You are on page 1of 18

Accepted Manuscript

Title: Simple synthesis of ZnO nanoflowers and its


photocatalytic performances toward the photodegradation of
metamitron

Author: Yan Xu Jingjie Jin Xianliang Li Yide Han Hao Meng


Tianyu Wang Xia Zhang

PII: S0025-5408(15)30236-1
DOI: http://dx.doi.org/doi:10.1016/j.materresbull.2015.11.062
Reference: MRB 8573

To appear in: MRB

Received date: 24-7-2015


Revised date: 20-10-2015
Accepted date: 23-11-2015

Please cite this article as: Yan Xu, Jingjie Jin, Xianliang Li, Yide Han, Hao Meng,
Tianyu Wang, Xia Zhang, Simple synthesis of ZnO nanoflowers and its photocatalytic
performances toward the photodegradation of metamitron, Materials Research Bulletin
http://dx.doi.org/10.1016/j.materresbull.2015.11.062

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Simple synthesis of ZnO nanoflowers and its photocatalytic
performances toward the photodegradation of metamitron

Yan Xu,a,* Jingjie Jin,a Xianliang Li,b,* Yide Han,a Hao Meng,a Tianyu Wang,a Xia Zhang a,*

a
Department of Chemistry, College of Science, Northeastern University, Shenyang, Liaoning 110819,
China.
b
College of Materials Science and Engineering, Shenyang University of Chemical Technology,
Shenyang, Liaoning 110142, China.

*Corresponding Author:

xuyanjlu@126.com (Yan Xu);

lixianliang007@163.com (Xianliang Li);

xzhang@mail.neu.edu.cn (Xia Zhang);

Fax: +86-024-83684533;

Tel.: +86-024-83684533.

1
Graphical abstract

Highlights

 Fabricationof ZnO nanoflowers assembled from thin and uniform nanosheets.

 The material possesses photocatalytic acitivity toward degradation of metamitron.

 The catalyst features excellent cycling stability for at least 5 cycle times.

 The promising mechanism of photocatalysis of metamitron is also discussed.

2
Abstract

Large-scale ZnO nanoflowers assembled from numerous thin and uniform nanosheets with a
thickness of around 20 nm, were successfully prepared through a facile one-step hydrothermal
synthesis route by using zinc acetate, sodium citrate and sodium hydroxide in water solution. The
method was simple, green and effective. The obtained ZnO nanoflowers exhibited remarkable
photocatalytic acitivity and good cycle stability for the degradation of metamitron under a 300W of
Osram® ultra-vitalux lamp light emitting UV and visible radiation over 300-600 nm. UV-visible
spectrophotometery was used to measure the rate of photodecomposition of metamitron. The results
indicate that about 97% of the metamitron disappeared in the suspension of flower-like ZnO
microspheres within four hours, and the degradation efficiency were not changed even after 5 cycle
times.

KEYWORDS: A. oxides; A. semiconductors; B. microstructure; B. chemical synthesis; D. catalytic


properties

3
1. Introduction
Due to the increasing contamination of organic pollutants to the environment, extensive research
efforts have been concentrated on the photocatalytic technology for their treatment [1-2]. Recently,
some new class of materials with enhanced photocatalytic activity have been designed and efficiently
synthesized through simple solvothermal method [3-8]. For example, Ti-doped BrOBr with Ag
decoration microsphere, and Fe-doped BrOBr hollow microsphere materials featured excellent
photocatalytic activity towards the degradation of Rhodamine B [3-4]. The Zn2GeO4/CFs (carbon
fibers) composites exhibited good photocatalytic activity for the oxidation of p-toluidine [6].
The agrochemical of metribuzin and metamitron, derived from the structure of
4-amino-1,2,4-triazin-5(4H)-one, have been widely used as herbicides. They are chemically stable and
can slowly penetrate through the soil and cause the severe contamination of underground resources of
drinking water. The traditional semiconductor catalysts such as TiO2, and ZnO have been frequently
used for the degradation of organic pollutants owing to their advantages of low cost, high
photocatalytic activity and photostability. Zinc oxide (ZnO), a widely known important n-type
semiconductor with a wide-band-gap of 3.37 eV, has been receiving broad attention due to its
distinguished performance in optoelectronics, catalysis, chemical sensors and transducers [9-11]. In the
past years, various approaches like sonochemical route [12], thermal evaporation [13], chemical
precipitation method [14], microwave hydrothermal [15], employ ionic liquids, and hydrothermal
method [16] have been developed to synthesize ZnO nanostructures with various morphologies,
including nanoflowers [17], nanowires [18], nanobelts [19], nanorods [20], and nanotubes [21] and so
on. Among them, three-dimensional (3D) nanostructures are of great interest to scientists owing to
their novel architectural nanostructures may provide good performance in photocatalysis and
photoluminescence. Especially, hierarchical 3D ZnO nanostructures with excellent photocatalytic
activity, and diverse properties for more potential applications are strongly expected to be prepared
through an economical method.
In our work, a hierarchical ZnO nanoflowers assembled from numerous nanosheets has been
successfully synthesized through a facile, eco-friendly and one-step hydrothermal synthesis method
using zinc acetate, sodium citrate, and sodium hydroxide in the water solution. The products possess
remarkable photocatalytic acitivity and excellent cycle stability for the degradation of metamitron in
aqueous solution under a 300 W of Osram® lamp light emitting UV and visible radiation over 300-600

