You are on page 1of 19

Author’s Accepted Manuscript

Growth and optical properties of hierarchical


flower-like ZnO nanostructures

Xiangzhou Lv, Xinchuan Liu, Qimeng Sun,


Yongqian Wang, Bing Yan

www.elsevier.com/locate/ceri

PII: S0272-8842(16)32175-7
DOI: http://dx.doi.org/10.1016/j.ceramint.2016.11.168
Reference: CERI14256
To appear in: Ceramics International
Received date: 6 September 2016
Revised date: 8 November 2016
Accepted date: 23 November 2016
Cite this article as: Xiangzhou Lv, Xinchuan Liu, Qimeng Sun, Yongqian Wang
and Bing Yan, Growth and optical properties of hierarchical flower-like ZnO
n a n o s t r u c t u r e s , Ceramics International,
http://dx.doi.org/10.1016/j.ceramint.2016.11.168
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 galley proof before it is published in its final citable 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.
Growth and optical properties of hierarchical flower-like ZnO

nanostructures
Xiangzhou Lva, Xinchuan Liua, Qimeng Suna, Yongqian Wanga,b*, Bing Yana,b*
a
Faculty of Materials Science and Chemistry, China University of Geosciences, Wuhan 430074,
China
b
Engineering Research Center of Nano-Materials of Ministry of Education, China University of
Geosciences, Wuhan 430074, China
cugwyq@126.com
125583005@qq.com
*
Correspondence author: Tel.: +86 13871379285
*
Correspondence author: Tel.: +86 13476140770

Abstract
The integration of low dimensions nanoscale building blocks into 3D architectures has attracted
great scientific attention. We have obtained the novel hierarchical flower-like ZnO nanostructures
self-assembled by nanorods via a facile hydrothermal method. The as-synthesized samples were
characterized with various technologies. The field emission scanning electron microscope
(FESEM) images indicated that hydroxide ions play a significant role on the formation of
hierarchical flower-like ZnO nanostructures. The X-ray diffraction (XRD) result proved that the
nanocrystals were well crystallized hexagonal wurtzite structure. A possible growth mechanism of
the nanostructures was proposed based on the effects of hydroxide ions. And the TEM imagines
provided some important evidence for the proposed growth mechanism. UV-Vis adsorption and
photoluminescence (PL) spectra results indicated that the obtained ZnO nanostructurs have a good
optical-absorption and photoluminescence property. The as-synthesized ZnO nanostructures
exhibited superior photocatalytic performance, which was higher than that of commercially
available ZnO.
Keywords: flower-like ZnO, hydrothermal, growth mechanism, optical property,
photodegradation

1. Introduction
It is an imperative to find efficient materials to cope with the environmental pollution and the
energy crisis under the increasingly severe ecological environment. The traditional methods
include physical adsorption, chemical oxidation [1,2], microbial processing [3,4] and so on.
However, these are limited to a certain extent for their poor efficiency, incomplete degradation of
pollutants, narrow range of application and high cost. A tremendous efforts have been devoted to
semiconductor-based photocatalyst [5-7] for their inexhaustibility, universality, and environmental
benignancy of solar energy since semiconductor oxide photocatalysts were first discovered [8].
Among these wide varieties of semiconductors,ZnO has been mostly studied due to its superior
properties with a high electron mobility and a large exciton binding energy of 60 meV [9] .
It has been reported that the properties and functions of ZnO can be greatly associated with its
microstructures [10], especially, their morphologies and dimensions. To date, ZnO with different
nanostructures has been fabricated via various methods [11-18]. The dimensions of ZnO have
been zero-dimension (0D) including quantum dots [19], nanoparticles [20]; one-dimension (1D)
nanorods [21], nanowires [22]; two-dimension (2D) nanosheets [23], nanobelts [24];
three-dimension (3D) nanospheres [25], nanoflower [26]. Recently, scientific researchers have
focused on the integration of low dimensions nanoscale building blocks into 3D architectures due
to their higher stability and more superior properties [27-29]. The flower-like nanostructures, as a
special 3D architecture of ZnO with a large surface area and a direct pathway for electrontransport
[30] could lead to unique properties and potential applications. Chen and his co-workers [31]
synthesized flower-like ZnO nanostructures assembled by nanosheets with the assistance of
sodium dodecyl sulfate (SDS). Shi et al. [32] and Sun et al. [33] reported flower-like ZnO
nanostructures comprised of sword-like nanorods and needle-like nanorods, respectively. However,
most of the flower-like ZnO nanostructures that have been prepared were in the form of ZnO
powder, which will bring some difficulties in recycling and practical applications. Although some
flower-like ZnO films have been reported, most of them are deposited on Si or ITO glass
substrates via high-cost and complicated method [34,35]. Hydrothermal synthesis with zinc
substrate has been rarely reported. In addition, most previous researches have focused much on the
properties and applications of synthesized flower-like nanostructures but neglected the exploration
of their growth mechanism, which has important guiding significance for the fabrication and
applications.
In this contribution, we have synthesized the novel hierarchical flower-like ZnO nanostructures
on zinc substrate through a facile hydrothermal method. The effects of the concentrations of
hydroxide ions on the formation of the hierarchical flower-like ZnO nanostructures have been
systematically investigated. The results indicated that there is a strong relationship between the
concentrations and the formed ZnO nanostructures. Furthermore, a reasonable growth process and
mechanism of the flower-like nanostructures has been proposed based on the researches we have
conducted and some previous studies, hoping to provide some guidance for the preparation of this
special nanostructures.

