Sunteți pe pagina 1din 15

RSC Advances

View Article Online


PAPER View Journal | View Issue

Morphological evolution of ZnO nanostructures


and their aspect ratio-induced enhancement in
Cite this: RSC Adv., 2014, 4, 29249
photocatalytic properties†
Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

Faheem Ahmed, Nishat Arshi, M. S. Anwar, Rehan Danish and Bon Heun Koo*

This work presented controllable growth of ZnO nanostructures with different aspect ratios by the
microwave irradiation method and investigated the photocatalytic degradation of methyl red (MR). X-ray
diffraction (XRD), high resolution transmission electron microscopy (HRTEM) and selected area electron
diffraction (SAED) measurements showed that all ZnO nanostructures were of a hexagonal phase
structure. It was revealed by field-emission scanning electron microscopy (FESEM) and transmission
electron microscopy (TEM) images that the morphology of ZnO can be effectively controlled as sheet-
like, rod-like, brush-like, flower-like, prism-like, and pyramid-like only by changing the molar ratio (zinc
acetate: KOH) and reaction time. With the increase of molar ratio and reaction time, modification in the
E2(high) and E1(LO) Raman modes was observed. The energy band gap was found to be tuned by the
aspect ratio of ZnO nanostructures. Photoluminescence spectroscopy revealed the low-intensity NBE
emission and high and broad defect-related emission for high aspect ratio (14) nanorods. BET surface
area porosity analysis confirmed the presence of a mesoporous network in all the nanostructures,
showed high surface area and a uniform pore-size distribution for high aspect ratio nanorods. A
terephthalic acid assay study confirmed the formation of hydroxyl radicals (OH) in MR dye solution
treated with a ZnO nanostructures photocatalyst. The photodegradation of MR under UV light irradiation
showed that ZnO nanorods with a high aspect ratio of 14 showed superior photodegradation (98%
degradation of MR within 60 min) than that of the lower aspect ratio nanostructures. The apparent
Received 20th March 2014
Accepted 9th June 2014
reaction rate constant for high aspect ratio (14) nanorods was higher than that of the lower aspect ratio
nanostructures. The enhancement in photocatalytic performance could be due to the high surface area
DOI: 10.1039/c4ra02470b
and enhanced charge separation and transfer efficiency of photoinduced charge carriers in the high
www.rsc.org/advances aspect ratio nanorods.

Nowadays, increased industrialization poses a threat to


Introduction human health in the form of environmental pollution. Envi-
With a wide and direct band gap of 3.37 eV with a large exciton ronmental remediation of organic pollutants by photocatalytic
binding energy of 60 meV, ZnO, which has been extensively degradation using wide bandgap semiconductors has attracted
used in eld emission,1 short-wavelength optoelectronics,2 considerable attention in recent times.10–13 Furthermore, these
electroacoustic transducers,3 gas sensors,4 transparent con- materials in the nanostructured form exhibit enhanced prop-
ducting coating materials,5 piezoelectric devices,6 and photo- erties, and therefore have potential applications such as in the
catalysis,7 has proven to be a unique functional material during fabrication of nanodevices. Various semiconductors, including
the past several years. Recently, the room temperature ferro- ZnO and TiO2, are being used in photocatalytic application
magnetism (RTFM) behaviour with a Curie temperature (Tc) because of their low cost, high photosensitivity, non-toxicity
value above room temperature in pure ZnO has been reported and environmentally friendly nature.14–16 ZnO has been estab-
by various groups,8,9 which has attracted enormous interest lished to be a more efficient photocatalyst than TiO2 because of
from researchers for possible application in spintronic devices. its high surface reactivity owing to its large number of active
surface defect states. It has also been shown that ZnO has high
reaction and mineralization rates17 because of its more efficient
hydroxyl-ion production.18 When ZnO is irradiated with UV
School of Nano and Advanced Materials Engineering, Changwon National University, light, electrons are excited from the valence band to the
Changwon, Gyeongnam, 641-773, Korea. E-mail: bhkoo@changwon.ac.kr; Fax: +82-
conduction band and electron–hole pairs are created. These
55-262-6486; Tel: +82-55-264-5431
† Electronic supplementary information (ESI) available. See DOI:
electron–hole pairs activate the surrounding chemical species,
10.1039/c4ra02470b and then promote the chemical reactions.19,20 The efficiency of

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 29249–29263 | 29249
View Article Online

RSC Advances Paper

ZnO as a photocatalyst has been limited due to the recombi- nanorods were studied. The nanorods were prepared by auto-
nation of photogenerated charge carriers, which is typically clave conditions at 120  C for 2–6 h. These reports showed that
faster than the rate of production of reactive oxidation species. the preparation method used for ZnO nanostructures is both
Thus, improving the charge separation efficiency needs a time- and energy-consuming and does not full the economic
rational design of the photocatalyst structure. As photocatalytic and industrial requirements of ZnO nanostructures-based
activity of metal-oxide nanostructures depends on structures, photocatalysts. Nevertheless, the morphological evolution of
including the morphology, surface area and surface defects, ZnO nanostructures from nanosheets to nanorods, nano-
one-dimensional (1D) nanostructures such as nanorods, nano- brushes, nanoowers, nanoprisms, and nanopyramids with
wires, and nanotubes are ideal candidates for the application in high aspect ratios and the enhancement of photocatalytic
photocatalysis since they offer a larger surface-to-volume ratio properties by aspect ratio have not been reported to the best of
than nanoparticulate thin lms.21 1D nanostructures with high our knowledge.
Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

aspect ratios have been shown to promote the separation of In this paper, we described the growth of ZnO with a diversity
photogenerated charge carriers due to an increased delocal- of well-dened morphologies with different aspect ratios via the
ization of electrons in the 1D nanostructures, which is impor- simple microwave-assisted solution route within a few minutes,
tant for improving the efficiency of the photocatalysts. and the photocatalytic activity of these nanostructures were
Therefore, it is very important to develop a shape and size- investigated by measuring the degradation of MR. The experi-
controlled synthesis method for ZnO nanostructures. In recent mental results showed that the as-obtained ZnO nanorods with
years, rapid progress has been made in the preparation of ZnO high aspect ratios exhibited excellent photocatalytic activity
nanostructures, including nanowires,22 nanosheets,23 nano- compared with the low aspect ratio nanostructures. This study
tubes24 and nanoowers.25 Numerous approaches were will provide the platform to tune the morphology of ZnO
employed to fabricate ZnO nanostructures, which can be nanostructures using a microwave irradiation method, in which
usually classied into two categories, including physical various dimensional nanostructures such as 1D, 2D, and 3D can
methods and the solution phase route. Physical methods be prepared simultaneously for future applications. Moreover,
include thermal evaporation, chemical vapour deposition and the detailed structural, morphological, and optical character-
metal organic vapour phase epitaxy.26–28 However, the physical ization of ZnO nanostructures will provide helpful information
process requires high temperature, a complex procedure, and on the structure–property relationship. We envisage that this
sophisticated equipment, which make the procedure expensive, work will help to develop new photocatalysts based on high
and thereby restricts possibilities of applications. In contrast, aspect ratio ZnO nanorods for practical application due to the
the solution phase approaches, including microemulsion, sol- excellent photocatalytic behaviour, and this simple and one-
vothermal, hydrothermal, precipitation, self-assembly and step method can suitably be scaled up for large-scale synthesis.
template-assisted sol–gel processes, can allow for the growth of
ZnO nanostructures at much lower temperatures (<200  C) and Experimental section
large-scale production.29–34 Among them, the hydrothermal
technique is a relatively simple, low temperature process, but it All the reagents involved in the experiments were of analytical
requires longer reaction time (from a few hours to several days). grade and utilized as-received without further purication. Zinc
Thus, a simple and fast route for the synthesis of ZnO materials acetate dihydrate (Zn(CH3COO)2$2H2O; 99.999%), and potas-
under ambient conditions to full economic and industrial sium hydroxide (KOH; 99.99%) were purchased from Sigma
requirements is still required. Recently, the synthesis of nano- Aldrich. The synthesis was carried out in a domestic microwave
structured materials via microwave irradiation has been intro- oven (Samsung, 750 W). In a typical experiment, different molar
duced. In comparison with conventional heating, microwave ratios (ZnM : KOHM; 1 : 10, 1 : 15, 1 : 20 and 1 : 25) of
heating has unique effects such as rapid and homogeneous Zn(CH3COO)2$2H2O to KOH were dissolved in 100 mL distilled
volumetric heating, high reaction rate, short reaction time, water in a round-bottomed ask, and then placed in a domestic
enhanced reaction selectivity, energy savings and low cost.9 The microwave oven. Microwave irradiation proceeded at 300 W
effect of some synthesis parameters on the morphology of ZnO (irradiation 12 s, stop 10 s) for different time intervals (10, 20,
nanostructures has been studied. Vayssieres et al. reported that 30, 40 and 60 min). Aer microwave processing, the solution
the size of the ZnO nanostructures was decreased, and the was cooled to room temperature. The resultant precipitate was
aspect ratio was increased with a decrease in the solution separated by centrifugation, and then washed several times
concentration.29 The reaction time in the solution reaction is with deionized water and absolute ethanol, and then dried in an
another important parameter reecting the thermodynamic oven at 80  C for 24 h.
and/or kinetics processes because of its close relation to the Methyl red (MR) was used as the test contaminant for the
change of concentration and reaction rate in the reaction photocatalysis experiments. A 10 mM aqueous solution of MR
process.35 Recently, work has been reported on the growth of was prepared in which 10 mg of ZnO catalysts were suspended.
ZnO nanorods using a hydrothermal method and their photo- Before illumination, the suspension was magnetically stirred in
catalytic activity dependent on aspect ratio.36 However, the the dark for 30 min to attain an adsorption–desorption equi-
growth of ZnO nanorods required hydrothermal treatment for librium. A high-pressure mercury lamp was used as a light
6–48 h in order to obtain nanorods with different aspect ratios. source. The beaker containing MR and ZnO catalysts was then
Similarly, in another report,37 photocatalytic properties of ZnO placed under the UV light for the photodegradation of MR. At

