You are on page 1of 15

Powder Technology 410 (2022) 117886

Contents lists available at ScienceDirect

Powder Technology
journal homepage: www.journals.elsevier.com/powder-technology

Hydrothermal synthesis of Nb5+-doped SrTiO3 mesoporous nanospheres


with greater photocatalytic efficiency for Cr(VI) reduction
Jing Zhu a, Yongcai Zhang a, *, Li Shen b, Jing Li c, *, Liangliang Li d, Fen Zhang e, Ya Zhang f, *
a
School of Chemistry and Chemical Engineering, Yangzhou University, Yangzhou 225009, China
b
Guangling College, Yangzhou University, Yangzhou 225009, China
c
School of Materials and Chemical Engineering, Xuzhou University of Technology, Xuzhou 221018, China
d
Jiangsu Weiming New Material Co., Ltd, Yangkougang Economic Development Zone, Changsha Town, Rudong County, Nantong City, Jiangsu Province 226400, China
e
School of Chemistry and Chemical Engineering, Liaocheng University, Liaocheng 252000, China
f
School of Environmental Science and Engineering, Yangzhou University, Yangzhou 225009, China

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

• Nb5+-doped SrTiO3 mesoporous nano­


spheres were synthesized by a hydro­
thermal method.
• 3% Nb-SrTiO3 has far higher photo­
catalytic activity than SrTiO3 in Cr(VI)
reduction.
• It outperforms many other metal ions
doped SrTiO3 and many reported
photocatalysts.
• It has good photocatalytic stability.
• Its higher activity mainly results from
enhanced separation and transfer of e−
and h+.

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

Keywords: The Nb5+-doping strategy was adopted to enhance the photocatalytic efficiency of SrTiO3 to overcome the
SrTiO3 limitation for its practical photocatalytic applications. Different amounts of Nb5+ doped SrTiO3 (Nb-SrTiO3)
Nb5+-doping mesoporous nanospheres were synthesized via a convenient hydrothermal approach. The photocatalytic
Mesoporous nanomaterials
reduction of aqueous Cr(VI) using the Nb-SrTiO3 products under xenon lamp irradiation was investigated. Re­
Hydrothermal synthesis
Photocatalytic reduction
sults indicated that 3% Nb-SrTiO3 was the optimum product, which exhibited a photocatalytic efficiency 3 times
Cr(VI) as much as that of SrTiO3 and a fair reuse stability. Moreover, 3% Nb-SrTiO3 also demonstrated higher photo­
catalytic efficiency than many other metal cations doped SrTiO3 synthesized via the same process and a number
of previously reported photocatalysts. The best photocatalytic activity of 3% Nb-SrTiO3 was found to be pre­
dominantly due to its most efficient separation and transfer of photogenerated holes and electrons. This study
suggests that Nb-SrTiO3 is a potential efficient photocatalyst for the remediation of Cr(VI)-contaminated water.

* Corresponding authors.
E-mail addresses: zhangyc@yzu.edu.cn (Y. Zhang), ljshan@xzit.edu.cn (J. Li), zhangya@yzu.edu.cn (Y. Zhang).

https://doi.org/10.1016/j.powtec.2022.117886
Received 12 July 2022; Received in revised form 20 August 2022; Accepted 23 August 2022
Available online 27 August 2022
0032-5910/© 2022 Elsevier B.V. All rights reserved.
J. Zhu et al. Powder Technology 410 (2022) 117886

1. Introduction photoresponse range of SrTiO3, and/or promote the photogenerated


charge separation of SrTiO3 [10,21–32]. As a result, the photocatalytic
Heavy metal pollution in water bodies has become a serious concern. efficiency of SrTiO3 can often be enhanced by element doping [21–32].
Chromium-containing compounds are widely used in electroplating, So far, a number of metal cations (such as Bi3+ [21], Eu3+ [22], Ce3+ and
metallurgy, leather, anticorrosion, coloring, and chemical Ce4+ [23], Na+ [24], Cr3+ [25,26], V4+ [27], K+ [28], La3+ and Fe3+
manufacturing, etc., making Cr(VI) one of the main heavy metal pol­ [29], Cu2+ [30], Al3+ [31], and La3+ [32]) have been tried to dope
lutants in water bodies. Cr(VI) is the most harmful form of chromium, SrTiO3 to improve its photocatalytic efficiency. However, the effect of
due to its great toxicity, high solubility, and strong mobility [1–3]. There metal cation doping on the photocatalytic efficiency of SrTiO3 depends
is a risk of getting cancer and other diseases from prolonged exposure to on the species of the doped metal cation, the content of the dopant, and
Cr(VI) [4]. Hence, Cr(VI) pollution in water bodies seriously threatens the synthesis method and synthesis conditions, etc. There are still many
human health, and its control and treatment should be given priority unsolved problems concerning the effect of metal cation doping. For
[5]. So far, many technologies for the decontamination of aqueous Cr example, it is hard to predict that which metal cation, how much content
(VI) have been reported, including biological reduction, chemical pre­ of the dopant, and what synthesis method and synthesis conditions can
cipitation, ion exchange, adsorption, chemical reduction, and photo­ achieve the maximum improvement of the photocatalytic efficiency of
catalytic reduction [5–10]. Among the many reported Cr(VI) treatment SrTiO3. The solution of these problems requires further experimental
technologies, photocatalytic reduction can achieve low cost and envi­ investigation and mechanism exploration.
ronmental friendliness at the same time, because it can recycle the Nb5+-doping has already been adopted to augment the photo­
photocatalyst, produce no secondary pollution, and utilize directly the catalytic activity of TiO2, because the radii of Nb5+ (64 pm) and Ti4+
natural sunlight energy to drive the Cr(VI) reduction reactions [6–10]. (60.5 pm) are relatively close [38–40]. Most studies showed that the
As a result, the photocatalytic reduction treatment of Cr(VI) pollutant in Nb5+-doping can narrow the band gap, improve the electrical conduc­
water has received wide attention recently [6–10]. However, up to date, tivity and suppress the h+-e− recombination of TiO2, resulting in the
the lack of high-performance, inexpensive, and environmentally increased photocatalytic activity of TiO2 [38–40]. Besides, Nb5+-doping
friendly photocatalysts still impedes the industrial application of the has also been employed to ameliorate the electrical, photoelectric and
desirable photocatalytic technology [11–16]. other properties of SrTiO3 [41–43]. However, as far as we know, the
Recently, perovskite structure SrTiO3 has been frequently studied as information on the photocatalytic property of Nb5+-doped SrTiO3 (Nb-
an independent photocatalyst or a component of composite photo­ SrTiO3) is still lacking till now. Considering that Nb5+ has a radius close
catalyst in various fields of photocatalytic applications, due to its many to that of Ti4+ and the positive effect of Nb5+-doping on the photo­
inherent advantages [17–37]. SrTiO3 is a n-type semiconductor material catalytic activity of TiO2 and the electrical and photoanode performance
with a band gap of 3.2–3.4 eV [17–19]. It is cost-effective, non-toxic and of SrTiO3 [38–43], Nb5+ is likely an effective dopant for enhancing the
resistant to photocorrosion, and has good durability [17–19]. Compared photocatalytic efficiency of SrTiO3.Therefore, this work aimed to
with the most studied TiO2 photocatalyst, SrTiO3 has a more negative explore the feasibility of synthesizing Nb-SrTiO3 as a high-performance
conduction band [20], which is favorable for the photocatalytic reduc­ photocatalyst. Nb-SrTiO3 mesoporous nanospheres were prepared by a
tion reactions. In addition, SrTiO3 possesses good structural tolerance convenient hydrothermal approach, and their photocatalytic reduction
for enhancing its photocatalytic efficiency through A-site or B-site of aqueous Cr(VI) under xenon lamp irradiation was researched. The
doping [21–32]. However, SrTiO3 itself has great limitation for photo­ effect of the content of Nb5+ dopant on the photocatalytic activity of Nb-
catalytic applications due to its relatively low photocatalytic efficiency, SrTiO3 was investigated, and the mechanism behind the improved
which mainly results from the high recombination rate of its photo­ photocatalytic efficiency of Nb-SrTiO3 was probed. Furthermore, to
generated holes (h+) and electrons (e− ) [21–38]. In order to raise the demonstrate the advantages of Nb-SrTiO3 for use as a photocatalyst, its
photocatalytic efficiency of SrTiO3, many modification strategies have photocatalytic activity was also compared with those of many other
been explored, such as element doping [10,21–32], heterojunction metal cations (including Nd3+, Zr4+, Cu2+, La3+, Sn4+ and Co2+) doped
compositing [33,34], and combination with noble metals or carbon SrTiO3 synthesized via the same process and a number of previously
materials [35–37]. As an often used modification strategy, element reported SrTiO3-based and non-SrTiO3-based photocatalysts.
doping can introduce the dopant energy level between the conduction
band and valance band of SrTiO3 [10,21–32]. This can expand the

