You are on page 1of 9

Applied Surface Science 365 (2016) 209–217

Contents lists available at ScienceDirect

Applied Surface Science


journal homepage: www.elsevier.com/locate/apsusc

Fabrication of BiOBr nanosheets@TiO2 nanobelts p–n junction


photocatalysts for enhanced visible-light activity
Yang Zhao a,1 , Xiang Huang a,b,1 , Xin Tan b , Tao Yu a,c,∗ , Xiangli Li d , Libin Yang e ,
Shucong Wang d
a
School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, PR China
b
School of Science, Tibet University, Lhasa 850000, PR China
c
Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin 300072, PR China
d
School of Environmental Science and Engineering, Tianjin University, Tianjin 300072, PR China
e
College of Chemical Engineering and Materials Science, Tianjin University of Science & Technology, Tianjin Key Laboratory of Marine Resources and
Chemistry, Tianjin 300457, China

a r t i c l e i n f o a b s t r a c t

Article history: The construction of p–n junction structure is a smart strategy for improving the photocatalytic
Received 26 October 2015 activity, since p–n junctions can inhibit the recombination of photo-induced charges. Herein, BiOBr
Received in revised form nanosheets@TiO2 nanobelts p–n junction photocatalysts were prepared by assembling BiOBr nanosheets
24 December 2015
on the surface of TiO2 nanobelts via a hydrothermal route followed by a co-precipitation process.
Accepted 26 December 2015
BiOBr@TiO2 p–n junction photocatalysts exhibited enhanced photocatalytic activity in photocatalytic
Available online 4 January 2016
H2 production over water splitting and photodegradation of Rhodamine B (RhB) under visible light irra-
diation. Mott–Schottky plots confirmed the formation of p–n junctions in the interface of BiOBr and TiO2 .
Keywords:
Photocatalysis
The enhanced photocatalytic performance can be ascribed to the 1D nanostructure and the formation of
p–n junction p–n junctions. This work shows a potential application of low cost BiOBr as a substitute for noble metals
Water splitting in photocatalytic H2 production under visible light irradiation.
1D nanostructure © 2016 Elsevier B.V. All rights reserved.
Heterostructure

1. Introduction surface, both visible-light activity and the separation efficiency of


photo-induced carriers can be improved.
Titanium dioxide (TiO2 ), one of the most useful semiconductor Many semiconductor materials have been used as secondary
materials, has been widely studied for photocatalytic application phase hybridized with TiO2 , such as CdS [4], CuO [5], NiO [6],
due to its advantages, such as a suitable band gap for redox MoS2 [7,8], CeO2 [9], BiVO4 [10] and so on. In the last ten years,
reactions, high chemical stability, low cost, and non-toxicity [1]. Bismuth oxyhalides BiOX (X = Cl, Br, I) has became a promising pho-
However, there are two bottlenecks hindering the applications of tocatalyst due to its special layered structure with [Bi2 O2 ] slabs
TiO2 as photocatalysts. One is high recombination rate of the photo- interleaved by double slabs of halogen atoms along the [0 0 1] direc-
induced electron–hole pairs, and the other is that pure TiO2 can tion. The internal static electric fields between the [Bi2 O2 ]2+ and
hardly absorb the visible light due to its wide band gap [1]. To date, anionic halogen layers are believed to improve the separation of
many strategies have been developed to solve the above problem. photo-induced electron–hole pairs [11]. Notably, BiOX is a p-type
An effective alternative is to couple TiO2 with narrow bad gap semi- semiconductor [12–14], and BiOBr and BiOI with the band gaps
conductor or metal to form p–n junctions, Schottky junctions, and of ∼2.7 eV and ∼1.7 respectively [11] have visible-light response.
band-structure matching [2,3]. Based on the interface engineer- Therefore, BiOX, especially BiOI and BiOBr, is an attractive pho-
ing with the attachment of visibly active photocatalyst on the TiO2 tocatalyst that can couple with TiO2 to form p–n junction and
improve visible-light response. Recently, effort has been spent
in studying the hybrid photocatalyst of TiO2 with BiOCl [15–17],
BiOBr [17–19], and BiOI [12,13,17,20–25]. These reports mainly
∗ Corresponding author at: School of Chemical Engineering and Technology,
focused on the hybrid of TiO2 nanoparticles with BiOX. However,
Tianjin University, and Collaborative Innovation Center of Chemical Science and
it is hard for TiO2 nanoparticles to control the coverage and dis-
Engineering (Tianjin), Tianjin 300072, PR China.
E-mail address: yutao@tju.edu.cn (T. Yu). tribution of the secondary phase on the surface, which would
1
These authors contributed equally to this work. restrict the diffusion of photo-induced carriers to the TiO2 surface

http://dx.doi.org/10.1016/j.apsusc.2015.12.249
0169-4332/© 2016 Elsevier B.V. All rights reserved.
210 Y. Zhao et al. / Applied Surface Science 365 (2016) 209–217

