You are on page 1of 5

Applied Surface Science 325 (2015) 112–116

Contents lists available at ScienceDirect

Applied Surface Science


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

Effect of oxygen vacancy on enhanced photocatalytic activity of


reduced ZnO nanorod arrays
Chao Wang, Di Wu, Peifang Wang, Yanhui Ao ∗ , Jun Hou, Jin Qian
Key laboratory of Integrated Regulation and Resource Development on Shallow Lakes, Ministry of Education, College of Environment, Hohai University,
Nanjing 210098, China

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

Article history: Partially reduced ZnO nanorod arrays (NRAs) were synthesized in a reducing reagent of NaBH4 . The
Received 24 July 2014 NaBH4 treatment substantially increased the density of oxygen vacancies. The obtained materials were
Received in revised form 25 October 2014 characterized by a field emission scanning electron microscopy, X-ray diffractometer, X-ray photoelec-
Accepted 2 November 2014
tron spectroscopy and UV–vis absorption spectra. The photocatalytic activity of the resulting samples
Available online 7 November 2014
was evaluated by degradation of methylene blue (MB) under UV light irradiation. The ZnO NRAs with a
partially reduced surface exhibited enhanced photocatalytic activity. The improvement was ascribed to
Keywords:
the introducing oxygen vacancies in ZnO by NaBH4 treatment. The increased surface oxygen vacancy
ZnO nanorod arrays
NaBH4
decreased the surface recombination centers, and improved the charge separation efficiency, thus
Oxygen vacancies enhances the photocatalytic activity. Furthermore, we found that moderate treatment enhanced pho-
Photocatalysis tocatalytic activity. However, with longer time of treatment, bulk oxygen vacancies appeared, which
Visible light would cause a reduced activity.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction in a ZnO crystal leads to electrical and optical changes [21]. Accord-
ing to the published work, oxygen vacancies not only enhanced UV
Zinc oxide (ZnO) has been recognized as a promising semicon- light photocatalytic activity of metal oxide, but also induced visible
ductor material for photocatalytic applications owing to its direct light-active activity [22–25]. This can be ascribed to oxygen vacancy
wide bandgap (Eg ∼3.37 eV), and large exciton binding energy defects induce new donor energy levels below its conduction band
(∼60 meV). In particular, 1D ZnO nanostructures such as nanowire, [26]. Recently, researchers have taken their attentions to the cre-
nanorod, nanotube are considered to be one of the most promising ation of oxygen vacancy to improve photocatalytic activity of ZnO
nanomaterials for their strong application potentials as compo- nanomaterials using different techniques [27–31]. However, these
nents of nano-scale electronic or optoelectronic [1,2]. Graugnard strategies need harsh synthetic conditions or expensive facilities
et al. studied the properties and applications of 1D ZnO nanostruc- which limit their practical application.
ture arrays [3–5]. However, problems such as increased carrier Hence, developing a simple and economical strategy to syn-
recombination centers and decreased incident photon to electron thesize a stable reduced ZnO photocatalyst with suitable localized
conversion efficiency (IPCE), which significantly limits its photo- states is still a great challenge. To resolve these issues, we have
electrochemical applications [6]. It is very important to improve the used a one-step method that can substantially increase the density
morphology and electronic structure of ZnO for effective separation of oxygen vacancies by reducing ZnO nanostructures in a reduc-
and transportation of photoexcited charge carriers [7,8]. ing reagent, low-cost NaBH4 is used as a reducer. The increased
Various techniques have been employed to solve the problems. surface oxygen vacancy decreases the surface recombination cen-
These techniques include non-metals doping [9,10], metals doping ters, enhances the electrical conductivity, and improves the charge
[11–13], surface modification via organic materials [14] and semi- transportation.
conductor coupling [15]. Apart from this, the defects of metal oxide
nanostructures are attracting an increasing interest in recent sev-
2. Experimental
eral years [16–20]. It has been reported that the existence of defects

