You are on page 1of 9

Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 270–278

Contents lists available at SciVerse ScienceDirect

Journal of the Taiwan Institute of Chemical Engineers


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

Photocatalytic activity of a nitrogen-doped TiO2 modified zeolite in the


degradation of Reactive Yellow 125 azo dye
Elida Cristina Ilinoiu a, Rodica Pode a,*, Florica Manea a, Liliana Andreea Colar a, Agnes Jakab a,
Corina Orha b, Cornelia Ratiu b,c, Carmen Lazau b, Paula Sfarloaga b
a
Department of Applied Chemistry and Engineering of Inorganic Compounds and Environment, ‘‘Politehnica’’ University of Timisoara Victoriei Sq. 2, 300006 Timisoara, Romania
b
Department of Condensed Matter, National Institute of Research-Development for Electrochemistry and Condensed Matter Timisoara, Plautius Andronescu 1, 300224 Timisoara,
Romania
c
National Institute for Research and Development in Microtechnologies, Bucharest, Romania

A R T I C L E I N F O A B S T R A C T

Article history: In this study, the hybrid material based on natural zeolite modified with nitrogen-doped TiO2
Received 7 May 2012 photocatalyst (Z-TiO2-N) was synthesized by solid-state reaction under microwave-assisted hydrother-
Received in revised form 16 July 2012 mal conditions. The surface characterization of hybrid material has been investigated by X-ray
Accepted 2 September 2012
diffraction (XRD), scanning electron microscopy (SEM)/energy dispersive X-ray (EDX) analysis, and FT-IR
Available online 27 December 2012
spectroscopy. The photocatalytic activity of Z-TiO2-N photocatalyst under UV and VIS irradiation was
determined for the degradation of Reactive Yellow 125 (RY 125) dye solution. Also, UV–VIS diffuse
Keywords:
reflectance (DRUV–VIS) spectroscopy was used to determine the light absorption properties of the
Nitrogen-doped TiO2-modified zeolite
Photocatalysis
hybrid material. The operational conditions using photocatalyst dose of 1 g/L at pH 3 were established as
UV irradiation optimal conditions for photocatalytic application in RY 125 dye solution degradation. The experimental
VIS irradiation studies revealed that the degradation efficiency expressed in terms of discoloration, aromatic ring-
RY 125 dye opening and mineralization depended strongly on the initial RY 125 dye solution concentrations. The
photocatalyst exhibited a good performance for discoloration and aromatic-ring opening processes at all
studied concentrations (25, 50 and 100 mg/L) but an effective mineralization was reached only for
25 mg/L RY 125 dye solution. Also, the degradation of low concentrations of RY 125 dye solution was
achieved using Z-TiO2-N catalyst under VIS irradiation.
ß 2012 Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

1. Introduction coagulation [7,8], adsorption [9–11], advanced oxidation processes


(AOPs) [12,13] and membrane filtrations [14] have been employed.
Industrial dyestuffs including textile dyes represent a major All these processes exhibit some specific advantages or drawbacks.
threat for environment [1]. The most typical characteristic is their Advanced oxidation processes can be regarded as a potential
color, which is responsible for a non-esthetic visual pollution [2,3]. alternative to decolorate and reduce recalcitrant wastewater loads
In this way, the presence of dyes in wastewaters is visible even at from textile dyeing industry [5] based on the generation of some
low concentrations. Another characteristic of these effluents very reactive species, such as hydroxyl radicals (OH), which
consists of their variable composition, which are in general oxidize quickly and non-selectively a broad range of pollutants
represented by relatively low biological oxygen demand (BOD) and [15–17].
high chemical oxygen demand (COD) values. Due to their complex Among AOPs, heterogeneous photocatalysis is considered to be
structure, most of dyes are recalcitrant [4] and possible carcino- one of the most emerging oxidation technology due to the lack of
genic, especially azo dyes through their aromatic amines resulted limitations in mass transfer, the ability to be carried out under
as degradation products [5,6]. All these characteristics make dyes ambient conditions (atmospheric oxygen is used as oxidant), and
to be ineffectively treated by biological methods. may lead to complete mineralization of organic carbon to CO2. The
In order to handle the dye removal from water, various methods most common semiconductor catalyst used in this process is
have been investigated. Among biodegradation, methods like titanium dioxide (TiO2) because it is inexpensive, wide-available,
non-toxic and water insoluble [18].
The oxidation process involving TiO2 consists of the absorption
* Corresponding author. Tel.: +40 256 403070; fax: +40 256 403060. of an energy photon that is equal to or grater than its band gap
E-mail address: rodica.pode@chim.upt.ro (R. Pode). width, and an electron (e) is promoted from the valence band (VB)

1876-1070/$ – see front matter ß 2012 Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jtice.2012.09.006
E.C. Ilinoiu et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 270–278 271