[3]. Additionally, particle-like photocatalyst is inconvenient for spectrophotometer using BaSO4 as a reference in the wavelength
recycling. One-dimensional (1D) nanostructures, such as nanobelts, of 200–800 nm. The photoluminescence (PL) spectra were acquired
nanorods, nanowires [7–9,26–33] and nanotubes [34,35], have at room temperature with an F-4600 fluorescence spectrometer
attracted tremendous attention because of their specific morphol- under the ultraviolet excitation of 260 nm.
ogy and novel properties. In comparison with nanoparticles, 1D
nanostructure can supply relatively large surface area for assem- 2.3. Photocatalytic tests
bling various secondary phase materials and transport charge
carriers rapidly along the axial direction [3]. The photodegradation of RhB under monochromatic visible-
Herein, we report the fabrication of p–n heterostructure BiOBr- light irradiation ( = 420 nm) was conducted to evaluate the
TiO2 photocatalysts (BiOBr@TiO2 ) by a simple co-precipitation photocatalytic properties of the BiOBr@TiO2 nanobelt heterostruc-
method. Firstly, TiO2 nanobelts were prepared via a typical hydro- tures. A 300 W Xe lamp (HXS-F/UV 300, Beijing NBet Technology
thermal route. Then BiOBr nanosheets were coupled on the surface Co., Ltd) was used as light source. In each experiment, 50 mg of
of TiO2 nanobelts. BiOBr@TiO2 was firstly applied to photocatalytic photocatalyst powder was dispersed in 50 mL of 20 mg L−1 RhB
H2 production over water splitting under visible light irradiation. solution. Then, the suspensions were kept in the dark for 0.5 h with
In addition, BiOBr@TiO2 exhibited excellent photocatalytic activity continuous stirring before irradiation. During the photodegrada-
in the photodegradation of RhB. tion, the light intensity was maintained at about 100 mW cm−2 .
The concentration variation of RhB was monitored by a UV–vis
spectrophotometer (T6, Beijing Purkinje General Instrument Co.,
2. Experimental
Ltd). For comparison, physically-mixed TiO2 nanobelts/BiOBr
nanosheets were employed for the photodegradation of RhB under
2.1. Preparation of TiO2 nanobelts and BiOBr@TiO2
the same experimental condition.
heterostructures
In addition, H2 evolution from water splitting was performed
with ethylene glycol as sacrificial agent. 0.08 g of photocatalyst was
The chemicals used in this work were of analytical reagent
dispersed in the 80 mL aqueous solution. Before irradiation, the sys-
grade. The TiO2 nanobelts were synthesized via a typical hydro-
tem was bubbled with N2 for 30 min to remove the air inside. The
thermal method. Typically, 0.4 g of P25 was added in 40 mL of 10 M
amount of H2 evolution was determined on a gas chromatograph
NaOH aqueous solution under continuous stirring for 30 min. Then
(GC-2014, Shimadzu, Tokyo).
the mixture was poured into a Teflon-lined stainless steel auto-
clave (50 mL volume). The autoclave was maintained at 180 ◦ C for
2.4. Photoelectrochemical measurements
48 h and then cooled to room temperature naturally. The obtained
powder was washed with deionized water several times, followed
Photoelectrochemical properties of the as-prepared samples
by an immersion process in 0.1 M of HCl aqueous solution for 24 h.
were evaluated by a standard three-electrode configuration using
After that, the wet products were washed with deionized water to
an electrochemical workstation (CHI660E, Shanghai Chenhua,
pH 7, and then dried at 80 ◦ C over night. Finally, the TiO2 nanobelts
China). 1 M of Na2 SO4 solution was used as the supporting
were obtained by heat treatment for 2 h at 600 ◦ C.
electrolyte. The working electrode was prepared by depositing
BiOBr@TiO2 nanobelt heterostructures at various mole ratios
photocatalyst powders on 1.0 cm × 1.0 cm fluoride-tin oxide (FTO)
of 0.5:1, 1:1 and 1.5:1 (note as BiOBr@TiO2 -0.5, BiOBr@TiO2 -
glass. A saturated calomel electrode (SCE) and platinum wire
1, BiOBr@TiO2 -1.5) were prepared by a simple co-precipitation
counter electrode (˚ 0.5 mm × 37 mm) were employed as the ref-
method. In a typical procedure, an appropriate amount of KBr
erence and counter electrodes, respectively.
was dissolved in 20 mL deionized water. Meanwhile, a stoichio-
metric amount of Bi(NO3 )3 ·5H2 O was added to 20 mL deionized
3. Results and discussion
water to form a white suspension S1. Then 40 mg as-prepared TiO2
nanobelts were dispersed into the KBr solution under magnetic stir-
Fig. 1 shows the XRD patterns of pure BiOBr, TiO2 nanobelts,
ring to form suspension S2. The suspension S2 was slowly poured
BiOBr@TiO2 -0.5, BiOBr@TiO2 -1 and BiOBr@TiO2 -1.5. The corre-
into the suspension S1. After vigorous stirring for 30 min, the
sponding XRD of TiO2 nanobelts shows a mixed phase of TiO2
resulting precipitate was collected by centrifugation and washed
(B) and anatase. It can be observed from Fig. 1A(a) that TiO2 (B)
thoroughly with deionized water and ethanol. Finally, BiOBr@TiO2
is the main phase having three major diffraction peaks at 14.19◦ ,
heterostructures were obtained after dried at 60 ◦ C. For compar-
24.93◦ and 28.61◦ (JCPDS Card No.46-1237). Some weak response
ison, pure BiOBr nanosheets were synthesized under the same
peaks at 25.28◦ , 37.80◦ and 48.04◦ can be identified as the tetragonal
conditions in the absence of TiO2 nanobelts.
anatase phase (corresponding to JCPDS Card No. 21-1272). TiO2 (B),
TiO2 anatase and tetragonal BiOBr phases appear in the BiOBr@TiO2
2.2. Characterization nanobelt heterostructures, indicating the successful coupling of the
BiOBr nanosheets onto the TiO2 nanobelt surfaces. More specially,
The crystal structure of the products was identified by X-ray the diffraction peaks of BiOBr are intensified gradually with the
diffraction (XRD, Rigaku D/MAX-2500) with Cu K˛ radiation. The loading of BiOBr increasing. In addition, anatase phase begins to
Raman spectra were recorded using a DXR Microscope Raman disappear from BiOBr@TiO2 -1, which can be attributed to the small
spectrometer with a laser excitation of 532 nm at room temper- quantity of anatase and the coverage of the BiOBr nanosheets. As
ature. X-ray photoelectron spectroscopy (XPS) was employed for shown in Fig. 1B(a), two peaks at 24.93◦ and 25.28◦ can be assigned
the measurement of composition and chemical states of the sam- to the plane (1 1 0) of TiO2 (B) and the plane (1 0 1) of anatase
ples. The morphology of the samples was observed using field respectively, confirming the mixed phase of TiO2 nanobelts. Careful
emission scanning electron microscope (FESEM, FEI Nanosem 430). observation from Fig. 1B shows that the two diffraction peaks are
TEM images were conducted on a Tecnai G2 F20 transmission converted gradually into one peak and shift slightly to the larger
electron microscope with an accelerating voltage of 200 kV. The value of the diffraction angle from curve a to curve e, because the
specific surface areas of the samples were measured and calcu- peaks ascribed to TiO2 (B) and anatase are weakened, while the
lated by a BET apparatus (Quantachrome Autosorb iQ). The UV–vis peak ascribed to the plane (1 0 1) of BiOBr at 25.16◦ (JCPDS Card No.
diffuse reflectance spectra were measured on a Shimadzu UV-2550 09-0393) [36] is intensified with the increase of BiOBr.
Y. Zhao et al. / Applied Surface Science 365 (2016) 209–217 211

