Sie sind auf Seite 1von 16

698 © IWA Publishing 2016 Water Science & Technology | 74.

3 | 2016

Facile one pot synthesis of zinc oxide nanorods and


statistical evaluation for photocatalytic degradation of a
diazo dye
Suvanka Dutta, Ananya Ghosh, Humayun Kabir and Rajnarayan Saha

ABSTRACT
Suvanka Dutta
In the present work zinc oxide nanorods (ZNRs) have been synthesized to estimate its photocatalytic
Ananya Ghosh
degradation potential on an industrially used diazo dye and optimization of the total treatment Rajnarayan Saha (corresponding author)
Department of Chemistry,
process has been designed. Response surface methodology (RSM) has been used to model the National Institute of Technology Durgapur,
Durgapur 713209,
operational parameters for this photocatalytic degradation. The crystallite size (101 plane) of the India
E-mail: rnsahanitd@gmail.com
synthesized ZNR has been found to be 20.99 nm having a band gap energy of 3.45 eV. At elevated pH,
Humayun Kabir
the rate of degradation of the photocatalyst was found to be higher than that of acidic pH. The
Department of Earth and Environmental Science,
independent variables of the model are time (9.6–122 min), pH (2–12.2), catalyst dose (0.2–0.4 g/L) National Institute of Technology Durgapur,
Durgapur 713209,
and dye concentration (88–512 mg/L). It was seen that the degradation efficiency was significantly India

affected by the initial dye concentration and the pH, the optimal values of the parameters being a pH
of 10.67, an initial concentration of 150 mg/L and ZnO dose of 0.37 g/L, the time taken being
88.52 min. The actual degradation efficiency of the dye reached 96.9% at optimized condition, which
is quite close to the predicted value of 98.07%.
Key words | kinetics, optimization, reactive dye, RSM, ZnO nanorod

INTRODUCTION

Textile industries are an ever increasing business not only regulated in most countries because they are toxic, persist-
providing occupation to over millions of individuals but ent, bio-accumulative and hormone disruptive and can
also causing one of the biggest water footprints on the even cause cancer (Absalan et al. ; Dasgupta et al.
earth, and dyeing poses an especially serious drawback ; Dutta et al. a, b).
which adds to the already prevailing problem of water pol- Photocatalysts offer great potential for environmental
lution of the world. One of the significant classes of purification and converting photon energy into chemical
synthetic organic dye is the azo dye, which is characterized energy (Xu et al. ; Lee et al. ). In the last decade het-
by the presence of one or more azo bonds (N¼N). These erogeneous photocatalysts such as ZnO, TiO2, WO3, ZnS,
dyes have become the most important commercial colorants ZrO2, SnO2 have been gaining huge importance due their
because of their wide color range, good fastness properties reusability, thermal stability and presence of multiple active
and color density. These dyes represent approximately sites. These photocatalysts are generally semiconductive in
50% of the worldwide production and are extensively used nature. The electronic structure of the semiconductors
in a number of industries with the textile industry being made them act as sensitizers for light-induced redox pro-
the largest consumer (Sanmuga et al. ). Due to their cesses, which is characterized by a filled valence band and
increasing popularity and unrestricted usage, azo dyes an empty conduction band (Hu et al. ; Srivastava ;
started to adversely affect the water resources, aquatic Ludi & Niederberger ; Park et al. ; Zhang et al.
organisms and the entire ecosystem integrity (Kumar et al. ). Zinc oxide is a II–VI semiconductor with direct band
). The wastewater released by the dyeing industries con- gap energy of about 3.4 eV and a large exciton binding
tains significant amounts of hazardous chemicals such as energy of about 60 meV which makes it thermally stable at
tributyltin, pentabromodiphenyl ether, phthalates, perfluor- room temperature (Wu & Liu ; Nair et al. ). More-
ooctane sulfonate and aniline that are banned or strictly over, being a II–VI semiconductor ZnO has a tendency to
doi: 10.2166/wst.2016.248
699 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

show polytypism in two forms: one is cubic zinc blende form due to the treatment with the deviation owing to random
and another is hexagonal wurtzite form (Hughes & Alivisatos errors, linked to the measurements of produced responses.
). Isotropic growth in zinc blende structure generates four In this work zinc oxide nanorods (ZNRs) were syn-
equivalent faces. Whereas anisotropic growth of wurtzite is thesized by simple alkaline hydrolysis method. A few
quite complicated and crystal growth along higher surface research works on factorial modeling of dye and different
energy plane (001) is greatly favored. Higher growth rate organics degradation by zinc oxide nanoparticles have
along the c-axis results in quite fast disappearance of basal been reported earlier (Soltani et al. ; Lee & Hamid
plane and it leads to formation of nanorods (Manna et al. ), but a comparatively less amount of work has involved
; Peng et al. ). Nanorod structures of ZnO are poten- ZNRs and statistical modeling. Herein, the work covers a
tially ideal functional components for nanoscale electronics simplistic one pot synthesis of ZNR requiring just two
and optoelectronics. Furthermore, nanorod structures of easily available chemicals, a kinetics study utilizing the
semiconductors have a tendency to increase the band gap nanorods and lastly the factorial design and process optim-
energy comparative to nanoparticles, due to quantum con- ization of photodegradation of the chosen dye.
finement effects, which is in turn beneficial for
photocatalytic application as it decrease the recombination
rate between holes and electron pair. Moreover, higher MATERIALS
surface to volume ratio in nanorods, comparative to nanopar-
ticles, enhances the surface adsorption, thereby helping to Zinc nitrate hexahydrate (ZnNO3.6H2O, 99%) and sodium
augment photocatalysis rate. ZnO nanoparticles are used hydroxide (NaOH) used for the synthesis of ZNR, were pur-
for multidisciplinary purposes including solar cells (Chakra- chased from Merck (India). The chemicals were used as-
barti & Dutta ), Gas sensors (Mohaghegh et al. ), received without further purification. Deionized water
optoelectronic devices (Farzana & Meenakshi ), and (18.2 mΩ) was used throughout the studies to prepare the
paints (Amorim et al. ) and as a photocatalyst for stock solutions and working solutions of dye. Details of
microbial degradation (Puvaneswari et al. ). The reason the model dye are given in Table 1.
behind the widespread popularity of ZnO as an efficient
photocatalyst is that it can absorbs over a larger segment Synthesis of ZNRs
of the solar spectrum, it is easily available, has low cost
and is relatively nontoxic (Klemola et al. ; Kant ; For the preparation of the ZNRs the following process was
Gnanaprakasam et al. , Thangavel et al. ). adopted: 0.7 M ZnNO3 solution was taken as the precursor.
Previously used conventional water treatment processes
were generally carried out using the one-factor-at-a-time Table 1 | Particulars of the selected dye
(OFAT) approach. This a method for designing experiments
involving the testing of factors, or causes, one at a time instead Name of dye Reactive Black 31