Fig. 1. (a) XRD patterns of SrTiO3, 1% Nb-SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3 from bottom to top; (b) Enlarged images of the (110) diffraction peaks of SrTiO3,
1% Nb-SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3 from bottom to top.

2
J. Zhu et al. Powder Technology 410 (2022) 117886

Fig. 2. XPS spectra of SrTiO3 and 3% Nb-SrTiO3.

2. Material and methods resulting solution was marked as A-solution. 10 mmol SrCl2⋅6H2O and 0,
0.1, 0.3 or 0.5 mmol NbCl5 were added in 10 mL 2 mol⋅L− 1 KOH aqueous
2.1. Material solution, and stirred for 1 h, the resulting solution was marked as B-
solution. B-solution was dropped slowly into A-solution under stirring,
The information of the reagents used is given in Table S1 in the and the mixed solution was further stirred for 0.5 h. The mixed solution
Supporting information. was sealed in a 50 mL Teflon-lined stainless steel autoclave, and heated
at 200 ◦ C for 8 h. After the autoclaves naturally cooled to room tem­
perature, the formed products were centrifugally separated, washed in
2.2. Synthesis
turn with 1 mol⋅L− 1 HNO3 and deionized water for at least three times,
and dried at 100 ◦ C for 8 h. For convenience, the products synthesized
3.4 mL tetrabutyl titanate was added to the mixed solvent of 10 mL
employing 0, 0.1, 0.3 and 0.5 mmol NbCl5 were named as SrTiO3, 1%
ethylene glycol and 10 mL deionized water and stirred for 1 h, the

3
J. Zhu et al. Powder Technology 410 (2022) 117886

Fig. 3. EDX elemental mapping photographs of 3% Nb-SrTiO3.

Nb-SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3, respectively. (VI)) was added and stirred to mix them well, then 300 mg photocatalyst
For comparison, other metal cations (including Nd3+, Zr4+, Cu2+, was added. The mixture was magnetically stirred under the dark envi­
La , Sn4+ and Co2+) doped SrTiO3 were also prepared via the same
3+
ronment for 75 min to ensure the adsorption-desorption equilibrium
processes, except that 0.3 mmol Nd2O3, ZrCl4, CuCl2⋅2H2O, LaCl3, between Cr(VI) and photocatalyst. Subsequently, the xenon lamp was
SnCl4⋅5H2O, or CoCl2 replaced 0.3 mmol NbCl5. turned on. At every 15 min irradiation, nearly 4 mL of the reaction so­
lution was suctioned, and filtered through two layers of cellulose acetate
2.3. Characterization and photocatalytic investigation membranes (the pore size is 0.22 μm) for the removal of the photo­
catalyst. The diphenylcarbazide spectrophotometry was used for the
The characterization of the synthesized products is detailed in the determination of the Cr(VI) concentrations in the filtrates.
Supporting information. The photocatalytic experiments were carried
out using a GHX-2 type photocatalysis device (Yangzhou University
Town Technology Co. Ltd). 300 mL K2Cr2O7 aqueous solution (20
mg⋅L− 1) was added to the reaction bottle, and 1 mL citric acid (CA)
aqueous solution (the mole number of CA is 6 times as much as that of Cr

4
J. Zhu et al. Powder Technology 410 (2022) 117886

Fig. 4. SEM images of (a) SrTiO3 and (b) 3% Nb-SrTiO3.

3. Results and discussion synthesized Nb-SrTiO3 products are free of impurities such as SrCO3,
TiO2, and Nb2O5. The cubic perovskite structure of SrTiO3 has been
3.1. Composition and structure analysis retained after the doping of Nb5þ. Nevertheless, in Fig. 1(b), it can be
noticed that the (110) diffraction peak of SrTiO3 shifts to the smaller 2θ
XRD was employed to analyze the phase of the prepared products. position with the doping of Nb5+. Moreover, the positional shift of the
The XRD spectra of SrTiO3, 1% Nb-SrTiO3, 3% Nb-SrTiO3, and 5% Nb- (110) diffraction peak of Nb-SrTiO3 becomes bigger with the increase of
SrTiO3 are respectively recorded in Fig. 1(a) from bottom to top. As seen Nb5+-doping amount. The shift of the diffraction peak positions of Nb-
in Fig. 1(a), all the diffraction peaks exhibited by the four samples SrTiO3 relative to those of SrTiO3 is likely because that the radius of
belong to cubic phase SrTiO3 (JCPDS Card No. 35–0734). Thus, the as- Nb5þ is bigger than that of Ti4+ (the radii of Nb5+ and Ti4+ are 64 pm

5
J. Zhu et al. Powder Technology 410 (2022) 117886

Fig. 5. TEM images of (a) SrTiO3 and (b) 3% Nb-SrTiO3; HRTEM images of (c) SrTiO3 and (d) 3% Nb-SrTiO3.