Fig. 1. XRD patterns of samples (a) pure TiO2 nanobelts, (b) BiOBr@TiO2 -0.5 (c) BiOBr@TiO2 -1, (d) BiOBr@TiO2 -1.5, and (e) pure BiOBr.

respectively. As shown in Fig. 4A, the as-prepared TiO2 nanobelts


are 50–200 nm wide, about 40 nm thick and of several micrometers
in length. Furthermore, it can be seen clearly that the surface of the
TiO2 nanobelts are smooth. Fig. 4 B is the typical SEM image of BiOBr
synthesized by the same route in the absence of TiO2 nanobelts.
The pure BiOBr exhibits flower-like 3D hierarchical structures
with an average diameter of ca. 1 ␮m, having radially grown
nanoplates which stacked and intercrossed with one another. After
co-precipitation procedure, the TiO2 nanobelts are decorated by
numerous thin BiOBr nanosheets and remain 1D nanostructure, as
shown in Fig. 4C–E. When the mass ratio of BiOBr to TiO2 was 0.5,
the density of BiOBr nanosheets on the TiO2 nanobelts was low,
so the smooth surface of TiO2 nanobelts still can be observed in
Fig. 4C. As the mass ratio was increased to 1, the density of the
Fig. 2. Raman spectra of TiO2 nanobelts, pure BiOBr and BiOBr@TiO2 -1.
nanosheets on the surface of TiO2 nanobelts increased significantly.
In particular, the nanosheets were distributed uniformly without
Fig. 2 shows Raman spectra of the as-prepared TiO2 nanobelts, aggregation, and the belt-like morphology was changed to be cylin-
BiOBr and BiOBr@TiO2 -1. The bands at 125, 247, 254, 297, 370, 411, drical shape, as shown in Fig. 4D. With the mass ratio was further
437, and 471 cm−1 are assigned to TiO2 (B) [32,37,38]. The band increased to 1.5, the petal-like BiOBr nanosheets can be observed,
around 145 cm−1 is attributed to the B1g vibration of anatase [38]. as shown in Fig. 4E. A careful observation reveals that BiOBr@TiO2 -
The above analysis shows that the as-prepared TiO2 nanobelts are of 1.5 has a larger size of diameter (about 900 nm) than BiOBr@TiO2 -1
mixed phase of anatase and TiO2 (B), which is well accordant with (about 500 nm), indicating more BiOBr nanosheets covered on the
XRD results. Carefully observation on the curve of BiOBr@TiO2 -1 surface of TiO2 nanobelts.
can find that the peak around 150 cm−1 is asymmetric and shifts Fig. 5A and B shows the typical TEM images of BiOBr@TiO2 -1.
to the band position of pure BiOBr, further confirming that BiOBr is BiOBr nanosheets still cover the TiO2 nanobelts after ultrasonic
successfully introduced on the surface of the TiO2 nanobelts. dispersion, confirming the stable combination between BiOBr and
XPS measurements were performed to explore the composi- TiO2 . The diameter of TiO2 nanobelts in Fig. 5A and B is about 65 nm
tion and elemental valence state of the BiOBr@TiO2 . As shown in and 180 nm respectively, which agrees well with the results of SEM.
Fig. 3A, the survey spectrum of BiOBr@TiO2 -1 shows the existence From the HRTEM images in Fig. 5C and D, the lattice fringes of BiOBr
of Ti, Bi, Br, O and C in the heterostructure. The peak for C 1s at nanosheets and TiO2 nanobelts can be seen clearly, suggesting the
284.8 eV is ascribed to adventitious carbon from the XPS instru- well-defined crystal structure. The fringes with the lattice spacing
ment. In Fig. 3B, two peaks at the binding energy of 159.3 eV and of 0.277 nm correspond to the (1 1 0) plane of BiOBr. Combining the
164.6 eV for pure BiOBr are assigned to Bi 4f7/2 and Bi 4f5/2 , respec- above results and the corresponding FFT pattern in Fig. 4D inset,
tively. However, the Bi 4f7/2 and Bi 4f5/2 peaks for BiOBr@TiO2 -1 the top and bottom surfaces of the BiOBr nanosheets are indenti-
have a slight shift to low binding energy, which can be ascribed to fied as {0 0 1} facets. The crystal structure of the mixed-phase TiO2
the strong chemical bonding between BiOBr and TiO2 [16,18,19]. As nanobelts can be confirmed in Fig. 5C. Two different fringes are
for the high-resolution spectra of Br 3d, the peaks located at 68.1 eV observed in TiO2 region. The main lattice spacing of ca. 0.506 nm
and 69.1 eV (Fig. 3C) are ascribed to Br 3d5/2 and Br 3d3/2 respec- matches well with the (2 0 1) plane of TiO2 (B). The other fringes
tively, corresponding to Br− [39,40]. The two symmetric peaks for with the spacing of 0.352 nm are ascribed to the (1 0 1) plane of
TiO2 nanobelts located at 458.5 eV and 464.2 eV are assigned to Ti anatase phase. Interestingly, cross fringes occur in the interface of
2p3/2 and Ti 2p1/2 , indicating that Ti in the sample is in the form of BiOBr and TiO2 (Fig. 5D), indicating the well combination of the
Ti4+ [23,24,41]. As for BiOBr@TiO2 -1, the peak of Ti 2p1/2 (Fig. 3D) two materials. In addition, the energy dispersive X-ray (EDX) ele-
has a broad bump due to the partial overlap between Ti 2p1/2 at mental mappings of Ti (Fig. 5H), Bi (Fig. 5I), Br (Fig. 5J), and O
464.4 eV and Bi 4d3/2 at 466.3 eV. The XPS results further confirm (Fig. 5G) further confirm the formation of the heterostructures,
the coexistence of TiO2 and BiOBr in the BiOBr@TiO2 heterostruc- in which TiO2 nanobelts act as the core and BiOBr nanosheets are
ture, and the BiOBr nanosheets were introduced on the surface of well-distributed outside. EDX analysis of BiOBr@TiO2 -1 presented
TiO2 nanobelts through chemical bond rather than a simple mixture in Fig. 5K indicates that the heterostructure is composed of Ti, Bi,
of the two materials. Br and O elements. The peaks of Cu and C are ascribed to the cop-
Fig. 4 presents the corresponding SEM images of pure per net and the carbon film used in the EDX characterization. The
TiO2 nanobelts, pure BiOBr, and BiOBr@TiO2 heterostructures, elemental content of BiOBr@TiO2 -1 is shown in Table S1.
212 Y. Zhao et al. / Applied Surface Science 365 (2016) 209–217