2. Experiment
2.1. Materials
All reagents used in the experiment were of analytical grade and employed without any extra
purification throughout the whole process. Deionized water was used for the preparation for all
solution.
2.2. Experimental procedure
In a typical experimental process, zinc foils (1.5 cm*1.5 cm) were polished by using a sandy
process to remove the oxide film on the surface. The pretreated foils were cleaned by sonication in
acetone, ethanol and deionized water for 10 min, respectively. 40 ml mixed solution comprised of
sodium hydroxide (NaOH, 0.85 M) and dihydrate zinc acetat (Zn(Ac)2·2H2O, 0.035 M) was
prepared. Subsequently, the mixed solution was transferred into a 100 ml Teflon-lined stainless
steel autoclave and the cleaned zinc piece was put into the solution in the autoclave, which was
maintained and heated at 70°C for 12h. After the hydrothermal treatment, the resulting zinc foil
was taken out and thoroughly rinsed with deionized water. Finally, it was transferred into a muffle
furnace to be calcined at 300°C for 2h before further characterization. Other samples were
obtained by changing the concentration ratio of sodium hydroxide and zinc acetate.
2.3. Characterization
We characterized the as-synthesized samples with various technologies. The general
morphologies and crystal structure of the synthesized samples were investigated with field
emission scanning electron microscopy (FESEM, SU8010, HITACHI) and scan transmission
electron microscopy (STEM, CM12, Philips). The samples were coated with 5~10 nm Au layer
before the FESEM imagining. Structure and crystal phase of the samples were characterized by
X-ray diffraction (XRD, D8 Advanced, Bruker AXS) using a Cu Kα radiation source. Ultraviolet–
visible diffuse reflectance spectra were recorded with an Ultraviolet–visible spectrophotometer
(UV-Vis, UV-3600, Shimadzu). And the room temperature photoluminescent spectra of the
samples were obtained using photoluminescence spectrophotometer (PL, F-4500, HITACHI) at an
excitation wavelength at 325 nm.
2.4. Photocatalytic test
Degradation efficiency of methylene blue was recorded as a significant index to evaluate the
photocatalytic performance of the samples. Before photocatalytic test, all mixed solution with
photocatalysts and methylene blue were placed in dark environment for 1h to reach adsorption
equilibrium, and the residual concentration of methylene blue was measured. In a typical test, a
250 W high pressure mercury lamp was used as light source. The lamp was placed 8 cm above the
liquid surface. Subsequently, synthesized samples were added into a 100 ml of 1*10-5 M
methylene blue aqueous solution respectively. The residual concentration of methylene blue was
measured by a UV-Vis spectrophotometer. Photocatalytic efficiency of the samples was calculated
by following equation:
Photocatalytic efficiency (%) = (C0 – C)/C0 * 100%
Where C is the concentration of methylene blue at each irradiated time, and C0 is the initial
concentration when adsorption-desorption equilibrium is achieved.

3. Results and discussion


3.1. Morphologies analysis
The effects of OH- on the formation of the hierarchical flower-like nanostructures was
systematically investigated by changing the concentration ratio of sodium hydroxide and dihydrate
zinc acetat. The results indicated that OH- play an important role on the morphology and
nanostructures of ZnO.
3.1.1. Effects of low concentrations of hydroxide ions
Fig. 1 shows the FESEM images of the samples synthesized on zinc substrate with relative low
concentrations of OH- at 70°C for 12h. As shown in Fig. 1a, the sample prepared with a
concentration of 0.035 M presented no regular morphology, and the whole nanostructures were
composed of rough porous nanosheets which aggregated a net-like structure. When the
concentration of OH- was 0.070 M, the net-like structure disappeared and nanoparticles randomly
distributed were formed (Fig. 1b). Using 0.105 M of OH-, agminated nanoparticles were observed
in some regions. Different from the nanoparticles in Fig. 1b, some of the nanoparticles presented a
regular cone shape. In addition, nanorods were also found to have formed, but they were isolated
from each other (Fig. 1c). When the concentration increased to 0.140 M, the flower-like
nanostructures began to form, as depicted in Fig. 1d. Moreover, it can be spotted that the
flower-like nanostructures were formed on a layer of ZnO nanorod arrays, which grown
perpendicular to the zinc substrate. From the inset of Fig. 1d we could find that a single flower
was made up of nanorods that were “fused” together. To conclude, low concentrations of OH- are
not suitable to form the hierarchical flower-like ZnO nanostructures. When the concentrations
increased from 0.070 M to 0.140 M, nanoparticles transformed into nanorods, and disorganized
and nonuniform flower-like nanostructures were first observed to have formed under the
concentration of 0.140 M.