2.1. Materials

∗ Corresponding author. Tel.: +86 25 83787330; fax: +86 25 83787330. All chemicals were of analytic reagent grade and used without
E-mail address: andyao@hhu.edu.cn (Y. Ao). further purification. All the aqueous solutions were prepared using

http://dx.doi.org/10.1016/j.apsusc.2014.11.003
0169-4332/© 2014 Elsevier B.V. All rights reserved.
C. Wang et al. / Applied Surface Science 325 (2015) 112–116 113

Fig. 1. SEM images of the pristine ZnO NARs. (A) Top view and (B) cross-section view.

deionized water. Ti plates (4 cm × 3 cm × 0.1 cm) were cleaned 2.4. Photocatalytic activity measurement
ultrasonically in acetone, ethanol, and deionized water for 30 min,
and dried at 60 ◦ C. Pickling was performed in a 1:3:6 (involume) The photocatalytic activity of the resulting samples was eval-
mixture of aqueous solution of HF (55 wt%), HNO3 (63 wt%) and uated by degradation of methylene blue (MB) under UV light
distilled water for 45 s. irradiation. The experiments were carried out in the closed com-
partment to avoid interference from ambient light. A ZnO film
(effective area was 3 cm × 3 cm) was inserted into 100 ml MB
2.2. Preparation of ZnO nanorod array films on Ti foil solution whose initial concentration was 10 mg/L. The solution
was stirred for 30 min to allow adsorption–desorption equilibrium
ZnO NRAs were synthesized by hydrothermal growth on Ti foil before light irradiation. After that, the ZnO film was illuminated
(4 cm × 3 cm) coated with a layer of ZnO nanoparticles. The pro- with an ultraviolet lamp (Model: MUA-165, with a peak wavelength
cedure consists of two steps: (1) modification of Ti foil with a of 365 nm) passing through the quartz window. Then, 1 mL solution
thin layer of densely and uniformly dispersed ZnO nanoparticles was collected and analyzed by a UV–vis spectrophotometer (Shi-
by spin coating. The seed layer was prepared using a mixed solu- madzu, UV-3600) at a wavelength of 665 nm at regular interval
tion of 0.5 M zinc acetate dehydrate (Zn(CH3 COO)2 ·2H2 O) and 0.6 M (15 min). Photocurrent responses were recorded using an electro-
diethanolamine (C4 H11 NO2 ) in ethanol. The resulting mixture was chemical station (Chenhua Instruments, CHI660D) operated in a
then agitated at 60 ◦ C for 30 min to yield a homogeneous and sta- three-electrode configuration with the ZnO films as the working
ble colloid solution, which served as coating solution. After 2 min electrode, a Pt foil as the counter electrode and a Ag/AgCl electrode
spin coating, the substrates were dried at room temperature for as the reference electrode. An ultraviolet lamp (Model: MUA-165,
120 min and then annealed at 400 ◦ C for 60 min. (2) Hydrothermal with a peak wavelength of 365 nm) was used as the light resource.
growth of ZnO nanorods in aqueous solution. The aqueous solu- 50 mL of Na2 SO4 (0.1 M) was used as the electrolyte. The applied
tion used for ZnO nanorods growth process was prepared using potential biases were +0.3 V in the irradiation process.
0.05 M zinc nitratehexahydrate (Zn(NO3 )2 ·6H2 O) and 0.05 M hex-
amethylenetetramine (C6 H12 N4 ). Those solutions were transferred
into the Teflon-lined stainless steel autoclaves with a volume of 3. Results and discussion
70 mL. The Ti plates coated with ZnO seed layer were vertically
immersed into the reaction solution. Subsequently, the autoclaves 3.1. Characterization of the samples
were sealed and heated to a constant temperature of 95 ◦ C for 2 h.
After taking out, the samples were cleaned ultrasonically in ethanol Fig. 1 shows the morphology of ZnO NRAs grown on Ti sub-
and distilled water for 30 min, then dried at 90 ◦ C. Finally the sam- strates. The top-view of the SEM image (Fig. 1A) illustrates that
ples were dipped in 0.1 M NaBH4 for 15 min, 30 min and 45 min well aligned ZnO nanorods grow uniformly in large scale. The
respectively at room temperature followed by rinsing with water cross-section image of as-prepared NARs shown in Fig. 1B indi-
and air drying. Samples N-15, N-30, N-45 were respectively gained. cates that ZnO nanorods grown on Ti substrates have an average
length of ∼800 nm and diameter of approximately 100 nm. Over-
all, the images show that ZnO NRAs of sharp tips are vertically
2.3. Characterization oriented from the Ti foil substrate. Due to the one-dimensional fea-
tures of the nanorods, directional charge transport is promoted [4].
The surface morphologies were characterized with a field emis- Fig. 2 shows that the reduced ZnO NARs after NaBH4 treatment
sion scanning electron microscopy (SEM, Hitachi, S-4800). The have kept their pristine structure (including orientations, lengths,
crystalline structures of as-prepared samples were deter-mined by diameters, etc.), indicating the NaBH4 treatment has not destroyed
an X-ray diffractometer (XRD, Shimadazu, XD-3A) over the 2Â range the structure of ZnO nanorods.
30–75◦ . The photoluminescence (PL) spectra in the wavelength Fig. 3 shows the XRD patterns of the as-prepared samples. The
range 350–600 nm has been recorded using 325 nm line of Xe samples gave similar XRD patterns indicating the nanorods in high
laser using a fluorescence spectrophotometer (F-7000). The UV–vis crystallinity. All the peaks can be indexed to the standard diffraction
absorption spectra of the samples were measured by a UV–vis pattern of hexagonal phase ZnO (JCPDS 36-1451). The much higher
spectrophotometer (Shimadzu, UV-3600). The surface chemical intensity of the 002 diffraction peaks for the samples indicates that
bonding of the samples was examined by X-ray photoelectron spec- excellent orientation in the c-axis direction ZnO NARs grown on
troscopy (XPS) with Mg-K␣ radiation (ESCALB MK-II). Ti foils [32,33]. Furthermore, the results indicate that the NaBH4
114 C. Wang et al. / Applied Surface Science 325 (2015) 112–116