to the conduction band (CB) generating an electron vacancy or hole company (Timisoara city, Romania). It was used without further
(h+) in the valence band. The electron and the hole can migrate to purification. The molecular structure of this dye is presented in
the catalyst’s surface, where they participate in redox reactions Fig. 1.
with different species adsorbed on the catalyst surface. Holes can The chemicals used for synthesis, i.e., titanium (IV) isoprop-
oxidize the dye and/or react with surface-bound H2O or HO to oxide (TTIP, 98%), ethanol, urea, hydrochloric acid (HCl), sodium
produce the hydroxyl radical (OH). The photo-generated electrons nitrate (NaNO3) and nitric acid (HNO3), were purchased from
could reduce the dye or react with electron acceptors, such as O2 ALDRICH Company. The dye solution pH was adjusted with 0.1 N
adsorbed on the catalyst surface or dissolved in water, which is H2SO4 and 0.1 N NaOH of analytical grade.
reduced to superoxide radical anion O2. The hydroxyl radicals Romanian zeolitic mineral from Mirsid, used as support for TiO2
(OH) and superoxide radical anion O2 are supposed to be the loading, was supplied by CEMACON Company, Romania. The
primary oxidizing species in the photocatalytic oxidation process mineral was powdered and sieved with a Multilab sieve shaker.
[18,19]. The process economics can be improved by using sunlight The diameter of grains size selected to carry out the experiments
instead of UV light for the semiconductor photo-excitation. ranged between 0.8 and 1.2 mm with the mass composition:
Methods of the TiO2 absorption spectrum shifting to the main 62.20% SiO2, 11.65% Al2O3, 1.30% Fe2O3, 3.74% CaO, 0.67% MgO,
part of solar spectrum (l > 360 nm) have been also investigated 3.30% K2O, 0.72% Na2O, 0.28% TiO2.
[20–24]. It has been found that one of the most feasible and the
simplest is nitrogen doping [25,26], which is based on a changed 2.2. Preparation of nitrogen-doped TiO2
mechanism that prevents the electron/hole recombination on the
semiconductor. Nitrogen-doped TiO2 (N-doped TiO2) nanocrystals were syn-
There are several limitations to use ‘‘bare’’ TiO2 in photo- thesized by sol–gel method. An amount of ethanol (30 mL) was
catalytic reactors. The fast aggregation of TiO2 in suspension leads mixed under stirring with 5 mL of titanium isopropoxide (TiO2
to effective surface area decrease and further its catalytic efficiency precursor), which was added dropwisely. After a few minutes of
reduction. In addition, a filtration step after photocatalytic reaction stirring, distilled water was added. After pH adjusting at 8 using
is required because of TiO2 suspension. Attempts have been made ammonia solution, 0.0463 g urea (N precursor) was added. After
to immobilize TiO2 on different supports like glass beds [27], silica 1 h of continuous stirring the obtained material was filtered,
[28], activated carbon [29], biodegradable polymer [30] and washed several times and dried at 60 8C for 1 h. The resulted solid
zeolites [31]. was calcinated at 400 8C for 2 h.
Combining TiO2 with the adsorption properties of organic
molecules onto some adequate adsorbents, e.g., clays, zeolites with 2.3. Preparation of photocatalyst
large surface areas, a synergism is gained, which leads to an
enhancement of the reaction rate. Among the mentioned supports, The preparation of the chemically modified zeolite presumes
zeolites are supposed to provide an effective separation of two stages to reach acid form (H form) by using 2 M HCl solution
photogenerated electrons and holes due to the electric field of and sodium form (Na form) with 2 M NaNO3 solution for a more
their framework [19,32]. efficient ion-exchange. Consequently, Na forms of clinoptilolite are
This study aimed to obtain a hybrid materials based on the expected to remove easily other cations in ion-exchange applica-
natural zeolite modified with nitrogen-doped TiO2 (Z-TiO2-N) tions. For this reason, Na forms are the forms most frequently
suitable for photocatalysis application under VIS irradiation. prepared for clinoptilolite functionalization [33].
Reactive Yellow 125 azo dye was chosen a target pollutant, taking The hybrid material based on natural zeolite modified with
into account its negative impact potential on environment. To the nitrogen-doped TiO2 (Z-TiO2-N) was prepared by solid-state
best of our knowledge, there is no information about the obtaining reaction under microwave-assisted hydrothermal conditions.
and application of this kind of catalyst for dye degradation. The The microwave synthesis was performed in a microwave reaction
catalyst was characterized structurally and the photocatalytic system Multiwave 300, produced by Anton Paar at 2.45 GHz for a
activity was determined for RY 125 azo dye degradation under continuous power of 1000 W. The temperature measurements are
both UV and VIS irradiation. recorded with an IR sensor depending on autoclave solution
content. Therefore, an amount of 5 g of natural zeolite as Na-form
2. Materials and methods was mixed with 40 mL distilled water and N-doped TiO2
nanocrystals (2 wt%) under continuous stirring for 4 h. The
2.1. Chemicals obtained solutions were introduced into a Teflon autoclave with
a 50% degree of fullness, for 30 min to 180 8C, under microwave
Reactive Yellow 125 (RY 125) (commercial grade) with a radiations. After autoclaving, the Z-TiO2-N hybrid material was
molecular weight of 681.5 g/mol was supplied by a local textile washed with distilled water and dried at 60 8C for 5 h.

Fig. 1. Molecular structure of Reactive Yellow 125 dye.


272 E.C. Ilinoiu et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 270–278

2.4. Characterization of the photocatalyst

The crystalline structure of the prepared samples was


determined by X-ray diffraction (XRD) using PANalytical
X’PertPRO MPD Difractometer with Cu tube. The morphological
characterization of the catalyst was carried out by scanning
electron microscopy (SEM) coupled with the energy dispersive
X-ray (EDX) analysis using Inspect S PANalytical model, and by
Fourier transform infrared spectrometry (FT-IR) using a JASCO
FT/IR-430 spectrometer. The light absorption properties of the
hybrid materials were studied by UV–VIS diffuse reflectance
(DRUV–VIS) spectroscopy, performed under ambient conditions
using Lambda 950 Perkin Elmer in the wavelength ranged
between 200 and 430 nm. The blank board was used as the
reference.
Fig. 2. XRD spectrum for natural zeolite (a) and Z-TiO2-N photocatalyst (b); inset:
2.5. Photocatalytic experiments XRD spectrum of TiO2-N.

