You are on page 1of 8

Journal of Water Process Engineering 25 (2018) 88–95

Contents lists available at ScienceDirect

Journal of Water Process Engineering


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

Enhanced photocatalytic degradation of textile wastewater using Ag/ZnO T


thin films

Ahmed M.A. Abdelsamada, , Tarek A. Gad-Allaha, Fawzy A. Mahmoudb,c, Mohamed I. Badawya
a
Water Pollution Research Dept., National Research Centre, 33 El Bohouth St, Dokki, Giza, 12622, Egypt
b
Solid State Physics Dept., Physics Division, National Research Centre, P.O. 12311, Dokki, Giza, Egypt
c
Solar cell Lab., Center of Excellence for Advanced Sciences, National Research Center, P.O. 12311, Dokki, Giza, Egypt

A R T I C LE I N FO A B S T R A C T

Keywords: In this work, photodegradation of a synthetic textile wastewater composed of a mixture of three reactive azo
ZnO dyes (RO 122, RR 195 and RY 145) was performed using silver doped zinc oxide (Ag/ZnO) thin films. The
Thin film photocatalyst preparation was conducted through sol-gel method and was characterized by X-ray diffraction
Silver (XRD), Scanning electron microscopy (SEM), Energy-dispersive X-ray spectroscopy (EDX) and ultraviolet-visible
Photocatalysis
(UV–vis) spectroscopy. The photocatalytic activity of Ag/ZnO thin film was assessed by measuring the deco-
Textile wastewater
Azo dyes
lorization rate constant of the dyes mixture spectrophotometrically. The chemical structure of each dye influ-
ences the photodegradation rate either in mixture or in individual dyes solutions. Several parameters which
influence photocatalytic degradation such as catalyst dose, dyes concentrations as well as addition of hydrogen
peroxide as an electron acceptor have been studied. The results revealed that these parameters should be op-
timized to obtain the maximum photocatalytic degradation.

1. Introduction Although photocatalysis possesses various advantages, many re-


searchers imputed the limited rate of photocatalytic degradation to the
Textile industry is one of the industries that consumes large rapid recombination of photogenerated electron-hole pairs [12,13].
amounts of fresh water in the processing operations. It also discharges Recently, Borthakur and Das (2018) have studied the photodegradation
colored water that obstruct the light transmittance in water, and efficiency of NiS2-rGO and CoS-rGO nanocomposite toward azo dye
therefore causes an imbalance in the ecosystem [1]. Currently, great Congo Red (CR) under natural sunlight irradiation [14]. They have
majority of synthetic dyes used in the industry belongs to azo deriva- found that the photocatalytic efficiency increased by 40% higher than
tives [2]. At the time that the strong stability of azo derivatives dyes NiS2 and CoS nanoparticles without reduced graphene oxide (rGO)
considers a great advantage in the dying process, discharging of these support and they have attributed that to the presence of rGO sheets
dyes to the aquatic environment causes many problems because of their which prevent the recombination of photo-generated electrons and
high toxicity and heavy color [3] Therefore, efficient treatment pro- holes. In our previous work [15] we proved the enhanced photo-
cesses are required to reuse the high amounts of consumed water and to catalytic activity of ZnO thin film when doped with silver. This was
preserve the environment as well. attributed to the ability of silver nanoparticles to reduce the electron-
Many researchers tried to develop effective and economical tech- hole recombination rate by attracting the positively charged holes and
niques for the treatment of dye-containing wastewater [4–6]. Photo- hence increases the lifetime of the electrons in the conduction band
catalysis had been attracting growing attention for the degradation of which in turn improves the photocatalytic properties of ZnO. On the
organic dyes [7,8] due to its fruitful features such as: ability for full other hand, several studies have used peroxides such as H2O2 success-
decomposition of various organic pollutants present in water and fully as strong oxidants and electron scavengers [16–18]. All of these
wastewater within short time, no formation of persistent by-products, researches have reported that addition of H2O2 enhances the rate
and accessibility of scaling up [9–11]. This technique is based on the constant of photodegradation reaction and they ascribed this im-
generation of powerful oxidizing species (e.g. hydroxyl radicals) during provement to reaction of H2O2 with the electrons of conduction band
the photon irradiation on the surface of a semiconducting material producing very reactive hydroxyl radicals which oxidize most of the
which decompose the organic pollutants. organic pollutants according to the equations:


Corresponding author.
E-mail address: ahmed.abdelsamad@daad.alumni.de (A.M.A. Abdelsamad).

https://doi.org/10.1016/j.jwpe.2018.07.002
Received 12 April 2018; Received in revised form 14 June 2018; Accepted 8 July 2018
Available online 17 July 2018
2214-7144/ © 2018 Published by Elsevier Ltd.
A.M.A. Abdelsamad et al. Journal of Water Process Engineering 25 (2018) 88–95

Fig. 1. a) XRD patterns, b) optical transmittance and c) band gap energies of ZnO and 6 wt.% Ag/ZnO thin films.