Fig. 2. SEM images of the ZnO NARs after NaBH4 treatment.

002 In our case, NaBH4 treatment is an effective method to introduce


Pristine ZnO NRAs surface defect (oxygen vacancy) into ZnO NARs [34,35].
Reduced ZnO NRAs Fig. 5 shows the PL spectra of the ZnO NARs samples. Clearly,
XRD Intensity (a.u.)

in comparison to the pristine ZnO NARs, the reduced ZnO NARs


101
show two new bands at 449.5 nm, 474.8 nm. The two new bands
100 103 are ascribed to the oxygen vacancies forming at the surface of ZnO
102 110
NARs [36,37]. The proposed new surface oxygen vacancies may act
as the radiate recombination sites and increase the opportunity of
electron transition from the energy level of oxygen vacancies to the
valence band [38].

3.2. Photocatalytic properties

30 35 40 45 50 55 60 65 70 75 Fig. 6(a) shows the photodegradation plot of MB under UV light


irradiation by different samples. The photocatalytic degradation of
2 theta (degree)
MB is used as a model reaction to compare the photocatalytic activ-
ity of the as-prepared ZnO NRAs. The initial concentration of MB is
Fig. 3. XRD patterns of the ZnO NARs before and after NaBH4 treatment.
10 mg/L and the adsorption time is 30 min. With the increase of
degradation time, the residual MB concentrations decrease gradu-
ally. It can be seen that the ZnO NARs being treated for 30 min by
treatment has no significant effect on the ZnO bulk nanocrystal
NaBH4 shows the highest photocatalytic activity. The degradation
structure.
efficiency for pristine ZnO NRAs, N-15, N-45 are little difference.
Fig. 4 shows oxygen 1s XPS profiles of the ZnO NARs before and
After 15 min UV irradiation, the degradation efficiency for pristine
after NaBH4 treatment, which slightly shifted to lower energy after
ZnO, N-15, N-45 samples are 29%, 28%, 27%, however, N-30 shows
reduction process. It shows the change of surface chemical bonding
enhanced degradation efficiency about 55%. After 60 min UV, the
of ZnO nanocrystals induced by NaBH4 treatment. The shift is in line
degradation percent values of MB respectively are about 65%, 70%,
with the transfer of electrons to the neighboring oxygen vacancies.