of all concurrently. In this recent era, the OFAT method is get- Molecular formula C29H18CuN6Na4O17S4
ting replaced by chemometric methods like response surface Molecular weight 1,006.25
methodology (RSM) founded on statistical design of exper- Chemical structure
iments (Daniel ; Bezerra et al. ; Kasiri & Khataee
; Vaez et al. ). In order to overcome the problem
faced in the application of OFAT approach, the optimization
of analytical methods is carried out using multivariate statisti-
cal techniques like RSM. It is a collection of mathematical and
statistical procedures to analyze the influence of independent
variables (inputs) on a specific dependent variable (response)
(Ahmadi et al. ; Dutta et al. a, b). RSM utilizes
an experimental strategy such as the central composite
Maximum 549 nm
design (CCD) for the purpose of fitting a model by least squares absorbance (λmax)
technique (Bezerra et al. ; Khayet et al. ). The data
CAS no. 12731-63-4
obtained is analyzed by analysis of variance (ANOVA). The
Supplier Colour Chem, Clariant, Mumbai, India
fundamental concept of ANOVA is to compare the variation
700 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

NaOH (1 M) was taken as the source of the hydroxide ions 1,000 mg/L and was prepared by dissolving the dye in deio-
required for the hydrolysis of the salt. A hundred millilitres nized water. The dye concentration in the stock solution, the
of this solution was added to a round-bottom flask and intermediate dye solutions and the treated samples were all
placed in an oil bath, maintained at a temperature of measured using a UV-Vis spectrophotometer (1601 Shi-
madzu) at a λmax value of 549 nm wavelength for RB 31 dye.
W
110 C by a thermostat; the whole setup was then mounted
on a magnetic stirrer and rotated vigorously. This helped The pH was adjusted by 0.1 N and 1 N HCl and 0.1 N and
homogenization of the reactant and even heat distribution 1 N NaOH.
within the synthesis vessel. During the synthesis, after the
basic solution gained some momentum, 50 mL of zinc Experimental setup
nitrate solution was added drop-wise, keeping the tempera-
ture constant at 110 C. Zinc hydroxide is formed at first
W
Photo-oxidation of RB 31 dye was conducted in a batch
and the solution turns turbid. Eventually with continuous process. This experimental unit is configured with an annu-
heating and stirring zinc hydroxide converts into zinc lar cylindrical glass reactor (Figure 1) with a quartz tube at
oxide. The reaction pathway is given below the middle of the reactor to accommodate the UV light
source of 125 watt; the jacket was provided for cooling
ZnNO3 þ NaOH ! ZnðOHÞ2 þNaþ þ NO
3 (1) purposes. Cooling medium (water) was continuously circu-
lated throughout the reaction time to prevent the UV bulb
ZnðOHÞ2 ! ZnO ↓ þH2 O (2) from fusing. The reactor is provided with two ports: one
serves as inlet for reactants, and the second one for collec-
The entire reaction was carried out for a span of 2 hours; tion of treated samples. The reactor is in full contact with
after that the final product was filtered with a membrane air and homogenization is done by a magnetic stirrer.
filter (0.2 μm pore size), washed with deionized water first The reactor is covered with a wooden box to prevent loss
and then with absolute ethanol several times, and finally of UV light. All the photo-induced experiments were per-
the residue was dried at a temperature of 60 C in a
W
formed in this cylindrical photo-reactor whose total
vacuum oven. volume is 500 mL. Samples were withdrawn from the
sampling port after certain fixed time intermissions and
were filtered (pore size <0.22 μm) to discard the suspended
Particle characterization
ZnO particles present in the solution. Then the sample
solutions were analyzed for remaining dye. Before illumi-
Field emission scanning electron microscopy (FE-SEM) analy-
nation, the suspension of RB 31 dye along with the
sis of the obtained ZNR was carried out by a Bruker XFlash
photocatalyst were magnetically stirred in the dark for
6160. That the nanorods were crystalline in nature was
30 min, after dispersing in an ultrasonic bath for 5 min,
proved by the X-ray diffraction (XRD) data, done by
to ensure the establishment of an adsorption–desorption
PANalytical XPERT PRO, the scan range of the sample
equilibrium between ZnO and RB 31. Mineralization
being 35–100 (2θ) with the scan speed being 1.2 min1.
W W

efficiency (ME) is calculated by determining the total


UV–visible (UV-Vis) absorption spectra of dispersed ZnO
organic carbon (TOC) at different time and the required
nanorods solution were recorded by a UV-Vis spectropho-
equation is given below ME (%)
tometer (1601 Shimadzu), for the evaluation of the band gap
energy of the sample. Fourier transform infrared (FT-IR) spec-
ðTOC0  TOCt Þ
tra of the sample were recorded by a Thermo Nicolet iS10 ME ð%Þ¼ ×100 (3)
TOC0
spectrometer using KBr pellets in the range 4,000–400 cm1.
The pH measurements of the dye solutions were carried out Here TOC0 and TOCt represent TOC of the dye solution
by an Orion star A214 Thermo Scientific pH meter. before photocatalysis process starts (t ¼ 0) and after time t
respectively.
Preparation of synthetic wastewater
RSM in experimental design
A synthetic wastewater solution was prepared using Reactive
Black 31 (RB 31) dye, which is widely used in the textile indus- RSM has been widely applied for the optimization of several
tries in India. The concentration of the stock dye solution was chemical and physical methods (Cao et al. ). In this
701 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

Figure 1 | Design of the photo-reactor setup.

work, the optimization of process variables was done by Table 2 | Coded variables for RSM

using RSM combined with the most commonly used form


Range and level
of factorial experimental design, CCD. In the RSM
method, several influencing variables were studied along Variable Factors 1 þ1 α þα

with the responses by changing them concurrently and car- X1 Time (min) 26.00 105.00 9.64 121.36
rying out a limited number of experiments. The outcome of X2 pH 3.50 10.70 2.01 12.19
the experiments were analyzed using Design Expert7 soft-
X3 Dose (g/L) 0.23 0.36 0.20 0.41
ware, and the regression model was suggested. For the
X4 Dye concentration 150.00 450.00 87.87 512.13
evaluation purpose, four independent operating parameters, (mg/L)
influencing the removal efficiency of the dye, were chosen:
time (X1), pH (X2), dose of ZnO (X3) and dye concentration
(X4), with the percentage degradation efficiency of RB 31 as response surface delineating the relationship between the
the response (dependent variable). Additional parameters response of interest, that is, percentage dye removal
like stirring rate, temperature and light intensity were efficiency (Y1), and the independent process factors, could
taken as constant. Table 2 shows the ranges and level of not be adequately modeled using a first order model.’ The
independent variables. approximation of the real response function therefore
In their study involving photocatalytic degradation, necessitated the usage of a second order polynomial
Chakraborty et al. () stated ‘the curvature of the true model, i.e. a quadratic model using a standard RSM
702 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