and 60.5 pm, respectively) [38–40]. According to the Bragg equation Sr 3d XPS spectra of SrTiO3 and 3% Nb-SrTiO3 can be fitted into the two
(2dsinθ = nλ), when Nb5þ cations occupy the positions of Ti4+ cations, peaks at about 134.6 eV (Sr 3d3/2) and 132.9 eV (Sr 3d5/2), which are in
the crystal unit cell of SrTiO3 becomes larger and its d-spacings increase, correspondence to Sr2+ [25–28]. The Ti 2p XPS spectra of SrTiO3 and 3%
leading to the decrease of the 2θ value. Because Nb5þ cations are doped Nb-SrTiO3 can be fitted into the two peaks at approximately 458.3 eV (Ti
in the lattice of SrTiO3, the as-synthesized Nb-SrTiO3 can be considered 2p3/2) and 464.0 eV (Ti 2p1/2), which indicate that Ti is of positive
as a form of solid solution as SrTi1-xNbxO3. tetravalence [25–28]. The O 1 s XPS spectra of SrTiO3 and 3% Nb-SrTiO3
The surface compositions of SrTiO3 and 3% Nb-SrTiO3 were deter­ can be fitted into the two peaks at about 529.4 eV and 530.7 eV, which
mined by means of XPS, and the results obtained are depicted in Fig. 2. are well matched with the lattice oxygen (OL) and surface hydroxyl
The survey spectra indicate that compared with SrTiO3, 3% Nb-SrTiO3 oxygen (OOH) [25–28], respectively. The Nb 3d XPS spectrum of 3% Nb-
contains one more element, Nb. The C element in both the samples is SrTiO3 can be fitted into the two peaks at 206.2 eV (Nb 3d5/2) and 208.9
associated with the adventitious carbon contaminant. The C 1 s XPS eV (Nb 3d3/2), suggesting that the doped Nb exists in the form of Nb5+
spectra can be fitted into the three peaks at about 284.6 eV, 286.0 eV and [38]. According to the previous reports [38–40,43], the doping of Nb5+
288.6 eV, which correspond to C–C, C–O, and C– – O, respectively. The into SrTiO3 may lead to two consequences due to the compensation for

6
J. Zhu et al. Powder Technology 410 (2022) 117886

Fig. 6. (a) N2 adsorption-desorption isotherms and (b) Pore size distribution plots of SrTiO3 and 3% Nb-SrTiO3.

4+
the extra positive charge of Nb+Ti: (i) Ti cations are reduced to Ti3+ above XRD, XPS and EDX elemental mapping characterization results
4+
cations; (ii) one Ti cation vacancy is generated by introducing every jointly substantiate the successful synthesis of Nb-SrTiO3.
four Nb5+ or the (NbTi-VTi)3− complex is formed [38], or the Sr defi­ The morphology and size of SrTiO3 and 3% Nb-SrTiO3 were revealed
ciency is created [43]. Considering that the Nb5+-doping causes no by means of SEM and TEM. From the SEM images of SrTiO3 and 3% Nb-
alteration of the valence state of Ti4+ in SrTiO3, our synthesized Nb- SrTiO3 displayed respectively in Fig. 4(a) and (b), it is observed that
SrTiO3 products may belong to the second scenario. both samples appear to be nanospheres with the sizes of about 45–140
The EDX analysis reveals that the actual doping contents of Nb (atom ± 5 nm. The doping of Nb5+ causes no remarkable change in the
%) in 1% Nb-SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3 are 0.8%, 3.1% morphology of SrTiO3. The nanospheres seem to be formed from the
and 5.5%, respectively. In addition, the EDX elemental mapping pho­ aggregation of primary nanoparticles. The primary particle sizes of
tographs of 3% Nb-SrTiO3 in Fig. 3 indicate that the Sr, Ti, O and Nb SrTiO3 and 3% Nb-SrTiO3 were estimated to be in turn 16 nm and 21 nm
component elements are evenly and coincidentally distributed in the 3% based on the peak widths at half height of their (110) diffraction peaks,
Nb-SrTiO3 product, confirming its solid solution structure. Thus, the by employing the Scherrer formula [44]. The morphology and size of

7
J. Zhu et al. Powder Technology 410 (2022) 117886

Fig. 7. (a) Comparison of the performance of SrTiO3,


1% Nb-SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3 in
the adsorption and photocatalytic reduction of
aqueous Cr(VI); (b) Determination of the k values for
the photocatalytic Cr(VI) reduction reactions in (a)
via the ln(Ci0/Cit) vs. ti plots. Note: C0 is the initial
concentration of the Cr(VI) solution used, Ct is the
concentration of the Cr(VI) solution after adsorption
and photocatalytic reaction for t min, ti is the irradi­
ation time, and Ci0 and Cit represent the concentration
of Cr(VI) in the solution after being irradiated for
0 and t min, respectively.

SrTiO3 and 3% Nb-SrTiO3 observed from their TEM images (Fig. 5(a) sizes of SrTiO3 and 3% Nb-SrTiO3 are centered on 3.5 nm and 3.3 nm,
and (b)) are in agreement with those observed from their SEM images. respectively. The BET specific surface area and total pore volume of
Besides, according to Fig. 5(c) and (d), the HRTEM images of both SrTiO3 were obtained to be in turn 256.1 m2/g and 0.43 cm3/g, whereas
SrTiO3 and 3% Nb-SrTiO3 show the lattice fringes with the adjacent those of 3% Nb-SrTiO3 were obtained to be 224.9 m2/g and 0.37 cm3/g,
spacing being 0.275–0.276 nm, which corresponds to the (110) crystal respectively. Apparently, the specific surface area and total pore volume
plane of cubic perovskite structure SrTiO3. of 3% Nb-SrTiO3 are smaller than those of SrTiO3. The specific surface
The specific surface areas and pore structures of SrTiO3 and 3% Nb- areas and total pore volumes of SrTiO3 and 3% Nb-SrTiO3 are in the
SrTiO3 were analyzed based on their N2 adsorption-desorption iso­ same order as their Cr(VI) adsorption capacities, but contrary to the
therms. According to Fig. 6(a), both the samples show type-IV isotherms, order of their photocatalytic activities (Fig. 7). These results suggest that
which are indicative of their mesoporous structures [45–47]. From the the specific surface areas and total pore volumes of SrTiO3 and 3% Nb-
pore size distribution plots in Fig. 6(b), it can be observed that the pore SrTiO3 can strongly affect their Cr(VI) adsorption capacities.

8
J. Zhu et al. Powder Technology 410 (2022) 117886

Fig. 8. Performance of 3% Nb-SrTiO3 in the photocatalytic reuse tests.