Fig. 1 FESEM images of ZnO prepared with the low concentrations of OH-,
(a) 0.035 M, (b) 0.070 M, (c) 0.105 M, (d) 0.140 M

3.1.2. Effects of high concentrations of hydroxide ions


Fig. 2 shows the FESEM images of the samples synthesized with relative high concentrations of
sodium hydroxide at 70°C for 12h. The morphology of the sample prepared with a concentration
of 0.55 M was shown in Fig. 2a, where a couple of flowers formed, but they were not
homogenously distributed. In some regions, flowers accumulated while in some other regions
there were no flower-like nanostructures existing at all. The inset in Fig. 2a indicated that the
“flower” were comprised of nanorods, but the length and diameter of these nanorods varied and
the shape of the end of nanorods was irregular. When the concentration of OH- reached 0.65 M
(Fig. 2b), the general morphology was similar to that of 0.55 M. The differences are that the
amount of flowers were much more and the diameter of nanorods became more average. When the
concentration was 0.75 M, the general morphology (Fig. 2c) has changed slightly compared with
that of Fig. 2a and b, and the length and diameter of the nanorods have been average. However,
with a concentration of 0.85 M, the morphology varied. It can be obviously observed that the
flowers were more uniformly distributed and they interacted with each other (Fig. 2d). The inset in
Fig. 2d is a magnifying image, from which one can find that a single flower was comprised of
many nanorods radiated from the center of the nanostructures with pretty good symmetry.
Additionally, the length and diameter of the nanorods were in consistence with each other. TEM
imagines of the sample prepared under this concentration is shown in Fig. 3. Well dispersed and
oriented ZnO nanorods agglomerate perpendicular to the substrate was found in Fig. 3a. The
diameter of these nanorods is 30 nm in average. And Fig. 3b is the image of the ZnO nanoflowers,
different from the flowers in the FESEM imagines, the symmetry and uniformity have decreased
sharply. This can be accountable for the dispersal process during the preparation for TEM
characterization has damage the self-assembled flower-like nanostructures. Fig. 3d is a
higher-resolution image of c, the red line in the white rectangular frame is a boundary of the two
different ZnO nanostructures: the nanorod arrays and the nanoflowers, which could be a
reasonable evidence for our proposed growth mechanism in the latter portion of our article. No big
differences have been found between the concentrations of 0.95 M and 0.85 M. However, when
the concentration increased to 1.05 M (Fig. 2f) and 1.15 M (Fig. 2g), the uniformity and symmetry
of the single flower decreased. This could be accountable for the effects of the excessive OH-,
which we would later discuss in the growth process and mechanism section concretely. When the
concentration was up to 1.50 M, flower-like nanostructures were not obvious to be observed
because of the disorder and muddled distribution of the nanorods (Fig. 2h). Furthermore, it can be
found from the red-circled region in the FESEM image besides Fig. 2g that the size of the flower
has been smaller and the number of the nanorods that made up for the petals was even less.
Interestingly, the hexagonal nanorods turned into cone-shaped (the FESEM image besides Fig. 2h).
In this high concentrations experiment, we can find that flower-like nanostructures have formed
under all various concentrations, which indicated that the high concentrations of OH- are more
easy and suitable for the formation of flower-like ZnO nanostructures compared with the low
concentrations.
Fig. 2 FESEM images of ZnO prepared under the various high concentrations of OH-: (a) 0.55 M,
(b) 0.65 M, (c) 0.75 M, (d) 0.85 M, (e) 0.95 M, (f) 1.05 M, (g) 1.15 M, (h) 1.50 M
Fig.3 TEM images of the synthesized sample with a concentration of 0.85 M NaOH. (a) ZnO
nanorod arrays (c) the formed ZnO nanoflower, (b) as well as (d) are the higher-resolution images
of (a) and (c), respectively

3.2. XRD analysis


Fig. 4 shows the XRD pattern of the hierarchical flower-like nanostructures fabricated with 0.85
M of OH- at 70°C for 12h. The diffraction peaks of Fig. 3 at 31.77°,34.45°,36.27°,47.55°,
56.58°,62.85°,66.38°,67.94°and 69.09°match well with (100), (002), (101), (102), (110),
(103), (200), (112) and (201) planes of the standard data for the wurtzite structure of ZnO (JCPDS
No.36-1451), which indicated that the synthesized ZnO exhibited hexagonal wurtzite crystalline
structure. The strong and sharp diffraction peaks as well as the symmetrical narrow full width at
half-maximum (fwhm) of the peaks indicate good crystallization of the sample. Moreover, except
the ZnO, no diffraction peaks from other impurities such as Zn, Zn(OH)2 are detected in the XRD
pattern, which conforms a purely wurtzite hexagonal phase ZnO.
Fig. 4 XRD pattern of the ZnO sample prepared with 0.85 M of OH-

