You are on page 1of 10

Res Chem Intermed

https://doi.org/10.1007/s11164-018-3268-5

Advanced ­Bi2O2.7/Bi2Ti2O7 composite film


with enhanced visible‑light‑driven activity
for the degradation of organic dyes

Wenzhang Fang1 · Li Zhou1 · Bin Shen1 · Yi Zhou1 · Qiuying Yi1 ·


Mingyang Xing1 · Jinlong Zhang1

Received: 24 February 2017 / Accepted: 30 October 2017


© Springer Science+Business Media B.V., part of Springer Nature 2018

Abstract Bi2O2.7/Bi2Ti2O7 composite photocatalyst films are synthesized by sol–


gel dip-coating. The ratio of adding Bi and Ti precursors can be controlled during
the preparation process. The phase structure is confirmed by X-ray diffraction. The
UV–visible diffuse reflectance spectrum shows that the composite catalysts present
light absorption in the visible region. The obtained B ­ i2O2.7/Bi2Ti2O7 composite
films possess superior photocatalytic degradation of rhodamine B, owing to the vis-
ible light response of B
­ i2O2.7 and the separation of photogenerated electrons and
holes between the two components. As a result, the B ­ i2O2.7/Bi2Ti2O7 (Bi/Ti = 1:1)
displays the highest photocatalytic activity under visible light or UV light irradiation
for the degradation of different organic dyes, including methyl blue, methyl orange
and acid orange 7.

Keywords Bi2O2.7 · Bi2Ti2O7 · Composite film · Photocatalysis · Organic dyes

Introduction

Titanium oxide ­(TiO2) is considered as one of the most promising catalysts since it was
applied to produce hydrogen from water splitting [1]. Owning to its abundance, non-
toxicity and chemical stability, T
­ iO2 can be widely used in catalysis, cosmetic, paints,
solar cells and antibacterial agents. However, the wide band gap of ­TiO2 (3.2 eV for
anatase phase) limits its applications to the UV region of the solar spectrum. Many

* Mingyang Xing
mingyangxing@ecust.edu.cn
* Jinlong Zhang
jlzhang@eucst.edu.cn
1
Key Laboratory for Advanced Materials and Institute of Fine Chemicals, East China University
of Science and Technology, 130 Meilong Road, Shanghai 200237, People’s Republic of China

13
W. Fang et al.

efforts have been made to extend its absorption edge to the visible region and increase
the photocatalytic performance under solar light, such as metal or non-metal doping
[2–6], self-reduction [7–10], noble metal deposit [11–13], ­TiO2-based compositing
[14–18], dye-sensitization [19], etc. Apart from ­TiO2, ­Bi2O3 is one of the potential vis-
ible light semiconductors with a direct band gap of 2.8 eV. There are five main poly-
morphic forms of ­Bi2O3, known as α, β, γ, δ [20] and ε-Bi2O3 [21] phases. In addition,
a transition phase ­(Bi2O2.7) was also observed in the transformation from α-, β- and
ε-Bi2O3 to δ-Bi2O3 [22]. Because of the narrow bandgap and suitable optical properties
of ­Bi2O3 [23, 24], ­Bi2O3 is reported to present high photodegradation of rhodamine B
(RhB) dye under visible light irradiation [25]. However, the photogenerated electron
and hole pairs on ­Bi2O3 tend to recombine easily, which greatly reduces its photocata-
lytic performance.
Bian et al. [26] reported the strong photosensitizing effect of B ­ i2O3 on ­TiO2 pho-
tocatalyst. Hou et  al. and Huo et  al. also reported that ­Bi2O3/TiO2 materials can be
easily activated by visible light owing to the photosensitizing effect of B ­ i2O3 quantum
dots [27, 28]. The photosensitizing of ­Bi2O3 can extend the solar light absorption edge
of ­TiO2 from the UV to the visible region, and the strong chemical bonding between
­Bi2O3 and T ­ iO2 facilitates the interface charge transfer in the heterostructures, which
finally results in the enhanced photocatalytic activity of ­Bi2O3/TiO2 materials under
visible light irradiation [26, 29–31].
Bi2Ti2O7 is one of phases of ­Bi2O3–TiO2 system which usually consists of five com-
pounds ­(Bi4Ti3O12, ­Bi2Ti4O11, ­Bi12TiO20, ­Bi8TiO14 and ­Bi2Ti2O7). The synthesis and
properties of these five bismuth titanate compounds have been extensively studied;
however, no investigation has been reported in the photocatalysis field [32]. In addi-
tion, it is difficult to identify a single phase of the bismuth titanate pyrochlore structure.
Hou et al. [33] reported the synthesis of ­Bi2Ti2O7 via metallorganic decomposition and
its dielectric and ferroelectric properties. However, Esquivel-Elizondo et al. [32] ques-
tioned the presence of B ­ i4Ti3O12 phase in this material, and the authors gave a brief
description of identifying the presence of impurities in bismuth titanate compounds
based on the color of the materials.
Although Bi- and Ti-based photocatalysts have been widely investigated and plenty
of applications have been achieved, powder samples are limited for practical applica-
tions because they are difficult to collect and reuse. Coating the catalysts on the sur-
face of glass substrates could be an effective way for solving these problems. Here,
we synthesized ­Bi2O2.7/Bi2Ti2O7 composite photocatalyst film via sol–gel dip-coating.
­Bi2O2.7 presents visible light absorption because of its narrow band gap, and the com-
posite of ­Bi2O2.7 and ­Bi2Ti2O7 contributes to the inhibition of photogenerated elec-
tron–hole recombination. As a result, the prepared ­Bi2O2.7/Bi2Ti2O7 composite films
show enhanced absorption of visible light and high photocatalytic performance under
visible light irradiation.

