Vous êtes sur la page 1sur 38

Accepted Manuscript

Title: Photocatalytic degradation of textile dyes on


Cu2 O-CuO/TiO2 anatase powders
Author: A. Ajmal I. Majeed R.N. Malik M. Iqbal M. Arif
Nadeem I. Hussain S. Yousuf Zeshan G. Mustafa M.I. Zafar
M. Amtiaz Nadeem
PII:
DOI:
Reference:

S2213-3437(16)30121-X
http://dx.doi.org/doi:10.1016/j.jece.2016.03.041
JECE 1043

To appear in:
Received date:
Revised date:
Accepted date:

7-12-2015
3-3-2016
27-3-2016

Please cite this article as: A.Ajmal, I.Majeed, R.N.Malik, M.Iqbal, M.Arif Nadeem,
I.Hussain, S.Yousuf, Zeshan, G.Mustafa, M.I.Zafar, M.Amtiaz Nadeem, Photocatalytic
degradation of textile dyes on Cu2O-CuO/TiO2 anatase powders, Journal of
Environmental Chemical Engineering http://dx.doi.org/10.1016/j.jece.2016.03.041
This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.

Photocatalytic degradation of textile dyes on Cu2O-CuO/TiO2 anatase


powders
A. Ajmala, I. Majeedb, R. N. Malika, M. Iqbala, M. Arif Nadeemb, I. Hussaina, S. Yousufa,
Zeshana, G. Mustafaa, M. I. Zafare, M. Amtiaz. Nadeema* NadeemMI@SABIC.com,
amtiaznadeem@yahoo.co.uk
a

Department of Environmental Sciences, Quaid-i-Azam University, Islamabad 42320,

Pakistan.
b

Department of Chemistry, Quaid-i-Azam University, Islamabad 42320, Pakistan.

Department of Chemistry, Lahore University of Management Sciences, Lahore, Pakistan.

Corresponding Author at: Present Adress: Fundamental Catalysis Division, Saudi Basic

Industries Corporation P.O Box 4545-4700, Thuwal 23955, Saudi Arabia.

Graphical Abstract

Highlights
Non noble metal supported Cu2O-CuO/TiO2 catalysts have been prepared and
characterized
Photodegradation of textile dyes on Cu2O-CuO/TiO2 is studied
TiO2 anatase is preferred over P25 as former can be obtained with much higher
surface area

Abstract
TiO2 anatase photocatalysts with ca. 2, 4 and 8 wt. % copper metal loading have been
employed for the photocatalytic degradation of textile dyes (one of the major contaminants in
effluent from textile industries around the world) and their degradation efficiencies compared
with benchmark Degussa P25. In order to investigate their surface and bulk composition, all
catalysts were characterized using XRD, UV-Vis, XPS and TEM techniques. TEM and XPS
results confirmed the presence of Cu in its oxide form (Cu2O-CuO). Under optimized
conditions (i.e. catalyst dose of 125 mg L-1 and initial dye concentration of 5mg L-1), Degussa
P25 showed 76% degradation efficiency for primary reactive blue 49 (RB 49) dye in 80
minutes. However, under the same optimized conditions, a complete primary RB 49 dye
removal was observed at ca. 4 wt. % Cu2O-CuO/TiO2 catalyst. Kinetics analysis of
photodegradation (kd) nearly followed a pseudo-first-order kinetic model with a regression
coefficient (R2) value in the range of 0.97-0.86 for 4 wt. % Cu2O-CuO/TiO2 catalyst. These
results indicate that non noble supported TiO2 anatase may be a better choice for
photocatalytic dye degradation as compared to other commercially available photocatalysts.
Keywords: Industrial effluent; Textile dyes; Photocatalytic activity; Cu2O-CuO/TiO2;
Degradation efficiency

1. Introduction
The discharge of synthetic and chemically active dyes in to industrial wastewater is a major
environmental concern due to their carcinogenic nature [1]. Further, the development of color
due to the presence of these dyes can block sunlight and decrease oxygen dissolution
resulting in low photosynthetic activity for aquatic biota [2]. It is, therefore, direly important
to treat these colored effluents before they enter into aquatic bodies. Azo dyes are the largest
group of synthetic dyes and are characterized by nitrogen to nitrogen double bonds (N=N).
Azo dyes are currently used in a diverse range of industries including paper, textile, food,
additives, cosmetics, xerography, laser materials and laser printing industries [3]. It should be
noted that nearly 15% of worlds overall toxic chemicals including dye products are directly
disposed into the natural environment without any prior treatment [4, 5].
Several treatment approaches have been reported in literature regarding decontamination and
handling of colored effluents such as ion flotation [6], coagulation [7], adsorption [8], and
sedimentation [9]. However, all of these techniques end up in producing secondary waste
product which needs further treatment. To combat this issue, a relatively new and promising
set of techniques, called Advanced Oxidation Processes (AOPs) have been developed and
employed for dye contaminated industrial effluents treatment. Heterogeneous photocatalysis,
a promising approach of AOPs, has successfully been developed and improved in recent
years for better treatment of dyes in effluents [10, 11]. AOPs are considered clean
technologies for the treatment of polluted waters that apply the concept of producing
hydroxyl radicals (HO) that attack the organic pollutants. The efficiency of AOPs is based on
the generation of these highly reactive radicals that are unselective and powerful oxidizing
species (Eo = 2.80 V) and can degrade indiscriminately micro-pollutants with reaction rate
5