3.3. Growth process and mechanism


Base on the results and some previous works [36,37], we proposed a growth process and
mechanism of the flower-like nanostructures and the effects of OH- on the growth were further
investigated carefully. Basically, the crystal formation in solution can be divided into two parts,
that is, crystal nucleation and crystal growth. According to our researches and the studies
previously [11,38-40], the growth process of ZnO crystallities in an alkaline solution can be
described as the following mechanism:
Zn (foil) + H2O → Zn(OH)2 → ZnO (1)
Zn (foil) + 2OH- → Zn(OH)2 + 2e- → ZnO (2)
2+ - 2-
Zn + 4OH →[Zn(OH)4] (3)
2- -
[Zn(OH)4] → ZnO + H2O + 2OH (4)
We think the process of the formation of the flower-like nanostructures can be involved in two
stages: the growth of nanorod arrays on zinc substrate and the formation of the flower-like
nanostructures on the formed nanorod arrays. The schematic diagram of the possible growth
process is shown in Fig. 5.
In the first stage some zinc atoms on the surface of zinc foil became extremely active under
relative high temperature and pressure, they reacted with water (Eq. 1) and OH- (Eq. 2) in the
solution to produce Zn2+ ions, which would reacted with OH- later to produce zinc hydroxide
(Zn(OH)2 ) precipitates on the zinc substrate. Of course we think the main route to produce Zn2+
should be Eq. 2. Under this special hydrothermal condition, the precipitates would transform into
ZnO nuclei in situ (Fig. 5a). Prolonging the reaction time, ZnO nuclei became much more, and
they would aggregate to reduce the surface energy, in which way the ZnO seeds layer was formed
(Fig. 5b). At the same time, Zn2+ from Zn(Ac)2·2H2O would react with OH- to generate the growth
unit of [Zn(OH)4]2- (Eq. 3). Afterwards, [Zn(OH)4]2- ions would diffuse and adsorb on the surface
on the ZnO seeds layer. Later [Zn(OH)4] 2- ions changed into very minute ZnO particles in situ
through thermal decomposition (Eq. 4). Umar [41] thinks the wurtzite hexagonal crystal structure
ZnO has the lowest energy faces of (0001) that ZnO particles would grow along the (0001)
direction at a most fast velocity than any other faces, which is along the c-axis, to form nanorod
arrays that are perpendicular to the zinc substrate. To the best of our knowledge, ZnO is a wurtzite

polar crystal with a positive polar plane (0001)-Zn and a negative polar plane ( 0001 )-O. For the

[Zn(OH)4]2- ions are negatively charged, they have a tendency to diffuse and adsorb along the
positive polar plane (0001) -Zn, resulting in the crystallization along the c-axis, which the nanorod
arrays were finally formed (Fig. 5c). In addition, from the FESEM images of Fig. 1d we can find
that the flower-like nanostructures were formed on the nanorods, which can be an evidence for the
formation of the nanorod arrays.
The growth of the nanorod arrays ceased when the nanorods reached a certain length. But the
excessive Zn2+ and OH- in the solution would continuously product [Zn(OH)4] 2-, which would
eventually transform into ZnO nuclei on the nanorods arrays. It should be noticed that this time the
nuclei were not evenly distributed, their distribution was irregular and disordered due to the
diversity of the length of each nanorods. Moreover, some nucleus would accumulate (Fig. 5d) and
the crystal growth process would continue along the c-axis of each nucleus, thus the flower-like
ZnO nanostructures assembled by nanorods were formed (Fig. 5e).
In addition to the hierarchical flower-like nanostructures, the formation of other structures
should also be explained. When the concentration of OH- was pretty low (0.035 M), rough porous
nanosheets aggregated a net-like structure was observed (Fig. 1a). The surface of the polished zinc
foil was rugged, and the more rugged the region was the lower the energy barrier of nucleation
was. In the absence of OH-, the growth unit [Zn(OH)4] 2- ions were not ample, they would first
diffuse to the regions of low energy barrier to achieve the lowest thermodynamic stability. The
inhomogeneous distribution of [Zn(OH)4] 2- resulted in such irregular structure. With the
concentrations of NaOH increasing, there would be more [Zn(OH)4] 2- ions. On the condition of
sufficient [Zn(OH)4] 2- ions, the amount of [Zn(OH)4] 2- ions that diffused to different regions was
average, which was benefit to the formation of the uniform flower-like nanostructures. However,
there would be excessive OH- ions in the solution with a pretty high concentration. For ZnO is an
amphoteric oxide, the formed flower-like ZnO nanostructures would be partially dissolved by the
excessive OH- ions. When the concentration was higher than the optimum concentration of 0.85 M,
the higher the concentrations were the more serious damage to the flower-like nanostructure
would be. Specially, while the concentration was up to 1.50 M, the flowers were fewer and
uniform distributed and the nanorods that comprised the flower turned into cone-shaped. That was
because except for the corrosive effect of the excessive OH-, excessive OH- also ensured ample
growth unit of [Zn(OH)4] 2- ions, the nucleation and growth velocities were pretty fast, which
resulted in the randomly distribution of nanorods and their being cone-shaped. In a word, the
amount of OH- is essential to the nanostructures of ZnO.
Fig. 5 Schematic illustration of the growth process and mechanism for the flower-like ZnO
nanostructures
3.4. Optical properties
3.4.1. UV-Vis
In order to investigate the optical properties of synthesized flower-like ZnO nanostructures, the
UV-Vis adsorption spectra of ZnO nanostructures fabricated under 0.85 M OH- are presented in
Fig. 6. It is clearly observed that the synthesized sample showed a strong adsorption in the UV
light region, which should be ascribed to its direct band gap of 3.37 eV. Based on the relationship
of the adsorption coefficient with incident photo energy (αhv)=B(hv-Eg)n, where α is the
adsorption coefficient, hv is the energy for the incident photo, Eg is the band gap of semiconductor,
n is 1/2 for a direct transition semiconductor, e.g. ZnO. Since α is proportional to the absorbance
values, the energy intercept of the curve of (αhv)2 vs hv gives Eg when the linear region is
extrapolated to the zero ordinate. The band gap was calculated to be 3.15 eV as shown in the inset
of Fig. 6.
Fig. 6 UV-Vis adsorption spectra of flower-like ZnO nanostructures with 0.85 M OH-