13
Advanced ­Bi2O2.7/Bi2Ti2O7 composite film with…

Experimental

Preparation of ­Bi2O2.7/Bi2Ti2O7 composite materials

Glass substrates were placed into HCl aqueous solution under ultrasonic treatment
for 30  min, followed by washing with distilled water. Then, they were placed in
ethanol under ultrasonic waves for 30  min and washed with distilled water again.
Finally, the glass substrates were placed in distilled water with ultrasonic treatment
for 30 min. After drying in the oven, substrates with clean surfaces were obtained.
Next, 150  mL ethanol was added to a beaker and stirred at 85  °C for a few
minutes. A certain amount of oxalic acid dehydrate ­(C2H2O4·2H2O) was added
to the beaker until the pH value of the solution reached 3–4.5. Then, 17.55  g
Bi(NO3)3·5H2O was added to the solution and stirred at 90 °C for 1 h. After cool-
ing to room temperature, 150 mL of a transparent solution of Bi(NO3)3·5H2O with a
concentration of 0.2 mol/L was obtained.
An amount of 9.968 g polyethylene glycol (PEG2000) was added into 100 mL of
4 mol/L HCl aqueous solution under ultrasonic treatment. When the PEG2000 was
dissolved, 7.83 mL T ­ iCl4 was added and the obtained solution was kept in a water
bath at 80 °C for 30 min. After aging for 12 h, a Ti-based sol was finally obtained.
A series of mixtures with different Bi/Ti molar ratios were prepared by mixing
different amounts of Bi precursor (V1 mL) and Ti precursor (V2 mL). The Ti precur-
sor was added into the Bi sol at a speed of one drop per second, and stored in the
chamber (25 °C, 40% RH). The molar ratios and amounts of added Bi and Ti precur-
sors are listed in Table 1.
Bi2O2.7/Bi2Ti2O7 composite film was synthesized by a sol–gel dip-coating method
from the mixture of Bi and Ti precursors. The extraction speed was controlled to
0.24 cm/s. Finally, the coated implants were dried in an oven at 100 °C or by irradia-
tion under UV light. The obtained samples after UV irradiation for 3 h were noted as
­Bi2O2.7/Bi2Ti2O7 (m:n), where m:n refers to the Bi/Ti molar ratio.