constants usually around 109 L mol1 s1 yielding CO2, H2O and inorganic ions as final
products. A higher biodegradability and/or lower toxicity of the reaction byproducts, in
comparison with the parent compounds, are desirable benefits of applying AOPs to treat
wastewaters. AOPs can be applied as post-treatment or pre-treatment of biological processes.
The integration of different AOPs in a sequence of complementary processes is also a
common approach to achieve a biodegradable effluent that can be further treated by a cheaper
and conventional biological process, reducing the residence time and reagent consumption in
comparison with AOPs alone [11]. Conventional AOPs can be classified as homogeneous
and heterogeneous processes, depending on whether they occur in a single phase or they
make use of a heterogeneous catalyst like metal supported catalysts, carbon materials or
semiconductors such as TiO2, ZnO, and WO3. While homogeneous processes are
characterized by chemical changes depending on the interactions between the chemical
reagents and target compounds only, heterogeneous processes also depend on the adsorption
of reactants and desorption of products that occur at the active sites of the catalyst surface. As
the reaction occurs, the products are desorbed and new species can adsorb on the active sites,
so that the surface characteristics and the pore structure of the catalyst strongly affect
catalysts efficiency and stability [12]. One of the AOPs is a photocatalysis process which
involves the complete photo-mineralization or oxidation by generation of hydroxyl radicals
that reacts with most of the organics and mineralizes them into CO2, H2O and inorganic ions
including PO43, NO3, and halide ions without further production of intermediate toxic
byproducts [13, 14]. Photocatalytic degradation generally involves the complete photomineralization or oxidation of the target contaminants into H 2O, CO2. Among numerous
types of heterogeneous photocatalysts, TiO2 mediated photocatalytic oxidation has received
much attention owing to its strong oxidizing power, long-term photostability and non-toxicity
[15].
6

Nevertheless, TiO2 possesses oxidizing property that can degrade contaminants efficiently, it
has a large band gap (it can only be excited under UV photons) and fast charge carrier
recombination limiting its practical application. About 45% of the entire solar spectrum falls
within UV range that can excite TiO2 based catalysts to exhibit photocatalysis with relatively
lower activities [16-19]. To overcome these limitations, researchers have coupled a variety of
semiconductors, metals, noble metals and non-metals with TiO2. These strategies can not
only bring part of absorption in visible light region but also reduce charge carrier
recombination due to Schottky barrier formation at TiO 2-metal interface [19, 20]. For
example, incorporation of transition metals into crystal structure of TiO2 may result into
narrowing the band gap by shifting the absorbance towards the visible region [19]. Cu (II) has
been reported to improve photocatalytic activity by modifying TiO2 valence band spectrum
and widening the visible light absorption ability. Chiang and co-workers [21] studied the
photocatalytic activity of photodeposited CuO onto the surface of Degussa P25 particles to
oxidize cyanide. It was found that CuO resulted into red shift of absorption spectrum of TiO2.
However, possible limitations of such modifications can be photo-corrosion and
recombination of prompted charges at metal surface [19].
The detailed investigations regarding TiO2 coupled with non-noble transition metals mediated
photodegradation is important for enhancing the overall dye treatment process at affordable
cost. This study has been designed with an aim to evaluate the efficiency of Cu supported
TiO2 based photocatalysts for efficient degradation of textile dyes under controlled
environmental conditions. The experimental parameters involved in controlling the extent of
dye photodegradation reaction have been systematically evaluated.

