysis
process, the photo-generated electrons/holes will transfer from the matrix semiconductor to the
coupled semiconductor through the electron acceptor/donor pair or an electron mediator. As a result,
the electrons/holes will accumulate on different semiconductors with higher redox potentiasl, and an
effective spatial separation of electron–hole pairs is also realized. Up to now, the Z-scheme
heterojunctions have been well developed, and various photocatalysts with well-designed Z-scheme
heterojunctions have been invented for the wastewater treatment. [75] For example, Wu et al. [76]
reported the fabrication of the Ag2CO3/Ag/AgNCO composite photocatalyst via a simple in situ ion
exchange method. The obtained composite photocatalyst possessed the Z-scheme heterojunction and
exhibited a highly efficient degradation ratio of rhodamine B and the reduction of Cr (VI) under the
driving of visible light. They proved that the significantly enhanced photocatalytic activity could be
attributed to the low resistance for the interfacial charge transfer and desirable absorption capability.
Recently, considering the relative high cost of the common used electron mediators (e.g., Pt, Ag, and
Au), a new generation of Z-heterojunctions without the electron mediators have been invented for
wastewater treatment, which is named as the direct Z-scheme system [59]. For example, Lu et al. [77]
synthesized a CuInS2/Bi2WO6 composite catalyst with a direct Z-scheme heterojunction via the in situ
Figure 12.
(
a
,
b
) Schematic illustrating the growth of TiO
2
nanosheets at different conditions. (
c
) UV-vis
images of the related samples. (
d
) Photocatalytic degradation efficiency of different catalysts for methyl
orange. (
e
) Schematic demonstrating the migration of electrons and holes in the surface heterojunction.
Adapted with permission from Reference [74]. Copyright (2018) Elsevier.
4.2.3. Z-Scheme Heterojunctions
Z-scheme heterojunctions were constructed to overcome the limitation of the lower redox
potential of the heterojunction systems. [
59
,
75
] In general, the Z-scheme heterojunction is composed
of two different semiconductors and an electron acceptor/donor pair. During the photocatalysis
process, the photo-generated electrons/holes will transfer from the matrix semiconductor to the
coupled semiconductor through the electron acceptor/donor pair or an electron mediator. As a
result, the electrons/holes will accumulate on different semiconductors with higher redox potentiasl,
and an effective spatial separation of electron–hole pairs is also realized. Up to now, the Z-scheme
heterojunctions have been well developed, and various photocatalysts with well-designed Z-scheme
heterojunctions have been invented for the wastewater treatment. [
75
] For example, Wu et al. [
76
]
reported the fabrication of the Ag
2
CO
3
/Ag/AgNCO composite photocatalyst via a simple in situ ion
exchange method. The obtained composite photocatalyst possessed the Z-scheme heterojunction and
exhibited a highly efficient degradation ratio of rhodamine B and the reduction of Cr (VI) under the
driving of visible light. They proved that the significantly enhanced photocatalytic activity could be
attributed to the low resistance for the interfacial charge transfer and desirable absorption capability.
Recently, considering the relative high cost of the common used electron mediators (e.g., Pt, Ag,
and Au), a new generation of Z-heterojunctions without the electron mediators have been invented for
Catalysts 2019,9, 122 16 of 32
wastewater treatment, which is named as the direct Z-scheme system [
59
]. For example, Lu et al. [
77
]
synthesized a CuInS
2
/Bi
2
WO
6
composite catalyst with a direct Z-scheme heterojunction via the in
situ hydrothermal growth of Bi
2
WO
6
on the surface of CuInS
2
networks. The obtained composite
photocatalysts with an optimal Z-scheme exhibited a superior visible light degradation performance
of the tetracycline hydrochloride in water than that of the pristine CuInS
2
and Bi
2
WO
6
. The improved
photocatalytic activity was attributed to the formed intimate interface contact, which ensured a good
interfacial charge transfer ability (Figure 13).
Catalysts 2018, 8, x FOR PEER REVIEW 16 of 32
hydrothermal growth of Bi2WO6 on the surface of CuInS2 networks. The obtained composite
photocatalysts with an optimal Z-scheme exhibited a superior visible light degradation performance
of the tetracycline hydrochloride in water than that of the pristine CuInS2 and Bi2WO6. The improved
photocatalytic activity was attributed to the formed intimate interface contact, which ensured a good
interfacial charge transfer ability (Figure 13).
Figure 13. Schematic illustrating the interfacial electron transfer process and possible photocatalytic
mechanism of CuInS2/Bi2WO6 with the Z-scheme heterojunction. Adapted with permission from
Reference [77]. Copyright (2019) Elsevier.
4.2.4. Semiconductor/Carbon Heterojunctions
Carbonaceous nanomaterials have been widely employed for the design of novel photocatalysts
due to their advantages of high surface area, good conductivity, and chemical stability. In general,
the most commonly used carbonaceous materials for combining with semiconductors involves the
carbon dots (CDs), carbon nanotubes (CNTs), and graphene [78].
The CDs as typical nanocarbon materials have been widely used to enhance the photocatalytic
activity of semiconductors owing to their intriguing optical and electronic properties, low chemical
toxicity, adjustable photoluminescence, and the distinct quantum effect [79]. For example, Long et al.
[80] used carbon dots (CDs) to couple with the BiOI with highly exposed (001) facets to form a
composition of CDs/BiOI. Furthermore, the obtained CDs/BiOI composite exhibited a greatly
improved photocatalytic activity for the degradation of organic dyes in water. It has been proved that
the incorporated CDs in the semiconductor formed a CDs/BiOI heterojunction, which was able to
construct numerous electron surface trap sites and was beneficial for enhancing the visible light
absorption range as well as the charge separation. Recently, Zhao et al. [81] reported the fabrication
of carbon quantum dots (CQDs)/TiO2 nanotubes (TNTs) composite via an improved hydrothermal
method. The CQDs were incorporated on the surface of the TNTs, and played a vital role in
improving the visible light photocatalytic performance of the composite. As shown in Figure 14, they
demonstrated that there were three advantages for the formation of CQDs/TiO2: i) the CQDs could
effectively trap the photo-generated electrons from TNTs and suppress the recombination of electron-
hole pairs, ii) the up-conversion photoluminescence property of CQDs could further improve the
visible light utilization efficiency of CQDs/TNTs, and iii) the hetero-structure formed between the
CQDs and the TNTs could prolong the life of the photogenerated electron and hole pairs.
Figure 13.
Schematic illustrating the interfacial electron transfer process and possible photocatalytic
mechanism of CuInS
2
/Bi
2
WO
6
with the Z-scheme heterojunction. Adapted with permission from
Reference [77]. Copyright (2019) Elsevier.
4.2.4. Semiconductor/Carbon Heterojunctions
Carbonaceous nanomaterials have been widely employed for the design of novel photocatalysts
due to their advantages of high surface area, good conductivity, and chemical stability. In general,
the most commonly used carbonaceous materials for combining with semiconductors involves the
carbon dots (CDs), carbon nanotubes (CNTs), and graphene [78].
The CDs as typical nanocarbon materials have been widely used to enhance the photocatalytic
activity of semiconductors owing to their intriguing optical and electronic properties, low chemical
toxicity, adjustable photoluminescence, and the distinct quantum effect [
79
]. For example,
Long et al. [80]
used carbon dots (CDs) to couple with the BiOI with highly exposed (001) facets
to form a composition of CDs/BiOI. Furthermore, the obtained CDs/BiOI composite exhibited a
greatly improved photocatalytic activity for the degradation of organic dyes in water. It has been
proved that the incorporated CDs in the semiconductor formed a CDs/BiOI heterojunction, which was
able to construct numerous electron surface trap sites and was beneficial for enhancing the visible
light absorption range as well as the charge separation. Recently, Zhao et al. [
81
] reported the
fabrication of carbon quantum dots (CQDs)/TiO
2
nanotubes (TNTs) composite via an improved
hydrothermal method. The CQDs were incorporated on the surface of the TNTs, and played a vital
role in improving the visible light photocatalytic performance of the composite. As shown in Figure 14,
they demonstrated that there were three advantages for the formation of CQDs/TiO
2
: (i) the CQDs
could effectively trap the photo-generated electrons from TNTs and suppress the recombination of
electron-hole pairs, (ii) the up-conversion photoluminescence property of CQDs could further improve
the visible light utilization efficiency of CQDs/TNTs, and (iii) the hetero-structure formed between the
CQDs and the TNTs could prolong the life of the photogenerated electron and hole pairs.
Catalysts 2019,9, 122 17 of 32
Catalysts 2018, 8, x FOR PEER REVIEW 17 of 32
Figure 14. Schematic illustration indicating the photocatalysis mechanism of the CQDs/TNTs
photocatalyst. Adapted with permission from Reference [81]. Copyright (2018) Elsevier.
CNTs are typical nanocarbon materials with highly sp
2
-ordered structures, and thus exhibit an
excellent metallic conductivity, which could form a Schottky barrier junction between the CNT and
semiconductors; as reported before, the Schottky barrier junction could effectively increase the
recombination time of electron–hole pairs [78,82]. Moreover, CNTs could accept electrons in the
composite system with semiconductors due to its large electron-storage capacity, which is beneficial
for retarding or hindering the electron–hole recombination. As a result, a variety of semiconductor-
CNT composite photocatalysts have been developed. For example, Miribangul et al. [83] prepared a
TiO
2
/CNT composite via a simple hydrothermal method. The influence of the CNT concentration in
the TiO
2
-CNT composites on their photocatalytic activity was investigated and the 0.3 wt% CNT
content in TiO
2
/CNT composite could offer the highest photocatalytic degradation of Sudan (I) in
UV–vis light. Apart from the TiO
2
, some of the other semiconductors can also be employed to
composite with CNT, such as the CNT/LaVO
4
composite photocatalyst developed by Xu et al. [84].
As shown in Figure 15, with the presence of CNT, the photocatalytic activity of a CNT/LaVO
4
composite was greatly improved due to the synergistic effect between CNT and LaVO
4
, therefore the
corresponding photocatalytic degradation rate of CNT/LaVO
4
composite for organic contaminant is
2 times that of pure LaVO
4
.
Figure 14.
Schematic illustration indicating the photocatalysis mechanism of the CQDs/TNTs
photocatalyst. Adapted with permission from Reference [81]. Copyright (2018) Elsevier.
CNTs are typical nanocarbon materials with highly sp
2
-ordered structures, and thus exhibit
an excellent metallic conductivity, which could form a Schottky barrier junction between the CNT
and semiconductors; as reported before, the Schottky barrier junction could effectively increase the
recombination time of electron–hole pairs [
78
,
82
]. Moreover, CNTs could accept electrons in the composite
system with semiconductors due to its large electron-storage capacity, which is beneficial for retarding
or hindering the electron–hole recombination. As a result, a variety of semiconductor-CNT composite
photocatalysts have been developed. For example, Miribangul et al. [
83
] prepared a TiO
2
/CNT composite
via a simple hydrothermal method. The influence of the CNT concentration in the TiO
2
-CNT composites
on their photocatalytic activity was investigated and the 0.3 wt % CNT content in TiO
2
/CNT composite
could offer the highest photocatalytic degradation of Sudan (I) in UV–vis light. Apart from the TiO
2
,
some of the other semiconductors can also be employed to composite with CNT, such as the CNT/LaVO
4
composite photocatalyst developed by Xu et al. [
84
]. As shown in Figure 15, with the presence of CNT,
the photocatalytic activity of a CNT/LaVO
4
composite was greatly improved due to the synergistic effect
between CNT and LaVO
4
, therefore the corresponding photocatalytic degradation rate of CNT/LaVO
4
composite for organic contaminant is 2 times that of pure LaVO4.
Catalysts 2018, 8, x FOR PEER REVIEW 17 of 32
Figure 14. Schematic illustration indicating the photocatalysis mechanism of the CQDs/TNTs
photocatalyst. Adapted with permission from Reference [81]. Copyright (2018) Elsevier.
CNTs are typical nanocarbon materials with highly sp
2
-ordered structures, and thus exhibit an
excellent metallic conductivity, which could form a Schottky barrier junction between the CNT and
semiconductors; as reported before, the Schottky barrier junction could effectively increase the
recombination time of electron–hole pairs [78,82]. Moreover, CNTs could accept electrons in the
composite system with semiconductors due to its large electron-storage capacity, which is beneficial
for retarding or hindering the electron–hole recombination. As a result, a variety of semiconductor-
CNT composite photocatalysts have been developed. For example, Miribangul et al. [83] prepared a
TiO
2
/CNT composite via a simple hydrothermal method. The influence of the CNT concentration in
the TiO
2
-CNT composites on their photocatalytic activity was investigated and the 0.3 wt% CNT
content in TiO
2
/CNT composite could offer the highest photocatalytic degradation of Sudan (I) in
UV–vis light. Apart from the TiO
2
, some of the other semiconductors can also be employed to
composite with CNT, such as the CNT/LaVO
4
composite photocatalyst developed by Xu et al. [84].
As shown in Figure 15, with the presence of CNT, the photocatalytic activity of a CNT/LaVO
4
composite was greatly improved due to the synergistic effect between CNT and LaVO
4
, therefore the
corresponding photocatalytic degradation rate of CNT/LaVO
4
composite for organic contaminant is
2 times that of pure LaVO
4
.
Figure 15.
Schematic illustration indicating the reaction mechanism of the photocatalytic procedure of
CNT/LaVO
4
composite catalyst. Adapted with permission from Reference [
84
]. Copyright (2019) Elsevier.