Characterization

XRD was carried out at room temperature on a Rigaku D/Max series machine
equipped with Cu Kα radiation. The machine was operated at 40 kV and 100 mA.
The diffraction patterns were recorded in the region of 10°–80° with a scanning

Table 1  List of the ratio and nBi:nTi V1 (mL) V2 (mL)


corresponding adding amount of
Bi- and Ti-based precursors 4:1 23 2
2:1 21.4 3.6
1:1 29.67 10
1:2 15 10
1:4 10.6 14.4
1:6 8.3 16.7

13
W. Fang et al.

speed of 8°/min. Scanning electron microscopy (SEM) was detected at 15  kV on


a JSM-6360 (JEOL). The glass substrate coated with the catalyst was cut into a
2  cm  ×  2  cm piece. A Fourier transform infrared spectroscopy (FTIR) spectrum
was obtained on a Nicolet Magna 550 in the region 400–4000 cm−1. The B ­ i2O2.7/
Bi2Ti2O7 composite sample was scraped off the substrate and mixed with KBr. A
UV–Vis DRS test was collected at room temperature on a Varian Cary 500 using
­BaSO4 as the reference from 400 to 800 nm. The optical property was obtained by
testing the transmittance of the sample on the Varian Cary 100. A blank glass sub-
strate was used as the reference.

Photocatalytic study

Photodegradation of RhB was carried out to evaluate the photocatalytic activi-


ties of the composite materials. Several other organic compounds were also used,
e.g., methyl blue (MB, 5  mg/L), methyl orange (MO, 5  mg/L) and acid orange 7
(AO7, 5  mg/L). The glass substrate coated with B ­ i2O2.7/Bi2Ti2O7 composite film
was placed in the photoreactor which contained 60  mL RhB solution (5  mg/L).
The mixture was stirred in the dark for 30 min to achieve an adsorption–desorption
equilibrium. A 500-W tungsten halogen lamp with a UV cut-off filter (λ > 420 nm)
was used to simulate the photoreaction under visible light. The lamp was placed
in a quartz jacket equipped with a recycling water-cooling system. To simulate
UV light irradiation, a 300-W high-pressure Hg lamp with a UV-cut optical filter
(λ < 365 nm) was employed. Sampling was carried out every 30 min. The mixture
was centrifuged at 12,000 rpm to remove the precipitate and finally analyzed by a
UV–Vis spectrophotometer (Shimadzu UV-2450).

Results and discussion

The XRD patterns of the synthesized composite materials starting with different Bi/
Ti ratios are shown in Fig.  1, which reveals that all the composite materials con-
sist of two materials, ­Bi2O2.7 (JCPDS 75-0993) [22] and ­Bi2Ti2O7 (JCPDS 32-0118)
[32–34]. The composite with the lowest Bi/Ti molar ratio shows the characteristic
peaks of the B­ i2Ti2O7 phase. With the increase of the Bi/Ti molar ratio, ­Bi2O2.7
begins to form along with the B ­ i2Ti2O7 pyrochlore phase (Fig. 1). Deng et al. [22]
reported that the tetragonal phase ­Bi2O2.7 is proved to be in existence by the trans-
formation from α-, β-, ε- to δ-Bi2O3 phase by density functional theory calculations.
Scanning electron microscopy is employed to investigate the morphology of
nano-materials because photocatalytic activities are usually affected by the mor-
phology of the samples. The SEM image of B ­ i2O2.7/Bi2Ti2O7 composite materials
with Bi/Ti = 1:1 shows that the clusters of the sample have a rough surface (Fig. 2).
Because the growing of the sample is not controlled by any surfactants, the sample
does not show any special morphology.
The FT-IR spectrum of ­Bi2Ti2O7/TiO2 composite material with Bi/Ti  =  1:1 is
shown in Fig. 3. The main absorption regions located at 500–700 and 812 cm−1 are

13
Advanced ­Bi2O2.7/Bi2Ti2O7 composite film with…

Fig. 1  XRD patterns of ­Bi2O2.7/Bi2Ti2O7 composite materials for different Bi/Ti molar ratios shown on
the top right 

Fig. 2  SEM image of B
­ i2O2.7/Bi2Ti2O7 (1:1) composite material

the typical Ti–O–Ti stretching vibration peaks, while the peaks at 536, 2929, 1378
and 872  cm−1 are ascribed to Bi–O peaks, which implies the composition of the
Bi- and Ti-based materials. The two peaks at 3419 and 1606 cm−1 are ascribed to
symmetrical stretching vibration and bending vibration of –OH, which should be
attributed to the adsorbed water on the surface of the catalyst.
It is widely accepted that the band gap of one semiconductor catalyst directly
affects its solar light response. The UV–Vis diffuse reflectance spectrum of
­Bi2O2.7/Bi2Ti2O7 composite catalysts shows that the samples present light absorp-
tion in the visible region (Fig. 4a). As calculated by the Kubelka–Munk function,
the band gap of B ­ i2O2.7/Bi2Ti2O7 (1:1) turns out to be 2.75 eV, which is smaller