All experiments regarding adsorption and photocatalysis


processes were carried out under magnetic stirring at 20 8C into
a RS-1 photo-catalytic reactor (Heraeus, Germany), which consists TiO2-N. The peaks recorded at 2u  108, 22.58, 308 correspond to
of a submerged lamp surrounded by a quartz shield. Solutions of the presence of clinoptilolite as major component of the natural
RY 125 dye at different concentrations (prepared from an initial zeolite [36], and also smaller amounts of illite, quartz and albite
stock solution of 10 g/L) were placed in the photoreactor and were identified [37]. The presence of anatase form of TiO2 was
irradiated with an UV light set between 280 and 360 nm when UV evidenced by the peak recorded at 2u = 25.28. No significant
photocatalytic process was employed and a VIS light was set changes of the main peaks characteristics of the natural zeolite
between 460 and 510 nm in the case of VIS photocatalytic process, were noticed, which informed about the stability of the zeolite in
for 240 min. A system of water recirculation maintained a constant the hydrothermal synthesis conditions under microwave field.
temperature of 20 8C for the whole period of photocatalytic Also, no phase transition of TiO2 occurred at the temperature of
experiments. 180 8C for 30 min.
The catalyst dose ranging between 0.5 and 2 g/L was placed into
400 mL RY 125 solution. Prior to irradiation, the suspension was 3.1.2. Scanning electron microscopy (SEM) and energy dispersive X-
stirred continuously in the dark for 30 min to reach the steady- ray (EDX) analysis results
state conditions. Samples were periodically collected from the The comparative SEM images and EDX spectra of the natural
reactor to evaluate the degree of RY 125 adsorption on the catalyst. zeolite modified as Na form, Z-TiO2-N photocatalyst and TiO2-N are
The concentration of the dye in the bulk solution at the end point of presented in Fig. 3.
the adsorption step was considered as the initial concentration The SEM image of natural zeolite shown in Fig. 3a presents the
value for the investigation of the photocatalytic process. At regular lamellar texture of clinoptilolite, according to the literature data
time intervals of irradiation, samples were collected and filtered [38] and EDX spectrum (inset of Fig. 3a) evidenced its major
through a Milipore filter (pore size of 0.45 mm) in order to remove elements. The results of morphological analysis for Z-TiO2-N
the zeolite from the aqueous solution. The concentration of RY 125 photocatalyst are in accordance with XRD, the lamelar structure of
solutions was measured with a Varian Cary 100 UV-VIS zeolite being observed also from SEM image (Fig. 3b). No changes
spectrophotometer. of the form and dimensions of TiO2 particles were noticed (see
The assessment of the photocatalysis performance was carried Fig. 3c), which are distributed randomized and form cluster
out as discoloration, aromatic ring-opening and mineralization agglomerate groups on the surface and in site of zeolite channels,
efficiencies. The discoloration and aromatic ring-opening efficien- but not into the pores. EDX spectrum (inset of Fig. 3b) allowed
cies were determined based on the absorbance recorded at 388 nm evidencing N and TiO2 on the natural zeolite. For comparison, SEM
and 225 nm, respectively. The absorbances were selected from image and EDX spectrum of TiO2-N synthesized by sol–gel method
UV–VIS spectrum profile (see Section 3.2.1). Total organic carbon and used to functionalize the natural zeolite are shown in Fig. 3c.
(TOC) parameter measured with Shimadzu TOC analyzer was used TiO2-N exhibited a very good stability during the hydrothermal
to assess the mineralization process. The experiments were treatment under microwave field, which allows obtaining Z-TiO2-
conducted at the pH values of 3, 6 and 9, respectively. The pH N photocatalyst.
of dye solutions was adjusted by adding diluted HCl and NaOH
solutions (analytical grade), using an Inolab pH-meter. 3.1.3. FT-IR spectroscopy results
Fig. 4 presents the FT-IR spectrum of Z-TiO2-N in comparison
3. Results and discussion with Z-Na to evidence the possible bonds of TiO2 with the zeolite.
The stretching vibration of Ti–O–Si and Ti–O–Al groups is noticed
3.1. Characterization of the catalyst by the absorption band ranged from 945 to 905 cm1 [39]. The
similar tetrahedral loadings bands from 450 to 900 cm1 recorded
3.1.1. XRD analysis of the photocatalyst for both materials informed that the zeolite network was not
In Fig. 2 are shown XRD spectra for natural zeolite (a) and Z- affected by the presence of TiO2-N.
TiO2-N photocatalyst (b) and for comparison, the inset presents
XRD spectrum for titanium dioxide doped with 2% N that was used 3.1.4. DRUV–VIS spectroscopy
for photocatalyst synthesis. DRUV–VIS patterns are examined to determine the light
The specific peaks corresponding to anatase TiO2 form recorded absorption quantification and absorption wavelength range
at 2u  25.28, 37.878, 48.018, 53.818 [34,35] were identified for correlated with band gap energy. The reflectance was converted
E.C. Ilinoiu et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 270–278 273

By comparative analysis of the absorption spectra for Z-Na and


Z-TiO2-N (Fig. 5), it can be noticed more intense absorption in UV
and a slight shifting of the absorption band to the visible range for
Z-TiO2-N, which allows to design it for further photocatalytic use
under UV and VIS irradiation.
The further experiments will be carried out to determine the
photocatalytic activity under UV and VIS irradiation of Z-TiO2-N for
the degradation of RY 125 dye solution.

