Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

http://pubs.acs.

org/journal/acsodf Article

BiFeO3‑Black TiO2 Composite as a Visible Light Active Photocatalyst


for the Degradation of Methylene Blue
Pravallika Banoth, Chinna Kandula, Praveen Kumar Lavudya, Saidulu Akaram,
Luis De Los Santos Valladares,* RajaniKanth Ammanabrolu, Ghanashyam Krishna Mamidipudi,
and Pratap Kollu*

Cite This: ACS Omega 2023, 8, 18653−18662 Read Online


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations

ABSTRACT: The application of a novel BiFeO3 (BFO)-black TiO2


Downloaded via 49.205.243.112 on June 18, 2023 at 10:52:21 (UTC).

(BTO) composite (called BFOT) as a photocatalyst for the degradation


of methylene blue is reported. The p−n heterojunction photocatalyst
was synthesized for the first time through microwave-assisted co-
precipitation synthesis to change the molar ratio of BTO in BiFeO3 to
increase the photocatalytic efficiency of the BiFeO3 photocatalyst. The
UV−visible properties of p−n heterostructures showed excellent
absorption of visible light and reduced electron−hole recombination
properties compared to the pure-phase BFO. Photocatalytic studies on
BFOT10, BFOT20, and BFOT30 have shown that they decompose
methylene blue (MB) in sunlight better than pure-phase BFO in 70 min.
The BFOT30 photocatalyst was the most effective at reducing MB when
exposed to visible light (97%). Magnetic studies have shown that BTO is
diamagnetic, and the BFOT10 photocatalyst exhibits a very weak antiferromagnetic behavior, whereas BFOT20 and BFO30 show
diamagnetic behavior. This study confirms that the catalyst has poor stability and weak magnetic recovery properties due to the non-
magnetic phase BTO in the BFO.

1. INTRODUCTION In recent years, black TiO2 functioned better as a


A promising strategy to address environmental problems and photocatalyst than pure white TiO2 because it could better
the energy shortage is the use of catalysts for the remediation separate photogenerated electron−hole pairs. The higher solar
of organic pollutants.1−6 Photocatalysts are widely employed to light absorption of black TiO2 was caused by Ti3+ and the
break down contaminants using ultraviolet or visible light oxygen vacancies that resulted in refs 14−16. Black TiO2
sources. Perovskite-based photocatalysts are a significant class (BTO) photocatalysts have recently received a lot of interest
of semiconductors that both UV and visible light can activate. due to their exceptional light-absorbing properties, which even
Combining the photocatalysts is one method for improving make the NIR region of the solar spectrum practical. Due to
their photoactivity because the efficiency of pure photo- the low band gap value, they showed light absorption even in
catalysts is frequently low. the near-infrared range (1.54 eV).17 Since then, numerous
A stable, non-toxic, and multiferroic visible light active successful photocatalysis and other photoelectrochemical
photocatalyst has recently been explored to exist in the form of applications have drawn significant research attention to
BiFeO3 (BFO). BFO has a band gap from 2.2 to 2.7 eV, which BTO.6 Two usually mentioned benefits are increased photo-
indicates that it strongly absorbs light in the visible range from catalytic efficiency and reduced recombination of photo-
200 to 800 nm.7,8 Excellent characteristics of BFO have been generated charge carriers. The heterostructure of multiferroic
seen in the degradation of pollutants. However, the main BFO and BTO nanoparticles is expected to be a superior
problem with the significant photogenerated electron−hole
pairs recombination rate is still restricting the use of BFO for Received: January 27, 2023
photocatalysis. To tackle this issue, BFO has received much Accepted: April 28, 2023
attention when coupled with other visible active semi- Published: May 15, 2023
conductors, which prevents the high recombination of
photogenerated electron−hole pairs and enhances photo-
catalytic activity.9−13
© 2023 The Authors. Published by
American Chemical Society https://doi.org/10.1021/acsomega.3c00553
18653 ACS Omega 2023, 8, 18653−18662
ACS Omega http://pubs.acs.org/journal/acsodf Article