Fig. 4. O1s XPS spectra of the pristine and reduced ZnO NARs. Fig. 5. Photoluminescence (PL) spectra of pristine and reduced ZnO NARs.
C. Wang et al. / Applied Surface Science 325 (2015) 112–116 115

Fig. 7. The mechanism of photocatalytic degradation of MB by reduced ZnO under


UV light irradiation.

can greatly enhance the photocatalytic properties [39–41]. Fig. 7


shows the degradation of MB by the reduced ZnO nanostructure
under UV light. Band gap excitation of the semiconductor results in
electron–hole separation. Highly reactive hydroxyl radicals can also
be formed either by the decomposition of water or by the reaction
of a hole with OH− . The hydroxyl radical and photogenerated holes
are extremely strong, non-selective oxidants that lead to the degra-
dation of MB at the surface of reduced ZnO. This can be attributed
to the high concentrations of oxygen vacancies created in reduced
ZnO.
Fig. 8 shows the transient photocurrent responses of the samples
(pristine ZnO NRAs, N-15, N-30 and N-45). The applied potential
biases were +0.3 V in the irradiation process. The rise and fall of
the photocurrents correspond well to the UV illumination being
switched on and off for the four samples. The response of N-30
was the highest among the samples. It shows that photocurrents
of all reduced ZnO NRAs are higher than that of pristine ZnO NRAs
and increase gradually with the increasing of the NaBH4 treatment
time from 15 to 30 min. The photocurrent of N-30 reaches a max-
imum. As the NaBH4 treatment time is further increased above 30
min, the photocurrent of the reduced ZnO NRAs decrease. The N-
45 shows an obvious decay. Kong and Li [42] reported that surface
oxygen vacancies serve as charge carrier traps as well as adsorption
sites where the charge transfer to adsorbed species, which can pre-
vent e− –h+ recombination. However, bulk oxygen vacancies tend to
act as charge carrier traps where e− –h+ recombination occurs. The
NaBH4 treatment introduces oxygen vacancies in ZnO, the number
Fig. 6. (a) Photodegradation of MB under UV light by different samples and (b) of oxygen vacancies is proportional to treatment time, but when
linear transform ln(C0 /C) = f(t) of the kinetic curves of MB disappearance for different
samples.

91% and 60% in the presence of pristine ZnO NRAs, N-15, N-30 and
N-45. The apparent first order kinetic equation ln(C0 /C) = kapp t is
used to fit experimental data in Fig. 6(a), where kapp is apparent
rate constant. The results showed that the results are fitted well
with the apparent first order because the relative constant are all
higher than 0.95. The kapp obtained from Fig. 6(b) are 0.017, 0.020,
0.040 and 0.015 min−1 in the presence of pristine ZnO NRAs, N-15,
N-30 and N-45, respectively. N-30 showed the best photocatalytic
activity. The apparent rate constant of N-30 was about 2.4 times as
that of pristing ZnO NRAs.
ZnO photocatalysis is based on the generation of an
electron–hole pair upon light irradiation. However, the photo-
induced electron–hole easily recombined in the photocatalysis
process. Therefore, it is important to reduce the surface recombi-
nation of electrons and holes. Oxygen vacancies can act as electron
acceptors during the photocatalytic reaction and thus reduce
the recombination rate. Oxygen vacancies can also facilitate the
adsorbed O2 to capture photo-induced electrons, synchronously Fig. 8. Photocurrent responses of different samples at +0.3 V under UV light irradi-
generating superoxide radical anions. Therefore, oxygen vacancies ation.
116 C. Wang et al. / Applied Surface Science 325 (2015) 112–116

fund of NSFC (No. 51422902), the National Natural Science Founda-


tion of China (No. 51108158), The outstanding youth fund of Jiangsu
Province (BK2012037), Research Fund for innovation team of Min-
istry of education (IRT13061), Fundamental Research Funds for the
Central Universities (2013B32114, 2013B14114).

References

[1] Z.L. Wang, J. Phys.: Condens. Matter 16 (2004) R829.