Fig. 2. SEM images and EDX patterns of a,c) ZnO and b,d) 6 wt%Ag/ZnO thin films, respectively.

89
A.M.A. Abdelsamad et al. Journal of Water Process Engineering 25 (2018) 88–95

2. Experimental

2.1. Materials

Zinc acetate dihydrate, silver nitrate, diethanolamine isopropanol


and hydrogen peroxide solution 30% (w/w) were purchased from
Sigma Aldrich Company. Reactive red 195 (RR195), reactive yellow
145 (RY145), and reactive orange 122 (RO122) dyes were kindly sup-
plied by a local factory for dyes (for the chemical structure, see Figure
S1 in Supplementary Information). NaOH and HCl solutions were used
for pH adjustment. Double-distilled water was used throughout this
study for the preparation of dyes solutions.

2.2. Catalyst preparation

The photocatalyst was prepared by sol–gel spin coating method


Fig. 3. Rate constant of photocatalytic degradation of RR195, RY145, and
RO122, catalyst dose: 1 square slide (6.25 cm2), [dye] = 10 mg/L, and pH = 7. [21]. ZnO solution was prepared from the mixture of zinc acetate de-
hydrate as a precursor solution, isopropanol as a solvent and dietha-
nolamine as a stabilizer. The molar ratio of diethanolamine: zinc
acetate dihydrate was maintained at 1:1. In a typical experiment, 8.78 g
(0.04 mol) of zinc acetate were dissolved in 50 mL of isopropanol and
stirred for 30 min to obtain solution A. Solution B was prepared by
dissolving 0.245 g (0.0014 mol) of silver nitrate in 50 mL of isopropanol
and stirred for 30 min. Both solutions were mixed together in order to
yield 6% Ag/Zn weight ratio. The thin film photocatalyst was deposited
by sol–gel spin coating method onto glass substrates pre-cleaned by
detergent, ethanol, acetone and finally with distilled water and then
sonicated, using Bandelin Sonorex RK100, for 10 min in water and then
dried. For film formation, the sol solution was dropped within 30 s onto
a glass substrate rotating at 2000 rpm using the spin coater, at room
temperature. The obtained films were then dried at 100 °C for 10 min in
order to remove the solvent. This procedure was repeated seven times
to obtain a homogeneous film, then annealed in air at 450 °C for 1 h in a
tubular furnace. Pure ZnO thin film photocatalyst was prepared ac-
cording to the same procedure without the addition of silver nitrate.

Fig. 4. UV/Vis. spectra of the used dyes before the photocatalytic degradation.
2.3. Thin-film characterization

H2 O2 + e− → −OH +•OH (1) The crystal structure of the prepared thin films was characterized
using a Phillips (PW3719) X'pert materials research X-ray dif-

OH +organic compound → mineralized compound (2) fractometer (XRD) using CuKα radiation (λ = 1.5418 Ǻ) radiation, and
a scanning range of 2θ between 25° and 70°. The thickness of films was
However, increasing of hydrogen peroxide concentration so much, measured using Dektak 150 Stylus profilmeter, which has capability to
may decrease the rate of photocatalytic degradation due to scavenging measure thin film thickness with an error ± 10 Angstroms. Optical
effect [19]. Although photocatalysts are usually used in powder form to transmittance spectra were recorded using UV-VIS-NIR spectro-
maintain high surface to volume ratio, they suffer from many limita- photometer Jasco 760 over the wavelength range between 200 and
tions during the application such as difficult separation of the powder 2500 nm.
from the system after degradation process and aggregation of catalyst From the transmission spectra, the absorption coefficient (α) was
particles during stirring which reduces the photocatalyst efficiency. calculated using the following formula:
Coating the photocatalysts on various substrates can overcome the lnT
problems associated with powder systems. Thin film-shape photo- α=−
d (3)
catalyst can improve the transmittance of light and prevent its scat-
tering leading to enhanced reaction efficiency. Moreover, it can be Where T is the optical transmittance and d is thickness of the film (≈
easily removed after photodegradation process [20]. 0.2 μm). By using the absorption coefficient (α), and for the direct
In this paper, we used 6 wt% Ag/ZnO thin film prepared by sol-gel transitions, the optical band gaps (Eg) of our film can be calculated by
spin coating method for degradation of three azo dyes (RR195, RY145, using the conventional Tauc equation:
and RO122) present together in a mixture as a synthetic textile was-
(αhν)2 = A (h ν−Eg ) (4)
tewater. All photocatalytic experiments were carried out under simu-
lated solar irradiation. Effects of different operational conditions, such Where A is a constant, Eg is the direct optical band gap energy, hν is the
as catalyst dose, initial concentration of the dyes and addition of oxi- incident photon energy
dants like hydrogen peroxide (H2O2), on the photocatalytic reaction The surface morphology of the films morphologies was examined by
have been investigated. scanning electron microscope (SEM, XL30TMP). Energy dispersive X-
ray analysis (EDX) by electron probe microanalyzer model JEOL
(JXA_840 A) which utilized was used to examine the composition of the
films.