13
W. Fang et al.

Fig. 3  FT-IR images of B
­ i2O2.7/Bi2Ti2O7 (1:1) composite material

Fig.  4  a UV–Vis diffuse reflectance spectrum of B


­ i2O2.7/Bi2Ti2O7 composite catalysts with different
Bi:Ti proportions. b Light transmittance spectrum of the samples: curve a, ­Bi2O2.7/Bi2Ti2O7 (1:1) film;
curve b, pure B
­ i2O3 film

than those of β-Bi2O3 (2.85 eV) and ­TiO2 anatase phase (3.2 eV). The light trans-
mittance spectrum of ­Bi2O2.7/Bi2Ti2O7 (1:1) film also shows two wide peaks in
the visible light area (Fig.  4b), indicating the effect of light interference on the

13
Advanced ­Bi2O2.7/Bi2Ti2O7 composite film with…

film. In general, an enhanced solar light absorption of the catalyst indicates a bet-
ter photocatalytic activity under visible light irradiation.
The photocatalytic degradation of organic compounds such as RhB under vis-
ible irradiation was introduced to investigate the photocatalytic performance of these
Bi- and Ti-based composite samples. Figure 5 shows that the photocatalytic activity
enhances gradually with the increase of Bi/Ti ratio, and the sample B ­ i2O2.7/Bi2Ti2O7
(1:1) shows the highest photocatalytic degradation of RhB under visible light irradi-
ation. However, additional injection of the Bi precursor does harm to the photocata-
lytic performance, and the activity starts to decrease while the Bi/Ti ratio increases
to 2:1. Sample ­Bi2O2.7/Bi2Ti2O7 (4:1) presents the lowest activity among all these
samples.
According to several papers in the literature, pure ­Bi2O3 can be activated under
visible light because of its narrow band gap compared to T ­ iO2 [26, 27]. However,
the photogenerated electrons and holes are easily recombined. As one of various
­Bi2O3 phases, B­ i2O2.7 has been seldom reported concerning its photocatalytic activ-
ity as a single photocatalyst. Therefore, doping ­Bi2O3 phases with other photocata-
lysts has been developed, and it has been proven to be an efficient way to achieve
enhanced visible light photocatalytic activities because of the photosensitizing effect
of ­Bi2O3 phases with narrow band gaps [35]. For ­Bi2O2.7/Bi2Ti2O7 composite film,
the electrons generated under visible light can transfer from the ­Bi2O2.7 to ­Bi2Ti2O7,
leading to the inhibition of the recombination of electrons and holes [27]. Finally,
the enhanced absorption of solar light and reduced recombination rate of electrons
and holes contribute to the increased photocatalytic performance under visible light.
In order to investigate the photocatalytic degradation rates of RhB of B ­ i2O2.7/
Bi2Ti2O7 composite films with various Bi/Ti ratios, a schematic illustration of the
composite materials was proposed (Fig. 6). While the percentage of ­Bi2O2.7 is low,
it is loaded on the surface of ­Bi2Ti2O7. Although the relatively narrow bandgap of

Fig. 5  Photodegradation of RhB for ­Bi2O2.7/Bi2Ti2O7 composite films with different Bi:Ti proportions
under visible light irradiation

13
W. Fang et al.

Fig. 6  Schematic illustration of the composite materials with different Bi/Ti ratios: a low Bi/Ti ratio; b
reasonable Bi/Ti ratio; c high Bi/Ti ratio