[2] W. Lee, M.-C. Jeong, J.-M. Myoung, Nanotechnology 15 (2004) 254.
[3] X. Wang, C.J. Summers, Z.L. Wang, Nano Lett. 4 (2004) 423–426.
[4] A. Du Pasquier, H. Chen, Y. Lu, Appl. Phys. Lett. 89 (2006) 253513.
[5] X.D. Wang, E. Graugnard, J.S. King, Z.L. Wang, C.J. Summers, Nano Lett. 4 (2004)
2223–2226.
[6] S. Rehman, R. Ullah, A. Butt, N. Gohar, J. Hazard. Mater. 170 (2009) 560–569.
[7] J. Zhang, D. Gao, G. Yang, Z. Zhu, J. Zhang, Z. Shi, Int. J. Mater. Mech. Eng. 1 (2012).
[8] L. Nehru, M. Umadevi, C. Sanjeeviraja, Int. J. Mater. Eng. 2 (2012) 12–17.
[9] D. Li, H. Haneda, J. Photochem. Photobiol. A: Chem. 155 (2003) 171–178.
[10] L. Di, H. Haneda, J. Photochem. Photobiol. A Chem. 160 (2003) 203–212.
[11] S. Deka, P. Joy, Solid State Commun. 134 (2005) 665–669.
[12] S. Colis, H. Bieber, S. Bégin-Colin, G. Schmerber, C. Leuvrey, A. Dinia, Chem. Phys.
Lett. 422 (2006) 529–533.
[13] V. Shinde, T. Gujar, C. Lokhande, R. Mane, S.-H. Han, Mater. Chem. Phys. 96
Fig. 9. Diffuse reflectance UV–vis spectra for different samples. (2006) 326–330.
[14] R. Qiu, D. Zhang, Y. Mo, L. Song, E. Brewer, X. Huang, Y. Xiong, J. Hazard. Mater.
156 (2008) 80–85.
the treatment is longer, bulk oxygen vacancies will be obtained. [15] G.S. Pozan, A. Kambur, Chemosphere 105 (2014) 152–159.
That is the reason why N-30 showed better photocurrent activity [16] M. MansoobáKhan, S. AliáAnsari, M. HwanáCho, Nanoscale 5 (2013)
than N-45. 6323–6326.
[17] M. MansoobáKhan, M. OmaisháAnsari, D. HungáHan, M. HwanáCho, J. Mater.
Fig. 9 shows the UV–vis absorption spectra of the pristine Chem. A 2 (2014) 637–644.
and reduced ZnO NRAs. It should be noted that NaBH4 reduction [18] S.A. Ansari, M.M. Khan, M.O. Ansari, S. Kalathil, J. Lee, M.H. Cho, RSC Adv. 4
strikingly improves overall UV-light absorption. It shows that the (2014) 16782–16791.
[19] S.A. Ansari, M.M. Khan, M.O. Ansari, J. Lee, M.H. Cho, New J. Chem. 38 (2014)
UV absorption is positively related to oxygen vacancies in ZnO 2462–2469.
before 30 min. The absorption of N-30 reaches a maximum. The [20] M.M. Khan, S.A. Ansari, D. Pradhan, D.-H. Han, J. Lee, M.H. Cho, Ind. Eng. Chem.
N-45 shows an obvious decay again. It is clear that extended treat- Res. 53 (2014) 9754–9763.
[21] Ü. Özgür, Y.I. Alivov, C. Liu, A. Teke, M. Reshchikov, S. Doğan, V. Avrutin, S.-J.
ment will cause a high concentration of bulk oxygen vacancies. Cho, H. Morkoc, J. Appl. Phys. 98 (2005) 041301.
This is why that 30 min sample shows the highest absorption [22] Y. Lai, M. Meng, Y. Yu, X. Wang, T. Ding, Appl. Catal. B: Environ. 105 (2011)
performance. 335–345.
[23] Y. Lv, C. Pan, X. Ma, R. Zong, X. Bai, Y. Zhu, Appl. Catal. B: Environ. 138 (2013)
All in all, the enhanced photocatalytic activity can be ascribed 26–32.
to the following factor: The NaBH4 treatment introduces oxygen [24] Y. Lv, W. Yao, X. Ma, C. Pan, R. Zong, Y. Zhu, Catal. Sci. Technol. 3 (2013)