4
nm. The promising mechanism of photocatalysis of metamitron is also provided.

2. Experimental

2.1. Materials

The analytical reagents include zinc acetate (C4H6O4Zn·2H2O, 99.9 %, Guangfu, Tianjin, China),
sodium citrate (C6H5O7Na3·H2O, 99.0 %, Yongda, Tianjin, China), sodium hydroxide (NaOH, 99.5 %,
Guoyao, Shanghai, China, http://www.sinoreagent.com), metamitron (C10H10N4O, 97%, Aldrich,
Shanghai, China) and pure water (commercial “Wahaha” purified drinking water, China). All of the
chemical reagents are used without further purification.

2.2. Synthesis

The ZnO nanoflowers were prepared by a one-step hydrothermal method, which is simple, green
and effective. Typically, 0.549 g of C4H6O4Zn·2H2O (2.5 mmol) and 0.735 g of sodium citrate (2.5
mmol) were dissolved into 40 mL deionized water under stirring until a clear aqueous solution was
obtained. Then, 4 mL of NaOH with concentration of 4.0 M was added into the above prepared
aqueous solution with rigorous stirring. The mixture was transferred into a 50 mL of Teflon liner
stainless steel autoclave and heated at 120 ºC for 8 h. After cooling to room temperature, the white
precipitation was separated from the solution by centrifugation, washed with distilled water and
absolute ethanol several times, and dried at 60 ºC for 2 h under an air atmosphere.

2.3. Photodegradation of metamitron


A 300 W of Osram® lamp emitting UV and visible radiation over 300-600 nm was used as the
light resource in the photodegradation of metamitron. In a typical test, 100 mg of as-prepared
photocatalyst ZnO powders was placed in 100 mL of a 13 mg L-1 metamitron solution. The reaction
mixture was stirred for 30 min under darkness to achieve the adsorption equilibrium for metamitron.
Then the solution was irradiated in a photochemical reaction chamber with continuous stirring. At a
certain time of 30 min interval, 3 mL of the treated solution was taken out and centrifuged at 9800 rpm
for 3 min. The relative concentration of metamitron in the treated solution was analyzed by UV-Vis
spectroscopy.
To explore the cyclic stability of catalysts, a reutilization experiment was performed to evaluate

5
the stability of photocatalyst. All precipitates above were retrived, washed with ethanol and acetone
several times, dried in vacuum, and allowed for continuous photodegradation experiments.

2.4. Characterization

The phase of the samples was characterized by powder X-ray diffraction (XRD) on a X’Pert Pro
MRDDY2094 diffractometer with Cu-Kα radiation (λ = 1.5418 Å). A scan rate of 0.0167º s-1 was
applied to record the pattern in the 2θ range of 5–70º. The morphology of the samples was observed
using Ultra Plus field-emission scanning electron microscope (SEM). N2 adsorption-desorption
isotherm was conducted on a Micromeritics ASAP-2020M volumetric gas sorption apparatus using
99.999% pure N2. The UV-Visible adsorption spectra were recorded using a Hitachi U-3010
UV-Visible spectrometer.