Fig. 3. XPS spectra of TiO2 nanobelts, pure BiOBr and BiOBr@TiO2 -1 (A) survey, (B) Bi 4f, (C) Br 3d, and (D) Ti 2p.

Fig. 6A shows the UV–vis diffuse reflectance spectra of the activity of the pure TiO2 nanobelts was negligible due to the wide
pure BiOBr, TiO2 nanobelts and BiOBr@TiO2 heterostructures. Pure band gap. Clearly, the sample TiO2 + BiOBr (physically-mixed TiO2
TiO2 nanobelts have almost no absorbance in the visible light nanobelts and BiOBr nanosheets) exhibits the lowest photocat-
range with an absorbance edge less than 400 nm, which can alytic performance with the degradation efficiency of 34.1% in
be assigned to the intrinsic band-gap absorption. As for pure 40 min, confirming the p–n junction effect. Moreover, the degra-
BiOBr, red shift of the optical absorption edge is observed clearly. dation rates of RhB over the BiOBr@TiO2 heterostructures are
The band gaps (Eg ) of the pure BiOBr and TiO2 nanobelts can higher than that over the pure BiOBr. Since the specific surface
be evaluated by ˛(hv) = A(hv − Eg )n/2 , where ˛, h, v, Eg and A area of the pure BiOBr is similar to those of BiOBr@TiO2 het-
are the absorption coefficient, Plank constant, light frequency, erostructures (Fig. S2), we believe that the enhanced photocatalytic
band gap energy and a constant, respectively. Among them, n activities of the BiOBr@TiO2 heterostructures are attributed to
depends on the characteristics of the transition in a semicon- the p–n junction effect rather than surface area. It is worth not-
ductor. The estimated band gaps are about 2.75 eV and 3.02 eV ing that BiOBr amount in the heterostructures has an influence
for pure BiOBr and TiO2 nanobelts (Fig. 6B), respectively, which on the photodegradation of RhB. When BiOBr@TiO2 -0.5 was used
is close to the previous report [11,27,38]. Further observation as photocatalyst under monochromatic light ( = 420 nm), 84%
from Fig. 6A shows that the absorption edges of BiOBr@TiO2 of RhB can be degraded in 40 min; while the degradation effi-
heterostructures lie ca. 420 nm–430 nm, and have a continuous ciency of RhB was increased to 92.8% over BiOBr@TiO2 -1 under
red-shift with the increase of the BiOBr nanosheets grown on TiO2 the same condition. This can be attributed to two major rea-
nanobelts. sons. One is that more BiOBr nanosheets grown on the surface
The photocatalytic activity of the BiOBr@TiO2 heterostructures of TiO2 are favorable for the absorption of visible light; the
was evaluated by the photodegradation of RhB under visible other is that the formation of more p–n junctions in BiOBr@TiO2 -
light irradiation. Considering that RhB itself with the maximum 1 facilitates the separation of photo-generated electrons and
absorption wavelength of 553 nm can absorb visible light and its holes, which leads to the enhanced photocatalytic activity. How-
degradation can be initiated from the excited dye [42], a monochro- ever, the degradation efficiency of RhB was decreased to 86.8%
matic visible light ( = 420 nm) was selected to eliminate the dye when BiOBr amount was continuously increased (BiOBr@TiO2 -
sensitization effect since it has a minimal spectral overlap with 1.5). The above results indicate that there is an optimum amount
RhB. As shown in Fig. S1, RhB over BiOBr@TiO2 heterostructures of BiOBr in the heterostructures. Excess BiOBr amount can
can be completely degraded only in 10 min under visible light also decrease the photocatalytic activity, because excess BiOBr
irradiation with an ultraviolet cutoff filter ( ≥ 420 nm), while it nanosheets grown on the surface of TiO2 nanobelts may act as
needs at least 40 min under monochromatic light ( = 420 nm) in the recombination center of photo-induced carriers, resulting in
Fig. 7A, confirming monochromatic light irradiation is favorable the rapid recombination of charges. To eliminate the adsorption
for the elimination of dye sensitization effect. Additionally, blank effect, the pseudo-first order model expressed by ln (C0 /C) = kt
test indicates that the exposure of RhB without photocatalysts was applied to the evaluation of the degradation rate, as shown
to visible light doesn’t stimulate self-photodegradation. Fig. 7A in Fig. 7B. The k values of the as-prepared photocatalysts are
shows the variation in RhB concentration over the pure BiOBr, in the order of BiOBr@TiO2 -1 (k = 0.061 min−1 ) > BiOBr@TiO2 -1.5
BiOBr@TiO2 heterostructures and BiOBr + TiO2 (physical mixture) (k = 0.048 min−1 ) > BiOBr@TiO2 -0.5 (k = 0.044 min−1 ) > pure BiOBr
under monochromatic light ( = 420 nm). The photodegradation (k = 0.033 min−1 ) > BiOBr + TiO2 (k = 0.009 min−1 ).
Y. Zhao et al. / Applied Surface Science 365 (2016) 209–217 213

Fig. 4. SEM images of (A) pure TiO2 nanobelts, (B) pure BiOBr, (C) BiOBr@TiO2 -0.5, (D) BiOBr@TiO2 -1, and (E) BiOBr@TiO2 -1.5.