design. Hence, CCD was nominated as a proficient statisti- analysis, the main and interaction effects, the contour plot
cal tool for fitting second order empirical response surface and ANOVA were performed. The experimental outcomes
models. CCD was chosen to investigate the joint effect of of CCD were fitted to the second order (quadratic) poly-
the four independent variables by a set of 30 experiments, nomial equation as follows:
including six replications at the center points. The results
of experiments of dye removal efficiency percentage under Y% ¼ b0 þ b1 x1 þ b2 x2 þ b3 x3 þ b4 x4 þ b12 x1 x2
several experimental circumstances are shown in Table 3. þ b13 x1 x3 þ b14 x1 x4 þ b23 x2 x3 þ b24 x2 x4 þ b34 x3 x4
To examine the accuracy of the fitted model, a series of stat- þ b11 x21 þ b22 x22 þ b33 x23 þ b44 x24 (4)
istical analyses such as the normal plot, the residual

Table 3 | Matrix of CCD with its predictive value and experimental results (RE%, percentage removal efficiency)

Square root transformed


Experimental plans RE%

Run Time pH Dose g/L Concentration of dye mg/L Actual RE% Actual Predicted

1 26.00 10.70 0.23 450.00 18.67 4.32 4.31


2 65.50 7.10 0.30 300.00 31 5.57 5.56
3 65.50 7.10 0.30 300.00 31.3 5.59 5.57
4 65.50 7.10 0.30 87.87 79.21 8.90 8.92
5 65.50 7.10 0.40 300.00 34.7 5.89 5.85
6 121.36 7.10 0.30 300.00 35.76 5.98 5.96
7 105.00 3.50 0.23 150.00 69.4 8.33 8.31
8 65.50 7.10 0.30 300.00 31.2 5.59 5.57
9 26.00 10.70 0.38 450.00 24.21 4.92 4.96
10 105.00 10.70 0.38 150.00 99.16 9.96 9.99
11 26.00 3.50 0.38 150.00 44.82 6.69 6.71
12 105.00 3.50 0.23 450.00 28.65 5.35 5.36
13 65.50 2.01 0.30 300.00 43.03 6.56 6.61
14 105.00 10.70 0.23 150.00 85.54 9.07 9.06
15 105.00 10.70 0.23 450.00 46.66 6.83 6.84
16 26.00 10.70 0.23 150.00 36.87 6.07 6.09
17 65.50 7.10 0.30 300.00 30.8 5.55 5.57
18 65.50 7.10 0.30 300.00 30.7 5.54 5.57
19 9.64 7.10 0.30 300.00 6.24 2.50 2.51
20 65.50 7.10 0.30 512.13 25.91 5.09 5.07
21 105.00 10.70 0.38 450.00 47.21 6.87 6.88
22 26.00 3.50 0.23 150.00 29.16 5.40 5.36
23 105.00 3.50 0.38 150.00 82.26 9.07 9.06
24 26.00 3.50 0.23 450.00 9.2 3.03 3.02
25 65.50 7.10 0.20 300.00 23.23 4.82 4.86
26 65.50 7.10 0.30 300.00 31 5.57 5.57
27 26.00 10.70 0.38 150.00 55.95 7.48 7.45
28 105.00 3.50 0.38 450.00 29.01 5.39 5.39
29 65.50 12.19 0.30 300.00 67.69 8.23 8.18
30 26.00 3.50 0.38 450.00 13.6 3.69 3.67
703 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

where Y represents the response variable (decolorization sample (ICSD file No: 98000-9340). Diffraction patterns cor-
efficiency), bi, bij, and bii are the regression coefficients for responding to impurities are found to be absent. This proves
linear and quadratic effects and the coefficients of the inter- that pure ZnO nanoparticles were synthesized. The crystallite
action parameters, respectively, and xi are the independent size (τ) of the as prepared nano powder can be calculated by
variables studied. Decolorization efficiency was calculated using Scherrer’s formula.
by the following formula:
τ ¼ 0:94λ=β cosθ (6)
C0  Ct
Y% ¼ × 100 (5)
C0 where λ is the wavelength of X-rays used (1.54060 Å), β is the
full width at half maximum and θ is the angle of diffraction.
where C0 is initial dye concentration and Ct is the dye con- Calculated crystallite size of prepared sample is found to be
centration at time t. around 20.99 nm (101 plane), which is in the order of nano
size.

Microstructural study by FE-SEM analysis


RESULTS AND DISCUSSION
The FE-SEM image of the nanorods is shown in Figure 3(a). The
Characterization of zinc oxide nanoparticles
FE-SEM image discloses that the diameters of the ZnO nano-
rods are in the range 25–50 nm, and length is in the range of
Phase and structural analysis by XRD 300–400 nm. In a near similar process of synthesis, spherical
particles have been reported previously (Lanje et al. ; Rad-
XRD diffraction pattern of the as-synthesized zinc oxide nano
zimska & Jesionowski ), whereas the little difference in our
powder is shown in Figure 2. The observed diffraction peaks process resulted in wurtzite nanorods. Energy-dispersive X-ray
of ZnO at 2θ ¼ 32.12, 34.76, 36.64, 47.8, 57, 63.12, 66.64,
W
(EDX) analysis (Figure 3(b)) reveals the presence of five
68.2, 69.9, 72.96 and 77.28 are associated with representa-
elements: Zn, O, C, Nb and N. Presence of Nb and N can be
tive planes (100), (002), (101), (102), (110), (103), (200), due to instrumental error. Carbon source is mainly the grid
(112), (201), (004) and (202). The presence of several peaks
base of the SEM analysis and a little residual washing ethanol.
in the XRD patterns of as-grown sample revealed that the
Atomic ratio between Zn and O is less than 1 (0.644) which
nanostructures were polycrystalline in nature. All the diffrac-
again corroborates the fact that the presence of washing ethanol
tions can be assigned to the standard pattern for the pure contributed to the excess oxygen. Major fraction of atoms are
hexagonal phase (wurtzite structure) of ZnO. The values
zinc and oxygen (<81%); therefore we can confirm the for-
have agreed well with the standard card for ZnO powder
mation of ZnO nanoparticles from EDX data. Verification of
formation of ZnO particle by EDX analysis was also reported
by Gnanaprakasam et al. ().