photocatalyst with increased photocatalytic activity compared Following the above procedure, we have synthesized three
to pure-phase BFO.18−21 different heterostructures by loading different ratios of TiO2
Furthermore, the shorter band gap of the heterostructure onto the BFO. The prepared heterostructures (1 − x) BFO-
interfaces may result in more electrons and holes and exhibit xBTO, where x = 0.10, 0.20, and 0.30, are named BFOT10,
higher photocatalytic efficiency.22,23 It is well known from the BFOT20, and BFOT30, respectively.
literature that recent studies on the BFO-based heterostructure 2.2. Characterizations. UV−visible spectroscopy (Jasco
have shown improved photocatalytic activity in organic dye V-670 UV−visible double beam spectrophotometer) was used
degradation under visible light exposure. Therefore, creating to characterize the BFO, BTO, and BFOT heterostructures
strong photocatalytic activity based on black TiO2 seems to following synthesis, with ethanol serving as a blank. Wave-
have potential. lengths ranging from 200 to 800 nm were used to record the
To the present author’s knowledge, there are no reports on absorption spectra. Analysis of the X-ray diffraction (XRD)
the composite of black TiO2 with BFO for organic dye data (PANalytical, X’Pert Powder Diffractometer) has shown
degradation under visible-light exposure. In the current work, that the pure phases of BFO and BTO have been formed, as
for the first time, BFO-BTO (black TiO2) heterostructures well as the creation of BFOT heterostructures. At a scan rate of
were synthesized through the microwave-co-precipitation 4 °/min and with a step size of 0.02°, the scanning range in this
method by mixing different molar ratios of BFO and the diffractogram was 20θ to 80θ [2θ value]. Field emission
black TiO2 in the as-prepared condition by following scanning electron microscopy (FESEM: Carl Zeiss Smart Sem)
microwave heating technique. The developed BFO-BTO and transmission electron microscopy were used to study the
composite (referred to by the acronym BFOT in the rest of morphology of synthesized samples. A vibrating sample
the paper) photocatalysts with a narrow band gap have shown magnetometer measured the magnetic properties of synthe-
outstanding photocatalytic activity for methylene blue sized heterostructured samples at room temperature (VSM,
elimination under sunlight irradiation. Model: LakeShore).
The novelty of this study is that it uses black TiO2 and 2.3. Photocatalysis. Samples were tested for their
multiferroic BiFeO3 to develop a bismuth-based composite for photocatalytic ability to degrade methylene blue, an example
decomposing an organic dye in the visible range. For the of a pollution dye. A lot of sunlight was present during the
photocatalysis application, there are no reports on this photodegradation process. A total of 40 mg of each
composite material. In previous works, bismuth-based photocatalyst powder was initially added to a 100 mL MB
composites made with white TiO2 had a low photodegradation solution at a concentration of 10 mg/L at a neutral pH. After
percentage and took a long time to break down the dye the magnetic churning, the slurry was exposed to heavy
completely. In our current study, however, bismuth-based sunlight. We used UV−visible spectroscopy to keep track of
composites made with black TiO2 showed a high photo- the dye solution’s decolorization.
degradation percentage in just 70 min of sunlight exposure.
3. RESULT AND DISCUSSION
2. EXPERIMENTAL SECTION 3.1. X-ray Diffraction Analysis of BFOT Heterostruc-
2.1. Materials Used for BFOT Heterostructures. For tures. In this study, BFOT heterostructure powders were
this procedure, ethanol, double-distilled water (DDW), pure
potassium hydroxide (KOH, 98%), pure ferric chloride (FeCl3·
6H2O, 98%), pure bismuth chloride (BiCl3, 98%), and
commercial white TiO2 are all used.
2.1.1. Preparation of BiFeO3 Micro Flowers. The pure-
phase BFO microwave flowers are synthesized in 3 min at 800
W by microwave-assisted-solvothermal (MWAST) method in
a domestic solo-microwave oven. The complete experimental
procedure can be found in our previous work.7
2.1.2. Preparation of Black TiO2 Nanoparticles. White
TiO2 pellets were made utilizing simple but meticulous
techniques. Ethylene glycol (EG) was added to the ultrafine
TiO2 powder, which was then crushed for 60 min. EG was
employed as a binder, and then, a pellet with the shape of a
circular disc was formed by pressing a powder mixture with a
hydraulic press at a pressure of 15 tonnes for roughly 2 min.
The pellet is then sintered for 2−6 h at 800−1000 °C.
Afterward, the white TiO2 pellets were irradiated by an Figure 1. XRD patterns of pure BFO, BTO, BFOT10, BFOT20, and
electron beam for 40−60 min to get the black TiO2 [BTO] BFOT30.
powder.
2.1.3. Synthesis of (1 − x) BFO-TiO2 (x = 0.10, 0.20, and
0.30) Heterostructures. BFO-TiO2 heterostructures are synthesized through a microwave-assisted-chemical coprecipi-
prepared by a simple microwave-assisted-coprecipitation tation method and with BTO content levels of 10%
method following microwave heating at 360 W for 3 min. (BFOT10), 20% (BFOT20), and 30% (BFOT30). The XRD
We dissolved a stoichiometric ratio of BFO and BTO powders patterns of BFO, BTO, and BFOT heterostructures are
in 100 mL of ethanol and sonicated it for 8 h at 80 °C to displayed in Figure 1.
evaporate the solvent and obtain the gel. Finally, the gel was The BFO displayed a Rhombohedral perovskite crystalline
heated at 360 W for 3 min in a domestic microwave oven. phase with the R3c space group, which perfectly matched the
18654 https://doi.org/10.1021/acsomega.3c00553
ACS Omega 2023, 8, 18653−18662
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 2. FE-SEM images of (a) pure BFO, (b) BTO, (c) BFOT10, (d) BFOT20, and (e) BFOT30 and EDAX spectra of (f) BFOT10, (g)
BFOT20, and (h) BFOT30.