3.2. UV photocatalytic degradation

3.2.1. UV–VIS spectra profile of RY 125 dye solution


The UV–VIS spectra profiles of RY 125 solution containing Z-
TiO2-N catalyst recorded during UV irradiation are presented in
Fig. 6. The spectrum of the dye before irradiation exhibits two
major absorption peaks, due to the different units and groups
present in the dye molecule, which absorb the light in different
regions: one in the visible region at 388 nm and the second one at
225 nm in the UV region. Based on the literature data [24,40], the
absorption in the UV region can be attributed to aromatic (benzene
and naphthalene) rings, whereas the peak observed in the VIS
region can be assigned to azo linkage containing chromophore
group.
It can be noticed that after 30 min adsorption of the dye on the
catalyst, a small decrease in absorbance of both peaks appeared.
When irradiation is employed, a rapid discoloration of the dye
solution occurs, as well as a significant decrease of the absorbance
recorded at 255 nm that is associated to aromatic ring-opening.
Both peaks became plat at the final time of irradiation, respectively
at 240 min, although the remaining absorbance intensity can be
attributed to the organic intermediaries resulted during the
photocatalytic process. It has to be mentioned that the discolor-
ation and aromatic ring-opening of the dye solution does not
assume a complete azo-dye degradation and mineralization.
Monitoring total organic carbon (TOC) parameter is required to
assess the mineralization degree.

3.2.2. pH influence
pH is one of the main factors that influences the rate of
degradation of some organic compounds in the photocatalytic
process [40]. Fig. 6 presents the discoloration and aromatic ring-
opening efficiencies of 100 mg/L RY 125 solution at different pH
values, i.e., pH 3, 6, and 9. It is obviously that the best discoloration
and aromatic ring-opening efficiencies are reached at pH 3, which
was selected as optimum pH for further experiments.
First of all, pH influence is related to the ionization state of the
catalyst surface including TiO2 and zeolite behavior in alkaline
and acidic media [3], as well as to that of the dye structure and
forms, e.g., acids and amines. pH changes can influence the
adsorption of the dye molecules onto the catalyst surface, an
important step for the photocatalytic oxidation. Based on the
results of the prior step of adsorption for 30 min required to reach
the steady-state, it can be noticed a low extent of the dye is
adsorbed on the Z-TiO2-N surface, with a maximum of adsorption
degree at pH 3. 10.2 mg dye/g catalyst was adsorbed at pH 3 in
comparison with 1.5 mg dye/g catalyst at pHs 6 and 9. This can be
explained by the repulsion decrease between the RY 125 anionic
azo dye, which contains the sulphonate groups as anionic form
Fig. 3. SEM images and EDX spectra (inset) of: Z-Na (a), Z-TiO2-N (b) and TiO2-N (c). within the whole range of investigated pH and the negatively
charged catalyst. This fact indicates that the photocatalytic
into absorbance using Kubelka–Munk (Eq. (1)): reaction occurred both on the catalyst surface and through
reaction of the dye with the photogenerated holes or hydroxyl
ð1  RÞ2 radicals in the bulk solution in the proximity of the catalyst
FðRÞ ¼ (1) surface.
2R
Also, taking into account the general mechanism of photo-
where R represents the reflectance. catalytic oxidation process, the generation of hydroxyl radicals can
274 E.C. Ilinoiu et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 270–278

Fig. 4. FT-IR spectra of: (a) Z-TiO2-N and (b) Z-Na.

be favored during the reaction of positive holes with hydroxyl


anions (OH).
From Fig. 7, it is noticed that at pH 3 and in comparison with
aromatic ring-opening, the discoloration process reaches its
maximum efficiency value in a shorter time of irradiation, and
no further changes in efficiency is observed after 150 min of
irradiation. At low pH value, the positive holes are considered the
major oxidative species that could form hydroxyl radicals (HO)
and interact with azo linkage, which is particularly susceptible to
be electrophilic attacked by hydroxyl radicals through decreasing
the electron densities at azo group [23]. Also, dissolved O2 itself can
accept conduction band electron and generate O2, which in turn
can be protonated to form HO2, favored by the acidic pH. At pH
values of 6 and 9, the efficiency of aromatic ring-opening is
exceeded by the discoloration efficiency only after 120 min
irradiation for pH 6, and after 180 min for pH 9. The discoloration
reaches its maximum efficiency after the formation of some
intermediary species, which require a longer time of reaction to be
Fig. 5. DRUV–VIS spectra of Z-Na (a) and Z-TiO2-N (b). further degraded. Also, the mineralization efficiency reaches the

Fig. 6. Changes of UV–VIS spectra of RY 125 dye solutions containing Z-TiO2-N Fig. 7. Efficiencies of RY 125 solution discoloration and aromatic ring-opening at pH
catalyst during UV irradiation: (1) initial RY 125 solution; (2) after 30 min 3, 6 and 9 during UV irradiation; initial dye concentration is 100 mg/L; 1 g/L catalyst
adsorption; (3) after 50 min irradiation; (4) after 120 min irradiation; (5) after loading; reaction temperature 20 8C. & – aromatic ring-opening, pH 3; & –
240 min irradiation; 100 mg/L initial dye concentration; 1 g/L catalyst loading; decolorisation, pH 3; ~ – aromatic ring-opening, pH 6; D – decolorisation, pH 6;  –
reaction temperature 20 8C; solution pH = 3. aromatic ring-opening, pH 9; o – decolorisation, pH 9.
E.C. Ilinoiu et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 270–278 275