29250 | RSC Adv., 2014, 4, 29249–29263 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances

certain time intervals, 5 mL of aliquot was sampled and aqueous solution was replaced by the 5  104 M terephthalic
centrifuged immediately to remove the photocatalyst particles. acid with a concentration of 2  103 M NaOH solution. The
Then, the supernatant solution was analyzed by monitoring the generated 2-hydroxy terephthalic acid (TAOH) was measured by
maximum absorption peak at 428 nm (lmax for MR) using a UV- a JASCO FP-6500 uorescence spectrophotometer using an
vis spectrophotometer (Agilent 8453). In order to estimate the excitation wavelength of 315 nm.
photocatalytic stability of the ZnO nanostructures, the time
courses of the photocatalytic degradation of MR using photo-
catalyst were conducted. Results and discussion
The phase purity of the as-obtained product was character-
ized by X-ray diffraction using a Phillips X'pert (MPD-3040) X-ray Fig. 1 shows typical XRD patterns of ZnO nanostructures
diffractometer with Cu Ka radiations (l ¼ 1.5406 Å) operated at prepared at different molar ratios of zinc acetate and KOH aer
Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

a voltage of 40 kV and a current of 30 mA. Field-emission 20 min of microwave irradiation. All the diffraction peaks can
scanning electron microscopy (FESEM) images were obtained be readily indexed using POWDER-X soware to the hexagonal
using a MIRA II LMH microscope. The elemental composition wurtzite ZnO phase in the standard data (JCPDS 89-1397) with
of ZnO was determined by energy dispersive X-ray spectroscopy calculated lattice parameters a and c of 0.324 and 0.520 nm,
(EDX, Inca Oxford, attached to the FESEM). Transmission respectively, and it can be clearly seen from the patterns that all
electron microscopy (TEM) micrographs, selected area electron of the samples showed a single-phase nature with a wurtzite
diffraction (SAED) pattern and high-resolution transmission structure except for the sample prepared at the 1 : 10 molar
electron microscopy (HRTEM) images were obtained using a FE- ratio. Zn(OH)2 phases are present in the sample for the 1 : 10
TEM (JEOL/JEM-2100F version) operated at 200 kV. To prepare molar ratio, which is in good agreement with the standard data
samples for TEM examination, ZnO nanostructures were JCPDS no. 38-0385 of Zn(OH)2. With the increase in molar ratio
dispersed in an ethanol solution, followed by an ultrasonic to 1 : 15 and 1 : 20, the diffraction peaks of Zn(OH)2 dis-
treatment for 10 min. A minute drop of ZnO suspension was appeared completely, and only the peaks of ZnO were materi-
cast onto a carbon-coated copper grid with subsequent drying alized. In addition, the intensity of the peaks increased with an
in air before transferring it to the microscope. In order to increase in the molar ratio. It can be clearly seen from Table 1
perform the phonon vibrational study of the ZnO nano- that with the increase in molar ratio, the FWHM corresponds to
structures, micro-Raman spectrometer (NRS-3100) was used decrease in (101) peak, which revealed the increase in the
with a 532 nm solid-state primary laser as an excitation source crystallinity. The crystallite sizes of the prepared samples esti-
in the backscattering conguration at room temperature. Room mated from X-ray line broadening of (101) peak using Scherrer's
temperature optical absorption spectra were recorded in the equation38 were found to be 25 nm, 28 nm, and 34 nm for
range of 200–800 nm using a UV-vis spectrophotometer (Agi-
lent-8453). The Brunauer–Emmett–Teller (BET) specic surface
area measurements were carried out by nitrogen adsorption
using an Autosorb®-1 (Quantachrome Instruments, Boynton
Beach, FL, USA). The photoluminescence measurements were
carried out using a luminescence spectrometer (JASCO, FP-
6500) with a xenon lamp as the excitation source at room
temperature. The excitation wavelength used was 325 nm.
Electrochemical and photoelectrochemical measurements were
performed on standard three-electrode quartz cells containing
0.2 M sodium sulphate (Na2SO4) of electrolyte solution and
measured by an electrochemical system (Versa STAT 3, Prince-
ton Research, America). The as-fabricated ZnO nanostructures
with various aspect ratios served as the working electrode.
Platinum wire was used as the counter electrode, and Ag/AgCl
(saturated with KCl) was used as the reference electrode. The
photoresponses of the photocatalysts as UV light on and off
were measured at 0.0 V. Electrochemical impedance spectra Fig. 1 XRD pattern of ZnO nanostructures for different molar ratios of
(EIS) were measured at 0.0 V over the frequency range of zinc acetate and KOH.
1–5000 Hz.
The generation of hydroxyl radicals (cOH) on the surface of
photo-illuminated ZnO nanostructures were detected by the PL Table 1 Change in FWHM ((101) peak) with the molar ratio
technique using terephthalic acid (TA) as a probe molecule.
Molar ratio FWHM (degree)
Terephthalic acid readily reacts with cOH to produce a highly
uorescent product, 2-hydroxyterephthalic acid, which exhibits 1 : 10 0.33
a PL signal at 425 nm. Experimental procedures were similar to 1 : 15 0.29
the measurement of photocatalytic activity, except that the MR 1 : 20 0.24

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 29249–29263 | 29251
View Article Online

RSC Advances Paper

samples prepared at molar ratios of 1 : 10, 1 : 15, and 1 : 20,


respectively.
Fig. 2 shows the XRD patterns of ZnO nanostructures
prepared at the 1 : 20 molar ratio of zinc acetate and KOH for
different reaction times. All the samples showed a single-phase
nature with a hexagonal wurtzite structure. No characteristic
peaks of any impurities are detected, which demonstrates that
all of the samples have high phase purity, and the sharpness of
the peaks indicates the high crystallinity of the as-prepared
ZnO. The intensity and FWHM (see Table 2) of the diffraction
peaks were found to increase and decrease, respectively with an
Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

increase in reaction time, which indicates that the increase in


Fig. 3 Raman spectra of ZnO nanostructures for different molar ratios
the reaction time improves the crystalline nature of prepared of zinc acetate and KOH.
ZnO nanostructures. The crystallite sizes of these nano-
structures estimated from X-ray line broadening of (101) peak
using Scherrer's equation38 were observed 26 nm, 29 nm, 32 nm,
and 39 nm for the samples prepared at 10 min, 20 min, 30 min, 522 cm1, which may be due to the Zn(OH)2 intermediate
and 60 min, respectively. Moreover, no remarkable shi of the phase. With the increase in molar ratio from 1 : 10 to 1 : 15 and
diffraction peaks among all of the samples reveals that lattice then 1 : 20, this mode disappeared, indicating the complete
expansion and/or shrinkage should be neglected. conversion of Zn(OH)2 into ZnO. These results are in good
Fig. 3 depicts the room-temperature Raman spectra of ZnO agreement with the XRD analysis. It can be clearly seen from
nanostructures at different molar ratios of zinc acetate and Fig. 3 that the intensity of E2(high) mode increases with the
KOH for 20 min. The Raman spectrum ZnO nanostructures increase in molar ratio, which is due to the fact that an increase
show conventional vibration modes39 of E2(high)  E2(low), in supersaturation resulted in the increase in the intensity of
A1(TO), E1(TO), E2(high) and longitudinal optical E1(LO), Raman modes.40 However, the intensity of E1(LO) mode
centered at 332 cm1, 381 cm1, 408 cm1, 439 cm1, and 580 remained the same for all samples.
cm1 respectively, as shown in Fig. 3. However, for the sample The Raman spectrum of ZnO nanostructures prepared at the
prepared at the 1 : 10 molar ratio, an extra mode appeared at 1 : 20 molar ratio for different reaction times is shown in Fig. 4.
The peak intensity related to the E2(high) mode and E1(LO)
mode increases. The E1(LO) mode is well known, and it is
related to defects such as oxygen vacancies and zinc interstitials
in ZnO;41 thus, it may be expected that the change in the E1(LO)
mode for ZnO nanostructures may be due to defects introduced
by increasing the reaction time with minimum defects exhibi-
ted in 20 min irradiated sample. The large surface area and high
surface roughness indicate pronounced enhancement of the
surface activity compared with that of the bulk crystals, and may
activate the normally forbidden E1(LO) mode. Another note-
worthy feature is that the E2(high) mode is shied to a lower
wavenumber, which may be mainly due to the distortion of the
lattice and the defects.42

Fig. 2 XRD pattern of ZnO nanostructures for different reaction times.