Active species trapping experiments were carried out to inves- electrons and holes. The photodegradation RhB and photocat-
tigate the photodegradation process of RhB. As shown in Fig. S3, alytic H2 evolution indicated that BiOBr@TiO2 heterostructures
the RhB degradation was significantly suppressed by the addition possessed excellent activity in both photo-oxidation and photo-
of benzoquinone (BQ, a ·O− 2 scavenger) [39,43]. However, when reduction reactions.
EDTA (a hole scavenger) [39,43] and AgNO3 (a electron scavenger) It is well-established that BiOBr is a p-type semiconductor
[19,39] were added, there was no apparent effect on the degra- [12–14] and TiO2 is an n-type semiconductor [1,12,13], which was
dation process. Additionally, trapping ·OH with 2-butyl alcohol further confirmed by the Mott–Schottky plots in Fig. S4. As shown in
exhibited much weaker inhibition to the RhB photodegradation Fig. S4, the BiOBr nanosheets show a negative slope, while the TiO2
than trapping ·O− −
2 . Therefore, it can be deduced that ·O2 was the nanobelts show a positive slope, indicating that BiOBr is a p-type
dominative active species for the RhB photodegradation. semiconductor and TiO2 is an n-type semiconductor. When p-type
The photocatalytic H2 -generation activity was measured to BiOBr nanosheets grew on the surface of n-type TiO2 nanobelts, a
evaluate the as-prepared catalysts under visible light irradiation number of p–n junctions would form at the interface. Fig. 9A shows
( ≥ 420 nm). It can be seen from Fig. 8 that TiO2 nanobelts show an inverted “V-shape” in the Mott–Schottky plots of BiOBr@TiO2 -1.
negligible H2 -generation activity due to the poor visible-light The co-existence positive and negative slopes with similar values
response. Additionally, the H2 -production of pure BiOBr is slight. confirmed the formation of p–n junctions [44,45]. The valence band
However, the H2 evolution over the BiOBr@TiO2 heterostructures (VB) edge positions of the pure BiOBr and TiO2 nanobelts deter-
was significantly enhanced. Especially, there is an optimal ratio mined by the XPS VB spectra were 1.96 eV and 2.8 eV respectively,
of BiOBr to TiO2 in terms of H2 production. The BiOBr@TiO2 -1 as shown in Fig. 9B. Therefore, the conduction band (CB) edges of
exhibits highest H2 -generation activity up to 122 ␮mol−1 g−1 h−1 , the as-prepared BiOBr and TiO2 can be calculated to be −0.79 eV
which is 2-fold of BiOBr@TiO2 -0.5 and BiOBr@TiO2 -1.5. The results and −0.22 eV according to the equation ECB = EVB − Eg , where Eg is
are similar to the photodegradation of RhB. This is because that the band gap energy obtained by UV–vis diffuse reflection spectra
low loading of BiOBr couldn’t absorb sufficient visible light, while (Fig. 6B). The p–n junction band structure and the possible charge
excess loading of BiOBr resulted in the rapid recombination of separation mechanism are illustrated in Fig. 9C and D. TiO2 is a
214 Y. Zhao et al. / Applied Surface Science 365 (2016) 209–217

Fig. 5. (A and B) Low-magnification TEM images of BiOBr@TiO2 -1, (C and D) HRTEM images of the designated square parts in A and B, (E–J) EDX elemental mappings of
BiOBr@TiO2 -1, and (K) EDX analysis of BiOBr@TiO2 -1.

Fig. 6. (A) UV–vis diffuse reflection spectra of pure BiOBr, TiO2 nanobelts and BiOBr@TiO2 heterostructures, and (B) (ahv)1/2 vs. photon energy.
Y. Zhao et al. / Applied Surface Science 365 (2016) 209–217 215

Fig. 7. (A) Photocatalytic degradation efficiencies of RhB and (B) kinetics of RhB decolorization under visible light irradiation ( = 420 nm).

each other, the Femi level difference between BiOBr and TiO2 causes
electron transfer from TiO2 to BiOBr until the Femi levels of BiOBr
and TiO2 reach an equilibrium [46–48]. As a consequence, the Femi
levels of BiOBr and TiO2 rise up and descend respectively, and an
inner electric field is established from n-TiO2 to p-BiOBr (Fig. 9D).
When the p–n junction photocatalysts were irradiated with visi-
ble light, the electrons were excited from the VB of BiOBr to its CB
because the BiOBr nanosheets with the band gap energy of 2.75 eV
can absorb visible light. Driven by the inner electric field, the photo-
induced electrons on the CB of p-BiOBr transferred to that of n-TiO2 ,
and holes remained on the VB of p-BiOBr. Therefore, photo-induced
carriers were effectively separated by the p–n junction and the
recombination of the electrons and holes was retarded, which was
Fig. 8. Comparison of the photocatalytic hydrogen-generation activities of pure
beneficial to the photocatalytic activity.
BiOBr, TiO2 nanobelts, and BiOBr@TiO2 heterostructures under visible light irra-
diation ( ≥ 420 nm). To verify the above proposed mechanism, the photoluminesce-
nce (PL) emission spectra were measured under the ultraviolet
typical n-type semiconductor with Femi level located close to CB, excitation of 260 nm. The higher the PL intensity, the higher recom-
while BiOBr is a p-type semiconductor whose Femi level lies close bination rate of photo-generated electrons and holes. A comparison
to VB, as shown in Fig. 9C. Before contact, the Femi level of BiOBr is of pure BiOBr nanosheets, TiO2 nanobelts and BiOBr@TiO2 -1 is
lower than that of TiO2 . Once p-type BiOBr and n-type TiO2 contact shown in Fig. 10A. It can be clearly seen that the emission

Fig. 9. (A) Mott–Schottky plots of BiOBr@TiO2 -1, (B) XPS valence-band spectra of pure BiOBr nanosheets and TiO2 nanobelts, and schematic illustrations for (C) the band
energy of BiOBr and TiO2 before contact and (D) the formation of p–n junction and the charge transfer and separation process under visible light irradiation.
216 Y. Zhao et al. / Applied Surface Science 365 (2016) 209–217