Nevertheless, the superior photocatalytic activity of 3% Nb-SrTiO3 is the samples approach 1. The k values obtained for SrTiO3, 1% Nb-
hard to be explained by its smaller specific surface area and total pore SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3 are 0.020, 0.029, 0.060 and
volume. Therefore, there are other greater influencing factors (such as 0.039 min− 1, respectively. The half-life (t1/2) values for the photo­
more efficient separation and transfer of photogenerated charges) catalytic reduction of Cr(VI) over SrTiO3, 1% Nb-SrTiO3, 3% Nb-SrTiO3
contributing to the superior photocatalytic activity of 3% Nb-SrTiO3. and 5% Nb-SrTiO3 were calculated to be 4.61, 4.23, 3.51 and 3.94 min,
respectively, according to the formula of t1/2 = ln2 k [49]. Clearly, the
3.2. Photocatalytic performance evaluation doping of Nb5+ can augment the photocatalytic activity of SrTiO3. The
photocatalytic activity of the most efficient 3% Nb-SrTiO3 is 3.0 times as
Fig. 7(a) shows a comparison of the performance of SrTiO3, 1% Nb- that of SrTiO3. Moreover, the photocatalytic activity of 3% Nb-SrTiO3 is
SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3 in the adsorption and photo­ higher than those of Nd3+, Zr4+, Cu2+, La3+, Sn4+ and Co2+ doped
catalytic reduction of aqueous Cr(VI). In addition, the photolysis result SrTiO3 synthesized via the same process (Fig. S1). In addition, 3% Nb-
in the absence of any photocatalyst sample is also given in Fig. 7(a). SrTiO3 is also more efficient than many previously reported SrTiO3-
Apparently, without the addition of any photocatalyst sample, the Cr based and non-SrTiO3-based photocatalysts in the photocatalytic
(VI) concentration in the solution hardly changed with time under both reduction of Cr(VI) (Table S2).
dark and irradiated conditions, indicating that the reduction of Cr(VI) by The reuse stability of a photocatalyst is crucial to its industrial
CA and photolysis is negligible. The adsorption equilibrium between application. Hence, further research was conducted to investigate the
each photocatalyst sample and Cr(VI) can be reached after mixing for 60 reuse stability of 3% Nb-SrTiO3 in the photocatalytic reduction of
min in dark environment, because the concentration of Cr(VI) in the aqueous Cr(VI). The conditions of the photocatalytic reuse experiments
solution remained unchanged when the mixing time was further were identical to those of the above photocatalytic activity assessment.
increased to 75 min. SrTiO3 exhibited obviously larger adsorption ca­ Each cycle comprised 75 min of adsorption in dark environment fol­
pacity for Cr(VI), compared with all the Nb-SrTiO3 samples. The lower lowed by 75 min of photocatalytic reduction under Xe lamp irradiation.
Cr(VI) adsorption capacities of the Nb-SrTiO3 samples are likely due to When one cycle ended, 3% Nb-SrTiO3 was recovered via centrifugation,
their decreased specific surface areas, which would be bound to reduce washed using deionized water for three times, and dried in an Electro-
the contact area between Nb-SrTiO3 and Cr(VI) solution. On the other thermostatic blast oven at 100 ◦ C for 8 h, then used for the following
hand, all the Nb-SrTiO3 samples exhibited much higher efficiencies than cycle. Fig. 8 shows the performance of 3% Nb-SrTiO3 in the photo­
SrTiO3 in the photocatalytic reduction of Cr(VI) when subjected to Xe catalytic reuse tests. The photocatalytic activity of 3% Nb-SrTiO3
lamp irradiation. After 75 min irradiation, the concentration of Cr(VI) in decreased with the rise in the cyclic number, but such decreases are very
the solution decreased to virtually zero when 3% Nb-SrTiO3 was minor, for example, above 93% of Cr(VI) in the solution is still reduced
employed as the photocatalyst, but 21.6% of Cr(VI) was still left in the in the fourth cycle. Besides, according to Fig. S2 and Fig. 9, the XRD
solution when SrTiO3 was used as the photocatalyst. In order to make a pattern and the Sr 3d, Ti 2p, Nb 3d and O 1 s XPS spectra of 3% Nb-
quantitative comparison of the photocatalytic activities of SrTiO3, 1% SrTiO3 after the photocatalytic reuse tests remained hardly changed
Nb-SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3, the reaction rate constant compared with those of the fresh 3% Nb-SrTiO3. This indicates that the
(k) values of their photocatalytic reduction of Cr(VI) were determined. crystal phase and elemental valences of 3% Nb-SrTiO3 have been
This was accomplished by utilizing the ln(Ci0/Cit) vs. ti plots (Fig. 7(b)) maintained during the photocatalytic reuse tests. Therefore, 3% Nb-
drawn based on the pseudo-first-order reaction kinetics [48]. According SrTiO3 has a fair reuse stability for photocatalytic application. Never­
to Fig. 7(b), the photocatalytic reduction of Cr(VI) over SrTiO3, 1% Nb- theless, the XPS survey spectrum of 3% Nb-SrTiO3 after the photo­
SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3 follows the pseudo-first-order catalytic reuse tests shows the presence of an additional element, Cr
reaction kinetics, because the correlation coefficient (R2) values for all

9
J. Zhu et al. Powder Technology 410 (2022) 117886

Fig. 9. XPS spectra of 3% Nb-SrTiO3 (a) before and (b) after the photocatalytic reuse tests.

(Fig. 9). The Cr 2p XPS spectrum was fitted into the two peaks located at during the reuse tests.
577.3 eV (Cr 2p3/2) and 585.5 eV (Cr 2p1/2) (Fig. 9), which correspond
to the Cr(III) species such as Cr(OH)3 [50]. So, Cr(VI) was reduced to Cr
3.3. Investigation on the photocatalytic activity enhancement mechanism
(III) via the photocatalysis of 3% Nb-SrTiO3. The deposition of Cr(OH)3
of Nb-SrTiO3
on the surface of 3% Nb-SrTiO3 would take up the surface active sites,
and is a main factor leading to the decrease of its photocatalytic activity
Generally, the photocatalytic activity of a photocatalyst material is

10
J. Zhu et al. Powder Technology 410 (2022) 117886

Fig. 10. (a) UV–Vis diffuse reflection spectra of SrTiO3, 1% Nb-SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3; (b) Determination of their Eg values using the Tauc
plot approach.

dependent on its specific surface area, light-absorbing ability, and sep­ light absorption particularly in the range of about 400–500 nm. The
aration and transfer efficiency of photogenerated carriers, etc. Consid­ band gap (Eg) values of various Nb-SrTiO3 samples and SrTiO3 were
ering that the orders of the specific surface areas and photocatalytic further determined according to the Tauc plot approach [51,52]. By
activities of SrTiO3 and 3% Nb-SrTiO3 are opposite (Fig. 6 and Fig. 7), extrapolation of the linear parts of the (αhν)2 vs. (hν) plots to the point
other factors (rather than the specific surface area) should play greater where (αhν)2 = 0 (Fig. 10(b)), the Eg values of SrTiO3, 1% Nb-SrTiO3,
roles in the superior photocatalytic activity of 3% Nb-SrTiO3. 3% Nb-SrTiO3 and 5% Nb-SrTiO3 were obtained to be 3.34, 3.32, 3.31
Fig. 10(a) displays the UV–vis diffuse reflection spectra of SrTiO3 and and 3.28 eV, respectively. The slightly decreased Eg of the Nb-SrTiO3
various Nb-SrTiO3 samples. It is observed that compared with the samples enable them to absorb a little more light, which may contribute
undoped SrTiO3, the Nb-SrTiO3 samples demonstrate increased visible- to their greater photocatalytic efficiencies than that of SrTiO3. However,