Table 2 Change in FWHM (101 peak) with the reaction time

Reaction time (min) FWHM (degree)

10 0.31
20 0.28
30 0.26
60 0.21 Fig. 4 Raman spectra of ZnO nanostructures for different reaction
times.

29252 | RSC Adv., 2014, 4, 29249–29263 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances

The morphological evolution of ZnO nanostructures in our directions, and each nanosheet consists of numerous ZnO
microwave-irradiation method is achieved by changing the nanorods. When the molar ratio is increased to 1 : 20, the
reaction time and molar ratio of zinc acetate and KOH. nanorods grow larger and exhibit improved crystal perfection,
Fig. 5 shows FESEM images of the basic ZnO nanostructures which is shown in Fig. 5(d). These nanorods have a regular
with different morphologies synthesized from different molar hexagonal shape and a at end with a diameter of 150 nm and
ratios of zinc acetate and KOH. The summarized morphologies a length of 2 mm (aspect ratio 14). Brush-like morphology
and reaction conditions are illustrated in Table 3. Fig. 5(a) consisting of nanorods with a diameter of 90 nm and a length
depicts morphology of ZnO without microwave process, of 4 mm (aspect ratio 20) was formed in the molar ratio of
agglomerated particles with diameter ranging from 90–150 nm 1 : 25 (Fig. 5(e)). These results demonstrate that the molar ratio
can be seen. It is obvious that the size of the nanostructures plays a crucial role in the fabrication of ZnO morphology.
increases with an increase in the molar ratio. When the molar The elemental composition of the ZnO nanostructures was
Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

ratio of the solution is low (1 : 10), the ZnO nanosheets with investigated using energy dispersive X-ray spectroscopy (EDX).
lateral dimensions of 1 mm and thicknesses of 30 nm (aspect The EDX plot as shown in Fig. 5(f) depicts peaks of Zn and O for
ratio 33) were formed (Fig. 5(b)). With an increase in the molar all of the ZnO nanostructures, which indicates that the ZnO
ratio to 1 : 15 (Fig. 5(c)), ZnO NRs (length 1 mm and diameter structures are composed of only Zn and O. No evidence of other
90–100 nm; aspect ratio 11) connected with nanosheets impurities was found, which also conrms the high purity of
(lateral dimension 1.5 mm, thickness 40–50 nm; aspect ratio the ZnO nanostructures.
37) were obtained. These nanosheets grow in different In order to further investigate the morphological evolution of
ZnO, by keeping the molar ratio xed at 1 : 20, reaction time-
dependent experiments were carried out. The ZnO sample
reacted for 10 min exhibited ower-like morphology with
several symmetric petals consisting of a number of aggregative
nanorods with lengths 2.5 mm and diameters of 250 nm
(aspect ratio 10) (Fig. 6(a)). It can be clearly seen that the
branches of the single product grow in different directions and
are composed of symmetric nanorods extending radially from
the centre. With the prolongation of the reaction time to 20
min, the shape of the ZnO nanorods changed from sword-like
sharp tips to at ends having lengths of 2 mm and diameters of
150 nm (aspect ratio 14) (Fig. 6(b)). Interestingly, the shape
of the ZnO nanorods (at end) again changed from hexagonal to
tapered with lengths and diameters of 1.5 mm, 400 nm
(aspect ratio 3.75), respectively, and started to separate from
the bunch of owers as the reaction time was increased to 30
min (Fig. 6(c)). When the reaction time is further increased to 40
min, the crystal shape developed into a prism-like morphology
having a length and diameter of 2.2 mm, 500 nm (aspect ratio
4.5), respectively (Fig. 6(d)). Finally, aer increasing the reac-
tion time to 60 min, well-dispersed at-tip semi-porous nano-
rods (pyramid-like) with lengths of 3 mm and diameters of
600 nm (aspect ratio 5) were formed as shown in Fig. 6(e).
Fig. 5 FESEM images of ZnO (a) without microwave process (b) 1 : 10, The semi-porous nature of the nanorods can be clearly seen
(c) 1 : 15, (d) 1 : 20 and (e) 1 : 25 molar ratio of zinc acetate and KOH. (f) from the high-magnication image in Fig. 6(f). This shape
EDX spectrum of ZnO nanostructures.

Table 3 Summarized morphologies and reaction conditions

KOH concentration (M) Molar ratio Zn2+ : OH Reaction time (min) Morphology Aspect ratio

0.5 1 : 10 20 Nanosheets 33
0.75 1 : 15 20 Nanosheets; nanorods 37; 11
1 1 : 20 20 Nanorods 14
1.25 1 : 25 20 Nanorods (brush-like) 20
1 1 : 20 10 Nanorods (ower-like) 10
1 1 : 20 20 Nanorods 14
1 1 : 20 30 Nanorods (ower-like) 3.75
1 1 : 20 40 Nanorods (prism-like) 4.5
1 1 : 20 60 Nanorods (pyramid-like) 5

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 29249–29263 | 29253
View Article Online

RSC Advances Paper


Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

Fig. 6 FESEM images of ZnO nanorods of (a) 10 min (b) 20 min, (c) 30 Fig. 7 TEM images of ZnO nanorods for (a) 20 min, (c) 30 min and (e)
min, (d) 40 min, and (e) 60 min and (f) high-magnification image of 60 60 min, respectively. Corresponding SAED patterns (b), (d) and (f), and
min microwave irradiation. HRTEM images are shown in the insets of (a), (c) and (e), respectively.

transformation can be explained in terms of the differences in is generally accepted that the nal morphology of ZnO crystals
the growth rates of various crystal faces. In general, the crystal is related to both their intrinsic crystal structure and external
plane with the highest growth rate disappears quickly, thus factors.44,45 The crystal formation process is divided into two
making the relative growth rate crucial to determine the stages of nucleation and growth. The overall reaction for the
morphology of the crystal. growth of ZnO nanocrystals may be expressed as follows:
Additional morphological characterization is achieved
Zn(CH3COO)2 + 2KOH / Zn(OH)2 + 2CH3COOK (1)
through the TEM, as shown in Fig. 7. Fig. 7(a) shows a typical
TEM image of a single ZnO nanorod prepared in 20 min in
Zn2+ + 2OH / Zn(OH)2 (2)
which a well-developed hexagonal phase with a at end can be
seen clearly. The TEM image of single nanorod prepared for 30 Zn(OH)2 + 2OH / [Zn(OH)4]2 (3)
min is shown in Fig. 7(c). From this TEM image, it can be seen
that the end of the ZnO nanorod is a sharp sword-like tip.
Fig. 7(e) shows the TEM image of a single ZnO semi-hollow
nanorod (pyramid-like). The atomic structures of ZnO nano-
structures were characterized by high-resolution transmission
electron microscopy (HRTEM). The HRTEM images in the
insets of Fig. 7(a), (c) and (e) show that all of the ZnO nanorods
are highly crystalline with a lattice spacing of about 0.26 nm,
corresponding to the distance between the (002) planes in the
ZnO crystal lattice. Moreover, the selected area electron
diffraction (SAED) patterns of the three ZnO structures
(Fig. 7(b), (d) and (f)) are indexed to hexagonal ZnO, indicating
that ZnO nanorods are single crystalline and have growth along
the [001] direction.
A schematic representation of the possible growth mecha-
nism of ZnO nanostructures is shown in Fig. 8. It is well known
that ZnO is a positively Zn2+-terminated (001) and negatively
O22-terminated (001) polar surfaces, which induces a net
dipole moment along the c axis.43 The growth habit of ZnO Fig. 8 A schematic representation of the possible growth mechanism
under solution conditions has been widely investigated, and it of ZnO nanostructures.