Fig. 10. (A) Photoluminescence (PL) spectra of pure BiOBr, TiO2 nanobelts and BiOBr@TiO2 -1.0 at room temperature with the excitation wavelength of 260 nm, and (B)
photocurrent responses of pure BiOBr, TiO2 nanobelts and BiOBr@TiO2 heterostructures in 0.1 M Na2 SO4 aqueous solution under visible light irradiation ( ≥ 420 nm).

intensity of BiOBr@TiO2 -1 is much lower than those of pure BiOBr over water splitting under visible light irradiation. It is believed
and TiO2 , suggesting that the recombination of photo-induced that the formation of p–n junctions could reduce the recombination
electron–hole pairs is hampered. To further investigate the separa- of photo-induced electron–hole pairs and that 1D TiO2 nanobelts
tion of electron–hole pairs, the transient photocurrent responses could supply abundant active sites for forming p–n junctions and
of pure BiOBr, TiO2 + BiOBr, and three BiOBr@TiO2 heterostruc- promote charge carriers separation.
tures were measured under visible light irradiation ( ≥ 420 nm). As
shown in Fig. 10B, all the samples generate photocurrent promptly,
Acknowledgements
and BiOBr@TiO2 -1 exhibits the highest photocurrent intensity.
Carefully observation indicates that the photocurrent intensities
This work was supported by the National Natural Science Foun-
are in the order of TiO2 + BiOBr < BiOBr@TiO2 -0.5 < BiOBr@TiO2 -
dation of China (21406164, 21466035), the National Key Basic
1.5 ≈ pure BiOBr < BiOBr@TiO2 -1.0. It is not surprising that
Research and Development Program of China (973 program, No.
TiO2 + BiOBr show the lowest photocurrent intensity due to the
2014CB239300, and 2012CB720100), Research Fund for the Doc-
intrinsic wide band gap of TiO2 , which is in accord with the result
toral Program of Higher Education of China (Nos. 20110032110037,
of photodegradation RhB. When BiOBr nanosheets were deposited
20130032120019).
on the surface of TiO2 nanobelts, the absorption capacity for visible
light was improved, which led to the enhanced photocurrent inten-
sities of BiOBr@TiO2 heterostructures. However, the photocurrent Appendix A. Supplementary data
of BiOBr@TiO2 -1.5 is weaker than that of BiOBr@TiO2 -1.0 probably
because excess BiOBr nanosheets would act as the recombination Supplementary data associated with this article can be found, in
centers, resulting in lower charge separation efficiency. To provide the online version, at http://dx.doi.org/10.1016/j.apsusc.2015.12.
additional evidence for the above mechanism, electrochemical 249.
impedance spectroscopy (EIS) were carried out under dark con-
dition. It can be seen from Fig. S5 that the arc radii of the samples
References
are arranged similar to the trend of the photocurrent. BiOBr@TiO2 -
1.0 shows the smallest arc radius, which further demonstrates that [1] J. Schneider, M. Matsuoka, M. Takeuchi, J.L. Zhang, Y. Horiuchi, M. Anpo, D.W.
the BiOBr@TiO2 -1.0 electrode has a higher separation efficiency Bahnemann, Understanding TiO2 photocatalysis: mechanisms and materials,
of photo-generated electron–hole pairs and faster charge-transfer Chem. Rev. 114 (2014) 9919–9986.
[2] Y.Q. Qu, X.F. Duan, Progress, challenge and perspective of heterogeneous
than the other samples at the solid-liquid interface. Therefore, photocatalysts, Chem. Soc. Rev. 42 (2013) 2568–2580.
the results confirm that the BiOBr@TiO2 heterostructures, espe- [3] J. Tian, Z.H. Zhao, A. Kumar, R.I. Boughton, H. Liu, Recent progress in design,
cially BiOBr@TiO2 -1.0, can effectively inhibit the recombination of synthesis, and applications of one-dimensional TiO2 nanostructured surface
heterostructures: a review, Chem. Soc. Rev. 43 (2014) 6920–6937.
photo-induced electron–hole pairs, which in turn leads to the high
[4] H. Tada, T. Mitsui, T. Kiyonaga, T. Akita, K. Tanaka, All-solid-state Z-scheme in
photocurrent response and excellent activity in photodegradation CdS–Au–TiO2 three-component nanojunction system, Nat. Mater. 5 (2006)
dyes and photocatalytic H2 production. 782–786.
[5] L. Zhao, T. Cui, Y.J. Li, B. Wang, J.H. Han, L. Han, Z.F. Liu, Efficient visible light
photocatalytic activity of p–n junction CuO/TiO2 loaded on natural zeolite,
RSC Adv. 5 (2015) 64495–64502.
[6] M.G. Wang, Y.M. Hu, J. Han, R. Guo, H.X. Xiong, Y.D. Yin, TiO2 /NiO hybrid
4. Conclusions
shells: p–n junction photocatalysts with enhanced activity under visible light,
J. Mater. Chem. A 3 (2015) 20727–20735.
In summary, the BiOBr@TiO2 hybrid photocatalysts have been [7] W.J. Zhou, Z.Y. Yin, Y.P. Du, X. Huang, Z.Y. Zeng, Z.X. Fan, H. Liu, J.Y. Wang, H.
successfully prepared with the growth of BiOBr nanosheets on the Zhang, Synthesis of few-layer MoS2 nanosheet-coated TiO2 nanobelt
heterostructures for enhanced photocatalytic activities, Small 9 (2013)
surface of TiO2 nanobelts. According XRD, Ramam spectra and TEM, 140–147.
the TiO2 nanobelts are of mixed phase of anatase and TiO2 (B). [8] M.L. Mao, L. Mei, D. Guo, L.C. Wu, D. Zhang, Q.H. Li, T.H. Wang, High
Then the optimal BiOBr loading content was determined to be the electrochemical performance based on the TiO2 nanobelt@few-layered MoS2
structure for lithium-ion batteries, Nanoscale 6 (2014) 12350–12353.
ratio of BiOBr to TiO2 1:1. Mott–Schottky plots indicated the for- [9] J. Tian, Y.H. Sang, Z.H. Zhao, W.J. Zhou, D.Z. Wang, X.L. Kang, H. Liu, J.Y. Wang,
mation of p–n junctions at the interface of BiOBr nonasheets and S.W. Chen, H.Q. Cai, H. Huang, Enhanced photocatalytic performances of
TiO2 nanobelts. Compared with the pure BiOBr, TiO2 nanobelts and CeO2 /TiO2 nanobelt heterostructures, Small 9 (2013) 3864–3872.
[10] Y. Hu, D.Z. Li, Y. Zheng, W. Chen, Y.H. He, Y. Shao, X.Z. Fu, G.C. Xiao, BiVO4 /TiO2
other BiOBr@TiO2 heterostructures, BiOBr@TiO2 -1.0 exhibited the nanocrystalline heterostructure: a wide spectrum responsive photocatalyst
best photocatalytic performance for the degradation of RhB and towards the highly efficient decomposition of gaseous benzene, Appl. Catal. B:
photocatalytic H2 production Environ. 104 (2011) 30–36.
Y. Zhao et al. / Applied Surface Science 365 (2016) 209–217 217