­Bi2O2.7 allows solar light absorption to the visible region, the low concentration
of ­Bi2O2.7 cannot introduce sufficient photogenerated electron and hole pairs in
the photocatalytic process, as exhibited in Fig. 6a. If extra Bi-based precursor was
injected, the obtained B
­ i2O2.7 will cover the surface of B
­ i2Ti2O7 (Fig.  6c). As the
result, although visible light absorption can be achieved, the photocatalytic perfor-
mance of the catalyst will be inhibited because of the high recombination rate of
photogenerated electron and hole pairs. Only when Bi and Ti precursors with a rea-
sonable ratio (e.g., 1:1) were added during the preparation, could both the enhance-
ment of solar light absorption and reduced electron–hole recombination rate be
achieved simultaneously.
We have investigated the effect of post-drying or post-UV irradiation pro-
cesses on the photocatalytic performance of the catalysts. Figure 7 shows that the

Fig. 7  Visible light-driven photocatalytic activity test for RhB degradation of ­Bi2O2.7/Bi2Ti2O7 compos-
ite materials (Bi/Ti = 1:1) with different post-processing conditions

13
Advanced ­Bi2O2.7/Bi2Ti2O7 composite film with…

Fig. 8  Photocatalytic activity test of B


­ i2O2.7/Bi2Ti2O7 composite materials (Bi/Ti  =  1:1) for different
dyes (AO7, RhB, MO, MB) under visible (a) and UV light (b) irradiation

samples treated by UV irradiation present better RhB degradation activity than


the film treated by thermal conditions. And the sample treated by UV irradiation
for 3 h shows the highest photocatalytic activity among all the samples. After UV
light irradiation, the dehydration could be realized on the gels which contained
massive –OH bonds, resulting in the densification of the films. Therefore, post-
UV irradiation treatment is a simple and effective way to obtain dense films at
room temperature.
Photocatalytic degradations of several other common organic compounds (AO7,
MO and MB aqueous solutions, 5 mg/L) were also tested under visible and UV light.
After irradiation under visible light for 150 min, RhB can be removed completely,
while AO7 and MO only show a low degradation rate of less than 8% (Fig. 8a). The
difference could be caused by the various sensitization degrees of these dyes under
visible light. Under UV light irradiation, the photodegradation rates of these dyes
show a huge difference (Fig. 8b). The sample is efficient for the photodegradation of
AO7, but inefficient for that of MB. The difference should be attributed to the differ-
ent adsorption abilities of various dyes.

Conclusions

Visible light-driven B
­ i2O2.7/Bi2Ti2O7 composite catalysts are firstly synthesized via
sol–gel dip-coating. The ratio of Bi/Ti can be easily controlled by the added amount
of Bi- and Ti-precursors. The obtained B ­ i2O2.7/Bi2Ti2O7 composite films present
superior photocatalytic degradation of rhodamine B under visible light irradiation
owing to its visible light response and the separation of photogenerated electron
and hole pairs between these two materials. Four common organic compounds have
been used as the target pollutants during the photodegradation tests under visible
and UV light. The composite catalyst also possesses considerable activities for the
photodegradation of AO7 and RhB under UV light irradiation.

13
W. Fang et al.

Acknowledgements  This work has been supported by National Nature Science Foundation of China
(21577036, 21377038, 21237003 and 21677048) and State Key Research Development Program of
China (2016YFA0204200) and sponsored by “Chenguang Program” supported by Shanghai Education
Development Foundation and Shanghai Municipal Education Commission (14CG30, 16JC1401400) and
the Fundamental Research Funds for the Central Universities (22A201514021).