29254 | RSC Adv., 2014, 4, 29249–29263 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances

[Zn(OH)4]2 / ZnO + H2O + 2OH (4) consumed. In this case, the Ostwald ripening will become
appreciable and the thermodynamics will dominate the reac-
when the KOH is rst introduced into the Zn2+ aqueous solu- tion process.35 The morphology of the ZnO nanostructures will
tion, the solution is observed to become turbid due to the be directed to lower the system energy. The hollow structure
formation of white Zn(OH)2 colloids (reactions (1) and (2)). In may have a lower energy than that of the solid structure due to
the solution environment, a part of the Zn(OH)2 colloids the higher energy at the polar surfaces.24 This may facilitate the
dissolves into Zn2+ and OH. Large quantities of ZnO nuclei formation of the ZnO semi-hollow nanorods (pyramid-like).
form when the concentrations of Zn2+ and OH ions exceed the The optical absorption spectra were employed to study the
critical value,46 and a subsequent crystal growth process effects of shape and aspect ratio on the optical properties of
develops (reactions (3) and (4)). Based on the above discussion, ZnO nanostructures. The inset of Fig. 9 shows the absorption
for the sample without the microwave process, large quantities spectra of the ZnO nanostructures with various aspect ratios.
Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

of ZnO nuclei rst aggregate instead of exhibiting anisotropic The absorption edge of the high aspect ratio (14) ZnO nanorods
growth due to a lack of active sites around the circumference of shis to shorter wavelengths; however, for low aspect ratio
ZnO nuclei at low temperatures. Moreover, the number of nanostructures, the absorption band is shied toward higher
growth units Zn(OH)4)2 is still not enough; consequently, only wavelengths.
dispersive ZnO nanoparticles are obtained (see Fig. 8(a)). Under Because the interband transition in ZnO is the allowed direct
the microwave process, at a lower molar ratio of 1 : 10, the OH transition, the optical-absorption coefficient a of ZnO can be
ligands are fewer, and the energy of the ions is high enough to expressed by50
make the high-energy surface smoother, and thus the ions will
diffuse to low-energy surfaces.47 Under these conditions, the (ahv) ¼ A(hn  Eg)1/2, (5)
growth is determined by dynamics instead of thermodynamics.
The 2D growth mode is more preferred under this nonequilib- where A is a constant, hn is the photon energy, and Eg is the
rium process.48 This is because of the fact that the growth optical bandgap energy. Fig. 9 shows the plots of (ahn)2 versus
velocities are in the order V{010} > V{001}, and the nanosheets (hn) for the ZnO nanostructures with various aspect ratios. The
(Fig. 8(b)) were formed at the 1 : 10 molar ratio of zinc acetate band gap (Eg) values of the nanostructures were determined
and KOH. The formation of the nanosheets is attributed to the from the intercept of (ahn)2 versus (hn) curves and found to be
close surface energy of 1D/2D crystals under proper experi- 3.47 eV for ZnO nanoower (A.R  10). Interestingly, the
mental conditions.49 When the molar ratio is increased to 1 : 15, increased bandgap of 3.48 eV was observed for high aspect ratio
the quantity of Zn(OH)4)2 increases, and the anisotropic (14) nanorods. In contrast, the band gap obtained decreased to
growth is favourable, resulting in the ZnO nanorods growing 3.45 eV and 3.43 eV for nanoprisms (A.R  3.7) and nano-
along the c-axis connected with the nanosheets (Fig. 8(c)). In the pyramids (A.R  5), respectively (see Fig. 9). These results
proper molar ratio of 1 : 20, only (001)-oriented ZnO nuclei grow indicated that the aspect ratio of ZnO nanostructures could give
preferentially along the c-axis, leading to the growth of (001)- rise to a negative/positive correction to the conduction and the
oriented hexagonal ZnO nanorods (Fig. 8(d)). For a higher molar valence-band edges, leading to a bandgap tuning.
ratio of 1 : 25, the OH ligands are large enough, and they Photocatalytic properties of the ZnO nanostructures (1 : 20
promote crystal growth along the c-axis direction (Fig. 8(e)). This molar ratio, 10–60 min reaction time) with different morphol-
anisotropy growth is supported by the abovementioned crys- ogies and aspect ratios (A.R) were examined by the
tallographic habit of ZnO; hence, the 1D growth of the nanorods
in the (001) direction is preferred.
With an increase in the reaction time from 10 min to 60 min,
as shown in Fig. 8, both the diameters and lengths of the ZnO
nanostructures increase. This indicates the transversal and
longitudinal growth of the single nanostructure with time.
Therefore, the different growth pattern should be considered.
Note that the uneven width change along the length for the
single nanostructure may be related to the anisotropy of the
ZnO material. The crystal planes, which have higher surface
energy, possess faster growth velocity. Because of the different
surface energy, the growth velocity of the (0001) planes is higher
than that of the (1000) planes. As a result, when different crystal
planes grow at different velocities, those surfaces with the faster
growth velocities ((001) planes) will continually decrease their
area and the surfaces with slower growth velocities ((100)
planes) will gradually dominate the morphology of crystal,
which results in the formation of 1D structure with uneven
width along its length. During the longer reaction time in our Fig. 9The plot of (ahn)2 versus hn for ZnO nanostructures. The inset
experiments, the zinc species in the solution can be largely shows the corresponding UV-vis absorption spectra.

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 29249–29263 | 29255
View Article Online

RSC Advances Paper

decomposition of MR. The characteristic absorption of MR at


428 nm was chosen to monitor the photocatalytic degradation
process. Fig. 10 illustrates the time-dependent absorption
spectra of MR aqueous solutions during UV light irradiation in
the presence of ZnO nanorods (A.R ¼ 14). As a control, the
absorbance peak of the MR solution was monitored under two
different conditions: (i) with a photocatalyst in the dark and (ii)
without photocatalysts under UV light illumination (ESI* 1†).
The change in the absorbance peak of MR under these condi-
tions is found to be negligible, indicating that there is no loss of
MR without an irradiated photocatalyst. For ZnO nanorods (A.R
Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

¼ 14) placed in an MR solution, the maximum absorption of the


MR solution was found to decrease with illumination time and
disappeared almost completely aer irradiation for about 60
min. In particular, the photocatalytic performance of 50% was Fig. 11 Temporal evolution of MR absorption spectra with various
rapidly achieved within 10 min of photoirradiation for ZnO ZnO nanostructure photocatalysts.
nanorods, whereas aer 60 min illumination of UV light, MR
was almost completely (98%) removed over ZnO nanorods. In
order to evaluate the relationship between photocatalysis and UV-light photocatalytic properties in the degradation of MR.
aspect ratio, nanoowers, nanoprisms, and nanopyramids were The catalytic activity of the these nanostructures possesses a
used as the contrast examples. Fig. 11 shows the relative sequence of nanorods > nanoowers > nanopyramids > nano-
concentration (C/C0) of MR as a function of time for various prisms for 60 min of irradiation.
nanostructures, where C is the concentration of MR at the The kinetic behavior of these photocatalysts was further
irradiation time (t) and C0 is the concentration of the dye before studied, and the results are shown in Fig. 12. There is an
irradiation. When the suspensions were magnetically stirred in obvious linear relationship between the value of ln(C0/C) and
the dark for 30 min to ensure the establishment of an adsorp- the irradiation time. The photocatalytic process can be regarded
tion/desorption equilibrium of MR on the sample surface, only as pseudo rst-order reaction, and the rate equation is
a slight decrease in the MR solution concentration was expressed as ln(C0/C) ¼ Kt, where t is the reaction time, K is the
observed. This demonstrates that the adsorption of MR on the apparent reaction rate constant, and C0 and C are the concen-
samples is limited aer the adsorption–desorption equilibrium trations of MR at the time of 0 and t, respectively. The apparent
is reached. A control experiment revealed negligible decolor- reaction rate constant K for the degradation of MR was calcu-
ization of the dye solution treated with photocatalysts in the lated to be 6.32  102 min1, 2.91  102 min1, 1.53  102
dark. The extent of decolorization was similar to the blank min1 and 1.05  102 min1, respectively, for ZnO nanorods
sample comprising a dye solution illuminated with UV light (A.R  14), nanoowers (A.R  10), nanopyramids (A.R  5) and
without a photocatalyst (ESI* 1†). It is clear from Fig. 11 that for nanoprisms (A.R  3.7), respectively. The inset of Fig. 12 shows
all ZnO nanostructures placed in the MR solution, the the relationship between k and different aspect ratios
concentration of the MR solution is observed to decrease with
irradiation time, indicating that all of the nanostructures show

Fig. 12 Kinetic relationship of ln(C0/C) versus irradiation time for


Fig. 10 UV-visible absorbance spectra of photodegradation of MR in various ZnO nanostructure photocatalysts. The inset shows the plot of
the presence of ZnO nanostructures. rate constant versus the aspect ratio of nanostructures.