11
J. Zhu et al. Powder Technology 410 (2022) 117886

Fig. 11. (a) Transient photocurrent response and (b) Electrochemical impedance spectra of SrTiO3, 1% Nb-SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3.

for the Nb-SrTiO3 samples (1% Nb-SrTiO3, 3% Nb-SrTiO3 and 5% Nb- The transient photocurrent response and electrochemical impedance
SrTiO3), their light absorption capacities are not in the same order as spectroscopy were employed to compare the photogenerated carrier
their photocatalytic activities, for example, 5% Nb-SrTiO3 has the separation and transfer efficiencies of SrTiO3 and various Nb-SrTiO3
smallest Eg value or the greatest light absorption ability, but 3% Nb- samples. Fig. 11(a) shows the transient photocurrent response spectra of
SrTiO3 has the highest photocatalytic activity. Hence, rather than the SrTiO3, 1% Nb-SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3. It is observed
light absorption ability, there are other major factors (such as the sep­ from Fig. 11(a) that all the Nb-SrTiO3 samples exhibited far stronger
aration and transfer efficiencies of h+ and e− ) contributing to the photocurrent response signals than that of SrTiO3, with the photocurrent
greatest photocatalytic activity of 3% Nb-SrTiO3. response signal of 3% Nb-SrTiO3 being the strongest. Commonly, the

12
J. Zhu et al. Powder Technology 410 (2022) 117886

Fig. 12. Mechanism for the photocatalytic reduction of Cr(VI) by 3% Nb-SrTiO3.

greater intensity of transient photocurrent response means the higher semiconductors are typically 0.1 V more negative than their flat-band
separation efficiency of h+ and e− [53–55]. Thus, compared to the potentials. Thus, the CB potentials of SrTiO3 and 3% Nb-SrTiO3 are in
undoped SrTiO3, all the Nb-SrTiO3 samples should have much higher turn − 0.40 V and − 0.26 V vs. NHE. Correspondingly, the valence band
efficiencies in separating h+ and e− . Thereinto, 3% Nb-SrTiO3 exhibited (VB) potentials of SrTiO3 and 3% Nb-SrTiO3 were calculated to be
the biggest photocurrent signal, so it is most efficient in separating h+ +2.94 V and + 3.05 V vs. NHE, respectively, by utilizing the equation of
and e− . The order of the photogenerated carrier separation efficiencies EVB = ECB + Eg. It is widely shared that heterogeneous photocatalytic
of SrTiO3, 1% Nb-SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3 is identical reduction of Cr(VI) to Cr(III) normally takes place through three suc­
to that of their photocatalytic activities. Hence, the enhanced photo­ cessive one-electron reduction steps, with the first electron reduction
generated carrier separation of the Nb-SrTiO3 samples is supposed to being the rate-determining step (Cr(VI) + e−CB = Cr(V)) [62]. Because the
play a major contributing role in the improvement of their photo­ CB potentials of SrTiO3 and 3% Nb-SrTiO3 are more negative than the
catalytic activities. Fig. 11(b) shows the electrochemical impedance reduction potential of Cr(VI)/Cr(V) couple (E⊖(Cr(VI)/Cr(V)) = +0.55
spectra of SrTiO3, 1% Nb-SrTiO3, 3% Nb-SrTiO3 and 5% Nb-SrTiO3. V, E⊖(Cr(V)/Cr(IV)) = 1.34 V, and E⊖(Cr(IV)/Cr(III)) = 2.10 V vs. NHE
From Fig. 11(b), the Nyquist semicircle radii of SrTiO3, 1% Nb-SrTiO3, [62]), the reduction of Cr(VI) to Cr(III) by e− in the CB of SrTiO3 and 3%
5% Nb-SrTiO3 and 3% Nb-SrTiO3 are observed to be in a descending Nb-SrTiO3 would be thermodynamically feasible. Fig. 12 depicts the
order. Normally, the shorter Nyquist semicircle radius means the smaller mechanism for the improved photocatalytic reduction of Cr(VI) by 3%
interfacial charge transfer resistance and thus the faster interfacial Nb-SrTiO3. Under Xe lamp irradiation, e− in the VB of 3% Nb-SrTiO3 are
charge transfer between the photocatalyst and the electrolyte [56–58]. excited to the CB, and simultaneously the same amount of h+ are created
The Nyquist semicircle radius of 3% Nb-SrTiO3 is the shortest, indicating in the VB. A portion of e− migrate directly to the photocatalyst surface to
that its charge transfer resistance is the smallest and thus it has the reduce Cr(VI) to Cr(III). Another portion of e− are captured by the
greatest efficiency in transferring h+ and e− . The interfacial charge doping level (Nb 4d), which can restrain the recombination of h+ and e− .
transfer rates of SrTiO3, 1% Nb-SrTiO3, 3% Nb-SrTiO3 and 5% Nb- The e− in the doping level can also reduce Cr(VI) to Cr(III). As a result,
SrTiO3 inferred from their electrochemical impedance spectra are in the 3% Nb-SrTiO3 has higher efficiency than SrTiO3 in the photocatalytic
same order as their photocatalytic activities. Therefore, the increased reduction of Cr(VI).
interfacial charge transfer rates of the Nb-SrTiO3 samples should also
play a main contributing role in the enhancement of their photocatalytic 4. Conclusions
activities. On the basis of the above results and analyses, the most
probable underlying reasons for the highest photocatalytic activity of A series of Nb-SrTiO3 mesoporous nanospheres were prepared via a
3% Nb-SrTiO3 are that it has the greatest efficiency in the separation and viable hydrothermal approach. 3% Nb-SrTiO3 had a smaller specific
transfer of h+ and e− . surface area and a slightly decreased band gap, compared with SrTiO3.
The semiconductor type and flat-band potentials of SrTiO3 and 3% For the photocatalytic reduction of aqueous Cr(VI) under Xe lamp
Nb-SrTiO3 were determined based on the Mott-Schottky plots shown in irradiation, all the Nb-SrTiO3 products exhibited higher efficiency than
Fig. S3. From Fig. S3, the Mott-Schottky plots of SrTiO3 and 3% Nb- SrTiO3, with 3% Nb-SrTiO3 being the most efficient. 3% Nb-SrTiO3 not
SrTiO3 are seen to demonstrate positive slopes, revealing that both the only showed a photocatalytic efficiency 3 times as much as that of
samples are of n-type semiconductors [59–61]. The flat-band potentials SrTiO3, but also was superior to many other metal cations (including
of SrTiO3 and 3% Nb-SrTiO3 are in turn − 0.54 V and − 0.40 V vs. Nd3+, Zr4+, Cu2+, La3+, Sn4+ and Co2+) doped SrTiO3 synthesized via
saturated calomel electrode (SCE), corresponding to − 0.30 V and − the same process and a number of previously reported photocatalysts. In
0.16 V vs. normal hydrogen electrode (NHE). According to the refer­ addition, 3% Nb-SrTiO3 also demonstrated a fair reuse stability. It was
ences [59–61], the conduction band (CB) potentials of n-type discovered that the greatest photocatalytic activity of 3% Nb-SrTiO3 was