3.4.2. PL spectra
Photoluminescence spectra are another method to investigate the optical property of the
synthesized nanostructures. The photoluminescence spectra were measured with excitation
wavelength at 325 nm using a xenon lamp as excitation source. Fig. 7a and b show the PL spectra
of the as-synthesized samples under the low and high concentrations of OH-, respectively. In Fig.
7a we could observe strong UV emissions from 380 nm to 400 nm and two relative weak blue
emission peaks at 446 nm and 464 nm, independently. The intensities of the UV emission peaks
decrease and the position of the UV emission peak is red shifted as the concentrations increase.
For the high concentrations of OH- in Fig. 7b, the UV emissions center at 393 nm, besides two
blue emissions are also presented at 448 nm and 464 nm, respectively. In general, the samples
prepared under different concentrations have UV and blue emissions nearly at the same
wavelength but the intensities vary.
A lot of researches have been carried out to investigate the photoluminescence mechanism of
ZnO so far [42-45]. It is generally accepted by most of the researchers that the UV emissions are
due to the near band edge (NBE) emission [41,46,47]. When ZnO is irradiated under a certain
wavelength light, the electrons in the valence band (VB) will be excited into the conduction band
(CB) to form electron-hole pairs. The excitons will later recombine to luminesce, for the band gap
of ZnO is 3.37 eV, the NBE emission peaks usually concentrate at 380 nm to 400 nm. And the
broad emission peaks in the visible region have been recognized as deep level emission (DPE),
which is ascribed to the intrinsic defects[42,48-50]. The intrinsic defects in ZnO involved
vacancies, interstitials, and anti-sites of Zn and of O, which can be classified into donor-type and
acceptor-type defects based on their physical properties[48,50-52]. Based on the first-principles
calculation [48,50,51,54], the doubly ionized zinc vacancies (VZn2-) are located about 0.56 eV
lower than the CBM (conduction-band minimum), that is about 2.81 eV higher than the VBM
(valance-band maximum)[51,52,54]. The electrons in the VZn2- can recombine with the
photogenerated holes in the VB under special excitation conditions [51]. And the transition energy
between the defects of VO2+ and VO+ is calculated to be 0.7-0.8 eV above the VBM
(E(2-/1-)=0.7-0.8 eV) [52-54]. VO2+, generated by surface trapping of photogenerated holes were
regarded as recombination centers, with which the electrons in the VO+ will combine [51,54].
According to the formula E=1242.375/λ, where λ is the wavelength, we can get the corresponding
emission wavelength of these transitions. The results are in accordance with the PL spectra,
indicating that the blue emissions are duo to the oxygen vacancies and zinc vacancies. For the low
concentrations series, the amount of OH- was inadequate, because of the lack of raw materials the
crystalline quality of ZnO was not so good. And more oxygen vacancies were likely to come into
being, which can be the main reason for the blue emissions. However, as the concentrations
increased the crystalline quality of ZnO became better, resulting fewer oxygen vacancies, causing
the red shift of the UV emission peak and the intensities decreasing. When the concentrations were
high, the amount of OH- was sufficient, which ensured reasonably good crystalline quality. The
intensities of the emission peaks have great association with the defect configurations and
concentrations, which are greatly determined by growth conditions [43,44]. So in the case of
adequate raw materials the intensities should have weak correlation with the concentrations
compared with the growth conditions. And the blue emission peaks in the range 446-468 nm in
different samples may be originated from oxygen vacancies and zinc vacancies, but the later
should be the main cause [54].