FT-IR spectroscopy

The FT-IR spectrum of the as-prepared ZNRs shows


(Figure 4) the absorption at 901 cm1, which is due to the
development of tetrahedral coordination of ZnO. Metal
oxides in general give absorption bands in their fingerprint
region, i.e. below 1,000 cm1, occurring due to interatomic
vibrations (Rao ; Nadeem & Ahmed ). The peak at
925 cm1 is due to overtone of water molecules (Mohan
et al. ). The band below 700 cm1 is associated with
metal oxide (ZnO) bond. In the IR region, ZnO generally
shows discrete absorption bands about wavenumbers of
464 cm1. The spectra should show a representative ZnO
Figure 2 | Powder XRD of nano ZnO particles at 2θ. absorption at 564 cm1, blue-shifted to 512 cm1 and there
704 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

Figure 3 | (a) FE-SEM analysis of ZnO nanorods; (b) EDX data of ZNRs.

will be an additional maximum at wavenumbers smaller than Optical study by UV-Vis spectroscopy
500 cm1, which cannot be identified by our spectrometer
(Wu et al. a, b). The fundamental mode of vibration The UV optical spectra of ZnO nanorods, suspended
at 3,399 cm1 corresponds to O-H stretching frequency, in deionized water, were obtained in the range of
which is due to the presence of the water molecules bound 200–800 nm by a 1601 Shimadzu spectrophotometer. Opti-
to the surface of the prepared sample. The peak at cal absorption spectrum of ZNRs is shown in Figure 5 (see
2,847 cm1 corresponds to C-H stretching frequency. The the inset in Figure 5). ZNR shows a small but sharp absorp-
presence of residual ethanol from the washing process can tion peak at 356 nm in the UV-Vis region, which is blue-
be the only reason behind this peak. The peak at shifted when compared to bulk ZnO particles having
1,635 cm1 corresponds to an overtone of water molecule. absorption at about 380 nm (Yong-hong et al. ; Wu
705 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

portion of the plot of (αhυ)2 vs. hυ with the abscissa axis


(Cao et al. ), where hυ and α accounts for the photon
energy and the absorption coefficient, respectively. α is
derived from the Lambert–Beer relation, α ¼ 2.303 A/d,
where d is the path length of the quartz cuvette and A is
the UV-Vis absorbance. The band gap has been calculated
to be about 3.45 eV. Quantum size effect could be the
reason behind increased band gap (Devaraj et al. ).
Enhanced band gap of ZNR compared with the zinc oxide
nanoparticles is useful for the photocatalytic process as it
will help to reduce the recombination rate.

Preliminary experiments

The time-dependent total scan in the absorbance spectra of


Figure 4 | FT-IR plot of nano ZnO nanorods. the dye solution throughout the reaction was also examined
with initial dye concentration of 100 mg/L and catalyst load-
ing of 0.2 g/L at pH 7. As seen in Figure 6, the reductions in
five absorbance peaks at 257, 322, 413, 472 and 571 nm indi-
cate the degradation of the dye molecules. As no new
absorption peaks appeared during the reaction, it supports
the hypothesis that, if any intermediate product/s are
formed during the dye degradation, they are also efficaciously
degraded towards a full mineralization process; around 90%
reduction in the absorbance under 549 nm is also a strong
sign of degradation of intermediates.
In order to assess the substantial role of photocatalysts in
the photodegradation process, photolysis and catalyst adsorp-
tion in the dark were performed. As shown in Figure 7, 31%

Figure 5 | UV-Vis absorption spectrum (inset) and band gap energy determination of the
prepared ZnO nanoparticle.

et al. a, b). This blue shift may be attributed to the


crystallite size of the synthesized nanorods, which is below
30 nm. The absorption in the near-UV region is because of
near band edge emission occurring due to exciton recombi-
nation. The band gap energy of ZNRs has been calculated
from the UV-Vis absorption spectrum recorded in their col-
loidal state in absorbance mode (Figure 5) using the Tauc
plot. The required equation is αhυ ¼ A (hυ  Eg)n’, where
A is a constant, hν is photon energy, Eg is the allowed
energy gap, n ¼ ½ for allowed direct transition and n ¼ 2
for allowed indirect transition. The direct band gap (Eg), rep- |
Figure 6 UV-Vis spectra of photocatalytic degradation of RB 31 dye under different time
resents the junction point of the extrapolated straight interval, ZnO loading ¼ 0.2 g/L, dye concentration ¼ 100 mg/L, pH ¼ 7.
706 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

Kinetics study

Before executing an experimental design, it is essential to


study the extent of photocatalytic degradation in RB 31 deco-
lorization. The rate of degradation of RB 31 in the presence of
ZNR, under UV illumination, was estimated at different pH
values with fixed dye concentration of 100 mg/L and catalyst
dosage of 0.2 g/L. The effect of pH on the dye photodegrada-
tion method is pretty challenging since it has various roles.
Due to the amphoteric nature of most semiconductor
oxides, pH of the dispersion becomes the significant par-
ameter dominating the rate of reaction taking place on the
semiconductor surface, as it affects the surface-charge proper-
ties of the photocatalysts. Moreover, industrial effluents are of
various pH ranges. Hence, the effect of pH on the rate of
Figure 7 | Photocatalytic degradation of RB 31 dye under various conditions, ZnO load-
ing ¼ 0.2 g/L, dye concentration ¼ 100 mg/L, pH ¼ 7. degradation should be examined. Here, the experiments
were performed in the pH range of 3–12. Decay models of kin-
etics in photocatalysis usually follow pseudo zero order,
of the dye solution was photolyzed by only UV irradiation. pseudo first order and second order models. Among these
The sorption of dye solution by ZNR was trivial (9.4%). the most common forms are zero and first order reaction.
Although 97.97% of the dye solution was degraded in the Surface saturation on photocatalyst results in the zero
presence of ZNR under UV illumination within 2 hours. order kinetics model (Gaya ). If a change in concen-
This indicated the synergistic effect between ZNR photocata- tration does not affect the reaction rate, it is denoted as a
lyst and UV light for the photocatalytic degradation. zero order reaction. A first order reaction rate depends on
Figure 8 presents the mineralization efficiency of TOC a single reactant. A bi-component reaction or a second
of RB 31 dye photocatalysis by ZnO nanorods with order reaction converts into a pseudo first order reaction
120 min of time. Results proved that 76.6% of TOC was when the concentration of one reactant remains more or
removed within this time limit. Therefore it can be inferred less the same throughout the reaction (Auguliaro et al.
that ZnO nanorods are very effective for RB31 dye degra- ). In the case of the photocatalysis process of a com-
dation as well as for the mineralization process. pound, the two reacting species (the compound and
photocatalyst) result in a single compound P (assuming):

CþD!P (7)

The rate equation can be written as

d ½C 
 ¼ k ½C ½D  (8)
dt

where k represents the observed first order reaction rate


constant (min1) and t is time. Now, if the concentration
of D remains the same throughout the reaction or
[D] > >> [C], or in other words [D] becomes constant:

d ½C 
 ¼ k 0 ½C  (9)
dt

Figure 8 | Plot of mineralization efficiency determined by removal of TOC of RB 31 dye by