3. Results and discussion

3.1. Structure and morphology

The phase of the products was characterized by XRD. As shown in Fig. 1, in the XRD pattern of
as-prepared hierarchical ZnO nanoflowers (Fig. 1a), the peaks observed at 2θ = 30.1º, 35.5º, 43.2º,
53.6º, 57.0º and 62.7º correspond well to the stimulated XRD pattern of hexagonal wurtzite crystal
structure of ZnO (JCPDS 36-1451) (Fig. 1b), indicating that no other impurities are found in the
products. The morphologies of the hierarchical ZnO structure obtained was investigated by SEM. Fig.
2a and 2b showed the SEM images and enlarged view of the hierarchical ZnO materials, respectively.
The sphere-like structure had sizes around 3 μm assembled from nanosheets of about 20 nm in
thickness and 1 μm in width. The self-assembled growth process of the flower-like ZnO nanostructures
has been reported before [22]. Quantitative characterization of the nanoporous structures was
conducted with N2 adsorption-desorption measurements (Fig. 2c). The hierarchical ZnO had a BET
surface area of 18.9 m2 g-1.

The optical property of as-prepared ZnO was carried out by measuring the diffuse reflectance
spectroscopy. Fig. 2d shows the diffuse reflectance spectrum as a function of wavelength, which was
taken in the range of 300-750 nm. The low reflectance values in the spectra indicate high adsorption in
the corresponding wavelength region, and it can be observed that the spectrum exhibits an intense

6
absorption below 400 nm to lower wavelength. The optical band gap of the as-prepared ZnO powder
(inset of Fig. 2d) was estimated from diffuse reflectance spectrum by plotting the value of (hvF(R))1/2
as a function of energy [23], where R is the magnitude of reflectance, and Kubelka-Munk function F(R)
is given by the relation F(R) = (1-R)2/2R. The optical band gap of the as-prepared ZnO nanoflowers is
3.20 eV, which is similar to the reported values of ~ 3.37 eV for ZnO [24].

3.2. Photocatalytic performances of samples


To investigate the photocatalytic activity of the prepared hierarchical ZnO for degradatoin of
metamitron, two different conditions were conducted under Osram® lamp light (300-600 nm) in
aqueous solution: self-degradation without the presence of catalyst and photocatalytic degradation in
the presence of catalyst. Fig. 3a shows that the solution of metamitron exhibits maximum adsorption at
306 nm with an isosbestic point of 285 nm appeared in the UV spectra. The results of self-degradation
and photocatalytic degradation of metamitron are shown in Fig. 3b. The self-degradation efficiency
after 240 min of the irradiation is about 30 %. However, when ZnO was added into the metamitron
solution and irradiated under Osram® lamp light (300-600 nm), a significant and rapid degradation of
metamitron was observed. The maximum degradation efficiency of metamitron on hierarchical ZnO is
about 97%. The degradation efficiency of metamitron was calculated according to the following
equations:

Co-C
degradation efficiency (%)  100 ,
Co

where Co and C are the initial and final concentrations of metamitron vs. time, respectively.

3.3. Mechanism of photocatalytic acitivity of sample

Fig. 4 shows the promising mechanism of heterogeneous photocatalysis of metamitron in the


presence of ZnO microspheres. Upon irradiation, valence band electrons are promoted to the
conduction band leaving a hole behind. Electron in the conduction band on the catalyst surface can
reduce oxygen molecular to superoxide anion. The holes at the ZnO valence band can oxide adsorbed
water to produce hydroxyl radicals. The superoxide anion and hydroxyl radicals then attack organic
metamitron compound to produce desamino metamitron (R-COO-) and (R’) products as has been
described before [25]. The mechanism of the heterogeneous photocatalysis in the presence of ZnO can
7
be described by the following (1)–(5) reactions:

ZnO + hv (UV) eCB- +hVB+ (eCB- +hVB+ heat) (1)

eCB- +O2 O2 - (2)

hVB+ + H2O H+ + OH (3)

(O2 - + H+ + OH H2O + O2) (4)

O2 - + R R-OO- (degraded product) (5)

OH + R R' (degraded product) (6)