[11] H.F. Cheng, B.B. Huang, Y. Dai, Engineering BiOX (X = Cl, Br, I) nanostructures [30] Y.J. Hwang, C. Hahn, B. Liu, P.D. Yang, Photoelectrochemical properties of TiO2
for highly efficient photocatalytic applications, Nanoscale 6 (2014) nanowire arrays: a study of the dependence on length and atomic layer
2009–2026. deposition coating, ACS Nano 6 (2012) 5060–5069.
[12] G.P. Dai, J.G. Yu, G. Liu, Synthesis and enhanced visible-light [31] J.H. Yi, L.L. Huang, H.J. Wang, H. Yu, F. Peng, AgI/TiO2 nanobelts monolithic
photoelectrocatalytic activity of p–n junction BiOI/TiO2 nanotube arrays, J. catalyst with enhanced visible light photocatalytic activity, J. Hazard. Mater.
Phys. Chem. C 115 (2011) 7339–7346. 284 (2015) 207–214.
[13] J.Q. Liu, L.L. Ruan, S.B. Adeloju, Y.C. Wu, BiOI/TiO2 nanotube arrays, a unique [32] D. Sarkar, K.K. Chattopadhyay, Branch density-controlled synthesis of
flake-tube structured p–n junction with remarkable visible-light hierarchical TiO2 nanobelt and tunable three-step electron transfer for
photoelectrocatalytic performance and stability, Dalton Trans. 43 (2014) enhanced photocatalytic property, ACS Appl. Mater. Interfaces 6 (2014)
1706–1715. 10044–10059.
[14] D.F. Hou, X.L. Hu, P. Hu, W. Zhang, M.F. Zhang, Y.H. Huang, Bi4 Ti3 O12 [33] Y.H. Sang, Z.H. Zhao, J. Tian, P. Hao, H.D. Jiang, H. Liu, J.P. Claverie, Enhanced
nanofibers-BiOI nanosheets p–n junction: facile synthesis and photocatalytic property of reduced graphene oxide/TiO2 nanobelt surface
enhanced visible-light photocatalytic activity, Nanoscale 5 (2013) heterostructures constructed by an in situ photochemical reduction method,
9764–9772. Small 10 (2014) 3775–3782.
[15] W.Q. Fan, X.Q. Yu, S.Y. Song, H.Y. Bai, C. Zhang, D. Yan, C.B. Liu, Q. Wang, W.D. [34] K. Lee, A. Mazare, P. Schmuki, One-dimensional titanium dioxide
Shi, Fabrication of TiO2 –BiOCl double-layer nanostructure arrays for nanomaterials: nanotubes, Chem. Rev. 114 (2014) 9385–9454.
photoelectrochemical water splitting, CrystEngComm 16 (2014) [35] P. Roy, S. Berger, P. Schmuki, TiO2 nanotubes: synthesis and applications,
820–825. Angew. Chem. Int. Ed. 50 (2011) 2904–2939.
[16] F.F. Duo, Y.W. Wang, C.M. Fan, X.M. Mao, X.C. Zhang, Y.F. Wang, J.X. Liu, Low [36] R. Li, X.Y. Gao, C.M. Fan, X.C. Zhang, Y.W. Wang, Y.F. Wang, A facile approach
temperature one-step synthesis of rutile TiO2 /BiOCl composites with for the tunable fabrication of BiOBr photocatalysts with high activity and
enhanced photocatalytic activity, Mater. Charact. 99 (2015) 8–16. stability, Appl. Surf. Sci. 355 (2015) 1075–1082.
[17] J. Yang, X.X. Wang, X.W. Lv, X.R. Xu, Y.J. Mi, J.L. Zhao, Preparation and [37] J. Dai, J. Yang, X.H. Wang, L. Zhang, Y.J. Li, Enhanced visible-light
photocatalytic activity of BiOX-TiO2 composite films (X = Cl, Br, I), Ceram. Int. photocatalytic activity for selective oxidation of amines into imines over
40 (2014) 8607–8611. TiO2 (B)/anatase mixed-phase nanowires, Appl. Surf. Sci. 349 (2015) 343–352.
[18] X.X. Wei, H.T. Cui, S.Q. Guo, L.F. Zhao, W. Li, Hybrid BiOBr-TiO2 [38] W.J. Zhou, L.G. Gai, P.G. Hu, J.J. Cui, X.Y. Liu, D.Z. Wang, G.H. Li, H.D. Jiang, D.
nanocomposites with high visible light photocatalytic activity for water Liu, H. Liu, J.Y. Wang, Phase transformation of TiO2 nanobelts and
treatment, J. Hazard. Mater. 263 (2013) 650–658. TiO2 (B)/anatase interface heterostructure nanobelts with enhanced
[19] X.J. Wang, W.Y. Yang, F.T. Li, J. Zhao, R.H. Liu, S.J. Liu, B. Li, Construction of photocatalytic activity, CrystEngComm 13 (2011) 6643–6649.
amorphous TiO2 /BiOBr heterojunctions via facets coupling for enhanced [39] Y.N. Huo, J. Zhang, M. Miao, Y. Jin, Solvothermal synthesis of flower-like BiOBr
photocatalytic activity, J. Hazard. Mater. 292 (2015) 126–136. microspheres with highly visible-light photocatalytic performances, Appl.
[20] X. Zhang, L.Z. Zhang, T.F. Xie, D.J. Wang, Low-temperature synthesis and high Catal. B: Environ. 111–112 (2012) 334–341.