Catalysts 2019,9, 122 18 of 32
Recently, graphene as a newly developed nanocarbon material has attracted numerous research
attention in the area of photocatalysts due to its extraordinary physical properties, including superior
charge transport ability, unique optical properties, high thermal conductivity, large specific surface area,
and good mechanical strength [
78
,
85
,
86
]. Up to now, a myriad of attempts has been carried out to couple
the graphene with various semiconductors to further improve their photocatalytic activity. According to
the previous reports, the first graphene composite semiconductor for photocatalysis was prepared
by Zhang and co-workers [
87
]. They fabricated a TiO
2
(P25)-graphene composite with a chemically
bonding structure via a one-step hydrothermal reaction. As reported, there are three contributions of
the graphene for the photocatalytic activity: (i) enhancing the adsorption capacity of pollutants, (ii)
extending light absorption range, and (iii) improving charge transportation and separation efficiency.
As a result, the photodegradation of the obtained TiO
2
(P25)-graphene composite for methylene blue was
significantly improved, and was superior to that of the bare P25 and the commonly reported P25-CNTs
composite. After that, numerous photocatalysts based on the composite graphene-semiconductors have
been invented for the treatment of wastewater. Most recently, in order to overcome the limitation of
the poor homogenous dispersion of graphene, as shown in Figure 16, Isari et al. [
88
] created a ternary
nanocomposite catalyst (Fe-doped TiO
2
/rGO) derived from Fe-doped TiO
2
and reduced graphene oxide
via a simple sol-gel method. They proved that the band gap of Fe-TiO
2
/rGO could be significantly
decreased compared with that of the pristine TiO
2
, and the obtained Fe-doped TiO
2
/rGO exhibited an
effective decontamination performance for rhodamine B in water.
Catalysts 2018, 8, x FOR PEER REVIEW 18 of 32
Figure 15. Schematic illustration indicating the reaction mechanism of the photocatalytic procedure
of CNT/LaVO
4
composite catalyst. Adapted with permission from Reference [84]. Copyright (2019)
Elsevier.
Recently, graphene as a newly developed nanocarbon material has attracted numerous research
attention in the area of photocatalysts due to its extraordinary physical properties, including superior
charge transport ability, unique optical properties, high thermal conductivity, large specific surface
area, and good mechanical strength [78,85,86]. Up to now, a myriad of attempts has been carried out
to couple the graphene with various semiconductors to further improve their photocatalytic activity.
According to the previous reports, the first graphene composite semiconductor for photocatalysis
was prepared by Zhang and co-workers [87]. They fabricated a TiO
2
(P25)-graphene composite with
a chemically bonding structure via a one-step hydrothermal reaction. As reported, there are three
contributions of the graphene for the photocatalytic activity: i) enhancing the adsorption capacity of
pollutants, ii) extending light absorption range, and iii) improving charge transportation and
separation efficiency. As a result, the photodegradation of the obtained TiO
2
(P25)-graphene
composite for methylene blue was significantly improved, and was superior to that of the bare P25
and the commonly reported P25-CNTs composite. After that, numerous photocatalysts based on the
composite graphene-semiconductors have been invented for the treatment of wastewater. Most
recently, in order to overcome the limitation of the poor homogenous dispersion of graphene, as
shown in Figure 16, Isari et al. [88] created a ternary nanocomposite catalyst (Fe-doped TiO
2
/rGO)
derived from Fe-doped TiO
2
and reduced graphene oxide via a simple sol-gel method. They proved
that the band gap of Fe-TiO
2
/rGO could be significantly decreased compared with that of the pristine
TiO
2
, and the obtained Fe-doped TiO
2
/rGO exhibited an effective decontamination performance for
rhodamine B in water.
Figure 16. Schematic illustration demonstrating the photocatalytic mechanism of Fe-TiO
2
/rGO
catalyst for contaminant under solar irradiation. Adapted with permission from Reference [88].
Copyright (2018) Elsevier.
5. Morphology Regulation of the Composite Photocatalysts
Apart from the intrinsic characteristics of semiconductors, the corresponding catalysts with
different morphologies may result in different photocatalytic activities and different application
processes [89]. Recently, morphology modification of the photocatalysts have attracted more and
more attention owing to further improvements in their application performance, not only for the
photocatalytic performance but also for the application techniques. With this in mind, in this section
Figure 16.
Schematic illustration demonstrating the photocatalytic mechanism of Fe-TiO
2
/rGO
catalyst for contaminant under solar irradiation. Adapted with permission from Reference [
88
].
Copyright (2018) Elsevier.
5. Morphology Regulation of the Composite Photocatalysts
Apart from the intrinsic characteristics of semiconductors, the corresponding catalysts with
different morphologies may result in different photocatalytic activities and different application
processes [
89
]. Recently, morphology modification of the photocatalysts have attracted more and
more attention owing to further improvements in their application performance, not only for the
photocatalytic performance but also for the application techniques. With this in mind, in this section
we will briefly summarize the recent achievements of the composite photocatalysts with different
morphologies of 0D, 1D, 2D, and 3D materials.
Catalysts 2019,9, 122 19 of 32
5.1. Nanoparticles (0D)
Generally, the 0D materials are characterized as spherically shaped with nano-scaled dimensions.
Nanoparticles as a typical 0D material have been widely used in the area of photocatalysis with the
merits of large surface area, simple synthesis methods, and easy to be functionalized [
90
]. Up to now,
several synthesis approaches have been invented, among which the sol-gel method, hydrothermal
method, and solvothermal method could be the most-used techniques for the fabrication of 0D
composite photocatalysts.
5.1.1. Sol-Gel Method
The sol-gel process is a commonly used and effective strategy for the preparation of various
inorganic materials, especially for the metal oxides based on the corresponding precursors, and it
has several merits including being low cost, processed at low-temperature, and the fine control
of the product’s chemical composition. Therefore, the sol-gel process is one of the most-used
techniques for the preparation of composite semiconductor photocatalysts [
58
]. For example,
Vaiano et al. [91]
immobilized the N-doped TiO
2
nanoparticles (NPs) on glass spheres via the sol-gel
method. Through regulating the synthesis conditions, and employing the Triton X-100 as the surface
active agent, the obtained N-doped TiO
2
NPs/glass spheres exhibited a good photocatalytic activity
for methylene blue and eriochrome black-T in water under UV and visible light irradiation. Moreover,
the composite catalyst was easy to be separated from the reaction mixture with a good durability.
Recently,
Chen et al. [92]
prepared a Ni-Cu-Zn ferrite@SiO
2
@TiO
2
composite via a simple sol-gel
method. With the immobilization of Ag and magnetic ferrite, the composite photocatalysts exhibited
comparatively good photodegradation performance for the methylene blue under a visible light source
with lower power. Moreover, the composite catalysts can be easily separated using a magnet and can
be reused well without significant loss of photocatalytic activity (Figure 17).
Catalysts 2018, 8, x FOR PEER REVIEW 19 of 32
we will briefly summarize the recent achievements of the composite photocatalysts with different
morphologies of 0D, 1D, 2D, and 3D materials.
5.1. Nanoparticles (0D)
Generally, the 0D materials are characterized as spherically shaped with nano-scaled
dimensions. Nanoparticles as a typical 0D material have been widely used in the area of
photocatalysis with the merits of large surface area, simple synthesis methods, and easy to be
functionalized [90]. Up to now, several synthesis approaches have been invented, among which the
sol-gel method, hydrothermal method, and solvothermal method could be the most-used techniques
for the fabrication of 0D composite photocatalysts.
5.1.1. Sol-Gel Method
The sol-gel process is a commonly used and effective strategy for the preparation of various
inorganic materials, especially for the metal oxides based on the corresponding precursors, and it has
several merits including being low cost, processed at low-temperature, and the fine control of the
product’s chemical composition. Therefore, the sol-gel process is one of the most-used techniques for
the preparation of composite semiconductor photocatalysts [58]. For example, Vaiano et al. [91]
immobilized the N-doped TiO2 nanoparticles (NPs) on glass spheres via the sol-gel method. Through
regulating the synthesis conditions, and employing the Triton X-100 as the surface active agent, the
obtained N-doped TiO2 NPs/glass spheres exhibited a good photocatalytic activity for methylene
blue and eriochrome black-T in water under UV and visible light irradiation. Moreover, the
composite catalyst was easy to be separated from the reaction mixture with a good durability.
Recently, Chen et al. [92] prepared a Ni-Cu-Zn ferrite@SiO2@TiO2 composite via a simple sol-gel
method. With the immobilization of Ag and magnetic ferrite, the composite photocatalysts exhibited
comparatively good photodegradation performance for the methylene blue under a visible light
source with lower power. Moreover, the composite catalysts can be easily separated using a magnet
and can be reused well without significant loss of photocatalytic activity (Figure 17).
Figure 17. (a) The cycling test of the catalytic performance. (b) Dye removal capacity by adsorption
and degradation for different cycles. (c) Digital photos demonstrating the recyclability by using an
external magnetic field. Adapted with permission from Reference [92]. Copyright (2017) Elsevier.
5.1.2. Hydrothermal Methods
The hydrothermal method is a wet-chemistry method for synthesizing single crystals. Through
the hydrothermal method, a great deal of crystalline phases that are not stable at the melting point
can be obtained [93]. As a result, numerous semiconductor nanoparticles with different surface
(a) (b)
(c)
Figure 17.
(
a
) The cycling test of the catalytic performance. (
b
) Dye removal capacity by adsorption
and degradation for different cycles. (
c
) Digital photos demonstrating the recyclability by using an
external magnetic field. Adapted with permission from Reference [92]. Copyright (2017) Elsevier.
5.1.2. Hydrothermal Methods
The hydrothermal method is a wet-chemistry method for synthesizing single crystals. Through the
hydrothermal method, a great deal of crystalline phases that are not stable at the melting point
can be obtained [
93
]. As a result, numerous semiconductor nanoparticles with different surface
morphologies and compositions can also be prepared using the hydrothermal method. As a
Catalysts 2019,9, 122 20 of 32
representative work, Wu et al. [
94
] fabricated the F-doped flower-like TiO
2
nanoparticle on the
surface of Ti via a low-temperature hydrothermal process. They reported that the presence of
HF in water and the hydrothermal reaction time play an important role in the formation of the
F-doped flower-like TiO
2
nanostructures. Through regulating the synthesis parameters, the obtained
F-doped TiO
2
flower-like nanomaterials exhibited a superior photoelectrochemical activity for the
photodegradation of organic pollutants compared with P-25. They also demonstrated that the
improved photoelectrochemical activity of the F-doped TiO
2
flower-like nanomaterials was mainly
due to the larger surface area and the enhanced visible light harvest capacity. Additionally, magnetic
composite photocatalysts can also be synthesized using the hydrothermal method, such as the
magnetic CoFe
2
O
4
/Ag/Ag
3
VO
4
photocatalysts fabricated by Jing and co-workers [
95
]. During this
study, the as-prepared CoFe
2
O
4
nanoparticles were dispersed in the solutions with AgNO
3
and
Na
3
VO
4
, and the mixture suspensions were hydrothermally treated to prepare CoFe
2
O
4
/Ag/Ag
3
VO
4
composites. Through controlling the weight ratios of CoFe
2
O
4
in the composite system, the optimal
CoFe
2
O
4
/Ag/Ag
3
VO
4
composite exhibited significantly improved photocatalytic activity toward
the degradation of various contaminants including methyl orange, tetracycline, and could even kill
Escherichia coli solely under the driving of visible light. Moreover, with the advantage of having a good
magnetic response property, the corresponding CoFe
2
O
4
/Ag/Ag
3
VO
4
composite could be facilely
collected from the water by applying an extra magnetic field (Figure 18).
Catalysts 2018, 8, x FOR PEER REVIEW 20 of 32
morphologies and compositions can also be prepared using the hydrothermal method. As a
representative work, Wu et al. [94] fabricated the F-doped flower-like TiO2 nanoparticle on the surface
of Ti via a low-temperature hydrothermal process. They reported that the presence of HF in water
and the hydrothermal reaction time play an important role in the formation of the F-doped flower-
like TiO2 nanostructures. Through regulating the synthesis parameters, the obtained F-doped TiO2
flower-like nanomaterials exhibited a superior photoelectrochemical activity for the
photodegradation of organic pollutants compared with P-25. They also demonstrated that the
improved photoelectrochemical activity of the F-doped TiO2 flower-like nanomaterials was mainly
due to the larger surface area and the enhanced visible light harvest capacity. Additionally, magnetic
composite photocatalysts can also be synthesized using the hydrothermal method, such as the
magnetic CoFe2O4/Ag/Ag3VO4 photocatalysts fabricated by Jing and co-workers [95]. During this
study, the as-prepared CoFe2O4 nanoparticles were dispersed in the solutions with AgNO3 and
Na3VO4, and the mixture suspensions were hydrothermally treated to prepare CoFe2O4/Ag/Ag3VO4
composites. Through controlling the weight ratios of CoFe2O4 in the composite system, the optimal
CoFe2O4/Ag/Ag3VO4 composite exhibited significantly improved photocatalytic activity toward the
degradation of various contaminants including methyl orange, tetracycline, and could even kill
Escherichia coli solely under the driving of visible light. Moreover, with the advantage of having a
good magnetic response property, the corresponding CoFe2O4/Ag/Ag3VO4 composite could be
facilely collected from the water by applying an extra magnetic field (Figure 18).
Figure 18. (a) Magnetic separation performance of the CoFe2O4/Ag/Ag3VO4 composite. (b) The
absorption spectra of methyl orange solutions over time in the presence of CoFe2O4/Ag/Ag3VO4
(under visible light λ 440 nm) and the UV–vis absorption spectrum of CoFe2O4/Ag/Ag3VO4. (c) The
evolution of the absorption spectra of methyl orange solutions over time in the presence of
CoFe2O4/Ag/Ag3VO4 (under visible light λ 550 nm). (d) Photocatalytic degradation of tetracycline
with different samples under visible light irradiation. Adapted with permission from Reference [95]
Copyright (2016) Elsevier.
5.2. Nanofibers/Nanorods (1D)
Recently, nanofibrous photocatalysts have been intensively studied owing to their unique long
aspect ratio, large surface area, and being easily functionalized. Up to now, various strategies had
been developed to synthesize the 1D materials with different morphology like: wires, belts, rods,
tubes, and rings [61,89,96], among which, the hydrothermal method and electrospinning are the
(a) (b)
(c) (d)
Figure 18.