prior crystallographic data (ICSD 98-019-1940) and showed and BTO phases are not as tightly coupled. According to
no impurity phases within the detection range of the technique Figure 2d, the contact between the two phases grows as the
according to XRD examination. Tetragonal symmetry was amount of BTO does, and the distribution of the grains is
found in the BTO XRD pattern, where anatase and rutile uniform with a uniform distribution of particle size. The
phases were present. The XRD patterns revealed all peaks microstructure of BFOT30 is finally formed of a large number
belonging to pure-phase BFO with a low intense trace of the of uniformly dispersed grains with an enormous number of
significant BTO peaks after adding 10% BTO (BFOT10). grain boundaries, as seen in Figure 2e; this is because 30%
More intense BTO peaks began to form as the loading ratio of BTO was added to the BFO. We can observe a stronger and
BTO reached 20% (BFOT20). After loading 30% BTO, the more consistent coupling by comparing the BFO and BTO
heterostructure sample BFOT30 finally displays all the high- heterostructures to BFOT20 and BFT30 heterostructures. It is
intensity peaks from BTO and the diffraction pattern for pure established from the EDAX spectra of BFOT10, BFOT20, and
BFO. The enhanced intensity of all the prominent BTO peaks BFOT30 in Figure 2f−h that the Bi, Fe, O, and Ti components
besides the BFO peaks, increasing the BTO molar ratio up to are the source of the signal peaks, demonstrating that BTO was
30%, confirms the higher BTO concentration. successfully loaded onto the BFO.
3.2. FE-SEM Analysis of BFOT Heterostructures. The 3.3. Optical Properties of BFOT Heterostructures. The
various microstructures of the samples were examined using an UV−vis absorption spectrum of all the prepared samples is
FE-SEM, as shown in Figure 2. Figure 2a shows that the pure- shown in Figure 3a−h. Each sample demonstrated excellent
phase BFO powder comprised many BFO micro flowers with visible light absorption. Figure 3a,b and the inset inside Figure
hundreds of BFO nano petals packed tightly onto them.24 A 3a depict the light absorption and energy band gap of the pure
considerable number of BTO nanoparticles of different sizes phases of BFO and BTO. A pure-phase BFO exhibits light
and shapes were formed, as demonstrated by the micro- absorption at 550 nm with a band gap of 2.25 eV (Figure 3a
structures of pure-phase BTO in Figure 2b. Following the and inset, respectively). In contrast, a pure-phase BTO exhibits
loading of BFO with various molar ratios of BTO particles, as light absorption at 690 nm with a band gap of 1.68 eV (Figure
shown in Figure 2c−e, the microstructure of the BFOT 3a and inset, respectively). The UV−vis absorption spectrum
heterostructure is significantly changed. It is noteworthy to of all three BFO-BTO heterostructures is shown in Figure 3c−
note that it is believed that BFO micro flowers were damaged e. From these light absorption properties, we can see that
after heating the BFOT heterostructures with high microwave BFOT10 is absorbing light at the 552 nm range (Figure 3c),
energy. The BFOT10 heterostructure can be shown in Figure BFOT20 is absorbing light at the 560 nm range (Figure 3d),
2c to be made up of several grains of various sizes and fewer and BFOT30 is showing strong absorption in the higher
grain boundaries, which supports the hypothesis that the BFO wavelength side, i.e., at 650 nm (Figure 3e). It is essential to
18655 https://doi.org/10.1021/acsomega.3c00553
ACS Omega 2023, 8, 18653−18662
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 3. UV−vis absorption spectroscopy of (a) BFO (band gap in the inset), (b) BTO, (c) BFOT10, (d) BFOT20, (e) BFOT30, and energy
band gap of (f) BFOT10, (g) BFOT20, and (h) BFOT30.

Table 1. Optical Energy Band Gaps Obtained from UV−Vis in the light absorption properties leads the production of
Absorption Studies higher no of electrons and holes in the composite
materials.25−30 Therefore, from this UV−vis spectral studies,
s. no sample the band gap (eV)
all the heterostructures BFOT10, BFOT20, and BFOT30 have
1 BFO (Figure 4)15 2.25 shown excellent optical absorption in the visible range,
2 BTO 1.68 confirming the capability of these heterostructures for the
3 BFOT10 2.24 photodegradation of organic dyes in the visible range.
4 BFOT20 2.21 The following formula determines the band gap of the
5 BFOT30 2.1 heterostructures.

discuss that with the increase in the BTO content in the BFO, ( h ) = A( h Eg )n
the light absorption properties of the composite materials in
the visible range have increased, i.e., BFOT10 showing α is the absorption coefficient, Eg represents the energy band
absorption at 552 nm, BFOT20 at 560 nm, and BFOT30 at gap, A is constant and depending on transition, we have to
650 nm, therefore shifting of the band absorption toward the consider n values, direct permitted transitions, n = 1/2, direct
higher wavelength side indicates the increase in the light forbidden transitions, n = 3/2. Permitted indirect transitions, n
absorption property in the visible range, so this enhancement = 2. Indirect prohibited transitions, n = 3.
18656 https://doi.org/10.1021/acsomega.3c00553
ACS Omega 2023, 8, 18653−18662
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 4. Nitrogen absorption/desorption isotherms of (a) BFOT10, (b) BFOT20, and (c) BFOT30.

Figure 5. Time-dependant UV−vis spectral changes of MB in the presence of (a) BFO, (b) BTO, (c) BFOT10, (d) BFOT20, and (e) BFOT30.