90
A.M.A. Abdelsamad et al. Journal of Water Process Engineering 25 (2018) 88–95

Fig. 5. Decolorization rate constants of a) RO 122, b) RR 195 and c) RY 145 using different doses of 6 wt.% Ag/ZnO & d) change in first-order rate constants with
catalyst dose, [dye] = 15 mg/L and pH = 7.

2.4. Photocatalytic measurement revealed three noticeable peaks at 2θ = 38.1, 44.2 and 64.4 which can
be assigned to the most intense diffraction planes of metallic silver
Photocatalytic efficiencies of the obtained thin films were evaluated (JCPDS card number 4-0783). It is well known that Ag can be in-
by measuring the color fading of a mixture from the three mentioned corporated into the ZnO lattice either as a substitute for Zn2+ ion or as
dyes. In a typical experiment, 75 mL of the dyes mixture were placed in an interstitial atom. This can be examined by monitoring the diffraction
a glass Petri dish. A catalyst of square geometry (6.25 cm2) was im- peak position of ZnO. The remarkable shift to lower or higher 2θ can be
mersed in the dyes solution adjusted at pH 7. The thin film was located ascribed to substitutional or interstitial incorporation respectively
to face the irradiation source (HQI-T 400 W/DayLight lamp, OSRAM [22,23]. On the other hand, Ag can be segregated to the grain bound-
Germany). The luminous flux of irradiation was 20,000 lm with lumi- aries of ZnO crystal without any observed positional shift in 2θ values,
nous efficacy of 82 lm/W. The distance between surface of the dye so- which indicates the formation of Ag cluster [24]. According to our re-
lution and the lamp was fixed at 15 cm. The typical set-up scheme is sults, it can be seen that incorporation of Ag resulted in a shift of the
presented in Figure S2 in the supplementary information. ZnO peak position to lower 2θ value (from 36.29 to 36.21 for ZnO and
After immersing the thin film into the dyes solution and prior to 6 wt% Ag/ZnO respectively). The reason could be because of the dif-
start irradiation, the solution was allowed to be in a contact with the ference in ionic radii between Zn2+ (0.72 A) and Ag+ (1.22 A) where
photocatalyst for 60 min in the dark to allow the adsorption–desorption the other part of Ag atoms segregated to the grain boundaries of ZnO
equilibrium on the photocatalyst surface. Then, the lamp was switched crystal which indicates the formation of Ag cluster which appeared as
on to start the photodegradation process. Certain quantities of aliquot separated Ag peaks. Fig. 1b shows the transmittance spectra for ZnO
were collected at regular intervals under dark and irradiation condi- and 6 wt%Ag/ZnO thin films. It is observed that, the absorption edge
tions. Color removal was monitored by measuring the light absorbance shifted to longer wavelength (red shift) after the insertion of Ag. As a
versus wavelength using UV–vis spectrophotometer (JASCO V630, result, the transmittance decreased from ≈ 93% to ≈ 60% by Ag ad-
Japan). In a control experiment, the dyes solution was directly irra- dition. This could be due to the grain boundary scattering and the
diated for 1 h in absence of the photocatalyst (i.e. photolysis process). surface plasmonic resonance (SPR) of Ag nanoparticles, which im-
proved the absorption of light in the visible region [25]. The SPR is very
3. Result and discussion clear in case of increasing the Ag content in the ZnO matrix as shown in
(Figure S3 in supplementary information).
3.1. Photocatalyst properties The optical band gap energies (Eg) of the thin films were determined
by extrapolation of the linear portion of (αhν)2 curve, where α is the
The crystal structures of ZnO and 6 wt% Ag/ZnO thin films were absorption coefficient versus the photon energy (hν) as presented in
confirmed using XRD as presented in Fig. 1a. The XRD pattern for ZnO Fig. 1c. Eg was found to be 3.30 eV for the pure ZnO thin film and
thin film can be indexed as a hexagonal wurtzite structure (JCPDS card reduced to 3.21 eV for 6 wt% Ag/ZnO thin film. The reduction of Eg by
number 36-1451). The XRD pattern for 6 wt% Ag/ZnO thin film addition of Ag may be indicated that Ag has substituted the Zn in the