(
a
) Magnetic separation performance of the CoFe
2
O
4
/Ag/Ag
3
VO
4
composite. (
b
) The absorption
spectra of methyl orange solutions over time in the presence of CoFe
2
O
4
/Ag/Ag
3
VO
4
(under visible light
λ
440 nm) and the UV–vis absorption spectrum of CoFe
2
O
4
/Ag/Ag
3
VO
4
. (
c
) The evolution of the
absorption spectra of methyl orange solutions over time in the presence of CoFe
2
O
4
/Ag/Ag
3
VO
4
(under
visible light
λ
550 nm). (
d
) Photocatalytic degradation of tetracycline with different samples under visible
light irradiation. Adapted with permission from Reference [95] Copyright (2016) Elsevier.
5.2. Nanofibers/Nanorods (1D)
Recently, nanofibrous photocatalysts have been intensively studied owing to their unique long aspect
ratio, large surface area, and being easily functionalized. Up to now, various strategies had been developed
to synthesize the 1D materials with different morphology like: wires, belts, rods, tubes, and rings [
61
,
89
,
96
],
among which, the hydrothermal method and electrospinning are the most-used techniques. Consequently,
Catalysts 2019,9, 122 21 of 32
in this part, we present the development of composite semiconductors with nanofibrous morphology
derived from the electrospinning method and hydrothermal method for the treatment of wastewater.
5.2.1. Hydrothermal Method
As mentioned above, the hydrothermal method is capable of synthesizing various inorganics
with different morphologies, including the nanofibrous materials. For example, Yang et al. [
97
] once
reported the fabrication of a novel TiO
2
nanofibers with a shell of anatase nanocrystals based on the
hydrothermal process. Actually, the whole fabrication process included three steps: First, the H
2
Ti
3
O
7
nanofibers were obtained from the anatase TiO
2
particles and NaOH solutions via hydrothermal
method. After that, the as-prepared H
2
Ti
3
O
7
nanofibers were treated using a dilute acid solution
under certain hydrothermal condition to generate the anatase nanocrystal shell on the outside. Finally,
the H
2
Ti
3
O
7
phase was converted to TiO
2
(B) phase after a heat treatment while the anatase nanocrystal
shell remained unchanged. Owing to the well-matched phase interfaces, which ensures the charge
transfer across the interfaces, the recombination of electron-hole pairs was effectively suppressed and
the corresponding photoactivity was significantly enhanced. Most importantly, they demonstrated
that these nanofibrous photocatalysts possess specific surface areas similar to the commercial P25
powder, and the fibril morphology endowed them with a good recyclability from water, which is
critically important in practical applications. Recently, our group also carried out a series of studies
on the synthesis of nanofibrous photocatalysts via employing the hydrothermal method such as the
bimetallic AuPd alloy nanoparticles deposited on MoO
3
nanowires [
98
]. As shown in Figure 19, MoO
3
nanowires were firstly prepared from Mo powder and H
2
O
2
via the hydrothermal method. Then,
the as-prepared MoO
3
nanowires were used as the substrates to synthesize the MoO
3
/Au-Pd bimetallic
alloy nanowires via a simple chemical reduction method. As expected, the MoO
3
/Au–Pd bimetallic
alloy nanowires exhibited a good photocatalytic degradation performance for trichloroethylene (TCE)
under the driving of visible light. Similarly, the composite nanowires could be easily separated from
the reaction slurry in a short time after the reaction.
Catalysts 2018, 8, x FOR PEER REVIEW 21 of 32
most-used techniques. Consequently, in this part, we present the development of composite
semiconductors with nanofibrous morphology derived from the electrospinning method and
hydrothermal method for the treatment of wastewater.
5.2.1. Hydrothermal Method
As mentioned above, the hydrothermal method is capable of synthesizing various inorganics
with different morphologies, including the nanofibrous materials. For example, Yang et al. [97] once
reported the fabrication of a novel TiO
2
nanofibers with a shell of anatase nanocrystals based on the
hydrothermal process. Actually, the whole fabrication process included three steps: First, the H
2
Ti
3
O
7
nanofibers were obtained from the anatase TiO
2
particles and NaOH solutions via hydrothermal
method. After that, the as-prepared H
2
Ti
3
O
7
nanofibers were treated using a dilute acid solution
under certain hydrothermal condition to generate the anatase nanocrystal shell on the outside. Finally,
the H
2
Ti
3
O
7
phase was converted to TiO
2
(B) phase after a heat treatment while the anatase nanocrystal
shell remained unchanged. Owing to the well-matched phase interfaces, which ensures the charge
transfer across the interfaces, the recombination of electron-hole pairs was effectively suppressed and
the corresponding photoactivity was significantly enhanced. Most importantly, they demonstrated
that these nanofibrous photocatalysts possess specific surface areas similar to the commercial P25
powder, and the fibril morphology endowed them with a good recyclability from water, which is
critically important in practical applications. Recently, our group also carried out a series of studies
on the synthesis of nanofibrous photocatalysts via employing the hydrothermal method such as the
bimetallic AuPd alloy nanoparticles deposited on MoO
3
nanowires [98]. As shown in Figure 19, MoO
3
nanowires were firstly prepared from Mo powder and H
2
O
2
via the hydrothermal method. Then, the
as-prepared MoO
3
nanowires were used as the substrates to synthesize the MoO
3
/Au-Pd bimetallic
alloy nanowires via a simple chemical reduction method. As expected, the MoO
3
/Au–Pd bimetallic
alloy nanowires exhibited a good photocatalytic degradation performance for trichloroethylene (TCE)
under the driving of visible light. Similarly, the composite nanowires could be easily separated from
the reaction slurry in a short time after the reaction.
Figure 19. Schematic illustrating the band structure of MoO
3
/Au-Pd composite photocatalyst and the
possible reaction mechanism. Adapted with permission from Reference [98]. Copyright (2018)
Elsevier.
5.2.2. Electrospinning Method
Electrospinning is considered as a promising way to synthesize nanofibers with several
advantages, such as easy operation, low cost, and scalable [99–101]. In general, there are four major
Figure 19.
Schematic illustrating the band structure of MoO
3
/Au-Pd composite photocatalyst and the
possible reaction mechanism. Adapted with permission from Reference [
98
]. Copyright (2018) Elsevier.
5.2.2. Electrospinning Method
Electrospinning is considered as a promising way to synthesize nanofibers with several advantages,
such as easy operation, low cost, and scalable [
99
101
]. In general, there are four major parts in an
electrospinning device: (i) an electrical power supplier, (ii) a metallic needle, (iii) syringes with the
Catalysts 2019,9, 122 22 of 32
polymer solution, and (iv) a conductive collector. Meanwhile, several process parameters, such as the
polymer-based solution concentration, the viscosity of solution, the flow rate of the syringe driver, and the
electric field power, could also be well-regulated to manipulate the morphology of fibers. During the
electrospinning process, the solution is injected through a metallic needle via a syringe with a constant
pump speed. At the same time, a voltage is applied on the metallic needle; therefore, the solution
droplet will be charged, and then a Taylor cone will be generated when the electronic force is enough
to overcome the surface tension. Following this, a liquid jet is formed between the grounded collector
and the needle. The generated jets will be stretched by an electrostatic repulsion force until it reaches the
collector; meanwhile, the solvent will rapidly evaporate during this process. Finally, the jets are solidified
and the corresponding nanofibers are collected on the collector [100].
As for the applications of photocatalysis, high specific surface area is required to provide more
active sites for the redox reaction. More specifically, electrospun nanofibers as forefront fibrous
materials have attracted considerable research attention in the area of photocatalysis due to its several
advantages of large surface area, extremely high aspect ratio, and ease of functionalization [
102
,
103
].
For example, Zhang et al. [
104
] reported the fabrication of a flexible and hierarchical mesoporous TiO
2
nanoparticle (TiO
2
NP) modified TiO
2
nanofiber composites via the combination of electrospinning and
in situ polymerization method. At first, flexible TiO
2
nanofibers were prepared via the electrospinning
and the subsequently consuming process with the dopant of yttrium. After that, the as-prepared
TiO
2
nanofibers were used as a template for the incorporation of TiO
2
NPs by utilizing a bifunctional
benzoxazine as the carrier through a calcination process in the N
2
atmosphere. The as-prepared
membranes exhibited remarkable photocatalytic activity towards organic dyes in water; moreover,
it could be reused well via simply rinsing with water, and without time-consuming separation
procedures owing to the long aspect ratio and good mechanical property of the composite nanofibers.
In recent years, our group has carried out several works on the design of electrospun nanofibrous
photocatalysts [
105
109
]. As a representative sample, a BiOCl
0.3
/BiOBr
0.3
/BiOI
0.4
/PAN composite
fibrous catalyst was fabricated via combining the electrospinning and sol-gel method [
109
]. As shown
in Figure 20, the obtained composite photocatalyst exhibited a typical fibril structure with a good
uniformity, and the corresponding field emission transmission electron microscope (FE-TEM) image
demonstrated a highly crystalline structure in the composite fibers with a clear lattice spacing relating to
the (112) plane of BiOCl, the (110) plane of BiOBr, and the (200) plane of BiOI; therefore, a heterojunction
structure was generated via a close contact of the composite semiconductors. After a visible-light driven
photocatalysis performance evaluation, it was found that the obtained BiOCl
0.3
/BiOBr
0.3
/BiOI
0.4
/PAN
fiber displayed the highest photocatalytic degradation performance of TCE. Moreover, it was concluded
that the improved visible-light driven photocatalytic activity is caused by the interfacial contact of a
heterojunction and the inhibition of the recombination rate of the electron–hole pairs.
Catalysts 2018, 8, x FOR PEER REVIEW 22 of 32
parts in an electrospinning device: (i) an electrical power supplier, (ii) a metallic needle, (iii) syringes
with the polymer solution, and (iv) a conductive collector. Meanwhile, several process parameters,
such as the polymer-based solution concentration, the viscosity of solution, the flow rate of the
syringe driver, and the electric field power, could also be well-regulated to manipulate the
morphology of fibers. During the electrospinning process, the solution is injected through a metallic
needle via a syringe with a constant pump speed. At the same time, a voltage is applied on the
metallic needle; therefore, the solution droplet will be charged, and then a Taylor cone will be
generated when the electronic force is enough to overcome the surface tension. Following this, a
liquid jet is formed between the grounded collector and the needle. The generated jets will be
stretched by an electrostatic repulsion force until it reaches the collector; meanwhile, the solvent will
rapidly evaporate during this process. Finally, the jets are solidified and the corresponding
nanofibers are collected on the collector. [100]
As for the applications of photocatalysis, high specific surface area is required to provide more
active sites for the redox reaction. More specifically, electrospun nanofibers as forefront fibrous
materials have attracted considerable research attention in the area of photocatalysis due to its several
advantages of large surface area, extremely high aspect ratio, and ease of functionalization [102,103].
For example, Zhang et al. [104] reported the fabrication of a flexible and hierarchical mesoporous
TiO2 nanoparticle (TiO2 NP) modified TiO2 nanofiber composites via the combination of
electrospinning and in situ polymerization method. At first, flexible TiO2 nanofibers were prepared
via the electrospinning and the subsequently consuming process with the dopant of yttrium. After
that, the as-prepared TiO2 nanofibers were used as a template for the incorporation of TiO2 NPs by
utilizing a bifunctional benzoxazine as the carrier through a calcination process in the N2 atmosphere.
The as-prepared membranes exhibited remarkable photocatalytic activity towards organic dyes in
water; moreover, it could be reused well via simply rinsing with water, and without time-consuming
separation procedures owing to the long aspect ratio and good mechanical property of the composite
nanofibers. In recent years, our group has carried out several works on the design of electrospun
nanofibrous photocatalysts [105–109]. As a representative sample, a BiOCl0.3/BiOBr0.3/BiOI0.4/PAN
composite fibrous catalyst was fabricated via combining the electrospinning and sol-gel method [109].
As shown in Figure 20, the obtained composite photocatalyst exhibited a typical fibril structure with
a good uniformity, and the corresponding field emission transmission electron microscope (FE-TEM)
image demonstrated a highly crystalline structure in the composite fibers with a clear lattice spacing
relating to the (112) plane of BiOCl, the (110) plane of BiOBr, and the (200) plane of BiOI; therefore, a
heterojunction structure was generated via a close contact of the composite semiconductors. After a
visible-light driven photocatalysis performance evaluation, it was found that the obtained
BiOCl0.3/BiOBr0.3/BiOI0.4/PAN fiber displayed the highest photocatalytic degradation performance of
TCE. Moreover, it was concluded that the improved visible-light driven photocatalytic activity is
caused by the interfacial contact of a heterojunction and the inhibition of the recombination rate of
the electron–hole pairs.
Figure 20. SEM image of pristine PAN fibers (a) and BiOClx/BiOBry/BiOIz composite fibers (b). (c) FE-
TEM image of BiOClx/BiOBry/BiOIz fibers. (d) Lattice-resolved image for BiOClx/BiOBry/BiOIz
(e)
(b) (e)
(a)
(c) (d)
Figure 20.
SEM image of pristine PAN fibers (
a
) and BiOCl
x
/BiOBr
y
/BiOI
z
composite
fibers (
b
). (
c
) FE-TEM image of BiOCl
x
/BiOBr
y
/BiOI
z
fibers. (
d
) Lattice-resolved image for
BiOCl
x
/BiOBr
y
/BiOI
z
nanofibers. (
e
) Schematic indicating the photocatalytic degradation of TCE.
Adapted with permission from Reference [109]. Copyright (2016) Elsevier.