18657 https://doi.org/10.1021/acsomega.3c00553
ACS Omega 2023, 8, 18653−18662
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 6. Effect of photocatalyst dosage on the photodegradation efficiency of heterostructures (a) BFOT10, (b) BFOT20, and (c) BFOT30.

Table 2. Influence of Catalyst Dosage (mg) on The obtained band energies of these heterostructures
Photodegradation (%) BFOT10 (552 nm), BFOT20 (560 nm), and BFOT30 (650
nm) are shown in Figure 3f−h, and the obtained band gap
dye catalyst catalyst dosage time degradation
Name name (mg) (min) (%) values were 2.24 (Figure 3f), 2.21 (Figure 3g), and 1.91 eV
(Figure 3h). The obtained band gap studies show a reduction
MB BFOT10 10 70 49
in the energy band gap of these heterostructures compared to
20 70 51
30 70 59
the pure-phase BFO, which could be due to the combination
40 70 65
of two different energy band gaps at the interface between two
MB BFOT20 10 70 52
semiconductors.31 As a result, BFOT30 was the most efficient
20 70 65
photocatalyst when absorbing visible-light photons among the
30 70 72
coupled photocatalysts. Optical energy band gaps obtained
40 70 86
from UV−vis absorption studies are shown in Table 1.
MB BFOT30 10 70 62
3.4. Brunauer−Emmett−Teller Analysis of BFOT
20 70 75 Heterostructures. Nitrogen adsorption−desorption iso-
30 70 80 therms were used to study the specific surface areas of the
40 70 97 BFOT samples made through microwave-assisted. Figure 4a−c
shows the Brunauer−Emmett−Teller (BET) surface area

Figure 7. Magnetic properties of (a) BFO, (b) BFOT heterostructure, (c) zoomed M−H loops for BTO, BFOT20, and BFOF30.

18658 https://doi.org/10.1021/acsomega.3c00553
ACS Omega 2023, 8, 18653−18662
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 8. Photodegradation efficiency of heterostructures (a) BFOT20 and (b) BFOT30 after four cycles.

phase BFO, pure-phase BTO, and BFOT heterostructures. The


total degradation of MB for the BFO, BTO, and BFOT
heterostructures took 70 min to complete. As shown in Figure
5a, the pure-phase BFO is showing a poor photodegradation
efficiency of 33% and a pure-phase BTO is exhibiting 78%
degradation in MB (Figure 5b), which is higher photo-
degradation efficiency compared to the pure-phase BFO.
Figure 5c−e depicts the photocatalytic performance of BFO-
BTO heterostructures. After adding 10% BTO into the BFO
[BFOT10], we observed less degradation (65%) in the MB
solution in the presence of the BFOT10 catalyst (Figure 5c)
when compared to the other two heterostructures [BFOT20
and BFOT30]. From Figure 5d,e, we can see that the efficiency
of the catalysts was increased by increasing the BTO content
into the BFO, where the BFOT20 (86%) and BFOT30 (97%)
Figure 9. Magnetic recovery test for (a) BFO and (b) BFOT30. catalysts exhibited enhanced degradation in MB after 70 min of
sunlight illumination, as shown in Figure 5d,e, respectively.
curves of the BFOT heterostructure nanoparticles made using The increased photocatalytic performance of the samples
microwave-assisted co-precipitation. All of the BFOT samples BFOT20 and BFOT30 demonstrate how efficiency increased
BFOT10, BFOT20, and BFOT30 that were made had a type for higher concentrations of BTO. The higher photo-
IV nitrogen isotherm with a hysteresis loop. The shape of the degradation efficiency (97%) of the BFOT30 heterostructure
hysteresis loops for the BFOT10, BFOT20, and BFOT30 is due to its microstructures and more significant specific
samples was all H3 type.32 The obtained specific surface areas surface-to-volume ratio than the BFOT10 and BFOT20
of the samples BFOT10, BFOT20, and BFOT30 are 46.10, heterostructures. The other reason for higher photodegrada-
48.98, and 54.77 m2/g, respectively. The higher specific surface tion efficiency is that BFOT30 has a narrower band gap than
area sample BFOT30 can enhance the degradation efficiency. the other two samples, making it easier for light to form more
3.5. Photocatalytic Activity of BFOT Heterostruc- electrons and holes at the surface of BFOT. As a result, its
tures. Figure 5a−e shows the photocatalytic activity of pure- efficiency is improved.33 This photodegradation test demon-

Figure 10. XRD patterns of a BFOT30 heterostructure before and after the photocatalysis test.