Here, k0 is the rate constant of the pseudo first order
ZNRs. reaction. Now, if within the limit of t ¼ 0 to t ¼ t,
707 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

concentration of C changes from C0 to Ct, upon integration necessitating a power transformation. The Box–Cox plot
within this limit has been utilized as a tool to describe the best suitable
power transformation applicable to response data. Most
ð Ct ðt
of the data transformations can be designated by the
dC=C ¼ k0 dt (10)
C0 0 power function, σ ¼ f (μα), where sigma (σ) is the standard
deviation, mu (μ) is the mean and alpha (α) is the power.
LnCt =C0 ¼ k0 t (11) Lambda (λ) is (1  α) in all these cases. If the standard
deviation connected with an observation is proportionate
to the mean elevated to the power, then transforming the
Now it can be written as
observation by 1  α (or l) as the exponent value gives a
scale satisfying the equal variance requisite of the statisti-
0
Cdye ¼ Cdye × ekt (12) cal model. The bottom point on the Box–Cox plot
indicates the value of λ that results in the minimum
C0dye is the initial concentration (mg/L) of dye, and Cdye residual sum of squares in the transformed model. The
is the concentration (mg/L) of dye at time t. Finally k indi- potential for progress is greatest when the range of the
cates the empirical rate constant (min1). The constant k, maximum to minimum response value is greater than
for the various reactions, was obtained by linear regression. three. According to the responses of the current work
R2 denotes the regression coefficient (1.0 denotes ideal first the best power transformation can be affected by an expo-
order kinetics; the larger the deviation from unity, the less nent of 0.53 which indeed recommends the square root
the equation fits the first order kinetics). Table 4 summarizes (0.5) transformation (Figure 9).
rate constants from all the diverse experiments executed. The experimental matrix design and the responses cen-
The rate of photocatalytic reaction increased with increasing tered on experiments projected by CCD for the
pH. The decrease in the photocatalytic degradation at acidic degradation of RB 31 are given in Table 3. On the foun-
pH may be because of dissolution of ZnO at low pH. At dation of the experimental design presented in Table 3, a
higher pH range, there is enough hydroxyl (OH) anions, quadratic response surface model was designated by the
which assist photogeneration of hydroxyl radicals (OH•). software prompted sequential model sum of squares studies,
so as to aptly estimate the response surface function for the
responses of concern. Similarly, the final regression model
Development of regression model equation and its
validation for optimization

Model fitting and statistical analysis

Development of the model initially necessitates accurate


power transformation of the responses. In this work, the
response ranges from 6.24% to 99.16%, and the ratio of
maximum to minimum is 15.891 (>10), thereby

Table 4 | Rate constants and regression coefficient of dye degradation process at differ-
ent pH

pH K R2

3 0.0204 0.944
5 0.0251 0.996
7 0.0331 0.983
9 0.0273 0.959
11 0.0311 0.931
12 0.0472 0.972
Figure 9 | Box–Cox plot for power transformation.
708 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

equation, which is fitted to the experimental data is pre- residual or real error) of 5,191.91 implies that the model is
sented in terms of the actual variables as: significant. Now, there could be only 0.01% chance that a
model F-value this large could arise due to any noise.
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Removal efficiency ¼ þ4:00726 þ 0:10840 time Values of ‘Prob > F’ less than 0.0500 show that model
 0:94603 pH þ 25:88086 dose  0:02321 concentration terms are significant. It can be inferred that A, B, C, D,
  AB, AC, AD, BD, CD, A2, B2, C2, and D2 are significant
þ 3:48039 × 10−4 × time × pH  0:053511 × time × dose
  model terms. Furthermore, values greater than 0.1000
 2:61821 × 10−5 × time × concentration specify the model terms are not significant. The ‘Lack of
   
þ 8:35066 × 10−3 × pH × dose þ 2:56923 × 10−4 Fit F-value’ of 4.09 implies that the lack of fit is not signifi-
× pH × concentration  0:016337 × dose × cocentration cant relative to the pure error. There is a 6.68% chance
  that a ‘Lack of Fit F-value’ could happen due to noise. The
 4:2721 × 10−4 × time2 þ 0:070272 × pH2
‘Predicted R-Squared’ value of 0.9989 is in rational agree-
 
 20:920019 × dose2 þ 3:16220 × 10−5 × concentration2 ment with the ‘Adjusted R-Squared’ value of 0.9996,
(13) endorsing decent predictability of the model. ‘Adequate Pre-
cision’ measures the signal to noise ratio. A ratio greater
Further, the statistical suitability of the quadratic than 4 is desirable. Here the ratio of 295.375 indicates an
response surface model was investigated by Fisher’s statisti- adequate signal. This model can be used to navigate the
cal test (F-test) for ANOVA (Markova-Deneva ). The design space. The statistical significance of the model was
ANOVA outcomes for the recommended RSM model are furthermore verified through plotting the experimental
summarized in Table 5. The model F-value (ratio of mean observations against statistically anticipated responses for
square because of regression to the mean square due to dye removal efficiency (Figure 10(a)). Data points narrowly

Table 5 | ANOVA summary for the empirical quadratic response surface model

P-value
Source Sum of squares Degrees of freedom Mean square F-value Prob > F

Model 93.24 14 6.66 5,191.91 <0.0001


A: time 29.68 1 29.68 23,136.28 <0.0001
B: pH 6.17 1 6.17 4,807.38 <0.0001
C: dose 2.45 1 2.45 1,906.31 <0.0001
D: concentration 37.10 1 37.10 28,922.02 <0.0001
AB 0.039 1 0.039 30.55 <0.0001
AC 0.37 1 0.37 288.90 <0.0001
AD 0.39 1 0.39 300.18 <0.0001
BD 0.31 1 0.31 240.10 <0.0001
CD 0.50 1 0.50 388.32 <0.0001
A2 4.15 1 4.15 3,232.90 <0.0001
2
B 7.74 1 7.74 6,035.08 <0.0001
C2 0.11 1 0.11 85.58 <0.0001
2
D 4.72 1 4.72 3,683.39 <0.0001
Residual 0.019 15 1.283 × 103
Lack of fit 0.017 10 1.714 × 103 4.09 0.0668
3
Pure error 2.097 × 10 5 4.194 × 104
Total 93.26 29
Standard deviation ¼ 0.036 Mean ¼ 6.13 Coefficient of variation ¼ 0.58
2 2
Predicted R ¼ 0.9989 Adjusted R ¼ 0.9996 Adequate precision ¼ 295.375
709 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

fairly evident that the obtained data points steadily appear


on a straight trend line, showing that there is no evident
dispersion.