vacancies in ZnO nanocrystals, which induce effective separation 3136–3146.
[25] X. Bai, L. Wang, R. Zong, Y. Lv, Y. Sun, Y. Zhu, Langmuir 29 (2013) 3097–3105.
and transportation of photoexcited charge carriers, thus causes an
[26] L. Schmidt-Mende, J.L. MacManus-Driscoll, Mater. Today 10 (2007) 40–48.
improved photocatalytic activity. But when the treatment is longer, [27] U.N. Maiti, S. Maiti, K.K. Chattopadhyay, CrystEngComm 14 (2012) 640–647.
bulk oxygen vacancies will appear, which would cause a reduced [28] Y. Liu, Z. Kang, Z. Chen, I. Shafiq, J. Zapien, I. Bello, W. Zhang, S. Lee, Cryst. Growth
photocatalytic activity. Therefore, N-30 (the sample dipped in 0.1 M Des. 9 (2009) 3222–3227.
[29] Y. Zheng, C. Chen, Y. Zhan, X. Lin, Q. Zheng, K. Wei, J. Zhu, Y. Zhu, Inorg. Chem.
NaBH4 for 30 min at room temperature) shows the highest photo- 46 (2007) 6675–6682.
catalytic activity. [30] T.-J. Kuo, C.-N. Lin, C.-L. Kuo, M.H. Huang, Chem. Mater. 19 (2007) 5143–5147.
[31] A.B. Patil, K.R. Patil, S.K. Pardeshi, J. Solid State Chem. 184 (2011) 3273–3279.
[32] M. Guo, P. Diao, S. Cai, J. Solid State Chem. 178 (2005) 1864–1873.
4. Conclusions [33] M. Guo, P. Diao, X. Wang, S. Cai, J. Solid State Chem. 178 (2005) 3210–3215.
[34] J. Liqiang, Q. Yichun, W. Baiqi, L. Shudan, J. Baojiang, Y. Libin, F. Wei, F. Honggang,
We have synthesized reduced ZnO NRAs by NaBH4 treatment. S. Jiazhong, Solar Energy Mater. Solar Cells 90 (2006) 1773–1787.
[35] M. Samadi, H.A. Shivaee, A. Pourjavadi, A.Z. Moshfegh, Appl. Catal. A: Gen. 466
The ZnO NRAs were prepared by a hydrothermal method on Ti foil. (2013) 153–160.
The ZnO NRAs with a partially reduced surface exhibited enhanced [36] K. Vanheusden, W. Warren, C. Seager, D. Tallant, J. Voigt, B. Gnade, J. Appl. Phys.
UV light absorption and photocurrent response. NaBH4 treatment 79 (1996) 7983–7990.
[37] T. Tachikawa, S. Yamashita, T. Majima, Angew. Chem. Int. Ed. 49 (2010)
increased surface oxygen vacancies that account for decreased 432–435.
surface recombination centers and favorable charge transfer rate. [38] M. Raja, N. Muthukumarasamy, D. Velauthapillai, R. Balasundaraprabhu, Super-
Photocatalytic activity of reduced ZnO NRAs depends on the NaBH4 lattices Microstruct. 72 (2014) 102–110.
[39] S.A. Ansari, M.M. Khan, S. Kalathil, A.N. Khan, J. Lee, M.H. Cho, Nanoscale
treatment time. (2013).
[40] S.A. Ansari, M.M. Khan, J. Lee, M.H. Cho, J. Ind. Eng. Chem. 20 (2014)
Acknowledgements 1602–1607.
[41] M.M. Khan, S.A. Ansari, M.O. Ansari, B.-K. Min, J. Lee, M.H. Cho, J. Phys. Chem. C
(2014).
We are grateful for grants from the National Science Fund for [42] M. Kong, Y. Li, X. Chen, T. Tian, P. Fang, F. Zheng, X. Zhao, J. Am. Chem. Soc. 133
Distinguished Young Scholars (No. 51225901), outstanding youth (2011) 16414–16417.

You might also like