To demonstrate the reaction mechanism of the degradation of metamitron over flower-like ZnO,
3.0 mM of isopropyl alcohol (a quencher of OH), 0.05 mM carnosine (a quencher of O2-) and 3.0
mM triethanolamine (a quencher of h+) were added, respectively [5, 26]. As shown in Fig. 5a, it was
found that the degradation of metamitron decreased in the presence of isopropyl alcohol, carnosine and
triethanolamine. The addition of carnosine into the catalytic reaction system can lead to an even more
significant deceleration compared with the addition of isopropyl alcohol and triethanolamine. The
degradation of metamitron can be well described by the pseudo-first order kinetics, ln(Co/C) = kt, with
the squares of linear correlation coefficients R2 larger than 0.98, where t is irradiation time, and k is a
pseudo-first-rate kinetic constant, representing the rate of the degradation reaction. The value of
ln(Co/C) as a function of time are shown in Fig. 5b. It should be mentioned that the k value of the
degradation reaction without isopropyl alcohol, carnosine and triethanolamine is 0.00918, which is
about 1.8 times, 16.3 times and 4.4 times larger than that of catalytic reaction with the addition of
isopropyl alcohol, carnosine, and triethanolamine, respectively. Therefore, it can be concluded that
OH, O2-, and h+ are all active intermediate species for the photodegradation of metamitron, and O2-
may play a more important role in the degradation of metamitron over the flower-like ZnO material
under the UV light.

A reutilization experiment was conducted to evaluate the performance of the photocatalysts due to
the significance of cycling stability of catalyst in its practical application. Fig. 6 showed that the
hierarchical ZnO sample exhibited high cycling stability, and the degradation efficiency were not
changed much even over 5-fold reuse.

8
4. Conclusion

We report a simple and eco-friendly one-step hydrothermal route method for the preparation of
hierarchical ZnO nanostructure, and its remarkable photocatalytic acitivity toward metamitron in
aqueous solution under a 300 W of Osram® lamp light (300-600 nm), as well as its excellent cycle
stability in the reutilization experiment. It revealed that about 97% of the metamitron disappeared in
the suspension of the flower-like ZnO microspheres within four hours, and the degradation efficiency
were nearly unchanged even over 5-fold reuse. The promising mechanism of photocatalysis of
metamitron is also discussed. Comparing with the photogenerated holes and OH species, O2- might
be a more efficient composite for the photodegradation of metamitron.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant No.
21401018) and the Fundamental Research Funds for the Central Universities (No. N130305003).

9
References

[1] A. Kudo, Y. Miseki, Chemical Society Reviews 38 (2009) 253–278.