Fig. 7 PL spectra of the synthesized samples under low concentrations(a) and high
concentrations(b) of OH-

3.5. Photocatalytic test


Fig. 8 shows the photo-degradation rate of commercially available ZnO and the synthesized
hierarchical flower-like ZnO nanostructures. It can be observed obviously that the
photo-degradation rate reached the maximum at 210 min for both the two catalyzers. However, the
photodegradation rate of the flower-like ZnO nanostructures was as high as 91.9% while the
maximum of the commercially available ZnO was only 66%. In addition, we can observe that for
the flower-like ZnO nanostructures the photo-degradation rate was positively correlated with the
time, which indicated the photodegradation reaction was carried out stably and at a uniform speed.
Therefore, we can conclude that the synthesized flower-like ZnO nanostructures have good
photocatalytic performance and have a great prospect in the application of environmental
governance.
Fig. 8 The photodegradation rate of the commercially available ZnO and the synthesized
hierarchical flower-like ZnO nanostructures

4. Conclusion
In summary, we have successfully synthesized the hierarchical flower-like ZnO nanostructures
on the zinc substrate through a facile hydrothermal method at a relative low temperature. The
concentration-dependent experiments indicate that the concentrations of OH- determine the
amount of [Zn(OH)4]2-, which in turn affects the growth process and the formations of ZnO
nanostructures. And the growth process of the synthesized nanostructures can be involved in two
stages: the growth of nanorod arrays on zinc substrate and the formation of the flower-like
nanostructures on the formed nanorod arrays. UV-Vis adsorption spectra indicate that the
flower-like ZnO nanostructures have a strong of UV light and the band gap of it is calculated to be
3.15 eV. In the PL spectra this synthesized nanostructures have a strong UV emission and two blue
emission peaks centering at about 446nm and 464nm, respectively. These above indicate that the
optical property of the flower-like ZnO nanostructures is pretty good. And the photo-degradation
rate of the flower-like ZnO nanostructure is as high as 91.9%, suggesting its great prospect in the
application of environmental governance.

Acknowledgement
This work was supported by National Training Programs of Innovation and Enterpreneurship
for Undergraduates (NO. 201610491013). The financial support was gratefully appreciated.
References