91
A.M.A. Abdelsamad et al. Journal of Water Process Engineering 25 (2018) 88–95

Fig. 6. Decolorization rate constant of a) RO122, b) RR195 and c) RY 145 at different dyes concentrations in the mixture, pH = 7, catalyst dose = 1 square slide
(6.25 cm2).

lattice, which confirmed by XRD results discussed above, that creates UV/Vis. spectra of the commercial dyes used in this study and the
oxygen vacancies due to the valence difference between Ag+ and Zn2+. photocatalytic degradation of their mixture are presented in Fig. 4 and
The created oxygen vacancies play an important role in the narrowing Figure S4 in supplementary information respectively. It is obvious that
of Eg because they act as trap centers that reduce the recombination there are overlap among the maximum absorption peaks of the studied
rate of photo-generated charge carriers by capturing the electrons [26]. dyes. In order to follow the degradation of each dye during the reaction
Furthermore, Ag states may interact with ZnO host states leading to course, mathematical equations have been derived using three solutions
creating energy levels in ZnO band gap which reduce its optical band of different dyes concentration according to the literature [30] as fol-
gap [27]. The reduction of band gap energy is expected to improve the lows:
photocatalytic properties of ZnO by widen the spectrum of the visible Solution (1): 5 mg/L RO 122 + 2 mg/L RY 145 + 1 mg/L RR 195.
light response. Solution (2): 2 mg/L RO 122 + 1 mg/L RY 145 + 5 mg/L RR 195.
SEM micrographs presented in Fig. 2a&b show the surface mor- Solution (3): 1 mg/L RO 122 + 5 mg/L RY 145 + 2 mg/L RR 195.
phology of ZnO and 6 wt%Ag/ZnO thin films. The SEM images clearly The absorption of each solution at the wavelength of maximum
show the deposition of Ag into the ZnO surface as a white cluster as absorption (λmax) of each dye can be represented as follows:
reported by other studies [28,29]. EDX analysis was used as a direct
Abs1 = a1 CA1 + b1 CB1 + c1 CC1 (5)
method to determine the thin film composition. The EDX patterns for
ZnO and 6 wt%Ag/ZnO thin films are shown in Fig. 2 c&d respectively. Abs2 = a1 CA2 + b1 CB2 + c1 CC 2 (6)
It is observed that only Zn and O elements were detected in pure ZnO
film; while Ag element was noticed, in addition to Zn and O elements, Abs3 = a1 CA3 + b1 CB3 + c1 CC 3 (7)
in case of 6 wt%Ag/ZnO.
Absn = a1 CAn + b1 CBn + c1 CCn (8)

Where Absn is the absorbance of solution (n) at λ1, CAn is the con-
3.2. Photocatalytic degradation
centration of RO 122 dye in solution (n), CBn is the concentration of RY
145 dye in solution (n) and CCn is the concentration of RR 195 dye in
According to our previous work, the addition of silver improved the
solution (n). Therefore, the contribution of each dye (a1, b1 and c1) can
photocatalytic activity of ZnO. As one can see in Fig. 3, the degradation
be estimated at each wavelength.
rate constant (k) of RO 122, RR 195 and RY 145 always higher in case
Accordingly, the following mathematical relationships could be es-
of 6 wt%Ag/ZnO than pure ZnO [13]. The enhancement in the photo-
timated.
catalytic activity could be due to better charge separation caused by
silver atoms. Based on these results, 6%Ag/ZnO was selected for the Abs420 nm = 7.2x10−3CA + 8.3x10−3CB + 4.3x10−3Cc (9)
photocatalytic degradation of a mixture of three azo dyes as a synthetic
textile waste water. Abs487 nm = 4.5x10−3CA + 13.5x10−3CB + 12.5x10−3Cc (10)

92
A.M.A. Abdelsamad et al. Journal of Water Process Engineering 25 (2018) 88–95

Fig. 7. Decolorization behavior in case of individual solutions or in a mixture of dyes solution, catalyst dose: 1 square slide (6.25 cm2), [dye] = 10 mg/L, and pH = 7.