13
J. Zhu et al. Powder Technology 410 (2022) 117886

ascribed to its most efficient separation and transfer of photogenerated [9] S. Shraavan, S. Challagulla, S. Banerjee, S. Roy, Unusual photoluminescence of
Cu–ZnO and its correlation with photocatalytic reduction of Cr(VI), Bull. Mater.
charge carriers. This study suggests that Nb5+-doping is an effective
Sci. 40 (2017) 1415–1420.
modifying means for improving the photocatalytic efficiency of SrTiO3, [10] G. Xing, L. Zhao, T. Sun, Y. Su, X. Wang, Hydrothermal derived nitrogen doped
and Nb-SrTiO3 is a potential high-performance photocatalyst for treat­ SrTiO3 for efficient visible light driven photocatalytic reduction of chromium(VI),
ing Cr(VI)-contaminated water. Future work can be directed at (i) SpringerPlus 5 (2016) 1132.
[11] M.B. Tahir, T. Nawaz, M. Sagir, M. Alzaid, H. Alrobei, K. Shahzad, A.M. Ali,
exploring new synthesis methods, optimizing the doping amount of S. Muhammad, Functionalized role of highly porous activated carbon in bismuth
Nb5+ and synthesis conditions of Nb-SrTiO3 to achieve the maximum vanadate nanomaterials for boosted photocatalytic hydrogen evolution and
photocatalytic efficiency, (ii) revealing in-depth the structure-activity synchronous activity in water, Int. J. Hydrog. Energy 46 (2021) 39778–39785.
[12] S. Roy, Photocatalytic materials for reduction of nitroarenes and nitrates, J. Phys.
relationship and underlying photocatalytic mechanism of Nb-SrTiO3, Chem. C 124 (2020) 28345–28358.
and (iii) exploring other photocatalytic applications of Nb-SrTiO3, such [13] F. Zhang, L. Shen, J. Li, Y. Zhang, G. Wang, A. Zhu, Room temperature
as photocatalytic CO2 reduction, water splitting, organic synthesis, N2 photocatalytic deposition of Au nanoparticles on SnS2 nanoplates for enhanced
photocatalysis, Powder Technol. 383 (2021) 371–380.
fixation, and organic pollutant degradation. [14] M.B. Tahir, S. Tufail, A. Ahmad, M. Rafique, T. Iqbal, M. Abrar, T. Nawaz, M.
Y. Khan, M. Ijaz, Semiconductor nanomaterials for the detoxification of dyes in real
wastewater under visible-light photocatalysis, Int. J. Environ. Anal. Chem. 101
CRediT authorship contribution statement (2021) 1735–1749.
[15] S. Roy, Tale of two layered semiconductor catalysts toward artificial
Jing Zhu: Investigation, Methodology, Writing – original draft. photosynthesis, ACS Appl. Mater. Interfaces 12 (2020) 37811–37833.
[16] M.B. Tahir, A. Ahmad, T. Iqbal, M. Ijaz, S. Muhammad, S.M. Siddeeg, Advances in
Yongcai Zhang: Conceptualization, Resources, Supervision, Writing –
photo-catalysis approach for the removal of toxic personal care product in aqueous
review & editing. Li Shen: Resources, Writing – review & editing. Jing environment, Environ. Dev. Sustain. 22 (2020) 6029–6052.
Li: Funding acquisition, Writing – review & editing. Liangliang Li: [17] S. Challagulla, R. Nagarjuna, S. Roy, R. Ganesan, Scalable free-radical
Investigation. Fen Zhang: Writing – review & editing. Ya Zhang: polymerization based sol–gel synthesis of SrTiO3 and its photocatalytic activity,
ChemistrySelect 2 (2017) 4836–4842.
Funding acquisition, Writing – review & editing. [18] Q. Kuang, S. Yang, Template synthesis of single-crystal-like porous SrTiO3
nanocube assemblies and their enhanced photocatalytic hydrogen evolution, ACS
Appl. Mater. Interfaces 5 (2013) 3683–3690.
[19] Y. Kim, M. Watanabe, J. Matsuda, J.T. Song, A. Takagaki, A. Staykov, T. Ishihara,
Declaration of Competing Interest Tensile strain for band engineering of SrTiO3 for increasing photocatalytic activity
to water splitting, Appl. Catal. B Environ. 278 (2020), 119292.
We declare that we have no known competing financial interests or [20] S. Cao, X. Ye, H. Hu, H. Jin, Y. Wang, J. Ye, Rational synthesis of SrTiO3 nanodots
anchored mesocrystalline anatase TiO2 submicrospheres for photocatalytic
personal relationships that could have appeared to influence the work reduction of CrVI, Sep. Purif. Technol. 275 (2021), 119096.
reported in this paper. [21] L. Pan, H. Mei, G. Zhu, S. Li, X. Xie, S. Gong, H. Liu, Z. Jin, J. Gao, L. Cheng,
L. Zhang, Bi selectively doped SrTiO3-x nanosheets enhance photocatalytic CO2
reduction under visible light, J. Colloid Interface Sci. 611 (2022) 137–148.
Data availability [22] H.J. Jang, S.J. Park, J.H. Yang, S. Hong, C.K. Rhee, D. Kim, Y. Sohn, Photocatalytic
and photoelectrocatalytic properties of Eu(III)-doped perovskite SrTiO3
Data will be made available on request. nanoparticles with dopant level approaches, Mater. Sci. Semiconduct. Proc. 132
(2021), 105919.
[23] E. Vento-Lujano, L.A. González, Defect-induced modification of band structure by
Acknowledgment the insertion of Ce3+ and Ce4+ in SrTiO3: a high-performance sunlight-driven