References
1. A. Fujishima, K. Honda, Nature 238, 37 (1972)
2. M. Xing, J. Zhang, F. Chen, Appl. Catal. B 89, 563 (2009)
3. R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science 293, 269 (2001)
4. G. Wu, J. Wang, D.F. Thomas, A. Chen, Langmuir 24, 3503 (2008)
5. P. Xu, T. Xu, J. Lu, S. Gao, N.S. Hosmane, B. Huang, Y. Dai, Y. Wang, Energy Environ. Sci. 3,
1128 (2010)
6. W. Choi, A. Termin, M.R. Hoffmann, J. Phys. Chem. 98, 13669 (1994)
7. W. Fang, M. Xing, J. Zhang, Appl. Catal. B 160, 240 (2014)
8. G. Panomsuwan, A. Watthanaphanit, T. Ishizaki, N. Saito, Phys. Chem. Chem. Phys. 17, 13794
(2015)
9. L.B. Mo, Y. Wang, Y. Bai, Q.Y. Xiang, Q. Li, W.Q. Yao, J.O. Wang, K. Ibrahim, H.H. Wang, C.H.
Wan, J.L. Cao, Sci. Rep. 5, 17634 (2015)
1 0. W.Z. Fang, Y. Zhou, C.C. Dong, M.Y. Xing, J.L. Zhang, Catal. Today 266, 188 (2016)
11. S. Sakthivel, M.V. Shankar, M. Palanichamy, B. Arabindoo, D.W. Bahnemann, V. Murugesan,
Water Res. 38, 3001 (2004)
1 2. M. Haruta, Cattech 6, 102 (2002)
13. V. Subramanian, E.E. Wolf, P.V. Kamat, J. Am. Chem. Soc. 126, 4943 (2004)
14. H. Hu, J. Ding, S. Zhang, Y. Li, L. Bai, N. Yuan, Nanoscale Res. Lett. 8, 1 (2013)
15. M. Xing, D. Qi, J. Zhang, F. Chen, B. Tian, S. Bagwas, M. Anpo, J. Catal. 294, 37 (2012)
16. B. Qiu, Y. Zhou, Y. Ma, X. Yang, W. Sheng, M. Xing, J. Zhang, Sci. Rep. 5, 8591 (2015)
17. Y. Chen, S. Mishra, G. Ledoux, E. Jeanneau, M. Daniel, J. Zhang, S. Daniele, Chem. Asain J. 9,
2415 (2014)
1 8. L. Zhou, L. Wang, J. Zhang, J. Lei, Y. Liu, Res. Chem. Intermed. 43, 2081 (2016)
19. M.K. Nazeeruddin, P. Péchy, T. Renouard, S.M. Zakeeruddin, R. Humphry-Baker, P. Comte, P.
Liska, L. Cevey, E. Costa, V. Shklover, L. Spiccia, G.B. Deacon, C.A. Bignozzi, M. Grätzel, J. Am.
Chem. Soc. 123, 1613 (2001)
2 0. H.A. Harwig, Z. Anorg. Allg. Chem. 444, 151 (1978)
21. A.F. Gualtieri, S. Immovilli, M. Prudenziati, Powder Diffr. 12, 90 (1997)
22. H.-Y. Deng, W.-C. Hao, H.-Z. Xu, Chin. Phys. Lett. 28, 056101 (2011)
23. H.T. Fan, X.M. Teng, S.S. Pan, C. Ye, G.H. Li, L.D. Zhang, Appl. Phys. Lett. 87, 231916 (2005)
24. X. Dong, Y. Shao, X. Zhang, H. Ma, X. Zhang, F. Shi, C. Ma, M. Xue, Res. Chem. Intermed. 40,
2953 (2014)
2 5. L. Zhou, W. Wang, H. Xu, S. Sun, M. Shang, Chem. Eur. J. 15, 1776 (2009)
26. Z. Bian, J. Zhu, S. Wang, Y. Cao, X. Qian, H. Li, J. Phys. Chem. C 112, 6258 (2008)
27. J. Hou, C. Yang, Z. Wang, S. Jiao, H. Zhu, Appl. Catal. B 129, 333 (2013)
28. Y. Huo, X. Chen, J. Zhang, G. Pan, J. Jia, H. Li, Appl. Catal. B 148–149, 550 (2014)
29. K. Su, Z. Ai, L. Zhang, J. Phys. Chem. C 116, 17118 (2012)
30. L. An, G. Wang, Y. Cheng, L. Zhao, F. Gao, Y. Tian, Res. Chem. Intermed. 41, 7449 (2015)
31. J. Chen, S. Qin, Y. Liu, F. Xin, X. Yin, Res. Chem. Intermed. 40, 637 (2014)
32. J.R. Esquivel-Elizondo, B.B. Hinojosa, J.C. Nino, Chem. Mater. 23, 4965 (2011)
33. Y. Hou, M. Wang, X.-H. Xu, D. Wang, H. Wang, S.-X. Shang, J. Am. Ceram. Soc. 85, 3087 (2002)
34. W. Wei, Y. Dai, B. Huang, J. Phys. Chem. C 113, 5658 (2009)
35. J. Zhu, S. Wang, J. Wang, D. Zhang, H. Li, Appl. Catal. B 102, 120 (2011)

13

You might also like