Abs540 nm = 3x10−3CA + 2.87x10−3CB + 18.1x10−3Cc (11) photodegradation process increases by adding more catalyst until
reaching a specific value; then the rate decreases. This performance is
It was important to measure the effect of the irradiation on the due to the high turbidity caused by photocatalyst particles in solution;
degradation of the dyes mixture in absence of the catalyst (photolysis). therefore, the light path inside the photoreactor could be obstructed by
In addition, adsorption of the dyes at the surface of catalyst had been the photocatalyst itself which defined as shielding effect [32]. Also, at
measured. The adsorption was determined by allowing the dye solution high concentration of powder photocatalyst, many small particles tend
(or a mixture of the three dyes) to be contacted with the catalyst in the to aggregate together forming larger particles with smaller surface area
dark for 60 min. The results (Figure S5 in supplementary information) and less active sites which also decline the photocatalytic activity. On
showed that both of photolysis and adsorption produced only ≈1% contrary, in case of thin film photocatalyst there is neither aggregation
color removal. Therefore, their influence on the dyes degradation could of catalyst particles nor shielding of photons. As a result, increasing the
be neglected. catalyst amount in case of thin film shape could lead to a continuous
increase in the reaction rate.
It is worth noting also that decolorization rate constants (Fig. 5d)
3.2.1. Effect of catalyst dose
follow the order: RO 122 > RR 195 > RY 145. This behavior depends
Catalyst dose is a crucial factor which should not exceed a specific
mainly on the chemical structure of each dye (see Figure S1 in the
limit during the wastewater treatment to avoid adding excess catalyst
supplementary information). For instance, RY 145 is the least degrad-
amount that has a negative effect on the performance of the treatment
able dye where it contains (eCONH2) group that make resonance with
process as well as on the final cost. In this study, catalyst dose was
azo group giving more stabilization. This decreases the ability of OH
expressed in terms of number of glass slides carrying the photoactive
radical to attack the azo group. RR195 dye contains five electophilic
thin film. Each thin film slide has square area 6.25 cm2. The change in
sulfo groups (eSO3Na+) compared to only four groups in case of
each dye concentration during the photodegradation reaction of the dye
RO122 causing difficulty for the addition reaction of %OH radical to the
mixture was calculated using Eqs. (9)–(11) and presented in Fig. 5(a–c).
azo group. Therefore, RO122 is more susceptible to %OH radical attack
The kinetics of this reaction was found to follow the pseudo-first order
than RR 195 that exists in the same solution.
model. The dependence of first-order kinetic constant on the photo-
catalyst dose is presented in Fig. 5d. It is obvious that the decolorization
rate constant increases with the increase of the photoactive thin film 3.2.2. Effect of dye concentration
area. A possible explanation is that by increasing the area of thin film in The effect of initial dyes concentration on the rate constant of the
the reaction medium, the amount of catalyst exposed to light increases, reaction and on the percentage of color removal is another important
leading to activation of higher number of active sites. This subsequently factor in the treatment of textile wastewater. Therefore, the effect of
produces more %OH radicals which, enhances the photocatalytic effi- initial dyes concentration in a mixture was studied by varying the initial
ciency [31]. concentration from 5 to 15 mg/L for each dye in the mixture. The ob-
In case of powder-shape photocatalyst, the rate constant of tained results are represented in Fig. 6. The rate of decolorization was

93
A.M.A. Abdelsamad et al. Journal of Water Process Engineering 25 (2018) 88–95

Fig. 8. Decolorization rate constant of a) RO 122, b) RR 195 and c) RY 145. catalyst dose: 1 square slide (6.25 cm2), [dye] = 10 mg/L, and pH = 7.

Therefore, the relative number of free radicals attacking the dye mo-
lecules decreases with increase the initial dye concentration.