photocatalyst, Appl. Surf. Sc. 569 (2021), 151044.
[24] J. Jiang, K. Kato, H. Fujimori, A. Yamakata, Y. Sakata, Investigation on the highly
The Science and Technology Project of Xuzhou (Grant No. active SrTiO3 photocatalyst toward overall H2O splitting by doping Na ion, J. Catal.
KC21286), and The National Natural Science Foundation of China 390 (2020) 81–89.
(22172140). [25] D. Yang, X. Zhao, X. Zou, Z. Zhou, Z. Jiang, Removing Cr (VI) in water via visible-
light photocatalytic reduction over Cr-doped SrTiO3 nanoplates, Chemosphere 215
(2019) 586–595.
Appendix A. Supplementary data [26] K. Aravinthkumar, I.J. Peter, G.A. Babu, M. Navaneethan, S. Karazhanov, C.
R. Mohan, Enhancing the short circuit current of a dye-sensitized solar cell and
photocatalytic dye degradation using Cr doped SrTiO3 interconnected spheres,
Supplementary data to this article can be found online at https://doi. Mater. Lett. 319 (2022), 132284.
org/10.1016/j.powtec.2022.117886. [27] Q. Zhou, N. Li, D. Chen, Q. Xu, H. Li, J. He, J. Lu, Efficient removal of Bisphenol a
in water via piezocatalytic degradation by equivalent-vanadium-doped SrTiO3
nanofibers, Chem. Eng. Sci. 247 (2022), 116707.
References [28] F. Fang, F. Xu, Z. Su, X. Li, W. Han, Y. Qin, J. Ye, K. Chang, Understanding targeted
modulation mechanism in SrTiO3 using K+ for solar water splitting, Appl. Catal. B
Environ. 316 (2022), 121613.
[1] M.B. Tahir, H. Kiran, T. Iqbal, The detoxification of heavy metals from aqueous
[29] M. Abdi, V. Mahdikhah, S. Sheibani, Visible light photocatalytic performance of La-
environment using nano-photocatalysis approach: a review, Environ. Sci. Pollut.
Fe co-doped SrTiO3 perovskite powder, Opt. Mater. 102 (2020), 109803.
Res. 26 (2019) 10515–10528.
[30] B. Li, J. Hong, Y. Ai, Y. Hu, Z. Shen, S. Li, Y. Zou, S. Zhang, X. Wang, G. Zhao, X. Xu,
[2] Y.C. Zhang, L. Yao, G. Zhang, D.D. Dionysiou, J. Li, X. Du, One-step hydrothermal
Visible-near-infrared-light-driven selective oxidation of alcohols over
synthesis of high-performance visible-light-driven SnS2/SnO2 nanoheterojunction
nanostructured cu doped SrTiO3 in water under mild condition, J. Catal. 399
photocatalyst for the reduction of aqueous Cr(VI), Appl. Catal. B Environ. 144
(2021) 142–149.
(2014) 730–738.
[31] S. Zong, L. Tian, X. Guan, C. Cheng, J. Shi, L. Guo, Photocatalytic overall water
[3] Y. Li, Y. Han, C. Wang, Fabrication strategies and Cr(VI) elimination activities of
splitting without noble-metal: decorating CoP on Al-doped SrTiO3, J. Colloid
the MOF-derivatives and their composites, Chem. Eng. J. 405 (2021), 126648.
Interface Sci. 606 (2022) 491–499.
[4] R. Nagarjuna, S. Challagulla, R. Ganesan, S. Roy, High rates of Cr(VI)
[32] P. Nunocha, M. Kaewpanha, T. Bongkarn, A. Phuruangrat, T. Suriwong, A new
photoreduction with magnetically recoverable nano-Fe3O4@Fe2O3/Al2O3 catalyst
route to synthesizing La-doped SrTiO3 nanoparticles using the sol-gel auto
under visible light, Chem. Eng. J. 308 (2017) 59–66.
combustion method and their characterization and photocatalytic application,
[5] T. Ge, Z. Jiang, L. Shen, J. Li, Z. Lu, Y. Zhang, F. Wang, Synthesis and application of
Mater. Sci. Semiconduct. Proc. 134 (2021), 106001.
Fe3O4/FeWO4 composite as an efficient and magnetically recoverable visible light-
[33] Y. Zhang, Y. Li, Y. Yuan, K. Lin, C-dots decorated SrTiO3/NH4V4O10 Z-scheme
driven photocatalyst for the reduction of Cr(VI), Sep. Purif. Technol. 263 (2021),
heterojunction for sustainable antibiotics removal: reaction kinetics, DFT
118401.
calculation and mechanism insight, Sep. Purif. Technol. 295 (2022), 121268.
[6] K. Shahzad, M.B. Tahir, M. Sagir, M.R. Kabli, Role of CuCo2S4 in Z-scheme MoSe2/
[34] X. Yu, J. Wang, X. Fu, H. Meng, Y. Zhu, Y. Zhang, Construction of Z-scheme
BiVO4 composite for efficient photocatalytic reduction of heavy metals, Ceram. Int.
SrTiO3/Ag/Ag3PO4 photocatalyst with oxygen vacancies for highly efficient
45 (2019) 23225–23232.
degradation activity towards tetracycline, Sep. Purif. Technol. 241 (2020),
[7] R. Nagarjuna, S. Challagulla, P. Sahu, S. Roy, R. Ganesan, Polymerizable sol–gel
116718.
synthesis of nano-crystalline WO3 and its photocatalytic Cr(VI) reduction under
[35] Z. Chen, H. Yin, R. Wang, Y. Peng, C. You, J. Li, Efficient electron transfer by
visible light, Adv. Powder Technol. 28 (2017) 3265–3273.
plasmonic silver in SrTiO3 for low-concentration photocatalytic NO oxidation,
[8] S. Challagulla, R. Nagarjuna, R. Ganesan, S. Roy, Acrylate-based polymerizable
Environm, Sci. Technol. 56 (2022) 3604–3612.
sol− gel synthesis of magnetically recoverable TiO2 supported Fe3O4 for Cr(VI)
photoreduction in aerobic atmosphere, ACS Sustain. Chem. Eng. 4 (2016) 974–982.