[1] Sun H, Liu S, Zhou G, et al. Reduced graphene oxide for catalytic oxidation of aqueous
organic pollutants [J]. ACS Applied Materials & Interfaces, 2012, 4(10): 5466-5471.
[2] Wenk J, Aeschbacher M, Salhi E, et al. Chemical oxidation of dissolved organic matter by
chlorine dioxide, chlorine, and ozone: effects on its optical and antioxidant properties [J].
Environmental Science & Technology, 2013, 47(19): 11147-11156.
[3] Tenca A, Cusick R D, Schievano A, et al. Evaluation of low cost cathode materials for
treatment of industrial and food processing wastewater using microbial electrolysis cells [J].
International Journal of Hydrogen Energy, 2013, 38(4): 1859-1865.
[4] Cheirsilp B, Louhasakul Y. Industrial wastes as a promising renewable source for production
of microbial lipid and direct transesterification of the lipid into biodiesel [J]. Bioresource
Technology, 2013, 142: 329-337.
[5] Bao Y, Wang C, Ma J. Trisodium citrate as bridging and suppressing agent to control synthesis
of ZnO hollow hierarchical microspheres and their photocatalytic properties [J]. Ceramics
International, 2016, 42(1): 1746-1755.
[6] An H R, Park S Y, Kim H, et al. Advanced nanoporous TiO2 photocatalysts by hydrogen
plasma for efficient solar-light photocatalytic application [J]. Scientific Reports, 2016, 6.
[7] Zhukovskyi M, Tongying P, Yashan H, et al. Efficient photocatalytic hydrogen generation from
Ni nanoparticle decorated CdS nanosheets [J]. ACS Catalysis, 2015, 5(11): 6615-6623.
[8] Fujishima A. Electrochemical photolysis of water at a semiconductor electrode [J]. Nature,
1972, 238: 37-38.
[9] Lieber C M, Wang Z L. Functional nanowires [J]. MRS bulletin, 2007, 32(02): 99-108.
[10] Xia Z, Sha J, Fang Y, et al. Purposed built ZnO/Zn5(OH)8AC2·2H2O architectures by
hydrothermal synthesis [J]. Crystal Growth & Design, 2010, 10(6): 2759-2765.
[11] Quan J, Colorado H A, Yeh P C, et al. Hybridized ZnO nanostructures on carbon-fiber
through combustion synthesis induced by joule heating [J]. Ceramics International, 2016,
42(11): 13053-13060.
[12] Yang P, Xiao X, Li Y, et al. Hydrogenated ZnO core-shell nanocables for flexible
supercapacitors and self-powered systems [J]. ACS nano, 2013, 7(3): 2617-2626.
[13] Dong Z, Lai X, Halpert J E, et al. Accurate control of multishelled ZnO hollow microspheres
for dye‐sensitized solar cells with high efficiency [J]. Advanced Materials, 2012, 24(8):
1046-1049.
[14] Zhao X, Li M, Lou X. Sol–gel assisted hydrothermal synthesis of ZnO microstructures:
morphology control and photocatalytic activity [J]. Advanced Powder Technology, 2014,
25(1): 372-378.
[15] Park S, An S, Ko H, et al. Synthesis of nanograined ZnO nanowires and their enhanced gas
sensing properties [J]. ACS applied Materials & Interfaces, 2012, 4(7): 3650-3656.
[16] Umar A, Algarni H, Kim S H, et al. Time dependent growth of ZnO nanoflowers with
enhanced field emission properties [J]. Ceramics International, 2016, 42(11): 13215-13222.
[17] Khan M F, Ansari A H, Hameedullah M, et al. Sol-gel synthesis of thorn-like ZnO
nanoparticles endorsing mechanical stirring effect and their antimicrobial activities: Potential
role as nano-antibiotics [J]. Scientific Reports, 2016, 6.
[18] Downing J M, Ryan M P, McLachlan M A. Hydrothermal growth of ZnO nanorods: The role
of KCl in controlling rod morphology [J]. Thin Solid Films, 2013, 539: 18-22.
[19] Yang S J, Nam S, Kim T, et al. Preparation and exceptional lithium anodic performance of
porous carbon-coated ZnO quantum dots derived from a metal–organic framework [J].
Journal of the American Chemical Society, 2013, 135(20): 7394-7397.
[20] Xiong H M. ZnO nanoparticles applied to bioimaging and drug delivery [J]. Advanced
Materials, 2013, 25(37): 5329-5335.
[21] Yang J, Wang X, Jiang T, et al. Controllable preparation, growth mechanism and the
properties research of ZnO nanocrystal [J]. Superlattices and Microstructures, 2014, 72:
91-101.
[22] Kim J H, Yer I H. Growth of ZnO nanowire arrays on Ga-doped ZnO transparent conductive
layers [J]. Ceramics International, 2015, 41(8): 10227-10231.
[23] Li H, Wang X. Three-dimensional architectures constructed using two-dimensional
nanosheets [J]. Science China Chemistry, 2015, 58(12): 1792-1799.
[24] Yu L, Liu S, Yang B, et al. Sn–Ga co-doped ZnO nanobelts fabricated by thermal evaporation
and application to ethanol gas sensors [J]. Materials Letters, 2015, 141: 79-82.
[25] Yang C, Li Q, Tang L, et al. Synthesis, photocatalytic activity, and photogenerated hydroxyl
radicals of monodisperse colloidal ZnO nanospheres [J]. Applied Surface Science, 2015, 357:
1928-1938.
[26] Sohila S, Rajendran R, Yaakob Z, et al. Photoelectrochemical water splitting performance of
flower like ZnO nanostructures synthesized by a novel chemical method [J]. Journal of
Materials Science: Materials in Electronics, 2016, 27(3): 2846-2851.
[27] Sun X, Li Q, Jiang J, et al. Morphology-tunable synthesis of ZnO nanoforest and its
photoelectrochemical performance [J]. Nanoscale, 2014, 6(15): 8769-8780.
[28] Zhan W, Kuang Q, Zhou J, et al. Semiconductor@ metal–organic framework core–shell
heterostructures: A case of ZnO@ ZIF-8 nanorods with selective photoelectrochemical