18659 https://doi.org/10.1021/acsomega.3c00553
ACS Omega 2023, 8, 18653−18662
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 11. FESEM image (a) and EDAX spectra (b) of a BFOT30 heterostructure after the photocatalysis test.

strates excellent results for the rapid photodegradation catalysts, there is no difference in the photodegradation
efficiency of heterostructures compared to pure-phase BFO, efficiencies for the first two cycles, but the catalysts’ poor
demonstrating these heterostructures’ abilities to degrade MB stability is demonstrated by a total 16% decline in BFOT20’s
in sunlight. Photocatalyst performance is determined by the photodegradation efficiency after two cycles (Figure 8b).
equation below. Both C0 and Ct represent MB’s initial and Figure 9a highlights how pure-phase BFO magnetic
time-corresponding concentrations, respectively. responses fail to recover particles when applied to magnetic
C0 Ct fields after the photodegradation experiment. Therefore,
Efficiency = × 100 additional filtering or centrifugation is required to separate
C0 the catalyst from the MB solution.25−30 To study the magnetic
3.6. Effect of Photocatalyst Concentration on Photo- recovery of the most effective photocatalyst, BFOT30, we
catalytic Efficiency. Different heterostructures with applied the magnetic field shown in Figure 9b. Figure 9b shows
BFOT10, BFOT20, and BFOT30 catalyst doses were used that BFOT30 cannot be recovered when a magnetic field is
to degrade the MB dye at a constant dosage of 10 mg/L (10, applied due to its diamagnetism, confirming its weak magnetic
20, 30, and 40 mg are shown in Figure 6a−c). From Figure 6a, recovery. The results of these stability and recovery studies
we can see that as increasing the catalyst BFOT10 indicate that the stability and magnetic recovery properties of
concentration from 10 to 40 mg, the photodegradation BFO did not improve, even when black TiO2 was added. When
percentage increased from 49 to 65%, similarly for BFOT20, a magnetic field was applied to BFOT, the magnetic response
photodegradation increased from 52 to 86% (Figure 6b), and may have been reduced and exhibited diamagnetic properties
for BFOT30 increased from 62 to 97% (Figure 6c) (all the due to the non-magnetic TiO2 loaded on weak magnetic phase
obtained values are shown in Table 2), this analysis is showing BFO.
that the photodegradation efficiency is substantially dependent 3.9. XRD and FESEM Analysis of the Most Active
on the catalyst concentration. The literature well established BFOT30 Catalyst after the Photocatalytic Test. The
that the microstructures of the catalysts and their micro- BFOF30 powder underwent XRD and FESEM analysis to see
dosing, starting dye dose, type, and pH value all significantly if it could be appropriately recovered or not at the beginning
impact the photodegradation efficiency.34−38 and end of the photocatalytic process and recovery test. The
3.7. Room-Temperature Magnetic Properties. The XRD patterns for BFOF30 before and after MB degradation
room-temperature magnetic properties of BTO and BFOT are shown in Figure 10. We discovered no differences in the
heterostructures are shown in Figure 7. Figure 7a,b shows that XRD patterns or traces from MB for the BFOF30 catalyst used
single-phase BFO has a high magnetization value of 1.25 emu/ in the photocatalysis experiment. Therefore, BFOF30 catalysts
g, while black TiO2 is diamagnetic when a magnetic field is that are centrifugally recovered retain their original structural
applied. The magnetic properties of BFOT have decreased
characteristics.
with an increase in the diamagnetic phase BTO loaded on
The FE-SEM image and EDAX of the BFOF30 catalyst after
weak ferromagnetic phase BFO (Figure 7b). When 10%
four cycles are shown in Figure 11a,b. We can see from the
diamagnetic BTO is added to BFO, the magnetization value is
reduced to 0.32 emu/g (BFOT10). After increasing the BTO microstructural analysis of the BFOT30 sample that it was
molar ratio to 20 and 30%, the heterostructures BFOT20 and composed of many grains and boundaries without any
BFOT30 show linear M−H curves, indicating they are agglomerations and other structures, confirming that no
diamagnetic [(Figure 7c) shows the close-ups of the M−H microstructures correspond to MB Figure 11a. We have
loops for BTO, BFOT20, and BFOT30]. observed that the BFOT30 catalyst’s microstructures are
3.8. Stability and Reusability Test for BFOT Photo- similar before and after the photocatalysis and recovery test,
catalysts. The stability of catalysts was investigated by indicating the stability of this sample’s microstructural
recycling the most productive powders, BFOT20 and characteristics even after four cycles. Figure 11b shows the
BFOT30 catalysts, four times, as shown in Figure 8a,b. EDAX spectra of the BFOT30 catalyst after photocatalysis and
According to Figure 8a, there is no difference in the recovery test; from these EDAX spectra, it is observed that the
photodegradation efficiencies of the BFOT20 catalysts for Bi, Fe, O, and Ti components are the source of the signal
the first two cycles, but the catalysts’ poor stability is peaks, and no other peaks corresponding to MB was found
demonstrated by a total 13% decline in BFOT20’s photo- demonstrating that BFOF30 catalysts that are centrifugally
degradation efficiency after two cycles. Similarly, for BFOT30 recovered retain their original microstructural characteristics.
18660 https://doi.org/10.1021/acsomega.3c00553
ACS Omega 2023, 8, 18653−18662
ACS Omega http://pubs.acs.org/journal/acsodf Article