Table 1
TOC efficiencies and mineralization coefficients for 100 mg/L RY 125 dye solution at
different pH values.

pH value hTOC (%) hTOC/h225 nm


3 35.79 0.59
6 20.18 0.43
9 12.77 0.29

highest value of 35.79% at pH 3 compared to 20.18% at pH 6 and


respective, 12.77% at pH 9 (see Table 1).
The effective mineralization performance was assessed as the
ratio between the TOC and the aromatic ring-opening efficiency
(hTOC/h225 nm). The value of this ratio close to 1 indicates that the
effective mineralization was achieved. In our study, the highest
ratio value of 0.59 is obtained for pH 3, which informed that the Fig. 9. Influence of initial dye concentration on discoloration and aromatic ring-
mineralization process is not effectively under the studied opening efficiency under UV irradiation; 1 g/L catalyst loading; reaction
conditions. temperature 20 8C; solution pH = 3; & – aromatic ring-opening, 100 mg/L
RY125; & – decolorisation, 100 mg/L RY125; ~ – aromatic ring-opening, 50 mg/
L RY125; D – decolorisation, 50 mg/L RY125;  – aromatic ring-opening, 25 mg/L
3.2.3. Effect of the catalyst dose
RY125; o – decolorisation, 25 mg/L RY125.
In order to determine the optimum catalyst loading, the
degradation of RY 125 dye was investigated using different
catalyst concentrations in the range of 0.5–2 g/L. The results are 3.2.4. Effect of the initial dye concentration
presented in Fig. 8, and it can be observed that increasing the The photodegradation efficiency depends on the initial dye
catalyst dose from 0.5 to 2 g/L lead to increasing the dye concentration based on two main aspects. The amount of dye
discoloration efficiency from 70.75% to 89.62%. This may be adsorbed on the catalytic surface increases with the dye
attributed to the enhanced generation of HO from the photo- concentration in relation with the available surface, which
catalyst surface based on increasing the photocatalyst amount in enhances the catalytic efficiency of the photocatalyst. On the
the reaction solution. It must be noticed that the same final other hand, higher concentration of the dye affects negatively the
discoloration efficiency after 240 min irradiation was reached for UV light penetrability in the dye solution. Also, the dye molecules
1 and 2 g/L photocatalyst, and the dye discoloration rate was may absorb a significant amount of light required for the catalyst
higher for 1 g/L photocatalyst. surface irradiation, which diminishes the catalytic activity [41].
A similar behavior is noticed in the case of aromatic ring- The influence of various initial dye concentrations on the
opening efficiency, where the efficiency increases with the photocatalytic discoloration and degradation has been investigat-
catalyst dose increasing up to 1 g/L. At higher catalyst dose of ed in the concentration range of 25–100 mg/L RY 125 dye. The
2 g/L the efficiency decreases to 63.11% versus 77.92% reached results are presented in Fig. 9, which are in accordance with other
at 1 g/L photocatalyst. This behavior might be due to the literature data [3,4]. The increasing dye concentration leads to
enhancement of light reflectance by the catalyst and decreas- decreasing degradation efficiency.
ing the light penetration [4]. A similar behavior under the Also, TOC monitoring for all initial dye concentrations showed
condition of the catalyst excess was reported in other studies that the best mineralization efficiency is obtained for 25 mg/L RY
[18]. Based on these results, the catalyst dose of 1 g/L was 125 (Table 2).
chosen as optimum for an efficient photodegradation of RY 125 Based on the values of the ratio between mineralization and
dye solution. degradation efficiency, the ratio value of 0.83 informs about an
effective mineralization process in the case of low dye concentra-
tion, in comparison with higher dye concentrations.

3.2.5. Photodegradation kinetics


The dependencies of the dye aromatic ring opening, discolor-
ation and mineralization rate on the irradiation time have been
described by the Langmuir–Hinshelwood kinetic model [42]
described by Eq. (2), which can be simplified to a pseudo-first-
order equation, as presented in Eq. (3):

dC kKC
r¼ ¼ (2)
dt 1 þ KC
 
C0
ln ¼ kKt ¼ kapp t (3)
Ct

Table 2
TOC efficiencies and mineralization coefficients for different initial RY 125 dye
concentrations.
Fig. 8. Influence of catalyst dose on discoloration and aromatic ring-opening
efficiency under UV irradiation; 50 mg/L initial RY 125 dye concentration; reaction Initial RY 125 dye concentration (mg/L) hTOC (%) hTOC/h225 nm
temperature 20 8C; solution pH = 3; & – aromatic ring-opening, 2 g Z-TiO2-N/L; & –
100 35.79 0.59
decolorisation, 2 g Z-TiO2-N/L; ~ – aromatic ring-opening, 1 g Z-TiO2-N/L; D –
50 48.00 0.61
decolorisation, 1 g Z-TiO2-N/L;  – aromatic ring-opening, 0.5 g Z-TiO2-N/L; o –
25 76.29 0.83
decolorisation, 0.5 g Z-TiO2-N/L.
276 E.C. Ilinoiu et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 270–278

Fig. 11. The photodegradation efficiencies of RY 125 dye solution under UV and VIS
irradiation. & – aromatic ring-opening by UV irradiation; & – decolorisation by UV
irradiation;  – aromatic ring-opening by VIS irradiation; o – decolorisation by VIS
irradiation.