29256 | RSC Adv., 2014, 4, 29249–29263 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances

nanostructures. It is clear from inset of Fig. 12 that the reaction was almost completely degraded. Therefore, the ZnO nanorods
rate constant for high aspect ratio (14) nanorods is higher than obtained in the present work are photocatalytically more
that of lower aspect ratio nanostructures, which reveals the superior than the others.
higher photocatalytic activities of higher aspect ratio (14) ZnO Photocatalytic stability is the main limitation in the devel-
nanorods. opment of photocatalysts for organic dye degradation. To esti-
Fig. 13 shows the percentage decolorization (degradation) of mate the photocatalytic stability of the ZnO nanostructure
MR dye as a function of aspect ratio. It can be clearly seen from photocatalysts, the time courses of photocatalytic degradation
Fig. 13 that the percentage decolorization increases with the of MR using high aspect ratio (14) nanorods were conducted
increase in aspect ratio of nanostructures. The higher photo- and are shown in Fig. 14. No noticeable degradation of MR was
catalytic performances of 98% was achieved within 60 min of observed in repeated runs for the photocatalytic reaction of 60
photo-irradiation for high aspect ratio (14) nanorods, whereas min, which reveals that ZnO nanorods have good stability and
Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

only 80%, 58%, and 44% degradation efficiency of MR was reusability performance and can be a potential candidate for
obtained with nanoowers (A.R ¼ 10), nanopyramids (A.R ¼ 5), practical photocatalysis applications.
and nanoprisms (A.R ¼ 3.7), respectively. This indicates that The activation energy has been calculated from the Arrhe-
high aspect ratio ZnO nanorod photocatalysts a more superior nius equation:53
to than that of lower aspect ratio nanostructures. In addition, by
comparing the photocatalytic performance of the high aspect k ¼ AeEa/RT, (6)
ratio ZnO nanorods obtained in this study with other reported
nanostructures, ZnO nanorods showed better photocatalytic where Ea is the activation energy, k is the rate constant, A is the
properties. For example, Comparelli et al.51 showed the degra- pre-exponential factor, T is the temperature and R is the gas
dation of MR under UV irradiation in the presence of nanosized constant, which equals to 8.314 J mol1 K1, respectively. In the
ZnO and TiO2. Their results revealed that ZnO degraded 50% of photocatalytic reaction, the activation energy is the energy
MR for 140 min irradiation time, whereas about 90% of MR was required to promote photoelectrons from the photo catalyst to
degraded with TiO2. In another report, Kanjwal et al.52 pre- be trapped at surface by adsorbed oxygen molecules.54
sented hierarchical nanostructures consisting of ZnO and TiO2 The photocatalytic decolorization of MR dye using ZnO
prepared by an electro-spinning process, followed by a hydro- nanostructures with different reaction temperatures ranging
thermal technique for use as a photocatalyst for dye degrada- from 300 K to 325 K was represented by a pseudo rst order, and
tion. The results showed that the introduced ZnO–TiO2 the activation energy (Ea) and pre-exponential factor (A) were
hierarchical nanostructure can eliminate almost all the MR dye calculated by plotting ln(k) versus 1/T and found by using the
within 90 min. They also showed that less than 30% of the MR following equation:
dye was removed using pure ZnO nanoowers, even aer 180 Slope ¼ Ea/R, Intercept ¼ ln(A). The activation energy and
min. However, in the case of pristine TiO2 nanobers, up to pre-exponential factor of various ZnO nanostructures were
50% of dye was removed aer 180 min. The ZnO-doped TiO2 obtained to be approximately 6.88 kJ mol1 and 1.8  102, 8.81
nanobers showed better results than the rst two catalysts and kJ mol1 and 5.5  102, 10.41 kJ mol1 and 2.4  103, and 11.35
removed more than 80% of dye aer 3 h. kJ mol1 and 4.9  103 for nanostructures with aspect ratios of
In the present work, ZnO nanorods degraded more than 50% 14, 10, 5 and 3.7, respectively. The activation energy for high
of MR within 10 min and for 60 min of UV light irradiation, MR aspect ratio (14) ZnO nanorods is lowest (6.88 kJ mol1);
therefore, the photocatalytic reaction is fast and ends at 60 min.
These results are consistent with the trend observed in the
photocatalytic activities of these nanostructures, showing the

Fig. 13 Percentage decolorization (degradation) of MR dye as a Fig. 14 The stability of ZnO nanorods (A.R ¼ 14) for photodegradation
function of aspect ratio on various ZnO nanostructure photocatalysts. of MR.

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 29249–29263 | 29257
View Article Online

RSC Advances Paper

maximum photocatalytic activity for high aspect ratio (14) ZnO The peak at 530.2 eV can be indexed to the O2 species in ZnO.
nanorods. The higher energy peak at 531.4 eV can be assigned to oxygen
The surface composition and chemical states of ZnO nano- vacancies or defects (OV), and the peak at 532.6 eV can be
structures with various aspect ratios were investigated by XPS indexed as chemically absorbed oxygen (OC) species.55,56
analysis, and the results are shown in Fig. 15(a) and (b). Because the physically absorbed hydroxyl groups on ZnO can
Fig. 15(a) shows the high-resolution XPS spectra of the ZnO be easily removed under the ultra-high vacuum condition of the
nanostructures recorded for the Zn 2p regions. The binding XPS system, ZnO nanorods with ne surface structures will not
energy of Zn 2p3/2 and Zn 2p1/2 of all ZnO nanostructures are give signicant signals in the XPS.57,58 Therefore, the distinct
1021.47 eV and 1044.25 eV (Fig. 15 (a)), which are very well signals of the hydroxyl groups observed should be due to
matched with the standard values of ZnO.54 The binding ener- hydroxyl groups, i.e. Zn–OH and H2O, strongly bound to surface
gies of Zn 2p3/2 and Zn2p1/2 at 1021.47 eV and 1044.25 eV show defects on ZnO; i.e., the hydroxyl groups in XPS are associated
Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

an obvious shi to high energy with an increase in the aspect with surface defects, and the visible hydroxyl groups should
ratio of the ZnO nanostructures. indicate the existence of surface defects on ZnO samples.
The O 1s spectra from the nanorods are shown in Fig. 15(b). According to the XPS results, the OV (531.4 eV) and OC (532.6 eV)
An asymmetric prole of the O 1s region reveals the existence of values of O 1s are higher for the high aspect ratio nanorods
three different O species in the sample. The O 1s prole can be (Fig. 15(b)), which indicates that higher aspect ratios of ZnO
deconvoluted using a Gaussian tting technique into three nanorods have higher amounts of surface defects and visible
different peaks located at binding energy values of 530.2 eV, hydroxyl groups. Furthermore, oxygen species around 532.6 eV
531.4 eV and 532.6 eV, respectively, (Fig. 15(b)) in all samples. are caused by the surface hydroxyl group. The presence of
surface hydroxyl groups facilitates the trapping of photoin-
duced electrons and holes, thus enhancing the photocatalytic
degradation process.
Fig. 16 shows the room-temperature PL spectra of the ZnO
nanostructures with various aspect ratios. A near-band-edge
(NBE) emission at 369 nm and four visible light emission
bands: two weak blue bands at 407 and 431 nm, a weak blue-
green band at 465 nm and a strong and broad green band at
587 nm have been observed. The NBE emission is attributed to
a well-known recombination of free excitons through an
exciton–exciton collision process of the wide bandgap ZnO59
and the weak blue, blue-green and broad green light emissions
are resulted from the recombination of a photo-generated hole
with a singly ionized charge state of specic defect.60 It is
generally complicated in experiments to distinguish PL bands
caused by zinc interstitials (Zni) and oxygen vacancies (VO).
Theoretically, Kohan et al.61 and Van de Walle62 calculated the
formation energies and electronic structure of native point

Fig. 15 XPS spectra of ZnO nanostructures with various aspect ratios: Fig. 16 Room-temperature PL spectra of ZnO nanostructures with
(a) Zn-2p and (b) O-1s. various aspect ratios.