3.2.3. Comparison between the photoctalytic degradation behaviors of


individual and mixed dyes
Fig. 7 shows the decolorization behavior of each dye in case of in-
dividual solutions compared to its behavior with other dyes in a mix-
ture. Generally, the decolorization rate constant of individual dyes is
higher than that when the dye exists in a mixture, in particular for the
dyes RY 145 and RR 195. In case of dyes mixture, there is a competition
among the three dyes for the reaction with the same amount of %OH
radicals produced by the photocatalyst. Therefore, each dye may react
with less number of %OH radicals than it may do when presents as in-
dividual.
It is worth noting that the difference in decolorization behavior
between individual and mixed dyes follows the order: RY 145 > RR
195 > RO 122. This order is the reverse to the dye susceptibility to %OH
Fig. 9. The change of rate constant with H2O2 concentration for the three dyes,
catalyst dose: 1 square slide (6.25 cm2), [dye] = 10 mg/L, and pH = 7.
attack deduced from the chemical structures. Since, RO 122 dye is the
most susceptible dye for %OH attack, it is expected that this dye will
consume the largest portion of photogenerated %OH and will not be
found to decrease with increasing the initial dyes concentration. Similar significantly affected by the presence of other dyes in the mixture.
results were observed by Rajabi et al [33] during the photocatalytic Other dyes are expected to obey the same rule.
degradation of methyl violet dye using ZnS quantum dots. A possible
explanation for this behavior is that as the initial dye concentration
3.2.4. Effect of hydrogen peroxide addition
increases, the path length of photons entering the solution decreases
Addition of H2O2 was considered by many researchers as a promoter
and the opposite is true in case of low initial dyes concentration [34].
for the photocatalysis process [36,37,38]. Thereby, the photocatalytic
By that means, the number of photons absorbed by the catalyst in-
degradation of the dyes mixture has been studied at different hydrogen
creases in diluted dye solutions [35]. Furthermore, the high initial dye
peroxide quantities in the range of 0.005 – 0.16 mL (20 −700 mg/L)
concentration requires large amount of catalyst to be degraded. Since
According to the results presented in Figs. 8 and 9, the decolorization
the illumination time and amount of catalyst are constant, the %OH
rate of dyes increases with the increase of H2O2 volume up to 0.010 mL
radicals generated on the surface of the catalyst is also constant.
(45 mg/L). Beyond this concentration, the rate decreases. The higher