Catalysts 2019,9, 122 23 of 32
5.3. Nanosheets (2D)
Semiconductor nanosheets are typical 2D nanomaterials and have attracted significant attention
in the research area of photocatalysis for their larger surface area and tunable structures. Up to
now, a great deal of semiconductor nanosheets have been synthesized via various strategies
for different applications. The hydrothermal process is one of the most used strategies for the
preparation of 2D semiconductor photocatalysts for the application of wastewater treatment [
110
].
Through the hydrothermal process, various nanosheets derived from a single semiconductor or
multi-semiconductors could be synthesized. For example, Chen et al. [
111
] prepared TiO
2
-based
nanosheets (TNS) via the alkaline hydrothermal treatment of commercial P25. They reported that
the as-prepared TNS exhibited much higher specific surface area and much stronger adsorption for
crystal violet molecules than the raw P25. Furthermore, the TNS could be effectively regenerated using
a H
2
O
2
-assisted photocatalysis process, showing great potential for dealing with the high-chroma
dye wastewater. Besides TiO
2
, various nanosheets derived from different semiconductors could also
be fabricated using a hydrothermal process, such as the WO
3
nanosheet/K
+
Ca
2
Nb
3
O
10
ultrathin
nanosheet synthesized by Ma et al. [
112
] via a facile hydrothermal assembly of WO
3
nanosheets and
ultrathin K
+
Ca
2
Nb
3
O
10
nanosheets. They demonstrated that the composite nanosheets possess
2D/2D heterojunctions and display remarkably enhanced photocatalytic activity compared to the
pristine WO
3
and K
+
Ca
2
Nb
3
O
10
nanosheets, which were mainly caused by the strongly coupled
hetero-interfaces that provided more active sites for reactions and band structure. Additionally,
some other methods, such as the solvothermal or photo-reduction methods, could also be employed
for the preparation of composite nanosheets. For example, Wang et al. [
113
] fabricated the Bi
12
O
15
Cl
6
nanosheets with a narrowed band gap via a simple and facile solvothermal method followed by
a simple thermal treatment. The obtained Bi
12
O
15
Cl
6
nanosheets exhibited a good photocatalytic
degradation performance of bisphenol A solely under the driving of visible light, and the reaction
rate of the composite nanosheets was 13.6 and 8.7 times faster than those of BiOCl and TiO
2
(P25), respectively. In addition, the as-prepared Bi1
2
O
15
Cl
6
nanosheets possessed good stability
and recyclability during the photocatalytic process. As shown in Figure 21, our group recently
developed a Pt/BiOI composite nanosheet via a photo-reduction method in ambient conditions [
114
],
where the as-prepared Pt/BiOI composites exhibited a flower-like structure and could effectively
photocatalytically degrade the rhodamine B and phenol under visible-light irradiation (
λ
> 420 nm),
where the degradation rate was superior to that of pure BiOI. Moreover, it was found that the content
of Pt in the composite plays a vital important role on the photoactivity, and it was found that the
optimal ratio of Pt to BiOI in the composite was 3%. It was concluded that the enhanced photocatalytic
activity of the Pt/BiOI composite was caused by the superior electron transfer ability with the presence
of an appropriate amount of Pt.
Catalysts 2019,9, 122 24 of 32
Catalysts 2018, 8, x FOR PEER REVIEW 24 of 32
Figure 21. (a,b) SEM images of the obtained Pt/BiOI composite nanosheets. (c,d) HR-SEM images of
BiOI nanosheets and Pt/BiOI composite nanosheets, respectively. (eg) HR-TEM images of Pt/BiOI
composite nanosheets. (h) Element mappings of Pt, Bi, O, and I for the Pt/BiOI composite nanosheets.
Adapted with permission from Reference [114]. Copyright (2017) Elsevier.
5.4. Frameworks (3D)
In recent years, there have been a great number of works reporting a new generation of
composite photocatalysts with 3D frameworks [115]. Actually, the 3D photocatalysts with well-
deigned frameworks show great advantages such as a large specific surface area, high adsorptive
capacity, good structure stability, good mass transfer ability, and a large number of exposed active
sites, which make them promising candidates for the highly efficient photodegradation of
contaminants in water. In general, the 3D composite photocatalysts could be obtained via two
approaches, which are to directly construct a photocatalyst with 3D frameworks (type I), or
compositing the photocatalysts with a template with 3D frameworks (type II).
Through employing the commonly reported synthesis methods of various 3D frameworks, such
as the sol–gel process, in situ assembly, and template methods, various 3D photocatalysts with
different characteristics have be fabricated [116]. The sol-gel process is a well-developed strategy to
synthesize aerogels, and some of the produced aerogels, such as the SiO
2
aerogels, have been
commercialized [117]. In general, there are two stages for the sol-gel process method: first, a precursor
(e.g., metal alkoxide) is subjected to the hydrolysis and condensation reactions to form a wet gel,
during which time, numerous networks are generated between the alkoxide groups; subsequently,
Figure 21.
(
a
,
b
) SEM images of the obtained Pt/BiOI composite nanosheets. (
c
,
d
) HR-SEM images of
BiOI nanosheets and Pt/BiOI composite nanosheets, respectively. (
e
g
) HR-TEM images of Pt/BiOI
composite nanosheets. (
h
) Element mappings of Pt, Bi, O, and I for the Pt/BiOI composite nanosheets.
Adapted with permission from Reference [114]. Copyright (2017) Elsevier.
5.4. Frameworks (3D)
In recent years, there have been a great number of works reporting a new generation of
composite photocatalysts with 3D frameworks [
115
]. Actually, the 3D photocatalysts with well-deigned
frameworks show great advantages such as a large specific surface area, high adsorptive capacity,
good structure stability, good mass transfer ability, and a large number of exposed active sites,
which make them promising candidates for the highly efficient photodegradation of contaminants in
water. In general, the 3D composite photocatalysts could be obtained via two approaches, which are to
directly construct a photocatalyst with 3D frameworks (type I), or compositing the photocatalysts with
a template with 3D frameworks (type II).
Through employing the commonly reported synthesis methods of various 3D frameworks, such as
the sol–gel process, in situ assembly, and template methods, various 3D photocatalysts with different
characteristics have be fabricated [
116
]. The sol-gel process is a well-developed strategy to synthesize
aerogels, and some of the produced aerogels, such as the SiO
2
aerogels, have been commercialized [
117
].
Catalysts 2019,9, 122 25 of 32
In general, there are two stages for the sol-gel process method: first, a precursor (e.g., metal alkoxide) is
subjected to the hydrolysis and condensation reactions to form a wet gel, during which time, numerous
networks are generated between the alkoxide groups; subsequently, the formed wet gels are sufficiently
dried to obtain aerogels. As a matter of course, a photocatalytic aerogel can be obtained using a
metal alkoxide precursor with an appropriate photocatalytic activity. For example,
Dagan et al. [118]
prepared a series of highly porous TiO
2
aerogels via the sol-gel method and they also proved that the
photocatalytic degradation performance of the TiO
2
aerogels for organic contaminants is much better
than that of a commercial TiO
2
(P25). Besides, various photoactive metal oxides, metal silylamide, or
their composite aerogels have been developed. However, due to the limitation of sol-gel processes,
some metal oxides or metal chalcogenides are not able to be synthesized into aerogels, and the obtained
aerogel photocatalysts usually suffer from low crystallinity. Therefore, a new generated strategy,
namely an assembly method, has been invented to construct aerogels based on various nanoscale units
with different morphologies and chemical properties. As reported before [
116
], there are three typical
steps for the assembly process: (i) fabrication of the building blocks, (ii) preparing the dispersion of
the building blocks with appropriated concentration, and (iii) solidified the suspension of building
blocks to form a 3D monolith. Based on this principle,
Heiligtag et al. [119]
developed a 3D framework
Au-TiO
2
photocatalysts with a preformed TiO
2
nanoparticles as the blocking units without using any
templates. Through modifying the surface of anatase TiO
2
nanoparticles with trizma, the nanoparticles
undergo an oriented attachment process during gelation and finally result in well-bonded networks.
Moreover, based on the above-mentioned aerogel synthesis methods, various phototcatalytic aerogels
can also be prepared via employing the preformed aerogels as the templates, such as a C
3
N
4
aerogel
that was fabricated by Kailasam and co-workers [
120
] via preparing a C
3
N
4
/SiO
2
composite aerogel
based on the sol-gel method at first, and then remove the SiO
2
via treating the composite in 4 M
NH4HF2(Figure 22).
Catalysts 2018, 8, x FOR PEER REVIEW 25 of 32
the formed wet gels are sufficiently dried to obtain aerogels. As a matter of course, a photocatalytic
aerogel can be obtained using a metal alkoxide precursor with an appropriate photocatalytic activity.
For example, Dagan et al. [118] prepared a series of highly porous TiO2 aerogels via the sol-gel
method and they also proved that the photocatalytic degradation performance of the TiO2 aerogels
for organic contaminants is much better than that of a commercial TiO2 (P25). Besides, various
photoactive metal oxides, metal silylamide, or their composite aerogels have been developed.
However, due to the limitation of sol-gel processes, some metal oxides or metal chalcogenides are
not able to be synthesized into aerogels, and the obtained aerogel photocatalysts usually suffer from
low crystallinity. Therefore, a new generated strategy, namely an assembly method, has been
invented to construct aerogels based on various nanoscale units with different morphologies and
chemical properties. As reported before [116], there are three typical steps for the assembly process:
(i) fabrication of the building blocks, (ii) preparing the dispersion of the building blocks with
appropriated concentration, and (iii) solidified the suspension of building blocks to form a 3D
monolith. Based on this principle, Heiligtag et al. [119] developed a 3D framework Au-TiO2
photocatalysts with a preformed TiO2 nanoparticles as the blocking units without using any
templates. Through modifying the surface of anatase TiO2 nanoparticles with trizma, the
nanoparticles undergo an oriented attachment process during gelation and finally result in well-
bonded networks. Moreover, based on the above-mentioned aerogel synthesis methods, various
phototcatalytic aerogels can also be prepared via employing the preformed aerogels as the templates,
such as a C3N4 aerogel that was fabricated by Kailasam and co-workers [120] via preparing a
C3N4/SiO2 composite aerogel based on the sol-gel method at first, and then remove the SiO2 via
treating the composite in 4 M NH4HF2 (Figure 22).
Figure 22. Schematic illustration indicating the synthesis process of porous carbon nitride and silica
aerogels based on the sol-gel method and the digital photos of corresponding aerogels. Adapted with
permission from Reference [120]. Copyright (2011) Royal Society of Chemistry.
As for the fabrication of type II 3D photocatalysts, an appropriate 3D porous substrate should
be prepared before loading the active substance on its frameworks. Considering the requirements for
high photoreactivity and good service performance, the aerogels/hydrogels derived from ceramics
or carbon are mostly preferred. For example, Li et al. [121] fabricated a ternary magnetic composite
of Fe3O4@TiO2/SiO2 aerogel via combining the sol-gel process and a hydrothermal treatment. During
the fabrication process, Fe3O4 microspheres were first synthesized via the hydrothermal method; after
that, Fe3O4@TiO2 core shell microspheres were fabricated via an in situ reaction method. The used
SiO2 aerogel was derived from the industrial fly ash via a common sol-gel method. Finally, the as-
prepared Fe3O4@TiO2 core shell microspheres and SiO2 aerogel were combined via the hydrothermal
method. According to their report, the obtained Fe3O4@TiO2/SiO2 aerogel exhibited an enhanced
photocatalytic activity for the degradation of rhodamine B dye under visible light irradiation, and
the aerogel could be facilely collected after the reaction due to its good magnetic separation
performance. Interestingly, as shown in Figure 23, Jiang et al. [122] recently developed a separation-
Figure 22.
Schematic illustration indicating the synthesis process of porous carbon nitride and silica
aerogels based on the sol-gel method and the digital photos of corresponding aerogels. Adapted with
permission from Reference [120]. Copyright (2011) Royal Society of Chemistry.
As for the fabrication of type II 3D photocatalysts, an appropriate 3D porous substrate should be
prepared before loading the active substance on its frameworks. Considering the requirements for
high photoreactivity and good service performance, the aerogels/hydrogels derived from ceramics or
carbon are mostly preferred. For example, Li et al. [
121
] fabricated a ternary magnetic composite of
Fe
3
O
4
@TiO
2
/SiO
2
aerogel via combining the sol-gel process and a hydrothermal treatment. During the
fabrication process, Fe
3
O
4
microspheres were first synthesized via the hydrothermal method; after that,
Fe
3
O
4
@TiO
2
core shell microspheres were fabricated via an in situ reaction method. The used SiO
2
aerogel was derived from the industrial fly ash via a common sol-gel method. Finally, the as-prepared
Catalysts 2019,9, 122 26 of 32
Fe
3
O
4
@TiO
2
core shell microspheres and SiO
2
aerogel were combined via the hydrothermal method.
According to their report, the obtained Fe
3
O
4
@TiO
2
/SiO
2
aerogel exhibited an enhanced photocatalytic
activity for the degradation of rhodamine B dye under visible light irradiation, and the aerogel could
be facilely collected after the reaction due to its good magnetic separation performance. Interestingly,
as shown in Figure 23, Jiang et al. [
122
] recently developed a separation-free PANI/TiO
2
3D hydrogel
for the continuous photocatalytic degradation of various contaminants in water. In their studies,
the PANI hydrogel with 3D frameworks was synthesized via the polymerization of aniline. During the
gelling process, the TiO
2
nanoparticles (P25) were incapsulated in the hydrogels. As a result,
the obtained PANI/TiO
2
composite hydrogel exhibited an intriguing capacity for removing organic
contaminants from water, which was mainly caused by the synergistic effect of adsorption enrichment
of hydrogel and the in situ photocatalytic degradation of TiO
2
. Moreover, the presented separation-free
characteristics in the obtained bulk materials indicate a good recyclability of the composite hydrogel.