29258 | RSC Adv., 2014, 4, 29249–29263 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances

defects in ZnO. Based on their results, oxygen and zinc vacan- the large specic surface area of high aspect ratio (14) ZnO
cies are the two most common defects in ZnO. In zinc-rich nanorods can improve the contact opportunity between the dye
conditions, the oxygen vacancies (VO) have lower formation and photocatalyst, which can accelerate the rate of the photo-
energy (1.2 eV) than the zinc interstitials (Zni) and will dominate catalytic reaction. The advantage of nanopores in the high
in the defect; however, in oxygen-rich conditions, zinc vacancies aspect ratio (14) ZnO nanorods might be helpful to understand
(VZn) ought to dominate. In our microwave aqueous growth the rapid photocatalysis performance. The nanoholes in ZnO
condition, Zn source supplies from zinc salts and the O comes nanorods provide ideal channels for easy and fast diffusion of
from the OH. This aqueous growth method can be classied as the dye molecules to contact the different interplanar surfaces
Zn-rich conditions due to the high solubility of the zinc salts. of the nanoparticles, which greatly increases the chances and
Therefore, PL spectra indicate that the oxygen vacancies (VO) velocities of the encounter of the produced electron–hole pair
may possibly be responsible for the green light emission, with the dye molecules and thus enhances the photocatalytic
Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

whereas zinc interstitials (Zni) and zinc vacancies (VZn) may be activity.
excluded in the synthesized ZnO rods. These results are in good The transient photocurrent responses of a photocatalysis
agreement with the earlier report.63 We predict that the rapid may directly correlate with the recombination efficiency of the
microwave heating might enhance the defects. photogenerated carriers.66–68 Fig. 17 shows the transient
Generally, the NBE peaks directly reect the separation photocurrent density responses of ZnO nanostructure elec-
situation of the photogenerated charge carriers; i.e., the trodes under UV light irradiation. A generation of photocurrent
stronger the NBE peaks, the higher the recombination rate of with good reproducibility for all samples is observed via various
photogenerated charge carriers, which probably results in a on–off cycles of intermittent illumination. This indicates that
slower photocatalytic activity.64 High aspect ratio (14) ZnO the electrode is stable, and the photocurrent is quite reversible.
nanorods showed a decrease in the NBE intensity and hence, a The photocurrent of the ZnO nanostructures electrode is found
higher separation rate and lower recombination rate of photo- to decrease with a decrease in aspect ratio, showing maximum
induced charge carriers, which leads to higher photocatalytic current for high aspect ratio (14) nanorods. The enhancement
activity. in the photocurrent of the ZnO nanorod (14) photocatalyst
Two impurity levels, which can enhance the electron–hole indicates an enhanced photoinduced electrons and holes
pair separation rate in ZnO nanorods, are generated in the separation.
presence of oxygen vacancies and interstitial oxygen defects.65 Complementary information regarding the interface charge
As seen in Fig. 16, the green light emission intensity is higher separation efficiency of the ZnO nanostructure photocatalysts
for high aspect ratio (14) nanorods; hence, the photocatalytic was investigated by the electrochemical impedance measure-
activity of the ZnO nanorods can be enhanced with aspect ratio. ments. Fig. 18 shows the representative electrochemical
It can be concluded that abundant surface oxygen vacancies or impedance spectra (EIS), presented as a sum of real impedance
defects exist in ZnO nanorods, which may play a vital role in (Z0 ) and imaginary impedance (Z00 ) in the form of a Nyquist plot
photocatalytic activity. for ZnO nanostructures. The radius of the arc on the EIS spectra
To study the textural properties and nature of porosity of the reects the interface layer resistance occurring at the surface of
samples, BET analysis was carried out, and the parameters of electrode. The smaller arc radius indicates the higher efficiency
surface area obtained. BET analysis of the samples showed a of charge transfer.69 Fig. 18 shows that the diameters of the arc
surface area of 28.2 m2 g1, 19.4 m2 g1, 9.6 m2 g1 and 6.1 m2 radius for the high aspect ratio (14) ZnO nanorod electrodes is
g1 for the ZnO nanostructures with aspect ratio of 14, 10, 5 and
3.7, respectively. The isotherm of the ZnO nanostructures with
various aspect ratios exhibit type IV isotherms, according to
IUPAC classications, characteristic of the porous materials,
which conrms the presence of mesoporous materials with a
narrow range of uniform pores. The pore volume distribution
determined by the BJH method was found to be 0.082 cm3 g1,
0.075 cm3 g1, 0.058 cm3 g1 and 0.046 cm3 g1 with an average
uniform pore-size distribution of 9.4 nm, 7.8 nm, 4.6 nm
and 3.7 nm for ZnO nanostructures with aspect ratio of 14, 10,
5 and 3.7, respectively.
The parameters that affect the photocatalytic activity include
the catalyst, band gap, surface area, porosity, crystal structure,
crystallinity, purity, density of surface, hydroxyl groups, e and h+
migration characteristics, surface acidity and size distribution.
The maximum photocatalytic degradation of MR was ach-
ieved for high aspect ratio (14) ZnO nanorods, in which the
surface area and pore volume were maximum. The amount of Fig. 17 The transient photocurrent density responses of ZnO nano-
dye adsorbed on the surface of the catalysts is directly related to structure photocatalyst electrodes with light on–off cycles under UV
the surface area, which is available for adsorption. Therefore, light irradiation.

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 29249–29263 | 29259
View Article Online

RSC Advances Paper

Thus, it can be concluded that high aspect ratio (14) nano-


rods could produce more cOH content than that of lower aspect
ratio nanostructures, thereby possessing higher photocatalytic
performance. Therefore, cOH experiments further conrm that
cOH is actually produced and indeed participates in photo-
degradation reactions.
The photocatalytic activity of a catalytic material is inu-
enced by several factors such as carrier recombination, size of
the particles, surface area, surface acidity, and presence of a
higher number of hydroxyl groups. In this work, ZnO nanorods
with high aspect ratios show a higher percentage of degradation
Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

as compared to the low aspect ratio nanostructures as well as


the previously reported work. The enhancement of photo-
Fig. 18 Nyquist plots for ZnO nanostructures with various aspect catalytic activity was most likely due to the relative increase of
ratios in light on/off cycles under UV light irradiation. active morphological surfaces because of the increased surface-
to-volume ratio, as well as the low recombination rate of the
electron–hole pairs generated during optical exposure, owing to
smaller than that of the lower aspect ratio nanostructure elec- largely available surface states. This is mainly due to changes in
trodes. These observations indicated that both the solid-state space-charge regions because this is well-dened along the
interface layer resistance and the charge-transfer resistance on longitudinal direction of the nanorods.72,73 The enhanced
the surface of ZnO have signicantly decreased for high aspect oxygen adsorption in the high aspect ratio ZnO nanorods is
ratio (14) nanorods. This suggests that a high aspect ratio (14) expected to be very effective in accepting photogenerated holes.
ZnO nanorods acts as nanoscale electrodes and accelerates This leads to a reduction in the recombination probability of
charge transfer, decreases the intragranular resistance, inhibits photogenerated carriers, which in turn forms oxygen species74
recombination between charge carriers, and enhances photo- and thus results in high photocatalytic activity. The superior
catalytic efficiency.70 This result is in good agreement with the photocatalytic performance of ZnO nanorods can be extended
transient photocurrent responses of these photocatalysts. to remove other organic pollutants from wastewater.
Fig. 19 presents the PL spectral changes observed during the Fig. 20 shows the schematic diagram of a UV light photo-
UV illumination of the different photocatalysts in the aqueous catalytic mechanism with ZnO nanorods. Photocatalytic
basic solution of terephthalic acid. Usually, PL intensity at degradation was generally operated by the action of hydroxyl
about 425 nm is proportional to the amount of produced radicals formed during the reaction.74 The photocatalytic
hydroxyl radicals.71 In general, hydroxyl radicals (cOH) has been mechanism is proposed in the following manner: when the ZnO
regarded as the dominant active species responsible for the nanostructures were illuminated with light, electrons were
photocatalytic oxidation reactions. Obviously, it can be seen excited from the valence band of ZnO to the conduction band,
that the intensity of the generated 2-hydroxy-terephthalic acid leaving a hole in the valence band. The hydroxyl groups present
(TAOH) of high aspect ratio (14) nanorods was stronger than on the surface of the ZnO react with the photogenerated hole to
that of lower aspect ratio nanostructures, indicating that the produce hydroxyl radicals. Similarly, peroxide (O2) is formed
formation rate of cOH radicals on high aspect ratio (14) nano- when the dissolved oxygen interacts with photogenerated elec-
rods was larger than that of the others. trons. This peroxide takes one proton to yield a superoxide
(HO2) followed by the formation of hydrogen peroxide (H2O2).
A hydroxyl radical was also produced by the attack of a photo-
generated electron to the hydrogen peroxide. These reactive
radicals and intermediate species react with dye and degrade
them into non-toxic organic compounds.

ZnO + hn / ZnO + e + h+ (7)

Fig. 19 Detection of cOH radical by fluorescence method using ter- Fig. 20 Schematic representation of UV light photocatalytic process
ephthalic acid as trapping reagent. in the presence of ZnO nanostructures.