Fig. 10. The pseudo-first-order kinetics of 25 mg/L RY 125 photodegradation on Z-


TiO2-N, pH = 3. Fig. 11 illustrates the photodegradation efficiency under VIS in
comparison with UV light, and it can be observed that the results
are comparable.
where r is the rate of dye degradation, discoloration and For an initial dye concentration of 25 mg/L, at pH 3 after
mineralization (mg/L/min), C0 is the initial dye concentration 240 min irradiation time, the aromatic-ring opening efficiency of
(mg/L), Ct is the concentration of the dye at time t (mg/L), t is the 91.64% and the discoloration efficiency of 91.60% were reached
irradiation time (min), k is the reaction rate constant (min1) and under UV irradiation, and the aromatic-ring opening efficiency of
kapp is the apparent rate constant (min1). 85.91% and the discoloration efficiency of 94.85% were achieved
The apparent rate constant for each dye degradation, discolor- under VIS irradiation.
ation and mineralization, kapp, was calculated from the intercept of The photocatalytic activity of the Z-TiO2-N hybrid material was
the plot of ln(C0/Ct) against time, t (Fig. 10). compared with Z-Na and TiO2-N for 25 mg/L RY 125 dye solution
The kinetics results presented in Tables 3 and 4 showed that the in terms of degradation efficiency, and the results are gathered in
rates of the aromatic ring-opening, discoloration and mineraliza- Table 5. It can be noticed a synergic effect of the hybrid material for
tion processes depend on pH and dye initial concentration. the photocatalytic degradation of RY 125 dye solution taking into
For all the experiments, the discoloration occurred the fastest, account the content of 2 wt%, TiO2.
and the mineralization process was the slowest. Though, the rate of Though, a significant difference between the two irradiation
the mineralization process was better for low dye concentration domains was noticed for the photodegradation rate. The dye
than for higher dye concentrations. photodegradation process under UV irradiation occurred faster
than VIS irradiation, the rate constants are gathered in Table 6.
3.3. VIS photocatalytic degradation Degradation efficiency under VIS light is achieved not only by
the catalyst absorption band shifting into the visible region,
Because ultraviolet photons are limited in both solar light and capable to generate pairs of electrons–holes, which can participate
indoor illumination, the VIS irradiation was performed in order to directly in the photocatalytic reactions, but also by the so-called
assess the effectiveness of the Z-TiO2-N in the photocatalytic ‘‘photosensitized oxidation’’ mechanism, which provides a differ-
degradation of RY 125 dye solution. ent pathway versus that involved under UV light irradiation
[19,43]. The photocatalytic oxidation is superior to the photo-
sensitizing oxidation mechanism, but the later can help in
improving the overall efficiency and can also provide a synergism
Table 3 effect when VIS light is involved. This aspect is proved by the
Apparent rate constants for discoloration, aromatic ring-opening and mineraliza-
tion (100 mg/L RY 125 dye solution).

pH kapp,225 nm R2 kapp,388 nm R2 kapp,TOC R2 Table 5


value (min1) (min1) (min1) Degradation and discoloration efficiencies for different types of catalysts.

3 0.0053 0.9844 0.0110 0.9894 0.0015 0.9872 Catalyst type h225 nm (%) h388 nm (%)
6 0.0043 0.9689 0.0044 0.9914 0.0011 0.9634
TiO2-N 92.23 98.10
9 0.0038 0.9823 0.0025 0.9973 0.0002 0.9956
Z-TiO2-N 85.91 94.85
Z-Na 35.85 36.03
Table 4
Apparent rate constants for discoloration, aromatic ring-opening and mineraliza-
tion (pH = 3). Table 6
2 2 2
Apparent rate constants for UV and VIS degradation processes at pH 3 for 25 mg/L
RY initial kapp,225 nm R kapp,388 nm R kapp,TOC R
RY 125 dye solution.
concentration (min1) (min1) (min1)
(mg/L) Irradiation kapp,225 nm R2 kapp,388 nm R2
type (min1) (min1)
100 0.0053 0.9844 0.0110 0.9894 0.0015 0.9872
50 0.0078 0.9851 0.0164 0.9917 0.0023 0.9761 UV 0.0196 0.9896 0.0239 0.9410
25 0.0196 0.9896 0.0239 0.9410 0.0060 0.9785 VIS 0.0071 0.9852 0.0126 0.9385
E.C. Ilinoiu et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 270–278 277