Catalysts 2018, 8, x FOR PEER REVIEW 26 of 32
free PANI/TiO
2
3D hydrogel for the continuous photocatalytic degradation of various contaminants
in water. In their studies, the PANI hydrogel with 3D frameworks was synthesized via the
polymerization of aniline. During the gelling process, the TiO
2
nanoparticles (P25) were incapsulated
in the hydrogels. As a result, the obtained PANI/TiO
2
composite hydrogel exhibited an intriguing
capacity for removing organic contaminants from water, which was mainly caused by the synergistic
effect of adsorption enrichment of hydrogel and the in situ photocatalytic degradation of TiO
2
.
Moreover, the presented separation-free characteristics in the obtained bulk materials indicate a good
recyclability of the composite hydrogel.
Figure 23. Schematic illustration demonstrating the synthesis process of the 3D PANI/TiO
2
composite
hydrogel. Adapted with permission from Reference [122]. Copyright (2015) Wiley.
6. Summary and Perspectives
In summary, in order to address the worldwide concerned issues of water pollutions, various
photocatalysis processes based on different photocatalysts have been developed; meanwhile,
numerous efforts have been made to further improve the photocatalytic activity of the catalysts based
on the semiconducors. In this review, the recent progress in the development of composite
semiconductor photocatalysts for wastewater treatment is presented, including the most-used
strategies to narrow the band gap of semiconductors, to retard the recombination of the photo-
generated electron-hole pairs, to enhance the visible light adsorption capacity, as well as to increase
the reaction ratio between the photocatalysts and contaminants. Moreover, the composite catalyts
with different morphologies and the corresponding photocatlytic performance were also
summarized.
Although great development of the photocatalysis process has been obtained, there are still
several problems yet to be addressed to further improve the practical application performance of the
photocatalysis. Therefore, some plausible perspectives for the developing trend of composite
photocatalysts for the wasterwater treatment are proposed based on the presented studies: i) the
existing synthesis methods are relative complex, high cost, and harmful to the environment to some
degree, thus a more facile, highly efficent, and green method is anticipated; ii) the mechanism of the
composite semicondutor photocatalysts are still confusing and some of them are unpersuasive,
therefore much more effort is needed for the basic studies of the catalytic mechanisms; and iii) the
practical use are limitted because the collection and reuse of the catalysts in water are still
inconvenient due to their small size and poor mechanical property, therefore novel photocatalysts
with easy collection property or new hybrid devices based on the composite of photocatalysts with
selected substrates (e.g., polymers, metals) are proposed. Finally, we anticipate that this review can
provide some useful guidance for the design of next generation of photocatlysts for the wastewater
remediation.
Funding: This work was supported by the Technology Development Program (S2598148) funded by the
Ministry of SMEs and Startups (MSS, Korea) and the Commercialization Promotion Agency for R&D Outcomes
Figure 23.
Schematic illustration demonstrating the synthesis process of the 3D PANI/TiO
2
composite
hydrogel. Adapted with permission from Reference [122]. Copyright (2015) Wiley.
6. Summary and Perspectives
In summary, in order to address the worldwide concerned issues of water pollutions, various
photocatalysis processes based on different photocatalysts have been developed; meanwhile, numerous
efforts have been made to further improve the photocatalytic activity of the catalysts based on the
semiconducors. In this review, the recent progress in the development of composite semiconductor
photocatalysts for wastewater treatment is presented, including the most-used strategies to narrow the
band gap of semiconductors, to retard the recombination of the photo-generated electron-hole pairs,
to enhance the visible light adsorption capacity, as well as to increase the reaction ratio between the
photocatalysts and contaminants. Moreover, the composite catalyts with different morphologies and
the corresponding photocatlytic performance were also summarized.
Although great development of the photocatalysis process has been obtained, there are still
several problems yet to be addressed to further improve the practical application performance of
the photocatalysis. Therefore, some plausible perspectives for the developing trend of composite
photocatalysts for the wasterwater treatment are proposed based on the presented studies: (i) the
existing synthesis methods are relative complex, high cost, and harmful to the environment to some
degree, thus a more facile, highly efficent, and green method is anticipated; (ii) the mechanism of
the composite semicondutor photocatalysts are still confusing and some of them are unpersuasive,
therefore much more effort is needed for the basic studies of the catalytic mechanisms; and (iii)
the practical use are limitted because the collection and reuse of the catalysts in water are still
inconvenient due to their small size and poor mechanical property, therefore novel photocatalysts with
easy collection property or new hybrid devices based on the composite of photocatalysts with selected
Catalysts 2019,9, 122 27 of 32
substrates (e.g., polymers, metals) are proposed. Finally, we anticipate that this review can provide
some useful guidance for the design of next generation of photocatlysts for the wastewater remediation.
Funding:
This work was supported by the Technology Development Program (S2598148) funded by the Ministry
of SMEs and Startups (MSS, Korea) and the Commercialization Promotion Agency for R&D Outcomes (COMPA)
funded by the Ministry of Science and ICT (MSIT) [2018_RND_002_0064, Development of 800 mAh/g pitch
carbon coating.
Acknowledgments:
This work was supported by the Technology Development Program (S2598148) funded
by the Ministry of SMEs and Startups (MSS, Korea) and the Commercialization Promotion Agency for R&D
Outcomes (COMPA) funded by the Ministry of Science and ICT (MSIT) [2018_RND_002_0064, Development of
800 mAh/g pitch carbon coating.
Conflicts of Interest: The authors declare no conflict of interest.
References
1.
Carpenter, S.R.; Caraco, N.F.; Correll, D.L.; Howarth, R.W.; Sharpley, A.N.; Smith, V.H. Nonpoint pollution
of surface waters with phosphorus and nitrogen. Ecol. Appl. 1998,8, 559–568. [CrossRef]
2. Jarup, L. Hazards of heavy metal contamination. Br. Med. Bull. 2003,68, 167–182. [CrossRef] [PubMed]
3.
Dudgeon, D.; Arthington, A.H.; Gessner, M.O.; Kawabata, Z.I.; Knowler, D.J.; Leveque, C.; Naiman, R.J.;
Prieur-Richard, A.H.; Soto, D.; Stiassny, M.L.J.; et al. Freshwater biodiversity: Importance, threats, status
and conservation challenges. Biol. Rev. 2006,81, 163–182. [CrossRef] [PubMed]
4.
Ribeiro, A.R.; Nunes, O.C.; Pereira, M.F.R.; Silva, A.M.T. An overview on the advanced oxidation processes
applied for the treatment of water pollutants defined in the recently launched Directive 2013/39/EU.
Environ. Int. 2015,75, 33–51. [CrossRef] [PubMed]
5.
Kolpin, D.W.; Furlong, E.T.; Meyer, M.T.; Thurman, E.M.; Zaugg, S.D.; Barber, L.B.; Buxton, H.T.
Pharmaceuticals, hormones, and other organic wastewater contaminants in US streams, 1999-2000: A national
reconnaissance. Environ. Sci. Technol. 2002,36, 1202–1211. [CrossRef] [PubMed]
6.
Schwarzenbach, R.P.; Escher, B.I.; Fenner, K.; Hofstetter, T.B.; Johnson, C.A.; von Gunten, U.; Wehrli, B.
The challenge of micropollutants in aquatic systems. Science 2006,313, 1072–1077. [CrossRef]
7.
Radjenovic, J.; Sedlak, D.L. Challenges and opportunities for electrochemical processes as next-generation
technologies for the treatment of contaminated water. Environ. Sci. Technol.
2015
,49, 11292–11302. [CrossRef]
8. Jiang, J.Q. The role of coagulation in water treatment. Curr. Opin. Chem. Eng. 2015,8, 36–44. [CrossRef]
9.
Lekang, O.I.; Bomo, A.M.; Svendsen, I. Biological lamella sedimentation used for wastewater treatment.
Aquac. Eng. 2001,24, 115–127. [CrossRef]
10.
Ali, I. Water treatment by adsorption columns: Evaluation at ground level. Sep. Purif. Rev.
2014
,43, 175–205.
[CrossRef]
11.
Oe, T.; Koide, H.; Hirokawa, H.; Okukawa, K. Performance of membrane filtration system used for water
treatment. Desalination 1996,106, 107–113. [CrossRef]
12.
Joss, A.; Zabczynski, S.; Gobel, A.; Hoffmann, B.; Loffler, D.; McArdell, C.S.; Ternes, T.A.; Thomsen, A.;
Siegrist, H. Biological degradation of pharmaceuticals in municipal wastewater treatment: Proposing a
classification scheme. Water Res. 2006,40, 1686–1696. [CrossRef] [PubMed]
13.
Mamba, G.; Mishra, A. Advances in magnetically separable photocatalysts: Smart, recyclable materials for
water pollution mitigation. Catalysts 2016,6, 34. [CrossRef]
14.
Lee, S.Y.; Park, S.J. TiO
2
photocatalyst for water treatment applications. J. Ind. Eng. Chem.
2013
,19, 1761–1769.
[CrossRef]
15.
Van der Hoek, J.P.; Bertelkamp, C.; Verliefde, A.R.D.; Singhal, N. Drinking water treatment technologies in
Europe: State of the art-challenges-research needs. J. Water Supply Res. Technol.-Aquac.
2014
,63, 124–130.
[CrossRef]
16.
Gogate, P.R.; Pandit, A.B. A review of imperative technologies for wastewater treatment I:
Oxidation technologies at ambient conditions. Adv. Environ. Res. 2004,8, 501–551. [CrossRef]
17.
Oturan, M.A.; Aaron, J.J. Advanced oxidation processes in water/wastewater treatment: Principles and
applications. a review. Crit. Rev. Environ. Sci. Technol. 2014,44, 2577–2641. [CrossRef]
Catalysts 2019,9, 122 28 of 32
18.
Pelaez, M.; Nolan, N.T.; Pillai, S.C.; Seery, M.K.; Falaras, P.; Kontos, A.G.; Dunlop, P.S.M.; Hamilton, J.W.J.;
Byrne, J.A.; O’Shea, K.; et al. A review on the visible light active titanium dioxide photocatalysts for
environmental applications. Appl. Catal. B-Environ. 2012,125, 331–349. [CrossRef]
19.
Comninellis, C.; Kapalka, A.; Malato, S.; Parsons, S.A.; Poulios, L.; Mantzavinos, D. Advanced oxidation
processes for water treatment: Advances and trends for R&D. J. Chem. Technol. Biot. 2008,83, 769–776.
20.
Sansaniwal, S.K.; Sharma, V.; Mathur, J. Energy and exergy analyses of various typical solar energy
applications: A comprehensive review. Renew. Sustain. Energy Rev. 2018,82, 1576–1601. [CrossRef]
21.
Mills, A.; Davies, R.H.; Worsley, D. Water-purification by semiconductor photocatalysis. Chem. Soc. Rev.
1993,22, 417–425. [CrossRef]
22.
Mills, A.; LeHunte, S. An overview of semiconductor photocatalysis. J. Photochem. Photobiol. A Chem.
1997
,
108, 1–35. [CrossRef]
23.
Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature
1972
,238,
37. [CrossRef] [PubMed]
24.
Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.L.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W.
Understanding TiO
2
photocatalysis: Mechanisms and materials. Chem. Rev.
2014
,114, 9919–9986. [CrossRef]
[PubMed]
25.
Tong, H.; Ouyang, S.X.; Bi, Y.P.; Umezawa, N.; Oshikiri, M.; Ye, J.H. Nano-photocatalytic materials:
Possibilities and challenges. Adv. Mater. 2012,24, 229–251. [CrossRef] [PubMed]
26.
An, H.Z.; Du, Y.; Wang, T.M.; Wang, C.; Hao, W.C.; Zhang, J.Y. Photocatalytic properties of BiOX (X = Cl, Br,
and I). Rare Metals 2008,27, 243–250. [CrossRef]
27.
Azeez, A.A.; Rhee, K.Y.; Park, S.J.; Hui, D. Epoxy clay nanocomposites—Processing, properties and
applications: A review. Compos. Part B-Eng. 2013,45, 308–320. [CrossRef]
28.
Dhand, V.; Mittal, G.; Rhee, K.Y.; Park, S.J.; Hui, D. A short review on basalt fiber reinforced polymer
composites. Compos. Part B-Eng. 2015,73, 166–180. [CrossRef]
29.
Chong, M.N.; Jin, B.; Chow, C.W.K.; Saint, C. Recent developments in photocatalytic water treatment
technology: A review. Water Res. 2010,44, 2997–3027. [CrossRef] [PubMed]
30. Ani, I.J.; Akpan, U.G.; Olutoye, M.A.; Hameed, B.H. Photocatalytic degradation of pollutants in petroleum
refinery wastewater by TiO
2
- and ZnO-based photocatalysts: Recent development. J. Clean. Prod.
2018
,205,
930–954. [CrossRef]
31.
Yu, Z.B.; Chen, X.Q.; Kang, X.D.; Xie, Y.P.; Zhu, H.Z.; Wang, S.L.; Ullah, S.; Ma, H.; Wang, L.Z.; Liu, G.; et al.
Noninvasively modifying band structures of wide-bandgap metal oxides to boost photocatalytic activity.
Adv. Mater. 2018,30, 1706259. [CrossRef] [PubMed]
32.
Choi, W.Y.; Termin, A.; Hoffmann, M.R. The role of metal ion dopants in quantum-sized TiO
2
:
Correlation between photoreativity and charge carrier recombination dynamics. J. Phys. Chem.
1994
,
98, 13669–13679. [CrossRef]
33.
Liu, G.; Wang, L.Z.; Yang, H.G.; Cheng, H.M.; Lu, G.Q. Titania-based photocatalysts-crystal growth, doping
and heterostructuring. J. Mater. Chem. 2010,20, 831–843. [CrossRef]
34.
Zhu, J.F.; Deng, Z.G.; Chen, F.; Zhang, J.L.; Chen, H.J.; Anpo, M.; Huang, J.Z.; Zhang, L.Z.