Response surface analysis

The performance of the quadratic model estimated


response surface with the RB 31 dye removal efficacy as
the response, when exposed to the fluctuations of the
model environment, was graphically presented by means
of the three-dimensional response surface plots shown in
Figure 11. These plots can be utilized to evaluate any vicis-
situdes in the response surface founded on a polynomial
function. In this process, two parameters are held as con-
stant and two other parameters are varied within the
experimental ranges (Soltani et al. ). The variation in
dye removal efficiency (%) as functions of ratios of pH to
time concentration to time, and dose to concentration
have been validated by the response surface plots displayed
in Figure 11(a)–11(c), respectively.
The results in Figure 11(a) establish that the rate of
decolorization increases with increase in pH, presenting
maximum rate of removal at pH 10. Using an azo dye
namely Acid Brown 14, Sakthivel et al. () and Akyol
et al. () reported an analogous trend with ZnO, observ-
ing that the acid–base property of the metal oxide surfaces
can have substantial consequences on their photocatalytic
activity. The reported pHpzc for ZnO is 8.5–9.0 (Zhou &
Keller ), this means that above this pH, ZnO surface
gets negatively charged as the surface will be covered
with adsorbed OH ions. These accumulated OH ions
on the particle surfaces and in the reaction medium
favors the formation of OH• radical, which is now widely
accepted as the principal oxidizing species responsible for
the dye decolorization process. Lizama et al. () also
found out that pH 11 was the optimum value for reactive
Blue-19 dye removal, in presence of ZnO and TiO2 as
the photocatalysts. They commented that textile processes
Figure 10 | (a) Predicted vs actual plot of dye degradation efficiency. (b) Normal prob-
ability plot of the residuals.
using reactive dyes produce effluents with high pH, suit-
able for the use of ZnO as a catalyst. The effect of initial
dye concentration on decolorization efficiency is depicted
fell at around the diagonal of the plot, which further vali- by Figure 11(b) at an initial pH 7.1 and catalyst dosage
dated the statistical consistency of this predictive response of 0.3 g/L. The plot shows that decreasing initial dye con-
surface model. Moreover, the appropriateness of the centration from 450 mg/L to 150 mg/L results in an
model can be estimated from the residuals, calculated by increase in the decolorization efficiency. As the initial
defining the variance between the experimental and pre- dye concentration decreases, more hydroxyl radicals are
dicted decolorization efficiency. Figure 10(b) illustrates the available to degrade the dye, such that the decolorization
internally studentized residuals and normal probability efficiency increases. Furthermore, it can be stated that
plot for decolorization efficiency. From the figure it is OH• concentration decreases as a result of decreasing
710 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

path length of the photons entering the dye solution to


excite the surface of the photocatalyst, as the initial dye
concentration increases (Zhu et al. ). Figure 11(c)
depicts the influence of ZnO loading and initial RB 31
dye concentration on the degradation efficiency while
keeping pH at 7.10 and time duration of 65.5 min. As illus-
trated in the plot, the degradation percentage increased
with increasing catalyst dosage. This is based on the fact
that as the dye molecules get adsorbed onto the catalyst
surface, competitive adsorption of the OH ions on the cat-
alyst active sites declines leading to the reduction in OH•
radical generation, which is the prime oxidant essential
for a high degradation efficiency (Zhu et al. ). On the
other hand, considering the Lambert–Beer law, as the
initial dye concentration increases, penetration of photons
into the solution decreases, which this results in lower
photon absorption by the catalyst particles and leads to
consequent lower photocatalytic reaction rate.

Process optimization and model validation and


confirmation

Process optimization was executed on the basis of the desir-


ability function to define the optimal conditions for the
degradation of RB 31 dye. Design Expert7 software was
used here to recognize the specific points that maximize the
desirability function (Table 6). The program utilizes five
options as an objective to build desirability indices: none,
maximum, minimum, target, and within range. The norms
for all variables in correspondence with percentage removal
efficiency are shown in Table 6. The importance gives added
weight to upper or lower bounds or accentuates target values.
As a higher degradation efficiency is the main goal for such
studies, an ‘importance’ value of 5 has been denoted as the
maximum goal. On the basis of the settings and limits desig-
nated above, the optimum situations for maximum
degradation efficacy (98.07%) were observed at pH value of
10.67, an initial dye concentration of 150 mg/L, ZnO dose
of 0.37 g/L and time duration of 88.52 min (Figure 12).
To test the suitability of the model in predicting the
maximum removal percentage of RB 31, an authentication
experiment was performed in a 500 mL reactor at the opti-
mum conditions. A mean of maximum degradation of
96.9% was obtained from three replicate tests, as shown in
Figure 11 | (a) Effects of pH and time on removal efficiency at 0.3 g/L catalyst dosing and Table 7. A good agreement between the predicted value
300 mg/L dye concentration; (b) effects of concentration and time on removal
and the experimental value confirms the validity of the
efficiency at 0.3 g/L catalyst dosing and pH of 7.1; (c) effects of dose and
concentration on removal efficiency at pH of 7.1 and reaction time of model in simulating the photocatalytic degradation of
65.5 min.
RB 31.
711 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

Table 6 | Optimization of individual responses to obtain overall desirability response

Name Goal Lower limit Upper limit Lower weight Upper weight Importance

Time (min) minimize 40.00 70.00 1 1 3


pH maximize 5.00 10.70 1 1 3
Catalyst dose (g/L) minimize 0.23 0.38 1 1 3
Dye concentration (mg/L) maximize 150.00 450.00 1 1 3
Removal efficiency (%) maximize 6.24 99.16 1 1 5

Figure 12 | Desirability ramp for numerical optimization.

CONCLUSION

Table 7 | Decolorization efficiency at optimum values of the process parameters In the present study ZnO nanorods were synthesized by simple
alkaline hydrolysis method at elevated temperature. Morpho-
Optimum value Experimental
Variables predicted value logical and optical analysis revealed that at (101) plane the
nanorods have crystallite diameter of 20.99 nm and the band
Time (min) 88.52 89.00
gap energy is 3.45 eV. Photocatalytic degradation of the
pH 10.70 10.60
model reactive dye RB 31 by ZnO nanorod under UV
ZnO dose (g/L) 0.37 0.37
irradiation in an aqueous medium is investigated here in the
Dye concentration 150.00 150.00
context of RSM. Kinetics study at different pH presented an
(mg/L)
increasing trend in the rate of reaction with respect to increas-
Removal efficiency (%) 98.07 96.90
ing pH. The model showed that the decolorization efficiency
712 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