4. CONCLUSIONS SRF. All major and significant developments and experiments


The BFOT heterostructure was successfully processed by were done at the University of Hyderabad, India. Hence, the
coprecipitation with the help of microwave for MB authors acknowledge funding from the Institute of Eminence
degradation. Microstructure analysis and UV−vis absorption project, UGC, UoH-IoE-RC2-21-017.
spectroscopy in visible light have shown that forming nano-
interfaces between the BFO and black TiO2 phases can change
the band gap and encourage changes in the work function at
■ REFERENCES
(1) Xing, Z.; Zhou, W.; Du, F.; Qu, Y.; Tian, G.; Pan, K.; Tian, C.;
the surface of the heterostructures, improving the creation of Fu, H. A floating macro/mesoporous crystalline anatase TiO2 ceramic
electron−hole pairs and impeding its recombination. After 70 with enhanced photocatalytic performance for recalcitrant wastewater
min in the sun, BFOT produces the best (97%) results, with degradation. Dalton Trans. 2014, 43, 790−798.
the highest deterioration in MB. The photocatalyst can only be (2) Jiang, Z.; Zhu, C.; Wan, W.; Qian, K.; Xie, J. Constructing
utilized for two cycles with a 16% reduction in efficiency, graphite-like carbon nitride modified hierarchical yolk−shell TiO2
according to the recycling test for the most effective BFOT30. spheres for water pollution treatment and hydrogen production. J.
It also cannot be retrieved using a magnetic field. A Mater. Chem. A 2016, 4, 1806−1818.
photocatalytic test showed that black TiO2 increased the (3) Xing, Z.; Zhou, W.; Du, F.; Zhang, L.; Li, Z.; Zhang, H.; Li, W.
activity of pure-phase BFO when added to it. According to our Facile synthesis of hierarchical porous TiO2 ceramics with enhanced
photocatalytic performance for micro polluted pesticide degradation.
findings, a simultaneous electron drain process between BFO
ACS Appl. Mater. Interfaces 2014, 6, 16653−16660.
and BTO and a direct Fenton-like mechanism contributed to (4) Liu, X.; Xing, Z.; Zhang, H.; Wang, W.; Zhang, Y.; Li, Z.; Wu, X.;
the BFOT heterostructure’s enhanced photocatalytic effi- Yu, X.; Zhou, W. Fabrication of 3 D Mesoporous Black TiO2/MoS2/
ciency. TiO2 Nanosheets for Visible-Light-Driven Photocatalysis. ChemSu-

■ AUTHOR INFORMATION
Corresponding Authors
sChem 2016, 9, 1118−1124.
(5) Chen, X.; Liu, L.; Yu, P. Y.; Mao, S. S. Increasing solar
absorption for photocatalysis with black hydrogenated titanium
Luis De Los Santos Valladares − Cavendish Laboratory, dioxide nanocrystals. Science 2011, 331, 746−750.
(6) Ullattil, S. G.; Narendranath, S. B.; Pillai, S. C.; Periyat, P. Black
Department of Physics, University of Cambridge, Cambridge TiO2 nanomaterials: a review of recent advances. Chem. Eng. J. 2018,
CB3 0HE, U.K.; Laboratorio de Cerámicos y 343, 708−736.
Nanomateriales, Facultad de Ciencias Físicas, Universidad (7) Sando, D.; Yang, Y.; Bousquet, E.; Carrétéro, C.; Garcia, V.;
Nacional Mayor de San Marcos, Lima 14-0149, Peru; Fusil, S.; Dolfi, D.; Barthélémy, A.; Ghosez, P.; Bellaiche, L.; Bibes, M.
orcid.org/0000-0001-5930-9916; Email: ld301@ Large elasto-optic effect and reversible electrochromism in multi-
cam.ac.uk ferroic BiFeO3. Nat. Commun. 2016, 7, 10718−10727.
Pratap Kollu − School of Physics, University of Hyderabad, (8) Banoth, P.; Sohan, A.; Kandula, C.; Kollu, P. Structural,