Hydrothermal doping method for preparation of Cr
3+
-TiO
2
photocatalysts with concentration gradient
distribution of Cr3+.Appl. Catal. B-Environ. 2006,62, 329–335. [CrossRef]
35.
Dvoranova, D.; Brezova, V.; Mazur, M.; Malati, M.A. Investigations of metal-doped titanium dioxide
photocatalysts. Appl. Catal. B-Environ. 2002,37, 91–105. [CrossRef]
36.
Bouras, P.; Stathatos, E.; Lianos, P. Pure versus metal-ion-doped nanocrystalline titania for photocatalysis.
Appl. Catal. B-Environ. 2007,73, 51–59. [CrossRef]
37.
Devi, L.G.; Kottam, N.; Murthy, B.N.; Kumar, S.G. Enhanced photocatalytic activity of transition metal ions
Mn
2+
, Ni
2+
and Zn
2+
doped polycrystalline titania for the degradation of Aniline Blue under UV/solar light.
J. Mol. Catal. A Chem. 2010,328, 44–52. [CrossRef]
38.
Serpone, N. Is the band gap of pristine TiO
2
narrowed by anion- and cation-doping of titanium dioxide in
second-generation photocatalysts? J. Phys. Chem. B 2006,110, 24287–24293. [CrossRef]
39.
Kudo, A.; Niishiro, R.; Iwase, A.; Kato, H. Effects of doping of metal cations on morphology, activity,
and visible light response of photocatalysts. Chem. Phys. 2007,339, 104–110. [CrossRef]
40.
Chen, J.H.; Yao, M.S.; Wang, X.L. Investigation of transition metal ion doping behaviors on TiO
2
nanoparticles.
J. Nanopart. Res. 2008,10, 163–171. [CrossRef]
Catalysts 2019,9, 122 29 of 32
41.
Dunnill, C.W.; Parkin, I.P. Nitrogen-doped TiO
2
thin films: Photocatalytic applications for healthcare
environments. Dalton Trans. 2011,40, 1635–1640. [CrossRef] [PubMed]
42.
Asahi, R.; Morikawa, T.; Irie, H.; Ohwaki, T. Nitrogen-doped titanium dioxide as visible-light-sensitive
photocatalyst: Designs, developments, and prospects. Chem. Rev.
2014
,114, 9824–9852. [CrossRef] [PubMed]
43.
Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light photocatalysis in nitrogen-doped titanium
oxides. Science 2001,293, 269–271. [CrossRef] [PubMed]
44.
Kitano, M.; Funatsu, K.; Matsuoka, M.; Ueshima, M.; Anpo, M. Preparation of nitrogen-substituted TiO
2
thin
film photocatalysts by the radio frequency magnetron sputtering deposition method and their photocatalytic
reactivity under visible light irradiation. J. Phys. Chem. B 2006,110, 25266–25272. [CrossRef] [PubMed]
45.
Salah, N.; Hameed, A.; Aslam, M.; Abdel-Wahab, M.S.; Babkair, S.S.; Bahabri, F.S. Flow controlled fabrication
of N doped ZnO thin films and estimation of their performance for sunlight photocatalytic decontamination
of water. Chem. Eng. J. 2016,291, 115–127. [CrossRef]
46.
Mikkelsen, N.J.; Pedersen, J.; Straede, C.A. Ion implantation—The job coater’s supplement to coating
techniques. Surf. Coat. Technol. 2002,158, 42–47. [CrossRef]
47.
Tang, G.Z.; Li, J.L.; Sun, M.R.; Ma, X.X. Fabrication of nitrogen-doped TiO
2
layer on titanium substrate.
Appl. Surf. Sci. 2009,255, 9224–9229. [CrossRef]
48.
Song, X.Y.; Li, W.Q.; He, D.; Wu, H.Y.; Ke, Z.J.; Jiang, C.Z.; Wang, G.M.; Xiao, X.H. The “Midas
Touch” transformation of TiO
2
nanowire arrays during visible light photoelectrochemical performance
by carbon/nitrogen coimplantation. Adv. Energy Mater. 2018,8, 1800165. [CrossRef]
49.
Chen, X.; Mao, S.S. Titanium dioxide nanomaterials: Synthesis, properties, modifications, and applications.
Chem. Rev. 2007,107, 2891–2959. [CrossRef]
50.
Lee, S.Y.; Park, J.; Joo, H. Visible light-sensitized photocatalyst immobilized on beads by CVD in a fluidizing
bed. Sol. Energy Mater. Sol. Cell 2006,90, 1905–1914. [CrossRef]
51.
Kafizas, A.; Crick, C.; Parkin, I.P. The combinatorial atmospheric pressure chemical vapour deposition
(cAPCVD) of a gradating substitutional/interstitial N-doped anatase TiO2thin-film; UVA and visible light
photocatalytic activities. J. Photochem. Photobiol. A Chem. 2010,216, 156–166. [CrossRef]
52.
Youssef, L.; Leoga, A.J.K.; Roualdes, S.; Bassil, J.; Zakhour, M.; Rouessac, V.; Ayral, A.; Nakhl, M.
Optimization of N-doped TiO
2
multifunctional thin layers by low frequency PECVD process. J. Eur.
Ceram. Soc. 2017,37, 5289–5303. [CrossRef]
53. George, S.M. Atomic layer deposition: An Overview. Chem. Rev. 2010,110, 111–131. [CrossRef]
54.
Vilhunen, S.H.; Sillanpaa, M.E.T. Atomic layer deposited (ALD) TiO
2
and TiO
2-x
N
x
thin film photocatalysts
in salicylic acid decomposition. Water Sci. Technol. 2009,60, 2471–2475. [CrossRef]
55.
Pore, V.; Heikkila, M.; Ritala, M.; Leskela, M.; Areva, S. Atomic layer deposition of TiO
2-x
N
x
thin films for
photocatalytic applications. J. Photochem. Photobiol. A Chem. 2006,177, 68–75. [CrossRef]
56.
Lee, A.; Libera, J.A.; Waldman, R.Z.; Ahmed, A.; Avila, J.R.; Elam, J.W.; Darling, S.B. Conformal nitrogen-doped
TiO2photocatalytic coatings for sunlight-activated membranes. Adv. Sustain. Syst. 2017,1, 1600041. [CrossRef]
57.
Albrbar, A.J.; Djokic, V.; Bjelajac, A.; Kovac, J.; Cirkovic, J.; Mitric, M.; Janackovic, D.; Petrovic, R.
Visible-light active mesoporous, nanocrystalline N,S-doped and co-doped titania photocatalysts synthesized
by non-hydrolytic sol-gel route. Ceram. Int. 2016,42, 16718–16728. [CrossRef]
58.
Rajoriya, S.; Bargole, S.; George, S.; Saharan, V.K.; Gogate, P.R.; Pandit, A.B. Synthesis and characterization
of samarium and nitrogen doped TiO
2
photocatalysts for photo-degradation of 4-acetamidophenol in
combination with hydrodynamic and acoustic cavitation. Sep. Purif. Technol.
2019
,209, 254–269. [CrossRef]
59.
Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction photocatalysts. Adv. Mater.
2017
,29,
1601694. [CrossRef]
60.
Baek, J.H.; Kim, B.J.; Han, G.S.; Hwang, S.W.; Kim, D.R.; Cho, I.S.; Jung, H.S. BiVO
4
/WO
3
/SnO
2
Double-heterojunction photoanode with enhanced charge separation and visible-transparency for bias-free
solar water-splitting with a perovskite solar cell. ACS Appl. Mater. Interfaces 2017,9, 1479–1487. [CrossRef]
61.
Han, H.S.; Han, G.S.; Kim, J.S.; Kim, D.H.; Hong, J.S.; Caliskan, S.; Jung, H.S.; Cho, I.S.; Lee, J.K.
Indium-tin-oxide nanowire array based CdSe/CdS/TiO
2
one-dimensional heterojunction photoelectrode for
enhanced solar hydrogen production. ACS Sustain. Chem. Eng. 2016,4, 1161–1168. [CrossRef]
62.
Yu, J.C.; Yu, J.G.; Ho, W.K.; Zhang, L.Z. Preparation of highly photocatalytic active nano-sized TiO
2
particles
via ultrasonic irradiation. Chem. Commun. 2001, 1942–1943. [CrossRef]
Catalysts 2019,9, 122 30 of 32
63.
Uddin, M.T.; Nicolas, Y.; Olivier, C.; Toupance, T.; Servant, L.; Muller, M.M.; Kleebe, H.J.; Ziegler, J.;
Jaegermann, W. Nanostructured SnO
2
-ZnO heterojunction photocatalysts showing enhanced photocatalytic
activity for the degradation of organic dyes. Inorg. Chem. 2012,51, 7764–7773. [CrossRef]
64.
Shirmardi, A.; Teridi, M.A.M.; Azimi, H.R.; Basirun, W.J.; Jamali-Sheini, F.; Yousefi, R.
Enhanced photocatalytic performance of ZnSe/PANI nanocomposites for degradation of organic and
inorganic pollutants. Appl. Surf. Sci. 2018,462, 730–738. [CrossRef]
65.
Lee, C.H.; Lee, G.H.; van der Zande, A.M.; Chen, W.C.; Li, Y.L.; Han, M.Y.; Cui, X.; Arefe, G.; Nuckolls, C.;
Heinz, T.F.; et al. Atomically thin p-n junctions with van der Waals heterointerfaces. Nat. Nanotechnol.
2014
,
9, 676–681. [CrossRef]
66.
Kawazoe, H.; Yanagi, H.; Ueda, K.; Hosono, H. Transparent p-type conducting oxides: Design and fabrication
of p-n heterojunctions. MRS Bull. 2000,25, 28–36. [CrossRef]
67.
Lu, M.X.; Shao, C.L.; Wang, K.X.; Lu, N.; Zhang, X.; Zhang, P.; Zhang, M.Y.; Li, X.H.; Liu, Y.C.
p-MoO
3
Nanostructures/n-TiO
2
nanofiber heterojunctions: Controlled fabrication and enhanced
photocatalytic properties. ACS Appl. Mater. Interfaces 2014,6, 9004–9012. [CrossRef]
68.
Zhang, L.P.; Jaroniec, M. Toward designing semiconductor-semiconductor heterojunctions for photocatalytic
applications. Appl. Surf. Sci. 2018,430, 2–17. [CrossRef]
69.
Wen, X.J.; Niu, C.G.; Zhang, L.; Zeng, G.M. Novel p-n heterojunction BiOI/CeO
2
photocatalyst for wider
spectrum visible-light photocatalytic degradation of refractory pollutants. Dalton Trans.
2017
,46, 4982–4993.
[CrossRef]
70. Zhang, Y.; Park, S.J. Formation of hollow MoO3/SnS2heterostructured nanotubes for efficient light-driven
hydrogen peroxide production. J. Mater. Chem. A 2018,6, 20304–20312. [CrossRef]
71.
Wei, Z.D.; Zhao, Y.; Fan, F.T.; Li, C. The property of surface heterojunction performed by crystal facets for
photogenerated charge separation. Comp. Mater. Sci. 2018,153, 28–35. [CrossRef]
72.
Yu, J.G.; Low, J.X.; Xiao, W.; Zhou, P.; Jaroniec, M. Enhanced Photocatalytic CO
2
-Reduction Activity of
Anatase TiO2by Coexposed {001} and {101} Facets. J. Am. Chem. Soc. 2014,136, 8839–8842. [CrossRef]
73.
Gao, S.J.; Wang, W.; Ni, Y.R.; Lu, C.H.; Xu, Z.Z. Facet-dependent photocatalytic mechanisms of anatase TiO
2
:
A new sight on the self-adjusted surface heterojunction. J. Alloys Compd. 2015,647, 981–988. [CrossRef]
74.
Lu, J.; Wu, J.; Xu, W.X.; Cheng, H.Q.; Qi, X.M.; Li, Q.W.; Zhang, Y.A.; Guan, Y.; Ling, Y.; Zhang, Z.
Room temperature synthesis of tetragonal BiOI photocatalyst with surface heterojunction between (001)
facets and (110) facets. Mater. Lett. 2018,219, 260–264. [CrossRef]
75.
Li, H.J.; Tu, W.G.; Zhou, Y.; Zou, Z.G. Z-scheme photocatalytic systems for promoting photocatalytic
performance: Recent progress and future challenges. Adv. Sci. 2016,3, 12. [CrossRef]
76.
Wu, X.S.; Hu, Y.D.; Wang, Y.; Zhou, Y.S.; Han, Z.H.; Jin, X.L.; Chen, G. In-situ synthesis of Z-scheme
Ag
2
CO
3
/Ag/AgNCO heterojunction photocatalyst with enhanced stability and photocatalytic activity.
Appl. Surf. Sci. 2019,464, 108–114. [CrossRef]
77.
Lu, X.Y.; Che, W.J.; Hu, X.F.; Wang, Y.; Zhang, A.T.; Deng, F.; Luo, S.L.; Dionysiou, D.D. The facile fabrication
of novel visible-light-driven Z-scheme CuInS
2
/Bi
2
WO
6
heterojunction with intimate interface contact by in
situ hydrothermal growth strategy for extraordinary photocatalytic performance. Chem. Eng. J.
2019
,356,
819–829. [CrossRef]
78.
Leary, R.; Westwood, A. Carbonaceous nanomaterials for the enhancement of TiO
2
photocatalysis. Carbon
2011,49, 741–772. [CrossRef]
79.
Li, J.; Liu, K.; Xue, J.; Xue, G.; Sheng, X.; Wang, H.; Huo, P.; Yan, Y. CQDS preluded carbon-incorporated 3D
burger-like hybrid ZnO enhanced visible-light-driven photocatalytic activity and mechanism implication.
J. Catal. 2019,369, 450–461. [CrossRef]
80.
Long, B.; Huang, Y.C.; Li, H.B.; Zhao, F.Y.; Rui, Z.B.; Liu, Z.L.; Tong, Y.X.; Ji, H.B. Carbon Dots Sensitized
BiOI with Dominant {001} Facets for superior photocatalytic performance. Ind. Eng. Chem. Res.