apparent rate constants determined as kapp, 225 nm = 0.0034 min1 [8] Szygula A, Gnibal E, Ruiz M, Sastre AM. Removal of an anionic dye (Acid Blue
92) by coagulation–flocculation using chitosan. J Environ Manage
and kapp, 388 nm = 0.0029 min1 for photolysis process application 2009;90:2979–86.
in the degradation of 25 mg/L RY 125 dye solution in the absence [9] El Qada EN, Allen SJ, Walker GM. Adsorption of basic dyes from aqueous
of the photocatalyst under VIS irradiation. solution onto activated carbons. Chem Eng J 2008;135:174–84.
[10] Eren E, Cubuk O, Ciftci H, Eren B, Caglar B. Adsorption of basic dye from
These results revealed that Z-TiO2-N exhibited useful char- aqueous solutions by modified sepiolite: equilibrium, kinetics and thermody-
acteristics for application in the photocatalytic degradation of RY namics study. Desalination 2010;252:88–96.
125 dye solution at medium levels of concentrations under solar [11] Qiu M, Qian C, Xu J, Wu J, Wang G. Studies on the adsorption of dyes into
clinoptilolite. Desalination 2009;243:286–92.
irradiation. [12] El-Desoky HS, Ghoneim MM, Zidan NM. Decolorization and degradation of
Ponceau S azo-dye in aqueous solutions by the electrochemical advanced
4. Conclusions Fenton oxidation. Desalination 2010;264:143–50.
[13] Riga A, Soutsas K, Ntampegliotis K, Karayannis V, Papapolymerou G. Effect of
system parameters and of inorganic salts on the decolorization and degrada-
The hybrid material based on natural zeolite modified with tion of Procion H-exl dyes. Comparison of H2O2/UV, Fenton, UV/Fenton, TiO2/
nitrogen-doped TiO2 photocatalyst was successfully synthesized UV and TiO2/UV/H2O2 processes. Desalination 2007;211:72–86.
[14] Kim TH, Park C, Kim S. Water recycling from desalinization and purification
by solid-state reaction under microwave-assisted hydrothermal process of reactive dye manufacturing industry by combined membrane
conditions. The presence of crystalline pure anatase form of TiO2 filtration. J Clean Prod 2005;13:779–86.
doped with nitrogen onto zeolite was proved by XRD and nitrogen [15] Dixit A, Tirpude AJ, Mungray AK, Chakraborty M. Degradation of 2,4 DCP by
sequential biological-advanced oxidation process using UASB and UV/TiO2/
presence was confirmed by EDX analysis. The DRUV–VIS spectra H2O2. Desalination 2011;272:265–9.
indicated that hybrid materials exhibited the absorption bands [16] Ghaly MY, Farah JY, Fathy AM. Enhancement of decolorization rate and COD
slightly shifted to the visible range. The photocatalytic activity of removal from dyes containing wastewater by the addition of hydrogen
peroxide under solar photocatalytic oxidation. Desalination 2007;217:74–84.
the zeolite modified with nitrogen-doped TiO2 photocatalyst
[17] Zhang G, Choi W, Han Kim S, Bong Hong S. Selective photocatalytic degrada-
under UV and VIS irradiation was determined for the degradation tion of aquatic pollutants by titania encapsulated into FAU-type zeolites. J
of RY 125 dye solution. The solution pH value of 3 and the Hazard Mater 2011;188:198–205.
photocatalyst dose of 1 g/L were established as optimal conditions [18] Soutsas K, Karayannis V, Poulios I, Riga A, Ntampegliotis K, Spiliotis X,
Papapolymerou G. Decolorization and degradation of reactive azo dyes via
for photocatalysis application in the degradation of RY 125 dye heterogeneous photocatalytic processes. Desalination 2010;250:345–50.
solution. All experimental studies including kinetics revealed that [19] Carp O, Huisman CL, Reller A. Photoinduced reactivity of titanium dioxide.
the degradation efficiency expressed in terms of discoloration, Prog Solid State Chem 2004;32:33–177.
[20] Morikawa T, Asahi R, Ohwaki T, Aoki K, Suzuki K, Taga Y. Visible-light
aromatic ring-opening and mineralization depended strongly on photocatalyst – nitrogen-doped titanium dioxide. R&D Rev Toyota CRDL
the initial RY 125 dye solution concentrations. The photocatalyst 2005;40:45–50.
exhibited a good performance for discoloration and aromatic ring- [21] Narayana RL, Matheswaran M, Aziz AA, Saravanan P. Photocatalytic decolour-
ization of basic green dye by pure and Fe, Co doped TiO2 under daylight
opening processes at all studied concentrations (25, 50 and illumination. Desalination 2011;269:249–53.
100 mg/L) but an effective mineralization was reached only for [22] Sathishkumara P, Anandana S, Maruthamuthub P, Swaminathanc T, Zhoud M,
25 mg/L RY 125 dye solution. The results of photocatalyst activity Ashokkumar M. Synthesis of Fe3+ doped TiO2 photocatalysts for the visible
assisted degradation of an azo dye. Colloids Surf A 2011;375:231–6.
under VIS irradiation revealed a practical utility of the hybrid [23] Sun J, Qiao L, Sun S, Wang G. Photocatalytic degradation of Orange G on
material based on natural zeolite modified with nitrogen-doped nitrogen-doped TiO2 catalysts under visible light and sunlight irradiation. J
TiO2 photocatalyst for the degradation of low concentrations of RY Hazard Mater 2008;155:312–9.
[24] Wawrzyniak B, Morawski AW. Solar light induced photocatalytic decomposi-
125 dye solution.
tion of two azo dyes on new TiO2 photocatalyst containing nitrogen. Appl Catal
B 2006;62:150–8.
Acknowledgements [25] Fujishima A, Zhanf X, Tryk DA. TiO2 photocatalysys and related surface
phenomena. Surf Sci Rep 2008;63:515–82.
[26] Thompson TL, Yates JT. Surface science studies of the photoactivation of TiO2-
This work was supported by the strategic grant POSDRU 2009 new photochemical processes. Chem Rev 2006;106:4428–53.
project ID 50783 of the Ministry of Labor, Family and Social [27] Behnajady MA, Modirshahla N, Daneshvar N, Rabbani M. Photocatalytic
degradation of an azo dye in a tubular continuous flow photoreactor with
Protection, Romania, co-financed by the European Social Fund – immobilized TiO2 on glass plates. Chem Eng J 2007;127:167–76.
Investing in People, and by the Sectorial Operational Programme [28] Huanga CH, Changa KP, Oua HD, Chiangb YC, Changc EE, Wang CF. Characteri-
Human Resources Development (SOP HRD), financed from the zation and application of Ti-containing mesoporous silica for dye removal
with synergistic effect of coupled adsorption and photocatalytic oxidation. J
European Social Fund and by the Romanian Government under the
Hazard Mater 2011;186:1174–82.
contract number POSDRU/89/1.5/S/63700. [29] Ying Shan A, Idaty Mohd Ghazi T, Abdul Rashid S. Immobilization of titanium
dioxide onto supporting materials in heterogeneous photocatalysis: a review.
References Appl Catal A 2010;389:1–8.
[30] Sökmen M, Tatlidil İ, Breen C, Clegg F, Kurtuluş Buruk C, Sivlim T, Akkan ŞA.
new nano-TiO2 immobilized biodegradable polymer with self-cleaning prop-
[1] Gomes Silva C, Wong W, Luis Faria J. Photocatalytic and photochemical erties. J Hazard Mater 2011;187:199–205.
degradation of mono-, di- and tri-azo dyes in aqueous solution under UV [31] Wang CC, Lee CK, Lyu MD, Juang LC. Photocatalytic degradation of C.I. Basic
irradiation. J Photochem Photobiol A Chem 2006;181:314–24. Violet 10 using TiO2 catalysts supported by Y zeolite: an investigation of the
[2] Elwakeel KZ, Rekaby M. Efficient removal of Reactive Black 5 from aqueous effects of operational parameters. Dyes Pigm 2008;76:817–24.
media using glycidyl methacrylate resin modified with tetraethelenepenta- [32] Nikazar M, Gholivand K, Mahanpoor K. Photocatalytic degradation of azo dye
mine. J Hazard Mater 2011;188:10–8. Acid Red 114 in water with TiO2 supported on clinoptilolite as a catalyst.
[3] Konstantinou IK, Albanis TA. TiO2-assisted photocatalytic degradation of azo Desalination 2008;219:293–300.
dyes in aqueous solution: kinetic and mechanistic investigations. A review. [33] Inglezakis VJ. The concept of capacity in zeolite ion-exchange systems. J
Appl Catal B 2004;49:1–14. Colloid Interface Sci 2005;281:68–79.
[4] Muruganandham M, Shobana N, Swaminathan M. Optimization of solar [34] Nagaveni K, Sivalingam G, Hegde MS, Madras G. Solar photocatalytic degra-
photocatalytic degradation conditions of Reactive Yellow 14 azo dye in dation of dyes: high activity of combustion synthesized nano TiO2. Appl Catal B
aqueous TiO2. J Mol Catal A Chem 2006;246:154–61. 2004;48:83–93.
[5] Bali U, Catalkaya E, Sengul F. Photodegradation of Reactive Black 5, Direct Red [35] Zhang X, Yao B, Zhao L, Liang C, Mao Y. Electrochemical fabrication of single-
28 and Direct Yellow 12 using UV, UV/H2O2 and UV/H2O2/Fe2+: a comparative crystalline anatase TiO2 nanowire arrays. J Electrochem Soc 2001; 148:
study. J Hazard Mater 2004;114:159–66. G398- G400.
[6] Rauf MA, Salman Ashraf S. Fundamental principles and application of hetero- [36] Zhao D, Zhou J, Liu N. Preparation and characterization of Mingguang paly-
geneous photocatalytic degradation of dyes in solution. Chem Eng J gorskite supported with silver and copper for antibacterial behavior. Appl Clay
2009;151:10–8. Sci 2006;33:161–70.
[7] Gaoa B, Liua B, Chena T, Yuea Q. Effect of aging period on the characteristics and [37] Rodriguez-Fuentes G, Ruiz-Salvador AR, Mir M, Picazo O, Quintana G, Delgado
coagulation behavior of polyferric chloride and polyferric chloride–polyamine M. Thermal and cation influence on IR vibrations of modified natural clin-
composite coagulant for synthetic dying wastewater treatment. J Hazard optilolite. Microporous Mesoporous Mater 1998;20:269–81.
Mater 2011;187:413–20.
278 E.C. Ilinoiu et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 270–278