increases with increasing pH and doses of ZnO nanorods, Chakrabarti, S. & Dutta, B. K.  Photocatalytic degradation of
whereas the removal efficiency is inversely related to initial model textile dyes in wastewater using ZnO as
semiconductor catalyst. J. Hazard. Mater. 112 (3), 269–278.
dye concentration. Optimization of the process by RSM
Chakraborty, S., Dasgupta, J., Farooq, U., Sikder, J., Drioli, E. &
based on CCD showed 96.9% removal efficiency under opti- Curcio, S.  Experimental analysis, modeling and
mized operational conditions. This is quite consistent with optimization of chromium (VI) removal from aqueous
the software-predicted 98.07% removal efficiency. The results solutions by polymer-enhanced ultrafiltration. J. Membr. Sci.
obtained indicated applicability of the present process for effi- 456, 139–154.
cient removal of dye from aqueous media. Daniel, C.  ‘One-at-a-Time Plans’. J. Am. Stat. Assoc. 68, 353–360.
Dasgupta, J., Singh, M., Sikder, J., Padarthi, V., Chakraborty, S. &
Curcio, S.  Response surface-optimized removal of
Reactive Red 120 dye from its aqueous solutions using
polyethyleneimine enhanced ultrafiltration. Ecotoxicol.
ACKNOWLEDGEMENTS
Environ. Saf. 121, 271–278.
Devaraj, R., Karthikeyan, K. & Jeyasubramanian, K.  Synthesis
Authors are grateful to the infrastructure provided by the and properties of ZnO nanorods by modified Pechini
National Institute of Technology (NIT) Durgapur. Assistance process. Appl. Nanosci. 3, 37–40.
for TOC analysis was given by Gargi Biswas and Dr Susmita Dutta, S., Ghosh, A., Satpathi, S. & Saha, R. a Modified
Dutta from Chemical Engineering Department, NIT Durga- synthesis of nanoscale zero-valent iron and its ultrasound-
assisted reactivity study on a reactive dye and textile industry
pur. Professor D. Mal from Department of Chemistry of IIT
effluents. Desalin. Water Treat. DOI: 10.1080/19443994.
Kharagpur provided assistance for FE-SEM and EDX analy- 2015.1096833.
sis. Indrani Thakur, Ram Chandra Maji and Jhilli Dasgupta, Dutta, S., Ghosh, A., Moi, C. S. & Saha, R. b Application of
fellow research scholars, helped us throughout the work. response surface methodology for optimization of reactive
azo dye degradation process by Fenton’s oxidation. Int. J.
Environ. Sci. Dev. 6 (11), 818–823.
Farzana, M. H. & Meenakshi, S.  Visible light-driven photoactivity
REFERENCES of zinc oxide impregnated chitosan beads for the detoxification of
textile dyes. Appl. Catal. A (General) 503, 124–134.
Absalan, G., Asadi, M., Kamran, S., Sheikhian, L. & Goltz, D. M. Gaya, U. I.  Heterogeneous Photocatalysis using Inorganic
 Removal of reactive red-120 and 4-(2-pyridylazo) Semiconductor Solids. Springer, Amsterdam.
resorcinol from aqueous samples by Fe3O4 magnetic Gnanaprakasam, A., Sivakumar, V. M., Sivayogavalli, P. L. &
nanoparticles using ionic liquid as modifier. J. Hazard. Mater. Thirumarimuruganv, M.  Characterization of TiO2 and ZnO
192, 476–484. nanoparticles and their applications in photocatalytic
Ahmadi, M., Vahabzadeh, F., Bonakdarpour, B., Mofarrah, E. & degradation of azo dyes. Ecotoxicol. Environ. Saf. 121, 121–125.
Mehranian, M.  Application of the central composite Hu, Y., Jiang, Z., Xu, C., Mei, T., Guo, J. & White, T. 
design and response surface methodology to the advanced Monodisperse ZnO nanodots: synthesis, characterization, and
treatment of olive oil processing wastewater using Fenton’s optoelectronic properties. J. Phys. Chem. C 111 (27), 9757–9760.
peroxidation. J. Hazard. Mater. 123 (1–3), 187–195. Hughes, S. M. & Alivisatos, P. A.  Anisotropic formation and
Akoyl, A., Yatmaz, H. C. & Bayramoglu, M.  Photocatalytic distribution of stacking faults in II–VI semiconductor
decolorization of Remazol Red RR in aqueous ZnO nanorods. Nano Lett. 13 (1), 106–110.
suspensions. Appl. Catal. B. 54, 19–24. Kant, R.  Textile dyeing industry an environmental hazard.
Amorim, C. C., Leão, M. M. D., Moreira, R. F. P. M., Fabris, J. D. & Nat. Sci. 4 (1), 22–26.
Henriques, A. B.  Performance of blast furnace waste for Kasiri, M. B. & Khataee, A. R.  Photooxidative decolorization of
azo dye degradation through photo-Fenton-like processes. two organic dyes with different chemical structures by UV/H2O2
Chem. Eng. J. 224, 59–66. process: experimental design. Desalination 270 (1–3), 151–159.
Augugliaro, V., Coluccia, S., Loddo, V., Marchese, L., Martra, G., Khayet, M., Zahrim, A. Y. & Hilal, N.  Modelling and
Palmisano, L. & Schiavello, M.  Photocatalytic oxidation of optimization of coagulation of highly concentrated industrial
gaseous toluene on anatase TiO2 catalyst: mechanistic aspects grade leather dye by response surface methodology. Chem.
and FT-IR investigation. Appl. Catal. B: Environ. 20 (1), 15–27. Eng. J. 167 (1), 77–83.
Bezerra, M. A., Santelli, R. E., Oliveira, E. P., Villar, L. S. & Escaleira, Klemola, K., Pearson, J. & Lindstrom-Seppä, P.  Evaluating
L. A.  Response surface methodology (RSM) as a tool for the toxicity of reactive dyes and dyed fabrics with the HaCaT
optimization in analytical chemistry. Talanta 76, 965–977. cytotoxicity test. AUTEX Res. J. 7 (3), 217–223.
Cao, J., Wu, Y., Jin, Y., Yilihan, P. & Huang, W.  Response Kumar, A. N., Reddy, C. N. & Mohan, S. V.  Biomineralization
surface methodology approach for optimization of the of azo dye bearing wastewater in periodic discontinuous
removal of chromium (VI) by NH2-MCM-41. J. Taiwan Inst. batch reactor: effect of microaerophilic conditions on
Chem. Eng. 45, 860–868. treatment efficiency. Bioresour. Technol. 188, 56–64.
713 S. Dutta et al. | Statistical evaluation for photocatalytic dye degradation by ZNR Water Science & Technology | 74.3 | 2016