14
J. Zhu et al. Powder Technology 410 (2022) 117886

[36] Y. Hu, Z. Shen, B. Li, S. Li, J. Yue, G. Zhao, M. Muhler, X. Wang, Solvent effects on [50] T. Ge, L. Shen, J. Li, Y. Zhang, Y. Zhang, Morphology-controlled hydrothermal
photocatalytic anaerobic oxidation of benzyl alcohol over Pt-loaded defective synthesis and photocatalytic Cr(VI) reduction properties of α-Fe2O3, Colloids Surf.
SrTiO3 nanoparticles, ACS Appl. Nano Mater. 4 (2021) 9254–9264. A Physicochem. Eng. Asp. 635 (2022), 128069.
[37] G. Venkatesh, S. Vignesh, M. Srinivasan, G. Palanisamy, N. Elavarasan, [51] X. Wang, L. Jiang, K. Li, J. Wang, D. Fang, Y. Zhang, D. Tian, Z. Zhang, D.
K. Bhuvaneswari, P. Ramasamy, M. Alam, M. Ubaidullah, M.K. Raza, Construction D. Dionysiou, Fabrication of novel Z-scheme SrTiO3/MnFe2O4 system with double-
and investigation on perovskite-type SrTiO3@ reduced graphene oxide hybrid response activity for simultaneous microwave-induced and photocatalytic
nanocomposite for enhanced photocatalytic performance, Colloids Surf. A degradation of tetracycline and mechanism insight, Chem. Eng. J. 400 (2020),
Physicochem. Eng. Asp. 629 (2021), 127523. 125981.
[38] S. Khan, H. Cho, D. Kim, S.S. Han, K.H. Lee, S. Cho, T. Song, H. Choi, Defect [52] Y.C. Zhang, Z.N. Du, M. Zhang, Hydrothermal synthesis of SnO2/SnS2
engineering toward strong photocatalysis of Nb-doped anatase TiO2: nanocomposite with high visible light-driven photocatalytic activity, Mater. Lett.
computational predictions and experimental verifications, Appl. Catal. B Environ. 65 (2011) 2891–2894.
206 (2017) 520–530. [53] Y. Zhang, F. Zhang, Z. Yang, H. Xue, D.D. Dionysiou, Development of a new
[39] J. Yang, X. Zhang, C. Wang, P. Sun, L. Wang, B. Xia, Y. Liu, Solar photocatalytic efficient visible-light-driven photocatalyst from SnS2 and polyvinyl chloride,
activities of porous Nb-doped TiO2 microspheres prepared by ultrasonic spray J. Catal. 344 (2016) 692–700.
pyrolysis, Solid State Sci. 14 (2012) 139–144. [54] S. Payra, S. Roy, From trash to treasure: probing cycloaddition and photocatalytic
[40] X. Yang, Y. Min, S. Li, D. Wang, Z. Mei, J. Liang, F. Pan, Conductive Nb-doped TiO2 reduction of CO2 over cerium-based metal− organic frameworks, J. Phys. Chem. C
thin films with whole visible absorption to degrade pollutants, Catal, Sci. Technol. 125 (2021) 8497–8507.
8 (2018) 1357–1365. [55] F. Zhang, Y. Zhang, Y. Wang, A. Zhu, Y. Zhang, Efficient photocatalytic reduction
[41] E. Drożdż, A. Koleżyński, The structure, electrical properties and chemical stability of aqueous Cr (VI) by Zr4+ doped and polyaniline coupled SnS2 nanoflakes, Sep.
of porous Nb-doped SrTiO3 – experimental and theoretical studies, RSC Adv. 7 Purif. Technol. 283 (2022), 120161.
(2017) 28898–28908. [56] Y. Zhang, H. Zhou, H. Wang, Y. Zhang, D.D. Dionysiou, Synergistic effect of
[42] J. Yin, J. Ye, Z. Zou, Enhanced photoelectrolysis of water with photoanode Nb: reduced graphene oxide and near-infrared light on MoS2-mediated electrocatalytic
SrTiO3, Appl. Phys. Lett. 85 (2004) 689–691. hydrogen evolution, Chem. Eng. J. 418 (2021), 129343.
[43] Q. Fu, H. Gu, J. Xing, Z. Cao, J. Wang, Controlling the A-site deficiency and oxygen [57] X. Guo, Y. Xu, Y. Cheng, Y. Zhang, H. Pang, Amorphous cobalt phosphate porous
vacancies by donor-doping in pre-reductive-sintered thermoelectric SrTiO3 nanosheets derived from two-dimensional cobalt phosphonate organic frameworks
ceramics, Acta Mater. 229 (2022), 117785. for high performance of oxygen evolution reaction, Appl. Mater. Today 18 (2020),
[44] M.B. Tahir, T. Nawaz, G. Nabi, M. Sagir, M. Rafique, A. Ahmed, S. Muhammad, 100517.
Photocatalytic degradation and hydrogen evolution using bismuth tungstate based [58] Y. Zhang, L. Hu, Y. Zhang, X. Wang, H. Wang, Snowflake-like Cu2S/MoS2/Pt
nanocomposites under visible light irradiation, Int. J. Hydrog. Energy 45 (2020) heterostructure with near infrared photothermal-enhanced electrocatalytic and
22833–22847. photoelectrocatalytic hydrogen production, Appl. Catal. B Environ. 315 (2022),
[45] S. Challagulla, S. Payra, C. Chakraborty, S.A. Singh, S. Roy, Understanding the role 121540.
of catalytic active sites for heterogeneous photocatalytic oxidation of methanol and [59] Z. Jiang, K. Chen, Y. Zhang, Y. Wang, F. Wang, G. Zhang, D.D. Dionysiou,
thermal reduction of NOx, Mol. Catal. 476 (2019), 110505. Magnetically recoverable MgFe2O4/conjugated polyvinyl chloride derivative
[46] S. Challagulla, S. Payra, C. Chakraborty, S. Roy, Determination of band edges and nanocomposite with higher visible-light photocatalytic activity for treating Cr(VI)-
their influences on photocatalytic reduction of nitrobenzene by bulk and exfoliated polluted water, Sep. Purif. Technol. 236 (2020), 116272.
g-C3N4, Phys. Chem. Chem. Phys. 21 (2019) 3174–3183. [60] F. Zhang, Y. Zhang, C. Zhou, Z. Yang, H. Xue, D.D. Dionysiou, A new high
[47] B. Soman, S. Challagulla, S. Payra, S. Dinda, S. Roy, Surface morphology and active efficiency visible-light photocatalyst made of SnS2 and conjugated derivative of
sites of TiO2 for photoassisted catalysis, Res. Chem. Intermed. 44 (2018) polyvinyl alcohol and its application to Cr(VI) reduction, Chem. Eng. J. 324 (2017)
2261–2273. 140–153.
[48] F. Zhang, Y. Zhang, G. Zhang, Z. Yang, D.D. Dionysiou, A. Zhu, Exceptional [61] P. Su, X. Zhang, X. Hao, H. Liu, Z. Jin, Co3O4 modified Mn0.2Cd0.8S with different
synergistic enhancement of the photocatalytic activity of SnS2 by coupling with shells forms p-n heterojunction to optimize energy/mass transfer for efficient
polyaniline and N-doped reduced graphene oxide, Appl. Catal. B Environ. 236 photocatalytic hydrogen evolution, Sep. Purif. Technol. 285 (2022), 120318.
(2018) 53–63. [62] M.I. Litter, Last advances on TiO2-photocatalytic removal of chromium, uranium
[49] Y. Wang, Y. Su, W. Fang, Y. Zhang, X. Li, G. Zhang, W. Sun, SnO2/SnS2 and arsenic, Curr. Opin. Green Sust. Chem. 6 (2017) 150–158.
nanocomposite anchored on nitrogen-doped RGO for improved photocatalytic
reduction of aqueous Cr(VI), Powder Technol. 363 (2020) 337–348.

15

You might also like