2015
,54,
12788–12794. [CrossRef]
81.
Zhao, F.F.; Rong, Y.F.; Wan, J.M.; Hu, Z.W.; Peng, Z.Q.; Wang, B. High photocatalytic performance of
carbon quantum dots/TNTs composites for enhanced photogenerated charges separation under visible light.
Catal. Today 2018,315, 162–170. [CrossRef]
82.
Di, J.; Li, S.X.; Zhao, Z.F.; Huang, Y.C.; Jia, Y.; Zheng, H.J. Biomimetic CNT@TiO
2
composite with enhanced
photocatalytic properties. Chem. Eng. J. 2015,281, 60–68. [CrossRef]
Catalysts 2019,9, 122 31 of 32
83.
Miribangul, A.; Ma, X.L.; Zeng, C.; Zou, H.; Wu, Y.H.; Fan, T.P.; Su, Z. Synthesis of TiO
2
/CNT Composites
and its photocatalytic activity toward sudan (I) degradation. Photochem. Photobiol.
2016
,92, 523–527.
[CrossRef]
84.
Xu, Y.G.; Liu, J.; Xie, M.; Jing, L.Q.; Xu, H.; She, X.J.; Li, H.M.; Xie, J.M. Construction of novel CNT/LaVO
4
nanostructures for efficient antibiotic photodegradation. Chem. Eng. J. 2019,357, 487–497. [CrossRef]
85.
Wang, Y.J.; Shi, R.; Lin, J.; Zhu, Y.F. Significant photocatalytic enhancement in methylene blue degradation
of TiO
2
photocatalysts via graphene-like carbon in situ hybridization. Appl. Catal. B-Environ.
2010
,100,
179–183. [CrossRef]
86.
Mamaghani, A.H.; Haghighat, F.; Lee, C.-S. Hydrothermal/solvothermal synthesis and treatment of TiO
2
for photocatalytic degradation of air pollutants: Preparation, characterization, properties, and performance.
Chemosphere 2019,219, 804–825. [CrossRef]
87.
Zhang, H.; Lv, X.J.; Li, Y.M.; Wang, Y.; Li, J.H. P25-Graphene composite as a high performance photocatalyst.
ACS Nano 2010,4, 380–386. [CrossRef]
88.
Isari, A.A.; Payan, A.; Fattahi, M.; Jorfi, S.; Kakavandi, B. Photocatalytic degradation of rhodamine B
and real textile wastewater using Fe-doped TiO
2
anchored on reduced graphene oxide (Fe-TiO
2
/rGO):
Characterization and feasibility, mechanism and pathway studies. Appl. Surf. Sci.
2018
,462, 549–564.
[CrossRef]
89.
Nasr, M.; Eid, C.; Habchi, R.; Miele, P.; Bechelany, M. Recent progress on titanium dioxide nanomaterials for
photocatalytic applications. ChemSusChem 2018,11, 3023–3047. [CrossRef]
90.
Giberman, D. Against zero-dimensional material objects (and other bare particulars). Philos. Stud.
2012
,160,
305–321. [CrossRef]
91.
Vaiano, V.; Sacco, O.; Sannino, D.; Ciambelli, P. Nanostructured N-doped TiO
2
coated on glass spheres for
the photocatalytic removal of organic dyes under UV or visible light irradiation. Appl. Catal. B-Environ.
2015
,
170, 153–161. [CrossRef]
92.
Chen, C.C.; Jaihindh, D.; Hu, S.H.; Fu, Y.P. Magnetic recyclable photocatalysts of Ni-Cu-Zn
ferrite@SiO
2
@TiO
2
@Ag and their photocatalytic activities. J. Photochem. Photobiol. A Chem.
2017
,334,
74–85. [CrossRef]
93.
Somiya, S.; Roy, R. Hydrothermal synthesis of fine oxide powders. Bull. Mater. Sci.
2000
,23, 453–460.
[CrossRef]
94.
Wu, G.S.; Wang, J.P.; Thomas, D.F.; Chen, A.C. Synthesis of F-doped flower-like TiO
2
nanostructures with
high photoelectrochemical activity. Langmuir 2008,24, 3503–3509. [CrossRef]
95.
Jing, L.Q.; Xu, Y.G.; Huang, S.Q.; Xie, M.; He, M.Q.; Xu, H.; Li, H.M.; Zhang, Q. Novel magnetic
CoFe
2
O
4
/Ag/Ag
3
VO
4
composites: Highly efficient visible light photocatalytic and antibacterial activity.
Appl. Catal. B-Environ. 2016,199, 11–22. [CrossRef]
96.
Cho, I.S.; Lee, C.H.; Feng, Y.Z.; Logar, M.; Rao, P.M.; Cai, L.L.; Kim, D.R.; Sinclair, R.; Zheng, X.L.
Codoping titanium dioxide nanowires with tungsten and carbon for enhanced photoelectrochemical
performance. Nat. Commun. 2013,4, 8. [CrossRef]
97.
Yang, D.J.; Liu, H.W.; Zheng, Z.F.; Yuan, Y.; Zhao, J.C.; Waclawik, E.R.; Ke, X.B.; Zhu, H.Y. An Efficient
Photocatalyst Structure: TiO
2
(B) Nanofibers with a Shell of Anatase Nanocrystals. J. Am. Chem. Soc.
2009
,
131, 17885–17893. [CrossRef]
98.
Zhang, Y.; Park, S.J. Bimetallic AuPd alloy nanoparticles deposited on MoO
3
nanowires for enhanced
visible-light driven trichloroethylene degradation. J. Catal. 2018,361, 238–247. [CrossRef]
99.
Li, D.; Xia, Y.N. Electrospinning of nanofibers: Reinventing the wheel? Adv. Mater.
2004
,16, 1151–1170.
[CrossRef]
100.
Greiner, A.; Wendorff, J.H. Electrospinning: A fascinating method for the preparation of ultrathin fibres.
Angew. Chem. Int. Edit. 2007,46, 5670–5703. [CrossRef]
101.
Reneker, D.H.; Chun, I. Nanometre diameter fibres of polymer, produced by electrospinning. Nanotechnology
1996,7, 216–223. [CrossRef]
102.
Wu, X.H.; Si, Y.; Yu, J.Y.; Ding, B. Titania-based electrospun nanofibrous materials: A new model for organic
pollutants degradation. MRS Commun. 2018,8, 765–781. [CrossRef]
103.
Agarwal, S.; Greiner, A.; Wendorff, J.H. Functional materials by electrospinning of polymers.
Prog. Polym. Sci.
2013,38, 963–991. [CrossRef]
Catalysts 2019,9, 122 32 of 32
104.
Zhang, R.Z.; Wang, X.Q.; Song, J.; Si, Y.; Zhuang, X.M.; Yu, J.Y.; Ding, B. In situ synthesis of flexible
hierarchical TiO
2
nanofibrous membranes with enhanced photocatalytic activity. J. Mater. Chem. A
2015
,3,
22136–22144. [CrossRef]
105.
Panthi, G.; Park, M.; Kim, H.Y.; Park, S.J. Electrospun Ag-CoF doped PU nanofibers: Effective visible light
catalyst for photodegradation of organic dyes. Macromol. Res. 2014,22, 895–900. [CrossRef]
106.
Panthi, G.; Park, M.; Park, S.J.; Kim, H.Y. PAN Electrospun nanofibers reinforced with Ag
2
CO
3
nanoparticles:
Highly efficient visible light photocatalyst for photodegradation of organic contaminants in waste water.
Macromol. Res. 2015,23, 149–155. [CrossRef]
107.
Saud, P.S.; Pant, B.; Park, M.; Chae, S.H.; Park, S.J.; El-Newehy, M.; Al-Deyab, S.S.; Kim, H.Y. Preparation and
photocatalytic activity of fly ash incorporated TiO
2
nanofibers for effective removal of organic pollutants.
Ceram. Int. 2015,41, 1771–1777. [CrossRef]
108.
Seong, D.B.; Son, Y.R.; Park, S.J. A study of reduced graphene oxide/leaf-shaped TiO
2
nanofibers for
enhanced photocatalytic performance via electrospinning. J. Solid State Chem.
2018
,266, 196–204. [CrossRef]
109.
Zhang, Y.F.; Park, M.; Kim, H.Y.; Ding, B.; Park, S.J. In-situ synthesis of nanofibers with various ratios of
BiOCl
x
/BiOBr
y
/BiOI
z
for effective trichloroethylene photocatalytic degradation. Appl. Surf. Sci.
2016
,384,
192–199. [CrossRef]
110.
Zhang, D.Q.; Li, G.S.; Li, H.X.; Lu, Y.F. The development of better photocatalysts through composition- and
structure-engineering. Chem. Asian J. 2013,8, 26–40. [CrossRef]
111.
Chen, F.T.; Fang, P.F.; Gao, Y.P.; Liu, Z.; Liu, Y.; Dai, Y.Q. Effective removal of high-chroma crystal violet
over TiO
2
-based nanosheet by adsorption-photocatalytic degradation. Chem. Eng. J.
2012
,204, 107–113.
[CrossRef]
112.
Ma, X.D.; Jiang, D.L.; Xiao, P.; Jin, Y.; Meng, S.C.; Chen, M. 2D/2D heterojunctions of WO
3
nanosheet/K
+
Ca
2
Nb
3
O
10-
ultrathin nanosheet with improved charge separation efficiency for significantly
boosting photocatalysis. Catal. Sci. Technol. 2017,7, 3481–3491. [CrossRef]
113.
Wang, C.Y.; Zhang, X.; Song, X.N.; Wang, W.K.; Yu, H.Q. Novel Bi
12
O
15
Cl
6
photocatalyst for the degradation
of bisphenol a under visible-light irradiation. ACS Appl. Mater. Interfaces 2016,8, 5320–5326. [CrossRef]
114.
Zhang, Y.F.; Park, S.J. Fabrication and characterization of flower-like BiOI/Pt heterostructure with enhanced
photocatalytic activity under visible light irradiation. J. Solid State Chem. 2017,253, 421–429. [CrossRef]
115.
Jiang, W.J.; Zhu, Y.F.; Zhu, G.X.; Zhang, Z.J.; Chen, X.J.; Yao, W.Q. Three-dimensional photocatalysts with a
network structure. J. Mater. Chem. A 2017,5, 5661–5679. [CrossRef]
116.
Wan, W.C.; Zhang, R.Y.; Ma, M.Z.; Zhou, Y. Monolithic aerogel photocatalysts: A review. J. Mater. Chem. A
2018,6, 754–775. [CrossRef]
117.
Pierre, A.C.; Pajonk, G.M. Chemistry of aerogels and their applications. Chem. Rev.
2002
,102, 4243–4265.
[CrossRef]
118.
Dagan, G.; Tomkiewicz, M. Titanium dioxide aerogels for photocatalytic decontamination of aquatic
environments. J. Phys. Chem. 1993,97, 12651–12655. [CrossRef]
119.
Heiligtag, F.J.; Rossell, M.D.; Suess, M.J.; Niederberger, M. Template-free co-assembly of preformed Au and
TiO2nanoparticles into multicomponent 3D aerogels. J. Mater. Chem. 2011,21, 16893–16899. [CrossRef]
120.
Kailasam, K.; Epping, J.D.; Thomas, A.; Losse, S.; Junge, H. Mesoporous carbon nitride-silica
composites by a combined sol-gel/thermal condensation approach and their application as photocatalysts.
Energy Environ. Sci. 2011,4, 4668–4674. [CrossRef]
121.
Li, Z.D.; Wang, H.L.; Wei, X.N.; Liu, X.Y.; Yang, Y.F.; Jiang, W.F. Preparation and photocatalytic performance
of magnetic Fe
3
O
4
@TiO
2
core-shell microspheres supported by silica aerogels from industrial fly ash.
J. Alloys Compd. 2016,659, 240–247. [CrossRef]
122.
Jiang, W.J.; Liu, Y.F.; Wang, J.; Zhang, M.; Luo, W.J.; Zhu, Y.F. Separation-free polyaniline/TiO
2
3D hydrogel
with high photocatalytic activity. Adv. Mater. Interfaces 2016,3, 1500502. [CrossRef]
©
2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
... Advanced oxidative processes have recently been stated to be the best method for decomposing organic matter in water and wastewater [20][21][22].These processes rely on the production of very reactive radicals, such as hydroxyl groups, which oxidize organic compounds and break them down into innocuous products, like H 2 O and CO 2 [22][23][24]. In recent years, many photocatalytic semiconductor nanomaterials developed with the help of heterogeneous photocatalysts are used as photocatalysts for the degradation of toxic compounds in water [25,26]. ...
Article
Full-text available
In this study, Fe3O4@SiO2@TiO2-CPTS-HBAP (FST-CH) nanoparticle was prepared for the simultaneous adsorption and photocatalytic degradation of aromatic chemical pollutants (Rhodamine B dye) in aqueous solution. FST-CH nanoparticle was characterized using scanning electron microscopy (SEM), Fourier-transform infrared spectroscopy (FTIR), Energy Dispersive X-Ray (EDX) Fluorescence Spectrometer and X-Ray Diffraction (XRD) spectroscopy. The photocatalytic activity of rhodamine B dye (RhB) was evaluated with a Kerman UV 8/18 vertical roller photoreactor. About 56% of RhB in aqueous medium was adsorbed by FST-CH nanoparticles with only 45 min of stirring in the dark, and about 77.01% was degraded or converted to other structures under the photoreactor for 120 min. The photocatalytic degradation of RhB (apparent rate constant: 0.0026 mg dm⁻³ min⁻¹) occurred by a pseudo-second order reaction. In addition, the recovery of the prepared magnetic FST-CH nanoparticle by an external magnetic field, exhibiting good magnetic response and reusability, shows that the obtained magnetic FST-CH nanoparticle is stable and maintains high degradation ratio and catalyst recovery even after four cycles. Thus, the prepared FST-CH nanoparticle can be highly recommended for its use in potential applications of water decontamination.