[38] Lazau C, Ratiu C, Orha C, Pode R, Manea F. Photocatalytic activity of undoped [41] Sabona N, Selvam K, Swaminathan M. Optimization of photocatalytic degra-
and Ag-doped TiO2-supported zeolite for humic acid degradation and miner- dation conditions of Direct Red 23 using nano-Ag doped TiO2. Sep Purif
alization. Mater Res Bull 2011;46:1916–21. Technol 2008;62:648–53.
[39] Li F, Jiang Y, Yu L, Yang Z, Hou T, Sun S. Surface effect of natural zeolite [42] Gupta AK, Pal A, Sahoo C. Photocatalytic degradation of a mixture of Crystal
(clinoptilolite) on the photocatalytic activity of TiO2. Appl Surf Sci Violet (Basic Violet 3) and Methyl Red dye in aqueous suspensions using Ag+
2005;252:1410–6. doped TiO2. Dyes Pigm 2006;69:224–32.
[40] Sahel K, Perol N, Chermette H, Bordes C, Derriche Z, Guillard C. Photocatalytic [43] Rajeshwar K, Osugi ME, Chanmanee W, Chenthamarakshan CR, Zanoni MVB,
discoloration of Remazol Black 5 (RB5) and Procion Blue MX-5B-Isotherm of Kajitvichyanukul P, Krishnan-Ayer R. Heterogeneous photocatalytic treatment
adsorption, kinetic and discoloration and mineralization. Appl Catal B of organic dyes in air and aqueous media. J Photochem Photobiol C
2007;77:100–9. 2008;9:171–92.

You might also like