29260 | RSC Adv., 2014, 4, 29249–29263 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances

OH + h+ / OH (8) nanocrystal, which indicates that photogenerated electrons can
ow in the direction of the crystal length. Increased delocal-
O2 + e / O2 (9) ization of electrons at 1-D nanostructured crystals can lead to a
remarkable decrease in charge carrier recombination proba-
O2 + H+ / HO2 (10) bility. Consequently, larger numbers of electrons and holes
exist on the active sites of the nanocrystal surface, resulting in
2HO2 / H2O2 + O2 (11)
higher activity.75 Therefore, in the present work, the higher
aspect ratio (14) ZnO nanorods exhibit higher photocatalytic
H2O2 + e / OH + OH (12)
activity.

Mashkour et al.75 reported the effect of H2O2 addition on the Conclusions


Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

reaction rate constant of photocatalytic degradation. Their


results showed that the apparent rate constant increased with In summary, highly crystalline ZnO with different morphologies
increasing the concentration of H2O2 due to the increase in the have been successfully prepared by a convenient microwave-
formation of cOH radicals, in addition to inhibiting the assisted solution route. XRD, HRTEM and SAED measurements
recombination of electron–hole pairs.76,77 showed that all of the ZnO nanostructures are of a hexagonal
In the present work, the reaction rate is higher for high phase structure. The exception of secondary phase formation
aspect ratio nanorods, whereas the reaction rate is lower for the (Zn(OH)2) was observed for the sample prepared using the 1 : 10
low aspect ratio. The reaction rate is dependent on the gener- molar ratio, which nally converted to ZnO upon increasing the
ation of hydroxyl radicals. In general, a hydroxyl radical (cOH) molar ratio. PL and XPS studies revealed the existence of defects
has been regarded as the dominant active species responsible in the ZnO nanorods. FESEM and TEM images revealed that the
for the photocatalytic oxidation reactions. Obviously, it can be morphology of ZnO can be effectively controlled as sheet-like,
seen that the intensity of the generated 2-hydroxy-terephthalic rod-like, brush-like, ower-like, prism-like, and pyramid-like,
acid (TAOH) of high aspect ratio (14) nanorods was stronger which was ascribed to the process parameter-dependent tuning
than that of the lower aspect ratio nanostructures, indicating of the nanostructures. The aspect ratio continued to change as a
that the formation rate of cOH radicals on high aspect ratio (14) consequence of changes in the reaction time and molar ratio,
nanorods was larger than that of the others. Thus, it can be showing maximum (14) for the nanorods prepared at the 1 : 20
concluded that high aspect ratio (14) nanorods could produce molar ratio for 20 min. It is suggested that the zinc species and
more cOH content than that of lower aspect ratio nano- environment, including OH and K+ as well as the reaction
structures, thereby possessing higher photocatalytic perfor- time, are the crucial factors for the morphologies of obtained
mance. Therefore, it is logical to state that, in the present work, ZnO. Such knowledge would allow for quick synthesis of ZnO
H2O2 formed in the photocatalytic reaction, which in turn with anticipated morphology simply by selecting the appro-
formed cOH radicals and degraded the organic dye, which is in priate molar ratio and reaction time. The photocatalytic prop-
good agreement with the earlier report.75 erties of ZnO nanostructures with different aspect ratios toward
It is well known that the photocatalytic process is determined MR under UV light irradiation were found to be attractive. The
by the separation of electron–hole pairs,78 and the separation of observed rates of degradation were found to be proportional to
photogenerated carriers could be due to the crystalline phase, the rates of formation of active hydroxyl radicals studied by the
the aspect ratio of the nanorods, and the electronic structure of emission spectra of 2-hydroxy-terephthalic acid. EIS study
the facets and defects.79 Among these factors, surface defects shows that the high aspect ratio (14) ZnO nanorods minimized
have to be considered for most cases, except for perfect single charge recombination and effectively enhanced the photo-
crystals. Because the current ZnO nanorods were synthesized via catalytic activity. The enhanced percentage degradation of MR
a microwave irradiation method at relatively low temperatures, with ZnO nanorods was found to be 98% within 60 min of UV
the diffusion lengths of atoms during the decomposition reac- light illumination, which was due to the higher aspect ratio.
tion of the precursor is relatively short. It is therefore reasonable This enhancement in photocatalytic activity may be attributed
to believe that a large number of surface defects exist in the nal to the easy separation of photogenerated charge carriers in the
products, which result in visible emissions (400–700 nm). higher aspect ratio nanorods, which resulted to the enhanced
Generally, the recombination probability of the photo- oxygen chemisorptions. For bridging between the laboratory
generated carriers would be high, and the separation of the and industry, considering the excellent photocatalytic perfor-
electron–hole is the rate-limiting step for the photocatalytic mance and facile preparation method, the prepared ZnO
process.78 The aspect ratio of nanostructures is expected to nanorods are believed to have potential applications in the eld
enhance the separation of photogenerated carriers. Therefore, of environmental remediation and photocatalysis.
decreased recombination centers with a lesser number of
interparticle junctions and high electron delocalization in high Acknowledgements
aspect ratio nanorods lead to higher efficiency.80 For 1-D
nanostructured crystals, the space-charge region is well con- This work was supported by the National Research Foundation
structed along the longitudinal direction of the ZnO of Korea (NRF) grant funded by the Korean Government (MSIP)
(no. 2011-0030058). This work was also supported by the MSIP

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 29249–29263 | 29261
View Article Online

RSC Advances Paper

(Ministry of Science, ICT & Future Planning), Korea. Under the 29 L. Vayssieres, Adv. Mater., 2003, 15, 464.
ITRC (Information Technology Research Centre) support 30 B. Liu and H. C. Zeng, J. Am. Chem. Soc., 2003, 125, 4430.
program supervised by the NIPA (National IT Industry Promo- 31 H. Zhang, X. Y. Ma, J. Xu, J. J. Niu and D. R. Yang,
tion Agency) (NIPA-2014-H0301-14-1016). Nanotechnology, 2003, 14, 423.
32 J. Zhang, L. D. Sun, J. L. Yin, H. L. Su, C. S. Liao and
Notes and references C. H. Yan, Chem. Mater., 2002, 14, 4172.
33 C. Pacholski, A. Kornowski and H. Weller, Angew. Chem., Int.
1 N. Saito, H. Haneda, T. Sekiguchi, N. Ohashi, I. Sakaguchi Ed., 2002, 41, 1188.
and K. Koumoto, Adv. Mater., 2002, 14, 418. 34 H. Zhang, D. Yang, X. Ma, Y. Ji, J. Xu and D. Que,
2 D. M. Bagnall, Y. F. Chen, Z. Zhu, T. Yao, S. Koyama and Nanotechnology, 2004, 15, 622.
T. Goto, Appl. Phys. Lett., 1997, 70, 2230. 35 B. Liu and H. Zeng, Langmuir, 2004, 20, 4196.
Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