... However, the wide bandgap of 3.2 eV restricts its reaction in the visible light spectrum, and the rapid recombination of photoexcited electron-hole pairs poses a challenge for applying TiO 2 as a photocatalyst [24,25]. To overcome the challenges associated with TiO 2 photocatalysts, many studies have applied methods such as impurity doping, heterojunction constructions, surface/microstructure modification, and sensitization [26][27][28][29][30][31][32]. However, it is difficult to find studies that irradiate electron beams on Cu-doped TiO 2 nanofibers. ...
Article
Full-text available
Titanium dioxide (TiO₂) is a widely studied material with many attractive properties such as its photocatalytic features. However, its commercial use is limited due to issues such as deactivation in the visible spectrum caused by its wide bandgap and the short lifetime of photo-excited charge carriers. To overcome these challenges, various modifications could be considered. In this study, we investigated copper doping and electron beam treatment. As-spun TiO2 nanofibers were fabricated by electrospinning a TiO2 sol, which obtained viscosity through a polyvinylpyrrolidone (PVP) matrix. Cu-doped TiO2 nanofibers with varying dopant concentrations were synthesized by adding copper salts. Then, the as-spun nanofibers were calcined for crystallization. To evaluate photocatalytic performance, a photodegradation test of methylene blue aqueous solution was performed for 6 h. Methylene blue concentration was measured over time using UV-Vis spectroscopy. The results showed that Cu doping at an appropriate concentration and electron-beam irradiation showed improved photocatalytic efficiency compared to bare TiO2 nanofibers. When the molar ratio of Cu/Ti was 0.05%, photodegradation rate was highest, which was 10.39% higher than that of bare TiO2. As a result of additional electron-beam treatment of this sample, photocatalytic efficiency improved up to 8.93% compared to samples without electron-beam treatment.
... UV light has been employed to break down the solute dye from the aqueous medium using a variety of semiconductor photocatalysts. Semiconductor oxide (e.g., TiO 2 , ZnO, ZrO 2 , SnO 2 , etc.), some semiconductors of sulfides (e.g., CdSe, MoS 2 , ZnS, CdS, etc.) can be utilized as photocatalysts used because of their low band gaps [9]. Due to their astonishing physical, optical, and optoelectronic properties, semiconductor oxides are among the materials that are most often studied [10,11]. ...
Article
Full-text available
The fast growth of human civilization with rapid industrialization causes severe detrimental effects, as a result, freshwater sources are depleting day by day threatening the future of human civilization to the water crisis. It is essential to preserve water by eliminating pollutants and conserving it. Dyes are one of such toxic pollutants which have been used in numerous industries like plastics, textiles, cosmetics, paper and pulp, leather, and food industries. The effectiveness of the photocatalytic removal of the dye malachite green was examined in the current study employing two distinct additives as photocatalysts: ZnO and ZnO–TiO2 nanoparticles. Nanosized ZnO and ZnO–TiO2 were prepared by utilizing microwave-assisted combustion and co-precipitation methods, respectively. The obtained metal oxide nanoparticles were characterized by XRD, FESEM, TGA, FTIR, and BET. The photocatalytic effects of the synthesized metal oxides were estimated in an aqueous solution by degrading malachite green (MG) dye. The dye solution with each metal oxide was subjected to irradiation under sunlight as well as UV light and monitored up to the stage of complete decolorization. The results indicated that ZnO–TiO2 exhibits the highest photocatalytic performance in comparison to ZnO and TiO2 nanoparticles in the presence of sunlight. Further, the effects of dose and pH variation were examined. The highest efficiency of the degradation of dye was found to be 92% at pH 5.8 under sunlight radiation.
Article
Full-text available
In this work, a simple sol-gel method was used to fabricate a ternary nanocomposite of Fe-doped TiO 2 decorated on reduced graphene oxide (Fe-doped TiO 2 /rGO). XRD, Raman shift, FT-IR, BET, DRS, EIS, TEM, FESEM, EDX and EDS techniques were applied for characterization of the structural, optical, and surface morphological properties of the synthesized catalysts. The DRS results of the photocatalysts showed a narrowing band gap by the introduction of Fe ions to the titania framework. The photocatalytic performance of the prepared samples was determined through the decontamination of rhodamine B under solar illumination. The optimum content of iron and graphene oxide and the effect of operational factors including pH, catalyst dosage and the initial concentration of rhodamine B were studied. The findings revealed that a 0.6 g Fe-doped TiO 2 /rGO nano-composite containing 3% Fe and 5% rGO, with an initial pH of 6 and rhodamine B concentration of 20 mg/L could achieve a removal of 91% after 120 min under solar illumination. TOC analyses were conducted to explore the rhodamine B mineralization rate; the data showed complete mineralization after 300 min. The effect of co-existing ions was examined and the results indicated that the degradation efficiency was significantly decreased by the addition of chloride and sulfate anions, although it slightly decreased in the presence of nitrate and phosphate anions. Furthermore, the addition of H 2 O 2 as an enhancer was investigated and the data demonstrated that the addition of 8 mM H 2 O 2 enhanced the photocatalytic efficacy to complete degradation. Finally, in the treatment of real textile wastewater, the concentrations of TOC and COD decreased from 930 mg/L and 1550 mg/L to 310 mg/L and 634 mg/L, respectively, after 390 min under similar operational conditions.
Article
Environmentally friendly photodegradation of refractory pollutants utilizing semiconductor photocatalysis evolution exhibits satisfactory efficiency and low energy consumption, but some of its unbefitting band position largely limits practical applications. The controllable band structure tuning via coupling low-dimensional materials with enormous potential in semiconductor photocatalysis, notably attractive carbon quantum dots (CQDs) induced for a better utilization of low-energy photoexcitation of promising inherent optical properties and photocatalytic performance on organic pollutant degradation. Here, a CQDs modified ZnO hybrid materials was synthesized and the CQDs preluded carbon-incorporated burger-like ZnO nanoparticles clusters were presented thus demonstrated a specific doping-effect. The dispersity improved by loading halloysite nanotube (HNTs) for efficiently photocatalytic activity enhancement. Furthermore, the band structure, excitation and photocatalytic reactive oxygen species (ROS) generation were detected to reveal the photochemical properties and plausible mechanism of CQDs hybrid nanomaterials system. It is observed that the superoxide radical (O2⁻), and singlet oxygen(¹O2) are the principal ROS agents under UV–vis excitations. This work not only displays a CQDs modified system of expanding photo-response range to visible light but also throws a way for quantum elemental incorporated on the influence of band-tuning for semiconductor- dominated photocatalysts.
Article
Photocatalytic oxidation (PCO) is a well-known technology for air purification and has been extensively studied for removal of many air pollutants. Titanium dioxide (TiO2) is the most investigated photocatalyst in the field of environmental remediation owed to its chemical stability, non-toxicity, and suitable positions of valence and conduction bands. Various preparation techniques including sol-gel, flame hydrolysis, water-in-oil microemulsion, chemical vapour deposition, solvothermal, and hydrothermal have been employed to obtain TiO2 materials. Hydro-/Solvothermal (HST) synthesis, focus of the present work, can be defined as a preparation method in which crystal growth occurs in a solvent at relatively low temperature (<200 °C) and above atmospheric pressure. This paper aims to provide a comprehensive and critical review of current knowledge regarding the application of HST synthesis for fabrication of TiO2 nanostructures for indoor air purification. TiO2 nanostructures are categorized from the morphological standpoint (e.g. nanoparticles, nanotubes, nanosheets, and hierarchically porous) and discussed in detail. The influence of preparation parameters including hydrothermal time, temperature, pH of the reaction medium, solvent, and calcination temperature on physical, chemical, and optical properties of TiO2 is reviewed. Considering the complex interplay among catalyst properties, a special emphasis is placed on elucidating the interconnection between various photocatalyst features and their impacts on photocatalytic activity.
Article
Crystal facet engineering of semiconductors has been demonstrated to be an important strategy to promote the separation of photogenerated electron-hole pair for photocatalytic activity, but the mechanism behind is still in debate. Here, density functional theory calculations are used to reveal the detailed property of surface heterojunction, one of popular concepts for charge separation, with three typical model systems of TiO2, Cu2WS4 and SrTiO3. The results demonstrate that the conduction band minima and valence band maxima among different facets indeed result in surface heterojunction, thermal-kinetically favorable for spatial charge separation. However, these surface heterojunctions, caused by the surface dangling bonds and the interaction between surface atoms, are only in few surface layers (about 1 nm) and may not be powerful enough to initiate the preferential flow of charges, and the other possible mechanism for charge separation should be additionally considered.
Article
Molybdenum trioxide (MoO3) has significant potential as an attractive semiconductor material owing to its low cost, nontoxicity, high electrochemical activity, and environmentally benign nature compared to other metal oxides. Nevertheless, its potential photocatalytic application has been hindered due to the high recombination rate of photogenerated electron-hole pairs, resulting in lower photocatalytic performance. In this work, a facile method was developed to synthesize SnS2/MoO3 hollow nanotubes. The as-prepared samples exhibit enhanced hydrogen peroxide production performance. A series of characterizations were conducted and the results prove that SnS2 nanosheets are uniformly dispersed on the surface of the hollow MoO3 nanotubes. Three-dimensional hetero-SnS2/MoO3 exhibited high charge transfer ability owing to its hollow structure and high surface area, which provides more active sites and improves the mobility of the radicals. Meanwhile, a synergistic effect between MoO3 and SnS2 is found to yield optimal hydrogen peroxide production performance. The as-prepared SnS2/MoO3 hollow nanotubes can generate best hydrogen peroxide production performance with good stability ( 95 % after four cycles).
Article
It is an urgent matter to eliminate antibiotics in waste water, due to the rapid emergence of antibiotic resistance. The search for low cost, high activity and stable novel photocatalysts has attracted great interest. Herein, we demonstrated the rational construction of CNT/LaVO4 nanostructures for efficient antibiotic photodegradation by a one-step hydrothermal method. The phase structures, chemical compositions, morphologies, and optical properties of the prepared samples were investigated via various characterization techniques. The optimized CNT/LaVO4 nanostructures exhibited efficient photodegradation activity with remarkable stability. The 0.1% CNT/LaVO4 showed the highest tetracycline degradation rate, which is 2 times that of pure LaVO4. Photoluminescence (PL), transient photocurrent response and electrochemical impedance spectroscopy (EIS) together verified that this design successfully expedites the separation and transfer of photogenerated charge carriers. Subsequently, by the electron spin resonance (ESR) spin-trap technique, free radical trapping experiments and mass spectrometry analysis (MS), the active species, intermediate product, photodegradation pathway and reaction mechanism during the photocatalytic process were identified. The antibacterial results showed that the degrading products have lower toxicity. The CNT/LaVO4 composite is a potential photocatalyst for improving the water quality.
Article
The pollution of pharmaceutical wastewater has attracted global attention. Photocatalysis is an attractive yet challenging method for the degradation of pharmaceutical residues. Fabricating efficient Z-scheme heterojunctions with intimate interface contact for enhancing the performance of photocatalysts is a challenge. Herein, novel visible-light-driven direct Z-scheme CuInS2/Bi2WO6 heterojunctions with intimate interface contact were successfully synthesized by in situ hydrothermal growth of Bi2WO6 directly on the surface of CuInS2 network-like microspheres, and the content of CuInS2 was optimized. The photocatalytic activity of optimal Z-scheme 15% CuInS2/Bi2WO6 for the degradation of tetracycline hydrochloride (TC•HCl) is more than three times that of bare CuInS2 and 17% higher than that of Bi2WO6, which is attributed that the intimate interface contact can assure excellent interfacial charge transfer abilities. Moreover, Z-scheme CuInS2/Bi2WO6 heterojunctions are highly stable with no inactivation in photocatalytic cycles. More importantly, real pharmaceutical industry wastewater can be efficiently disposed by Fenton-aided photocatalysis of Z-scheme CuInS2/Bi2WO6 heterojunctions with COD removal efficiency being 90.5%, which is much higher than the reported results, and thus sets a new performance benchmark for practical application in real pharmaceutical wastewater treatment.
Article
A novel ternary Ag2CO3/Ag/AgNCO Z-scheme heterojunction with excellent visible-light-driven photocatalytic performance was fabricated. The [rad]O2⁻ and, h⁺ were proved to be the photocatalytic active substances for the degradation of Rhodamine B (RhB) in Ag2CO3/Ag/AgNCO. Compared to pristine Ag2CO3 and AgNCO, the enhanced photocatalytic activities of Ag2CO3/Ag/AgNCO could be attributed to the low resistance for interfacial charge transfer and desirable absorption capability. In addition, the Z-scheme photocatalytic mechanism is proved by the discussion of the band structure. This work could offer a new insight into the design and fabrication of advanced materials with Z-scheme structures for the efficient treatment of persistent pollutants in wastewater.
Article
Environmental pollution by Petroleum refinery wastewater (PRW) raises vital call for attention to scientist and industrialists due to its impact on human and eco-system. This manuscript reviews methods on modification of semiconductors, effect of operating parameters, reusability/stability, along with recent development on TiO2- and ZnO-based photocatalysts towards degradation of PRW pollutants. TiO2 and ZnO can be successfully modified to visible/solar light responsive photocatalysts. In-depth knowledge of reaction mechanism between photocatalysts and sorbates, structure formation in modified photocatalysts with respect to generation/mobilization of photo-charged carriers, informs the choice of composition of photocatalysts. Petroleum refineries are yet to benefit from heterogeneous photocatalysis due to some setbacks such as: recovery of photocatalysts for reuse, inability of the catalyst to degrade high concentration of pollutants, inability to handle the complex nature of PRW, and its reusability. Thus, a multifunctional and cost effective photocatalyst with reusability/stability needs more exploration to enable commercialization of the photocatalysts.