[2] H. Tong, S.X. Ouyang, Y.P. Bi, N. Umezawa, M. Oshikiri, J.H. Ye, Advanced Materials 24 (2012)
229–251.
[3] G.H. Jiang, R.J. Wang, X.H. Wang, X.G. Xi, R.B. Hu, Y. Zhou, S. Wang, T. Wang, W.X. Chen, ACS
Applied Materials & Interfaces 4 (2012) 4440–4444.
[4] G.H. Jiang, X.H. Wang, Z. Wei, X. Li, X.G. Xi, R.B. Hu, B.L. Tang, R.J. Wang, S. Wang, T. Wang,
W.X. Chen, Journal of Materials Chemistry A 1 (2013) 2406–2410.
[5] G.H. Jiang, W. Zhen, H. Chen, X.X. Du, L. Li, Y.K. Liu, Q. Huang, W.X. Chen, RSC Advances 5
(2015) 30433–30437.
[6] G.H. Jiang, B.L. Tang, H. Chen, Y.K. Liu, L. Li, Q. Huang, W.X. Chen, RSC Advances 5 (2015)
25801–25805.
[7] Y.Q. Qiu, B.Z. Lin, F.C. Jia, Y.L. Chen, B.F. Gao, P.D. Liu, Materials Research Bulletin 72 (2015)
235–240.
[8] Y.K. Cui, F.P. Wang, M. Zubair Iqbal, Z.Y. Wang, Y. Li, J.H. Tu, Materials Research Bulletin 70
(2015) 784–788.
[9] M.H. Huang, S. Mao, H. Feick, H.Q. Yan, Y. Wu, H. Kind, E. Weber, R. Russo, P.D. Yang, Science
292 (2001) 1897–1899.
[10] H. Kind, H. Yan, B. Messer, M. Law, P.D. Yang, Advanced Materials 14 (2002) 158–160.
[11] N. Chantarat, Y.W. Chen, S.Y. Chen, C.C. Lin, Nanotechnology 20 (2009) 395201 (1-5).
[12] S.H. Jung, E. Oh, K.H. Lee, Y. Yang, C.G. Park, W.J. Park, S.H. Jeong, Crystal Growth and
Design 8 (2008) 265–269.
[13] X.X. Yang, W. Lei, X.B. Zhang, B.P. Wang, C. Li, K. Hou, Y.K. Cui, Y.S. Di, Thin Solid Films
517 (2009) 4385–4389.
[14] S. Chakraborty, A.K. Kole, P. Kumbhakar, Materials Letters 67 (2012) 362–364.
[15] X.T. Su, H. Zhao, F. Xiao, J.K. Jian, J.D. Wang, Ceramics International 38 (2012) 1643–1651.
[16] R.X. Shi, P. Yang, X.B. Dong, Q. Ma, A.Y. Zhang, Applied Surface Science 264 (2013) 162–170.
[17] S. Ashoka, G. Nagaraju, C.N. Tharamani, G.T. Chandrappa, Materials Letters 63 (2009) 873–876.
[18] X. Xia, J. Tu, Y. Zhang, X. Wang, C. Gu, X.B. Zhao, H.J. Fan, ACS Nano 6 (2012) 5531–5538.
[19] J.B. Shen, H.Z. Zhuang, D.X. Wang, C.S. Xue, H. Liu, Crystal Growth and Design 9 (2009)
10
2187–2190.
[20] S.Y. Gao, H.J. Zhang, X.M. Wang, R.P. Deng, D.H. Sun, G.L. Zheng Journal of Physical
Chemistry B 110 (2006) 15847–15852.
[21] B. Liu, H.C. Zeng, Nano Research 2 (2009) 201–209.
[22] W. Zhao, X.Y. Song, Z.L. Yin, C.H. Fan, G.Z. Chen, S.X. Sun, Materials Research Bulletin 43
(2008) 3171–3176.
[23] S. Cimitan, S. Albonetti, L. Forni, F. Peri, D. Lazzari, Journal of Colloid and Interface Science
329 (2009) 73–80.
[24] D.G. Thomas, Journal of Physics and Chemistry of Solids 15 (1960) 86–96.
[25] W.U. Palm, M. Millet, C. Zetzsch, Chemosphere 35 (1997) 1117–1130.
[26] G.I. Klebanov, Yu.O. Teselkin, I.V. Babenkova, I.N. Popov, G. Levin, O.V. Tyulina, A.A. Boldyrev,
Yu.A. Vladimirov, Biochemistry and Molecular Biology International 43 (1997) 99–106.

11
Fig. 1 XRD patterns of (a) as-prepared ZnO microspheres and (b) simulated hexagonal wurtzite crystal
structure of ZnO (JCPDS 36-1451).

12
Fig. 2 (a) SEM image and (b) enlarged view of the as-prepared hierarchical ZnO nanoflowers; (c) N2
adsorption and desorption; (d) UV-Vis diffuse reflectance spectrum at room temperature and the line of
(hvF(R))1/2 versus hv (inset) for the as-prepared ZnO nanoflowers.

13
Fig. 3 (a) UV-vis spectra of degradation as a function of irradiation time in the presence of hierarchical
ZnO microspheres; (b) comparison of self-degradation (black curve) and photocatalytic degradation of
metamitron with ZnO (red curve).

14
Fig.4 A possible photocatalytic schematic for photodegradation of metamitron over ZnO nanoflowers.

15
Fig. 5 (a) Photodegradation ratio and (b) the linear relationship between ln(Co/C) and irradiation
time for metamitron without and with the addition of isopropyl alcohol, carnosine, and
triethanolamine.

16
Fig. 6 Five repeated process of using the hierarchical ZnO microspheres for the photodegradation
of metamitron before and after light irradiation (t = 240 min).

17

You might also like