Lanje, A. S., Sharma, S. J., Ningthoujam, R. S., Ahn, J. S. & Pode, Sanmuga, P. E., Senthamil, S. P. & Umayal, A. N. 
R. B.  Low temperature dielectric studies of zinc oxide Biodegradation studies on dye effluent and selective remazol
(ZnO) nanoparticles prepared by precipitation method. Adv. dyes by indigenous bacterial species through spectral
Powder Technol. 24, 331–335. characterization. Desalin. Water Treat. 55 (1), 241–251.
Lee, K. M. & Hamid, S. B. A.  Simple response surface Soltani, R. D. C., Rezaee, A., Godini, H., Khataee, A. R. &
methodology: investigation on advance photocatalytic Hasanbeiki, A.  Photoelectrochemical treatment of
oxidation of 4-chlorophenoxyacetic acid using UV-active ammonium using seawater as a natural supporting
ZnO photocatalyst. Materials 8, 339–354. electrolyte. Chem. Ecol. 29, 72–85.
Lee, K. M., Lai, C. W., Ngai, K. S. & Juan, J. S.  Recent Soltani, R. D. C., Rezaee, A., Khataee, A. R. & Safari, M. 
developments of zinc oxide based photocatalyst in water Photocatalytic process by immobilized carbon black/ZnO
treatment technology: A review. Water Res. 88, 428–448. nanocomposite for dye removal from aqueous medium:
Lizama, C., Freer, J., Baeza, J. & Mansilla, H. D.  Optimized optimization by response surface methodology. J. Ind. Eng.
photodegradation of Reactive Blue 19 on TiO2 and ZnO Chem. 20, 1861–1868.
suspensions. Catal. Today 76, 235–246. Srivastava, V. C.  Photocatalytic oxidation of dye bearing
Ludi, B. & Niederberger, M.  Zinc oxide nanoparticles: wastewater by iron doped zinc oxide. Ind. Eng. Chem. Res.
chemical mechanisms and classical and non-classical 52, 17790–17799.
crystallization. Dalton Trans. 42, 12554–12568. Thangavel, S., Thangavel, S., Raghavan, N., Krishnamoorthy, K. &
Manna, L., Scher, E. C. & Alivisatos, A. P.  Synthesis of soluble Venugopal, G.  Visible-light driven photocatalytic
and processable rod-, arrow-, teardrop-, and tetrapod-shaped degradation of methylene-violet by rGO/Fe3O4/ZnO ternary
CdSe nanocrystals. J. Am. Chem. Soc. 122 (51), 12700–12706. nanohybrid structures. J. Alloys Compounds 665, 107–112.
Markova-Deneva, I.  Infrared spectroscopy investigation of Vaez, M., Moghaddam, A. Z. & Alijani, S.  Optimization and
metallic nanoparticles based on copper, cobalt, and nickel modeling of photocatalytic degradation of azo dye using a
synthesized through borohydride reduction method. J. Univ. response surface methodology (RSM) based on the central
Chem. Technol. Metall. 45, 351–378. composite design with immobilized titania nanoparticles.
Mohaghegh, N., Tasviri, M., Rahimi, E. & Gholami, M. R.  Ind. Eng. Chem. Res. 51 (11), 4199–4207.
Nano sized ZnO composites: preparation, characterization Wu, J. J. & Liu, S. C.  Low-temperature growth of well-aligned
and application as photocatalysts for degradation of AB92 ZnO nanorods by chemical vapor deposition. Adv. Mater. 14,
azo dye. Mater. Sci. Semiconduct. Proc. 21, 167–179. 215–218.
Mohan, R., Krishnamoorthy, K. & Kim, S.-J.  Diameter Wu, L., Wu, Y., Pan, X. & Kong, F. a Synthesis of ZnO
dependent photocatalytic activity of ZnO nanowires grown by nanorod and the annealing effect on its photoluminescence
vapor transport technique. Chem. Phys. Lett. 539–540, 83–88. property. Opt. Mater. 28, 418–422.
Nadeem, M. Y. & Ahmed, W.  Optical properties of ZnS thin Wu, L., Wu, Y., Shi, Y. & Wej, H. b Synthesis of ZnO
films. Turk. J. Phys. 24, 651–659. nanorods and their optical absorption in visible-light region.
Nair, M. G., Nirmala, M., Rekha, K. & Anukaliani, A.  Structural, Rare Metals 25 (1), 68–73.
optical, photo catalytic and antibacterial activity of ZnO and Xu, T., Zhang, L., Cheng, H. & Zhu, Y.  Significantly enhanced
Co doped ZnO nanoparticles. Mater. Lett. 65 (12), 1797–1800. photocatalytic performance of ZnO via grapheme
Park, H., Chang, S., Jean, J., Cheng, J. J., Araujo, P. T., Wang, M., hybridization and the mechanism study. Appl. Catal. B:
Bawendi, M. G., Dresselhaus, M. S., Bulović, V., Kong, J. & Environ. 101, 382–387.
Gradecak, S.  Graphene cathode-based ZnO nanowire Yong-hong, N., Xian-wen, W., Jian-ming, H. & Yin, Y. 
hybrid solar cells. Nano Lett. 13 (1), 233–239. Hydrothermal preparation and optical properties of ZnO
Peng, X., Manna, L., Yang, W., Wickham, J., Scher, E., nanorods. Mater. Sci. Eng. B 121, 42–47.
Kadavanich, A. & Alivisatos, A.  Shape control of CdSe Zhang, W., Wang, S., Wang, Y., Zhu, Z., Gao, X., Yang, J. &
nanocrystals. Nature 404, 59–61. Zhang, H. X.  Zno@ZnS core/shell microrods with
Puvaneswari, N., Muthukrishnan, J. & Gunasekaran, P.  enhanced gas sensing properties. RSC Adv. 5, 2620–2629.
Toxicity assessment and microbial degradation of azo dyes. Zhou, D. & Keller, A. A.  Role of morphology in the aggregation
Ind. J. Exp. Biol. 44, 618. kinetics of ZnO nanoparticles. Water Res. 44, 2948–2956.
Radzimska, A. K. & Jesionowski, T.  Zinc oxide – from Zhu, H. Y., Xiao, L., Jiang, R., Zeng, G. M. & Liu, L. 
synthesis to application: a review. Materials 7 (4), 2833–2881. Adsorption removal of Congo red onto magnetic cellulose/
Rao, C. N. R.  Chemical Applications of Infrared Spectroscopy. Fe3O4/activated carbon composite: equilibrium, kinetic and
Academic Press, New York. thermodynamic studies. Chem. Eng. J. 172, 494–502.
Sakthivel, S., Neppolian, B., Shankar, M. V., Arabindoo, B., Zhu, H., Jiang, R., Fu, Y., Guan, Y., Yao, J., Xiao, L. & Zeng, G.
Palanichamy, M. & Murugesan, V.  Solar photocatalytic  Effective photocatalytic decolorization of methyl orange
degradation of azo dye: comparison of photocatalytic efficiency utilizing TiO2/ZnO/chitosan nanocomposite films under
of ZnO and TiO2. Sol. Energ. Mat. Sol. Cell 77, 65–82. simulated solar irradiation. Desalination 286, 41–48.

First received 21 December 2015; accepted in revised form 9 May 2016. Available online 1 June 2016

Das könnte Ihnen auch gefallen