94
A.M.A. Abdelsamad et al. Journal of Water Process Engineering 25 (2018) 88–95

reaction rate after the addition of hydrogen peroxide could be due to recent updates, Catalysts 2 (2012) 572–601.
the increase in the concentration of %OH radical generated from H2O2 [10] J.M. Monteagudo, A. Durán, R. Culebradas, I. San Martín, A. Carnicer, Optimization
of pharmaceutical wastewater treatment by solar/ferrioxalate photo-catalysis, J.
itself. This is because H2O2 produces reactive %OH radicals rapidly Environ. Manage. 128 (2013) 210–219.
under illumination (Eq. (12]) [39]. The oxidizing power of these radi- [11] I.K. Konstantinou, T.A. Albanis, Photocatalytic transformation of pesticides in
aqueous titanium dioxide suspensions using artificial and solar light: intermediates
cals is strong enough to break the azo groups in the reactive dyes and degradation pathways, Appl. Catal., B 42 (2003) 319–335.
leading to the formation of ammonium and nitrate ions. [12] L. Andronic, A. Duta, Influence of pH and H2O2 on dyes photodegradation, physica
status solidi (c) 5 (2008) 3332–3337.
H2 O2 + hν → 2 •OH (12) [13] B.J.P.A. Cornish, L.A. Lawton, P.K.J. Robertson, Hydrogen peroxide enhanced
photocatalytic oxidation of microcystin-LR using titanium dioxide, Appl. Catal., B
At higher H2O2 dosage, a remarkable decrease has been observed in 25 (2000) 59–67.
the decolourization rate and therefore on the degradation efficiency of [14] P. Borthakur, M.R. Das, Hydrothermal assisted decoration of NiS2 and CoS nano-
particles on the reduced graphene oxide nanosheets for sunlight driven photo-
the three dyes. This could be explained as at high H2O2 concentration, catalytic degradation of azo dye: effect of background electrolyte and surface
the %OH radicals can scavenge the hydrogen peroxide producing HO2• charge, J. Colloid Interface Sci. 516 (2018) 342–354.
[15] M.I. Badawy, F.A. Mahmoud, A.A. Abdel-Khalek, T.A. Gad-Allah, A.A. Abdel
radicals which are less active than %OH. Therefore, the number of %OH Samad, Solar photocatalytic activity of sol–gel prepared Ag-doped ZnO thin films,
radicals which is responsible for degradation of organic molecules de- Desalin. Water Treat. 52 (2014) 2601–2608.
creases according to (Eq. ([13]]) [40]. [16] D. Georgiou, P. Melidis, A. Aivasidis, K. Gimouhopoulos, Degradation of azo-re-
active dyes by ultraviolet radiation in the presence of hydrogen peroxide, Dyes
H2 O2 + •OH →HO2• + H2 O (13) Pigm. 52 (2002) 69–78.
[17] M. Muruganandham, Sobana N, M. Swaminathan, Solar assisted photocatalytic and
Conclusively, in order to keep the efficiency of the added hydrogen photochemical degradation of reactive black 5, J. Hazard. Mater. 137 (2006)
1371–1376.
peroxide, it was necessary to adjust the proper dosage of hydrogen [18] H. Ye, S. Lu, Effect of hydrogen peroxide on the structure and photocatalytic ac-
peroxide. The obtained results consistent with the previously reported tivity of titania, Res. Chem. Intermed. 41 (2015) 139–149.
studies [18,38]. [19] S. Alahiane, S. Qourzal, M. El Ouardi, A. Abaamrane, A. Assabbane, Factors influ-
encing the photocatalytic degradation of reactive yellow 145 by TiO2-coated Non-
woven fibers, Am. J. Anal. Chem. 5 (2014) 445–454.
4. Conclusions [20] A.Y. Shan, T.I. MohdGhazi, S. Abdul Rashid, Immobilisation of titanium dioxide
onto supporting materials in heterogeneous photocatalysis: a review, Appl. Catal., A
389 (2010) 1–8.
6 wt % Ag/ZnO thin film was used as a photocatalyst for the de- [21] Y. Kim, W. Tai, S. Shu, Effect of preheating temperature on structural and optical
gradation of dyes mixture composed of three azo dyes (RO 122, RR 195 properties of ZnO thin films by sol–gel process, Thin Solid Films 491 (2005)
153–160.
and RY 145) as a synthetic textile wastewater. There are several para- [22] C. Karunakaran, V. Rajeswari, P. Gomathisankar, Antibacterial and photocatalytic
meters affect the photocatalytic degradation process such as catalyst activities of sonochemically prepared ZnO and Ag–ZnO, J. Alloy. Compd. 508
(2010) 587–591.
dose, dye concentration and addition of H2O2. The results showed that [23] Y. Jin, Q. Cui, K. Wang, J. Hao, Q. Wang, J. Zhang, Investigation of photo-
increased catalyst dose produced more surface area available for %OH luminescence in undoped and Ag-doped ZnO flowerlike nanocrystals, J. Appl. Phys.
radicals and consequently increases the rate of reaction. The increase of 109 (2011) 53521–53525.
[24] Ö.A. Yıldırım, H.E. Unalan, C. Durucan, Highly efficient room temperature synth-
dye concentration decreases the rate of reaction because of the photons esis of silver-doped zinc oxide (ZnO:Ag) nanoparticles: structural, optical, and
entering the solution can be interrupted by the large number of dye photocatalytic properties, J. Am. Ceram. Soc. 96 (2013) 766–773.
molecules presented in the solution leading to a reduction in the rate of [25] N. Tarwal, P.S. Patil, Enhanced photoelectrochemical performance of Ag–ZnO thin
films synthesized by spray pyrolysis technique, Electrochim. Acta 56 (2011)