3 F. Quaranta, A. Valentini, F. R. Rizzi and G. Casamassima, 36 A. Leelavathi, G. Madrasa and N. Ravishankar, Phys. Chem.
J. Appl. Phys., 1993, 74, 244. Chem. Phys., 2013, 15, 10795.
4 N. Golego, S. A. Studenikina and M. Cocivera, J. Electrochem. 37 L. Zhang, H. Yang, J. Ma, L. L. X. Wang, L. Zhang and
Soc., 2000, 147, 1592. S. T. X. Wang, Appl. Phys. A, 2010, 100, 1061.
5 D. S. Ginley and C. Bright, Mater. Res. Bull., 2000, 25, 15. 38 B. D. Cullity and S. R. Stock, Elements of X-ray Diffraction,
6 Z. L. Wang and J. H. Song, Science, 2006, 312, 242. Prentice Hall, New Jersey, 2001.
7 F. Lu, W. P. Cai and Y. G. Zhang, Adv. Funct. Mater., 2008, 18, 39 R. Cusco, E. Alarcon-Llado, J. Ibanez, L. Artus, J. Jimenez,
1047. B. G. Wang and M. J. Callahan, Phys. Rev. B: Condens.
8 Q. Xu, H. Schmidt, S. Zhou, K. Potzger, M. Helm, H. Hochmuth, Matter Mater. Phys., 2007, 75, 165202.
M. Lorenz, A. Setzer, P. Esquinazi, C. Meinecke and 40 G. D. Nagy and E. J. Casey, Zinc–Silver Oxide Batteries, Wiley,
M. Grundmann, Appl. Phys. Lett., 2008, 92, 082508. 1971, pp. 29–36.
9 F. Ahmed, S. Kumar, N. Arshi, M. S. Anwar, B. H. Koo and 41 J. J. Wu and S. C. Liu, J. Phys. Chem. B, 2002, 106, 9546.
C. G. Lee, Funct. Mater. Lett., 2011, 4, 1. 42 Y. Q. Huang, M. D. Liu, Z. Li, Y. K. Zeng and S. B. Liu, Mater.
10 Q. Xiang and J. Yu, J. Phys. Chem. Lett., 2013, 4, 753. Sci. Eng., B, 2003, 97, 111.
11 W. J. Ong, L. L. Tan, S. P. Chai, S. T. Yonga and 43 Z. L. Wang, X. Y. Kong and J. M. Zou, Phys. Rev. Lett., 2003,
A. R. Mohamed, Nanoscale, 2014, 6, 1946. 91, 185502.
12 F. Fresno, R. Portela, S. Suárez and J. M. Coronado, J. Mater. 44 J. Zhang, L. D. Sun, C. S. Liao and C. H. Yang, Chem.
Chem. A, 2014, 2, 2863. Commun., 2002, 3, 262.
13 J. Ran, J. Zhang, J. Yu, M. Jaroniec and S. Z. Qiao, Chem. Soc. 45 X. Y. Zhang, J. Y. Dai, H. C. Ong, N. Wang, H. L. Chan and
Rev., 2014, DOI: 10.1039/C3CS60425J. C. L. Choy, Chem. Phys. Lett., 2004, 393, 17.
14 A. L. Manoj and W. A. Daoud, RSC Adv., 2013, 3, 4130. 46 H. Zhang, D. R. Yang, Y. J. Ji, X. Y. Ma, J. Xu and D. L. Que,
15 X. Pan, M. Q. Yang, X. Fu, N. Zhanga and Y. J. Xu, Nanoscale, J. Phys. Chem. B, 2004, 108, 3955.
2013, 5, 3601. 47 H. J. Fan, A. S. Barnard and M. Zacharias, Appl. Phys. Lett.,
16 W. J. Ong, L. L. Tan, S. P. Chai, S. T. Yong and 2007, 90, 143116.
A. R. Mohamed, ChemSusChem, 2014, 7, 690. 48 B. Q. Cao, W. P. Cai, H. B. Zeng and G. T. Duan, J. Appl. Phys.,
17 I. Poulios, D. Makri and X. Prohaska, Global NEST J., 1999, 1, 55. 2006, 99, 073516.
18 E. R. Carraway, A. J. Hoffman and M. R. Hoffmann, Environ. 49 B. Illy, B. A. Shollock, J. L. M. Driscoll and M. P. Ryan,
Sci. Technol., 1994, 28, 786. Nanotechnology, 2005, 16, 320.
19 D. S. Bohle and C. J. Spina, J. Am. Chem. Soc., 2009, 131, 4397. 50 J. I. Pankove, Optical Processes in Semiconductors, Prentice-
20 J. C. Wang, P. Liu, X. Z. Fu, Z. H. Li, W. Han and X. X. Wang, Hall Inc., Englewoord Cliffs, NJ, 1971.
Langmuir, 2009, 25, 1218. 51 R. Comparelli, P. D. Cozzoli, M. L. Curri, A. Agostiano,
21 S. Baruah, R. F. Raque and J. Dutta, NANO: Brief Reports and G. Mascolo and G. Lovecchio, Water Sci. Technol., 2004, 49,
Reviews, 2008, 3, 8. 183.
22 M. H. Huang, S. Mao, H. Feick, H. Yan, Y. Wu, H. Kind, 52 M. A. Kanjwal, N. A. M. Barakat, F. A. Sheikh, S. J. Park and
E. Weber, R. Russo and P. Yang, Science, 2001, 292, 1897. H. Y. Kim, Macromol. Res., 2010, 18, 233.
23 Y. Yan, P. Liu, J. G. Wen, B. To and M. M. Al-Jassim, J. Phys. 53 F. H. Hussein, A. H. Ibrahim and S. A. Shakir, Zanco, 1988,
Chem. B, 2003, 107, 9701. 2(4), 83.
24 H. Yu, Z. Zhang, M. Han, X. Hao and F. Zhu, J. Am. Chem. 54 S. A. Ansari, M. M. Khan, S. Kalathil, A. Nisar, J. Lee and
Soc., 2005, 127, 2378. M. H. Cho, Nanoscale, 2013, 5, 9238.
25 Y. Peng, A. W. Xu, B. Deng, M. Antoniettia and H. Colfen, 55 J. H. Zheng, Q. Jiang and J. S. Lian, Appl. Surf. Sci., 2011, 257,
J. Phys. Chem. B, 2006, 110, 2988. 5083.
26 Z. W. Pan, Z. R. Dai and Z. L. Wang, Science, 2001, 291, 1947. 56 X. G. Han, et al., J. Phys. Chem. C, 2008, 113, 584.
27 H. Saitoh, M. Saitoh, N. Tanaka, Y. Ueda and S. Ohshio, Jpn. 57 A. Iwabuchi, C. K. Choo and K. Tanaka, J. Phys. Chem. B,
J. Appl. Phys., 1999, 38, 6873. 2004, 108, 10863.
28 W. I. Park, D. H. Kim, S. W. Jung and G. C. Yia, Appl. Phys. 58 R. Wang, N. Sakai, A. Fujishima, T. Watanabe and
Lett., 2002, 80, 4232. K. Hashimoto, J. Phys. Chem. B, 1999, 103, 2188.

29262 | RSC Adv., 2014, 4, 29249–29263 This journal is © The Royal Society of Chemistry 2014
View Article Online

Paper RSC Advances

59 J. J. Wu and S. C. Liu, Adv. Mater., 2002, 14, 215. 70 X. Bai, L. Wang, R. Zong, Y. Lv, Y. Sun and Y. Zhu, Langmuir,
60 K. Vanheusden, W. L. Warren, C. H. Seager, D. R. Tallant, 2013, 29, 3097.
J. A. Voigt and B. E. Gnade, J. Appl. Phys., 1996, 79, 7983. 71 T. Hirakawa and Y. Nosaka, Langmuir, 2002, 18, 3247.
61 A. F. Kohan, G. Ceder, D. Morgan and C. G. Van de Walle, 72 D. Li, R. Shi, C. Pan, Y. Zhu and H. Zhao, CrystEngComm,
Phys. Rev. B: Condens. Matter Mater. Phys., 2000, 61, 15019. 2011, 13, 4695.
62 C. G. Van de Walle, Phys. B, 2001, 899, 308. 73 W. H. Leng, Z. Zhang, J. Q. Zhang and C. N. Cao, J. Phys.
63 C. H. Hung and W. T. Whang, Mater. Chem. Phys., 2003, 82, Chem. B, 2005, 109, 15008.
705. 74 L. S. Zhang, K. H. Wong, D. Q. Zhang, C. Hu, J. C. Yu,
64 J. Liqiang, Q. Yichun, W. Baiqi, L. Shudan, J. Baojiang, C. Y. Chan and P. K. Wong, Environ. Sci. Technol., 2009, 43,
Y. Libin, F. Wei, F. Honggang and S. Jiazhong, Sol. Energy 7883.
Mater. Sol. Cells, 2006, 90, 1773. 75 M. S. Mashkour, A. F. Alkaim, L. M. Ahmed and
Published on 09 June 2014. Downloaded by University of Zurich on 10/07/2014 15:39:08.

65 J. Wang, Q. Li, L. Peng and M. Malac, Microsc. Microanal., F. H. Hussein, Int. J. Chem. Sci., 2011, 9, 969.
2009, 15, 1218. 76 U. I. Gaya, A. Abdullah, Z. Zainal and M. Z. Hussein, Int.
66 H. Liu, S. Cheng, M. Wu, H. Wu, J. Zhang, W. Li and C. Cao, J. Chem., 2010, 2, 180.
J. Phys. Chem. A, 2000, 104, 7016. 77 F. H. Hussein and R. Rudham, J. Chem. Soc., Faraday Trans.
67 W. H. Leng, Z. Zhang, J. Q. Zhang and C. N. Cao, J. Phys. 1, 1987, 83, 1631.
Chem. B, 2005, 109, 15008–15023. 78 J. M. Herrmann, Top. Catal., 2005, 34, 49.
68 H. Park and W. Choi, J. Phys. Chem. B, 2003, 107, 3885–3890. 79 J. Chang and E. R. Waclawik, CrystEngComm, 2012, 14, 4041.
69 W. H. Leng, Z. Zhang, J. Q. Zhang and C. N. Cao, J. Phys. 80 H. J. Yun, H. Lee, J. B. Joo, W. Kim and J. Yi, J. Phys. Chem. C,
Chem. B, 2005, 109, 15008. 2009, 113, 3050.

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 29249–29263 | 29263

S-ar putea să vă placă și