Hyderabad, Telangana 500046, India; CASEST, School of dielectric, magnetic, and ferroelectric properties of bismuth ferrite
Physics, University of Hyderabad, Hyderabad, Telangana (BiFeO3) synthesized by a solvothermal process using hexamethyle-
500046, India; orcid.org/0000-0001-6250-9396; netetramine (HMTA) as precipitating agent. Ceram. Int. 2022, 48,
Email: pratapk@uohyd.ac.in 32817−32826.
(9) Morozov, M. I.; Lomanova, N. A.; Gusarov, V. V. Specific
Authors features of BiFeO3 formation in a mixture of bismuth (III) and iron
Pravallika Banoth − School of Physics, University of (III) oxides. Russ. J. Gen. Chem. 2003, 73, 1676−1680.
Hyderabad, Hyderabad, Telangana 500046, India (10) Valant, M.; Axelsson, A. K.; Alford, N. Peculiarities of a solid-
state synthesis of multiferroic polycrystalline BiFeO3. Chem. Mater.
Chinna Kandula − School of Physics, University of Hyderabad,
2007, 19, 5431−5436.
Hyderabad, Telangana 500046, India (11) Liu, G.; Wang, T.; Zhou, W.; Meng, X.; Zhang, H.; Liu, H.;
Praveen Kumar Lavudya − School of Physics, University of Kako, T.; Ye, J. Crystal-facet-dependent hot-electron transfer in
Hyderabad, Hyderabad, Telangana 500046, India plasmonic-Au/semiconductor heterostructures for efficient solar
Saidulu Akaram − CASEST, School of Physics, University of photocatalysis. J. Mater. Chem. C 2015, 3, 7538−7542.
Hyderabad, Hyderabad, Telangana 500046, India (12) Liu, G.; Wang, T.; Zhou, W.; Meng, X.; Zhang, H.; Liu, H.;
RajaniKanth Ammanabrolu − School of Physics, University of Kako, T.; Ye, J. Crystal-facet-dependent hot-electron transfer in
Hyderabad, Hyderabad, Telangana 500046, India plasmonic-Au/semiconductor heterostructures for efficient solar
Ghanashyam Krishna Mamidipudi − School of Physics, photocatalysis. J. Mater. Chem. C 2015, 3, 7538−7542.
University of Hyderabad, Hyderabad, Telangana 500046, (13) Wang, X.; Mao, W.; Zhang, J.; Han, Y.; Quan, C.; Zhang, Q.;
India; CASEST, School of Physics, University of Hyderabad, Yang, T.; Yang, J.; Li, X. A.; Huang, W. Facile fabrication of highly
Hyderabad, Telangana 500046, India efficient g-C3N4/BiFeO3 nanocomposites with enhanced visible light
photocatalytic activities. J. Colloid Interface Sci. 2015, 448, 17−23.
Complete contact information is available at: (14) Tian, L.; Xu, J.; Just, M.; Green, M.; Liu, L.; Chen, X. Broad
https://pubs.acs.org/10.1021/acsomega.3c00553 range energy absorption enabled by hydrogenated TiO2 nanosheets:
from optical to infrared and microwave. J. Mater. Chem. C 2017, 5,
Notes 4645−4653.
The authors declare no competing financial interest. (15) Li, K.; Xu, J.; Yan, X.; Liu, L.; Chen, X.; Luo, Y.; He, J.; Shen,
D. Z. The origin of the strong microwave absorption in black TiO2.
■ ACKNOWLEDGMENTS
We acknowledge the facilities provided by the University
Appl. Phys. Lett. 2016, 108, 183102.
(16) Xia, T.; Zhang, C.; Oyler, N. A.; Chen, X. Hydrogenated TiO2
nanocrystals: a novel microwave absorbing material. Adv. Mater. 2013,
Grants Commission-Networking Resources (UGC-NRC) of 25, 6905−6910.
the School of Physics, University of Hyderabad, India. C.K. (17) Jiang, X.; Zhang, Y.; Jiang, J.; Rong, Y.; Wang, Y.; Wu, Y.; Pan,
thanks CSIR (CSIR file no. 09/414(2018)/2019-EMR-I) for C. Characterization of oxygen vacancy associates within hydrogenated