degradation. Finally, the optimum quantity of H2O2 (45 mg/L) causes 6510–6516.
an increase in the rate or photodegradation reaction while by addition [26] Oxygen vacancy-mediated ZnO nanoparticle photocatalyst for degradation of me-
thylene Blue, Appl. Sci. 8 (353) (2018), https://doi.org/10.3390/app8030353.
of larger quantity, the rate decreases. [27] J. Wang, Z. Wang, B. Huang, Y. Ma, Y. Liu, X. Qin, X. Zhang, Y. Dai, Oxygen va-
cancy induced band-gap narrowing and enhanced visible light photocatalytic ac-
Acknowledgments tivity of ZnO, ACS Appl. Mater. Interfaces 4 (2012) 4024–4030.
[28] J. Rashid, M.A. Barakat, N. Salah, S.S. Habib, Ag/ZnO nanoparticles thin films as
visible light photocatalysts, RSC Adv. 4 (2014) 56892–56899.
The authors acknowledge National Research Center (NRC) for the [29] D.Y. Wang, J. Zhou, G.Z. Liu, Effect of Li-doped concentration on the structure,
optical and electrical properties of p-type ZnO thin films prepared by sol–gel
financial support.
method, J. Alloys Compd. 481 (2009) 802–805.
[30] T.A. Gad-Allah, S. Kato, S. Satokawa, T. Kojima, Treatment of synthetic dyes was-
Appendix A. Supplementary data tewater utilizing a magnetically separable photocatalyst (TiO2/SiO2/Fe3O4): para-
metric and kinetic studies, Desalination 244 (2009) 1–11.
[31] S. Mosleh, M.R. Rahimi, M. Ghaedi, K. Dashtian, S. Hajatic, Photocatalytic de-
Supplementary material related to this article can be found, in the gradation of binary mixture of toxic dyes by HKUST-1 MOF and HKUST-1-SBA-15 in
online version, at doi:https://doi.org/10.1016/j.jwpe.2018.07.002. a rotating packed bed reactor under blue LED illumination: central composite de-
sign optimization, RSC Adv. 6 (2016) 17204–17214.
[32] Ö.E. Kartal, M. Erol, H. Oǧuz, Photocatalytic destruction of phenol by TiO2Powders,
References Chem. Eng. Technol. 24 (2001) 645–649.
[33] H. Rajab, M. Farsi, Effect of transition metal ion doping on the photocatalytic ac-
tivity of ZnS quantum dots: synthesis, characterization, and application for dye
[1] J. Li, D. Wang, D. Yu, P. Zhang, Y. Li, Performance and membrane fouling in an decolorization, J. Mol. Catal. A: Chem. 399 (2015) 53–61.
integrated membrane coagulation reactor (IMCR) treating textile wastewater, [34] V. Vaiano, M. Matarangolo, O. Sacco, D. Sannino, Photocatalytic treatment of
Chem. Eng. J. 240 (2014) 82–90. aqueous solutions at high dye concentration using praseodymium-doped ZnO cat-
[2] F. Deniz, S. Karaman, Removal of an azo-metal complex textile dye from colored alysts, Appl. Catal., B 209 (2017) 621–630.
aqueous solutions using an agro-residue, Microchem. J. 99 (2011) 296–302. [35] R.J. Davis, J.L. Gainer, G. O’Neal, I.W. Wu, Photocatalytic decolorization of was-
[3] C. Racles, M. Zaltariov, M. Iacob, M. Silion, M. Avadanei, A. Bargan, Siloxane-based tewater dyes, Water Environ. Res 66 (1994) 50–53.
metal–organic frameworks with remarkable catalytic activity in mild environ- [36] T. Fei, Z. Rongshu, S. Kelin, O. Feng, C. Gang, Synergistic photocatalytic degrada-
mental photodegradation of azo dyes, Appl. Catal., B 205 (2017) 78–92. tion of phenol using precious metal supported titanium dioxide with hydrogen
[4] Q. Zhong, Q. Yue, Q. Li, X. Xu, B. Gao, Preparation, characterization of modified peroxide, Environ. Eng. Sci. 33 (2016) 185–192.
wheat residue and its utilization for the anionic dye removal, Desalination 267 [37] S. Wardhani, M.F. Rahman, D. Purwonugroho, R.T. Tjahjanto, C.A. Damayanti,
(2011) 193–200. I.O. Wulandari, Photocatalytic degradation of methylene Blue using TiO2-natural
[5] A. Hasanbeigi, L. Price, A review of energy use and energy efficiency technologies zeolite as a photocatalyst, J. Pure Appl. Chem. Res. 5 (2016) 19–27.
for the textile industry, Renew. Sustain. Energy Rev. 16 (2012) 3648–3665. [38] C.C. Wong, W. Chu, The hydrogen peroxide-assisted photocatalytic degradation of
[6] V. Khandegar, A.K. Saroha, Electrocoagulation for the treatment of textile industry alachlor in TiO2 suspensions, Environ. Sci. Technol. 37 (2003) 2310–2316.
effluent – a review, J. Environ. Manage. 128 (2013) 949–963. [39] B. Neppolian, H.C. Choi, S. Sakthivel, B. Arabindoo, V. Murugesan, Solar light in-
[7] K. Saeed, I. Khan, T. Gul, M. Sadiq, Efficient photodegradation of methyl violet dye duced and TiO2 assisted degradation of textile dye reactive blue 4, Chemosphere 46
using TiO2/Ptand TiO2/Pd photocatalysts, Appl. Water Sci. 7 (2017) 3841–3848. (2002) 1173–1181.
[8] R.M. Kakhki, R. Tayebee, S. Hedayat, Phthalhydrazide nanoparticles as new highly [40] N. Daneshvar, D. Salari, A.R. Khataee, Photocatalytic degradation of azo dye acid
reusable organic photocatalyst in the photodegradation of organic and inorganic red 14 in water on ZnO as an alternative catalyst to TiO2, J. Photochem. Photobiol.,
contaminants, Appl. Organomet. Chem. 32 (2018) e4033. A 162 (2004) 317–322.
[9] M. Lazar, S. Varghese, S. Nair, Photocatalytic water treatment by titanium dioxide:

95

You might also like