visible-light-induced photocatalytic activity of BiOI/TiO2 heterostructures, J. [40] X.J. Shi, X. Chen, X.L. Chen, S.M. Zhou, S.Y. Lou, Y.Q. Wang, L. Yuan, PVP
Phys. Chem. C 113 (2009) 7371–7378. assisted hydrothermal synthesis of BiOBr hierarchical nanostructures and
[21] Y.Y. Li, J.S. Wang, B. Liu, L.Y. Dang, H.C. Yao, Z.J. Li, BiOI-sensitized TiO2 in high photocatalytic capacity, Chem. Eng. J. 222 (2013) 120–127.
phenol degradation: a novel efficient semiconductor sensitizer, Chem. Phys. [41] J.Y. Zhang, F.X. Xiao, G.C. Xiao, B. Liu, Self-assembly of a Ag
Lett. 508 (2011) 102–106. nanoparticle-modified and graphene-wrapped TiO2 nanobelt ternary
[22] Z. Liu, X.X. Xu, J.Z. Fang, X.M. Zhu, J.H. Chu, B.J. Li, Microemulsion synthesis, heterostructure: surface charge tuning toward efficient photocatalysis,
characterization of bismuth oxyiodine/titanium dioxide hybrid nanoparticles Nanoscale 6 (2014) 11293–11302.
with outstanding photocatalytic performance under visible light irradiation, [42] S.Y. Baea, S.J. Kim, S.H. Lee, W.Y. Choi, Dye decolorization test for the activity
Appl. Surf. Sci. 258 (2012) 3771–3778. assessment of visible light photocatalysts: realities and limitations, Catal.
[23] D.Y. Wu, H.Y. Wang, C.L. Li, J. Xia, X.J. Song, W.S. Huang, Photocatalytic Today 224 (2014) 21–28.
self-cleaning properties of cotton fabrics functionalized with p-BiOI/n-TiO2 [43] T.B. Li, G. Chen, C. Zhou, Z.Y. Shen, R.C. Jin, J.X. Sun, New photocatalyst
heterojunction, Surf. Coat. Technol. 258 (2014) 672–676. BiOCl/BiOI composites with highly enhanced visible light photocatalytic
[24] C.X. Liao, Z.J. Ma, G.P. Dong, J.R. Qiu, BiOI nanosheets decorated TiO2 performances, Dalton Trans. 40 (2011) 6751–6758.
nanofiber: tailoring water purification performance of photocatalyst in [44] J.T. Li, F.K. Meng, S. Suri, W.Q. Ding, F.Q. Huang, N.Q. Wu,
structural and photo-responsivity aspects, Appl. Surf. Sci. 314 (2014) Photoelectrochemical performance enhanced by a nickel oxide-hematite p–n
481–489. junction photoanode, Chem. Commun. 48 (2012) 8213–8215.
[25] L.Y. Wang, W.A. Daoud, BiOI/TiO2 -nanorod array heterojunction solar cell: [45] F.K. Meng, J.T. Li, S.K. Cushing, M.J. Zhi, N.Q. Wu, Solar hydrogen generation by
growth, charge transport kinetics and photoelectrochemical properties, Appl. nanoscale p–n junction of p-type molybdenum disulfide/n-type
Surf. Sci. 324 (2015) 532–537. nitrogen-doped reduced grapheme oxide, J. Am. Chem. Soc. 135 (2013)
[26] X.D. Wang, Z.D. Li, J. Shi, Y.H. Yu, One-dimensional titanium dioxide 10286–10289.
nanomaterials: nanowires, nanorods, and nanobelts, Chem. Rev. 114 (2014) [46] J. Zhang, S.Z. Qiao, L.F. Qi, J.G. Yu, Fabrication of NiS modified CdS nanorod p–n
9346–9384. junction photocatalysts with enhanced visible-light photocatalytic
[27] Z.H. Zhao, J. Tian, Y.H. Sang, A. Cabot, H. Liu, Structure, synthesis, and H2 -production activity, Phys. Chem. Chem. Phys. 15 (2013) 12088–12094.
applications of TiO2 nanobelts, Adv Mater. 27 (2015) 2557–2582. [47] Y.B. Chen, Z.X. Qin, X.X. Wang, X. Guo, L.J. Guo, Noble-metal-free
[28] I.S. Cho, Z.B. Chen, A.J. Forman, D.R. Kim, P.M. Rao, T.F. Jaramillo, X.L. Zheng, Cu2 S-modified photocatalysts for enhanced photocatalytic hydrogen
Branched TiO2 nanorods for photoelectrochemical hydrogen production, production by forming nanoscale p–n junction structure, RSC Adv. 5 (2015)
Nano Lett. 11 (2011) 4978–4984. 18159–18166.
[29] J.Z. Chen, H.B. Yang, J.W. Miao, H.-Y. Wang, B. Liu, Thermodynamically driven [48] W.Z. Wang, J. Wang, Z.Z. Wang, X.Z. Wei, L. Liu, Q.S. Ren, W.L. Gao, Y.J. Liang,
one-dimensional evolution of anatase TiO2 nanorods: one-step hydrothermal H.L. Shi, p–n junction CuO/BiVO4 heterogeneous nanostructures: synthesis
synthesis for emerging intrinsic superiority of dimensionality, J. Am. Chem. and highly efficient visible-light photocatalytic performance, Dalton Trans. 43
Soc. 136 (2014) 15310–15318. (2014) 6735–6743.

You might also like