18661 https://doi.org/10.1021/acsomega.3c00553
ACS Omega 2023, 8, 18653−18662
ACS Omega http://pubs.acs.org/journal/acsodf Article

TiO2: a positron annihilation study. J. Phys. Chem. C 2012, 116, (37) Ajibade, P. A.; Oluwalana, A. E.; Andrew, F. P. Morphological
22619−22624. studies, photocatalytic activity, and electrochemistry of platinum
(18) Hu, Z.; Li, Q.; Li, M.; Wang, Q.; Zhu, Y.; Liu, X.; Zhao, X.; Liu, disulfide nanoparticles from bis (morpholinyl-4-carbodithioato)-
Y.; Dong, S. Ferroelectric memristor based on Pt/BiFeO3/Nb-doped platinum (II). ACS Omega 2020, 5, 27142−27153.
SrTiO3 heterostructure. Appl. Phys. Lett. 2013, 102, 102901. (38) Salama, A.; Mohamed, A.; Aboamera, N. M.; Osman, T. A.;
(19) Qu, T. L.; Zhao, Y. G.; Yu, P.; Zhao, H. C.; Zhang, S.; Yang, L. Khattab, A. Photocatalytic degradation of organic dyes using
F. Exchange bias effects in epitaxial Fe3O4/BiFeO3 heterostructures. composite nanofibers under UV irradiation. Appl. Nanosci. 2018, 8,
Appl. Phys. Lett. 2012, 100, 242410. 155−161.
(20) Ren, Y.; Nan, F.; You, L.; Zhou, Y.; Wang, Y.; Wang, J.; Su, X.;
Shen, M.; Fang, L. Enhanced photoelectrochemical performance in
reduced graphene oxide/BiFeO3 heterostructures. Small 2017, 13,
1603457.
(21) Lee, H.; Joo, H. Y.; Yoon, C.; Lee, J.; Lee, H.; Choi, J.; Park, B.;
Choi, T. Ferroelectric BiFeO3/TiO2 nanotube heterostructures for
enhanced photoelectrochemical performance. Curr. Appl. Phys. 2017,
17, 679−683.
(22) Alferov, Z. I. The history and future of semiconductor
heterostructures. Semiconductors 1998, 32, 1−14.
(23) Bagvand, N.; Sharifnia, S.; Karamian, E. A visible-light-active
BiFeO3/ZnS nanocomposite for photocatalytic conversion of green-
house gases. Korean J. Chem. Eng. 2018, 35, 1735−1740.
(24) Banoth, P.; Sohan, A.; Kandula, C.; Kanaka, R. K.; Kollu, P.
Microwave-Assisted Solvothermal Route for One-Step Synthesis of
Pure Phase Bismuth Ferrite Microflowers with Improved Magnetic
and Dielectric Properties. ACS Omega 2022, 7, 12910−12921.
(25) Gadhi, T. A.; Hernández, S.; Castellino, M.; Chiodoni, A.;
Husak, T.; Barrera, G.; Allia, P.; Russo, N.; Tagliaferro, A. Single
BiFeO3 and mixed BiFeO3/Fe2O3/Bi2Fe4O9 ferromagnetic photo-
catalysts for solar light driven water oxidation and dye pollutants
degradation. J. Ind. Eng. Chem. 2018, 63, 437−448.
(26) Vanga, P. R.; Mangalaraja, R. V.; Ashok, M. Structural,
magnetic and photocatalytic properties of La and alkaline co-doped
BiFeO3 nanoparticles. Mater. Sci. Semicond. Process. 2015, 40, 796−
802.
(27) Tseng, W. J.; Lin, R. D. BiFeO3/α-Fe2O3 core/shell composite
particles for fast and selective removal of methyl orange dye in water.
J. Colloid Interface Sci. 2014, 428, 95−100. Recommended by ACS
(28) Kuo, T. R.; Liao, H. J.; Chen, Y. T.; Wei, C. Y.; Chang, C. C.;
Chen, Y. C.; Chang, Y. H.; Lin, J. C.; Lee, Y. C.; Wen, C. Y.; Li, S. S.; Persulfate Activation of Photocatalysts Based on S-Scheme
et al. Extended visible to near-infrared harvesting of earth-abundant Bi3NbO7/BiOBr0.75I0.25 Nanosheet Heterojunctions for
FeS 2−TiO 2 heterostructures for highly active photocatalytic Degradation of Organic Pollutants
hydrogen evolution. Green Chem. 2018, 20, 1640−1647.
(29) Li, Y.; Lv, K.; Ho, W.; Zhao, Z.; Huang, Y. Enhanced visible- Yingyue Hu, Yi Du, et al.
light photo-oxidation of nitric oxide using bismuth-coupled graphitic JUNE 10, 2023
ACS APPLIED NANO MATERIALS READ
carbon nitride composite heterostructures. Chin. J. Catal. 2017, 38,
321−329.
(30) Gibertini, M.; Koperski, M.; Morpurgo, A. F.; Novoselov, K. S. Efficient p–n Heterojunction Photocatalyst Composed of
Magnetic 2D materials and heterostructures. Nat. Nanotechnol. 2019, Bismuth Oxyiodide and Layered Titanate
14, 408−419. Taya Ko Saothayanun, Makoto Ogawa, et al.
(31) Alferov, Z. I. The history and future of semiconductor DECEMBER 02, 2022
heterostructures. Semiconductors 1998, 32, 1−14. INORGANIC CHEMISTRY READ
(32) Chakraborty, S.; Pal, M. Enhanced and ultrafast dye
contaminated wastewater remediation using molybdenum doped Calcination-Induced Oxygen Vacancies Enhancing the
bismuth ferrite nanoparticles. Beilstein Arch. 2019, 1, 2019103. Photocatalytic Performance of a Recycled Bi2O3/BiOCl
(33) Zhang, N.; Chen, D.; Niu, F.; Wang, S.; Qin, L.; Huang, Y. Heterojunction Nanosheet
Enhanced visible light photocatalytic activity of Gd-doped BiFeO3
nanoparticles and mechanism insight. Sci. Rep. 2016, 6, 26467. Peng Li, Daoguang Teng, et al.
(34) Nguyen Thi Thu, T.; Nguyen Thi, N.; Tran Quang, V.; Nguyen DECEMBER 07, 2022
ACS OMEGA READ
Hong, K.; Nguyen Minh, T.; Le Thi Hoai, N. Synthesis, character-
isation, and effect of pH on degradation of dyes of copper-doped
TiO2. J. Exp. Nanosci. 2016, 11, 226−238. “Quasi-In Situ Synthesis of Oxygen Vacancy-Enriched
(35) Azeez, F.; Al-Hetlani, E.; Arafa, M.; Abdelmonem, Y.; Nazeer, Strontium Iron Oxide Supported on Boron-Doped Reduced
A. A.; Amin, M. O.; Madkour, M. The effect of surface charge on Graphene Oxide to Elevate the Photocatalytic Destruction...
photocatalytic degradation of methylene blue dye using chargeable Rajaraman Preetha, Aruljothy John Bosco, et al.
titania nanoparticles. Sci. Rep. 2018, 8, 7104−7109. MAY 10, 2023
(36) Sajid, M. M.; Shad, N. A.; Javed, Y.; Khan, S. B.; Zhang, Z.; LANGMUIR READ
Amin, N.; Zhai, H. Morphological effects on the photocatalytic
performance of FeVO4 nanocomposite. Nano-Struct. Nano-Objects Get More Suggestions >
2020, 22, 100431.

18662 https://doi.org/10.1021/acsomega.3c00553
ACS Omega 2023, 8, 18653−18662

You might also like