response [J]. Journal of the American Chemical Society, 2013, 135(5): 1926-1933. J. Am.
Chem. Soc. 135(2013);1926–1933.
[29] Zhao X, Lou F, Li M, et al. Sol-gel-based hydrothermal method for the synthesis of 3D
flower-like ZnO microstructures composed of nanosheets for photocatalytic applications [J].
Ceramics International, 2014, 40(4): 5507-5514.
[30] Q.F. Zhang, C.S. Dandeneau, X.Y. Zhou, G.Z. Cao, ZnO nanostructures fordye-sensitized
solar cells, Advanced Materials. 21 (2009) 4087–4108.
[31] Chen X, Liu J, Jing X, et al. Self-assembly of ZnO nanosheets into flower-like architectures
and their gas sensing properties [J]. Materials Letters, 2013, 112: 23-25.
[32] R.X. Shi, P. Yang, J.R. Wang, A.Y. Zhang, Y.N. Zhu, Y.Q. Cao, Q. Ma, Growth offlower-like
ZnO via surfactant-free hydrothermal synthesis on ITO substrateat low temperature,
CrystEngComm 14 (2012) 5996–6003.
[33] H.Y. Sun, Y.L. Yu, J. Luo, M. Ahmad, J. Zhu, Morphology-controlled synthesis ofZnO 3D
hierarchical structures and their photocatalytic performance,CrystEngComm 14 (2012) 8626–
8632.
[34] Ahmed N A, Hammache H, Makhloufi L, et al. Effect of electrodeposition duration on the
morphological and structural modification of the flower-like nanostructured ZnO [J]. Vacuum,
2015, 120: 100-106.
[35] Park J K, Kim Y J, Yeom J, et al. The topographic effect of zinc oxide nanoflowers on
osteoblast growth and osseointegration [J]. Advanced materials, 2010, 22(43): 4857-4861.
[36] Wang J X, Sun X W, Yang Y, et al. Hydrothermally grown oriented ZnO nanorod arrays for
gas sensing applications [J]. Nanotechnology, 2006, 17(19): 4995.
[37] Zhang H, Wu R, Chen Z, et al. Self-assembly fabrication of 3D flower-like ZnO hierarchical
nanostructures and their gas sensing properties[J]. CrystEngComm, 2012, 14(5): 1775-1782.
[38] Liu J, Lu R, Xu G, et al. Development of a seedless floating growth process in solution for
synthesis of crystalline ZnO micro/nanowire arrays on graphene: Towards high performance
nanohybrid ultraviolet photodetectors [J]. Advanced Functional Materials, 2013, 23(39):
4941-4948.
[39] Ko S H, Lee D, Kang H W, et al. Nanoforest of hydrothermally grown hierarchical ZnO
nanowires for a high efficiency dye-sensitized solar cell [J]. Nano letters, 2011, 11(2):
666-671.
[40] Xu S, Wang Z L. One-dimensional ZnO nanostructures: solution growth and functional
properties [J]. Nano Research, 2011, 4(11): 1013-1098.
[41] Umar A, Ribeiro C, Al-Hajry A, et al. Growth of highly c-axis-oriented ZnO nanorods on
ZnO/glass substrate: growth mechanism, structural, and optical properties [J]. The Journal of
Physical Chemistry C, 2009, 113(33): 14715-14720.
[42] Zhang L, Yin L, Wang C, et al. Origin of visible photoluminescence of ZnO quantum dots:
defect-dependent and size-dependent [J]. The Journal of Physical Chemistry C, 2010, 114(21):
9651-9658.
[43] Najafi M, Haratizadeh H, Ghezellou M. The effect of annealing, synthesis temperature and
structure on photoluminescence properties of Eu-doped ZnO nanorods [J]. Journal of
Nanostructures, 2015, 5(2): 129-135.
[44] Verma D, Kole A K, Kumbhakar P. Red shift of the band-edge photoluminescence emission
and effects of annealing and capping agent on structural and optical properties of ZnO
nanoparticles [J]. Journal of Alloys and Compounds, 2015, 625: 122-130.
[45] Boudjouan F, Chelouche A, Touam T, et al. Effects of stabilizer ratio on photoluminescence
properties of sol-gel ZnO nano-structured thin films [J]. Journal of Luminescence, 2015, 158:
32-37.
[46] Yang J, Wang Y, Jiang T, et al. ZnO/Er2O3 core–shell nanorod arrays: Synthesis, properties
and growth mechanism [J]. Applied Surface Science, 2015, 325: 117-123.
[47] Biroju R K, Giri P K, Dhara S, et al. Graphene-assisted controlled growth of highly aligned
ZnO nanorods and nanoribbons: growth mechanism and photoluminescence properties [J].
ACS applied materials & interfaces, 2013, 6(1): 377-387.
[48] Kohan A F, Ceder G, Morgan D, et al. First-principles study of native point defects in ZnO[J].
Physical Review B Condensed Matter, 2000, 61(22):15019-15027.
[49] Zhang D H, Xue Z Y, Wang Q P. The mechanisms of blue emission from ZnO films deposited
on glass substrate by rf magnetron sputtering[J]. Journal of Physics D: Applied Physics, 2002,
35(21): 2837.
[50] Oba F, Choi M, Togo A, et al. Point defects in ZnO: an approach from first principles[J].
Science and Technology of Advanced Materials, 2011, 12(3):1462-1470.
[51] Djurišić A B, Leung Y H. Optical properties of ZnO nanostructures[J]. small, 2006, 2(8‐9):
944-961.
[52] Xu P S, Sun Y M, Shi C S, et al. The electronic structure and spectral properties of ZnO and
its defects [J]. Nuclear Instruments and Methods in Physics Research Section B: Beam
Interactions with Materials and Atoms, 2003, 199: 286-290.
[53] Tuomisto F, Saarinen K, Look D C, et al. Introduction and recovery of point defects in
electron-irradiated ZnO[J]. Physical Review B, 2005, 72(8): 085206.
[54] Liu J, Lee S, Ahn Y H, et al. Tailoring the visible photoluminescence of mass-produced ZnO
nanowires [J]. Journal of Physics D: Applied Physics, 2009, 42(9): 095401.

You might also like