2. Experimental
2.1 Catalyst preparation
7

TiO2 anatase (Anatase > 99 %, crystalline size 2-10 nm) and copper chloride (CuCl2. 2H2O,
99.0% purity) were purchased from Sigma Aldrich while Degussa P25 was obtained from
Evonik industries AG (63403 Hanau), Germany. Ammonium hydroxide (NH4OH, ACS
reagent, 28.0-30.0% NH3 basis) and 0.1 M HCl (HCl, ACS reagent, 37%) were used for pH
controlled experiments. All reagents used were of analytical grade and were used without any
further purification. Copper was deposited on TiO2 nanoparticles using a combination of
impregnation and precipitation-deposition method [22]. Briefly, 1 g of TiO2 (Anatase) and
pre-determined amount of CuCl2 salt for 2, 4, 8 wt. % Cu loadings were thoroughly dissolved
in 100 mL of deionized water and kept at room temperature for 24 hours to allow Cu ions to
impregnate into TiO2. Then the solution pH was raised to 8.5 using NH4OH solution. The
final solution was stirred for one hour at 80 C and allowed to stabilize for 24 hours at room
temperature. The solution was agitated and filtered (Grade 2 filter paper, Whatman; particle
retention 8m) and the obtained precipitates were repeatedly washed with deionized water.
Catalysts were dried at 90 C for three hour and annealing at 400 C for two hours.
2.2 Characterization Technique
Powder X-ray diffraction (PXRD) spectrum of bare, Cu supported TiO2 anatase and Degussa
P25 composites were obtained on a Shimadzu XRD-6000 instrument by using Cu-K
radiation ( = 1.52 A) operating at a generator voltage of 40 kV and an emission current of
40 mA. X-ray diffraction patterns were acquired over 2 range of 1080 at a sampling scan
width of 5/min and a step time equal to 0.5 s [23]. Dimensional morphology and size of
synthesized Cu supported TiO2 and commercially available Degussa P25 were measured
using Philips model PW 6030 CM12/STEM electron microscope operating at 120 kV with
filament emission current of 10 mA.

XPS spectra of photocatalysts were collected on a Kratos Axis Ultra spectrometer using
monochromatized Al K Xrays (h = 1486.6 eV) equipped with a hemispherical electron
energy analyzer operated in the hybrid lens mode at 90 take off angle with respect to the
specimen surface. The binding energy scale was calibrated using adventitious hydrocarbon
referencing (C1s = 285.0 eV). Survey spectra were collected at an analyzer pass energy of 80
eV over the binding energy range from 0 to 1200 eV (5 sweeps/25 ms dwell time). The
relative concentration of the identified elements were calculated using peak areas of Ti2p,
O1s and Cu2p core level signals. Relative sensitivity factors, based on instrument modified
Schofield cross-sections were used in the quantification procedure (7.81, 2.93 and 25.39
respectively). The chemical states of Ti, O and Cu were determined from narrow scan spectra
taken at pass energy of 20 eV over the Ti2p, O1s, and Cu2p regions (30 sweeps/150 ms dwell
time), respectively.
2.3 Photodegradation experiments
Three synthetic textile dye samples including C.I Reactive Blue 49 (RB 49), C.I Reactive
Red 24 (RR 24) and C.I Reactive Yellow 160 (RY 160) were obtained from Sundar Industrial
Estate (SIE) textile garment zone, Lahore, Pakistan (C.I. = Colour Index). Table 1 shows the
physio-chemical characteristics of textile dye samples. The chemical structures of dyes are
given in Fig.1.
Photocatalytic reactions were carried in an open top rectangular Pyrex photoreactor having
2.5 (Height) 6 (Width) 12 (Length) dimensions placed in a bench top fume hood. UV
irradiating tubes (length = 12 ) were obtained from Phillips Company having an output
potential 2 5 = 30 watt at = 365 nm and were assembled at a distance of 12 cm from the
Pyrex reactor containing reaction solution. UV-Vis spectrophotometer (DR5000TM, Hach
Company, USA) was used to monitor the dye degradation or removal process. Typically,
9

during the course of reaction, 1mL of dye samples was withdrawn using a micro pipette with
an interval of 10 min, centrifuged and filtered to ensure catalyst removal from the dye
solution. Maximum dye absorbance wavelength (

max)

was used for absorbance

measurement. The dye degradation efficiency was calculated using equation 1 [24].
Dye removal efficiency (%) = [A At / A] 100

(1)

Where, A is the initial absorbance of the dye solution before reaction and At is the dye
absorbance during reaction at time t.
In a typical degradation experiment, 250 mL of each of RB 49, RR 24 and RY 160 dye
solutions (5 mgL-1) was reacted for 2 h under UV light exposure with optimized amount (125
mgL-1) of catalysts (Cu supported TiO2 and Degussa P25) at natural dye pH 6.5-7.5. Air was
continuously supplied (60 mL/min) to facilitate atmospheric oxygen for efficient oxidation of
dyes. Temperature for all reactions was maintained at room temperature (30.0 0.1 oC).
Relative dye adsorption ability of a given photocatalyst was measured under similar
conditions except the mixture was stirred under dark conditions for 30 min. The smaller
interval of time was used to prevent any contribution from photocatlaytic dye degradation
under possible presence of diffused light. The control solution was prepared by the similar
method without any catalyst addition having exactly same initial dye concentration.
Experiments to study effect of catalyst and dye amount were performed under similar
conditions except irradiation time was reduced to 80 min. Catalyst amount was varied from
25-150 mgL-1 while that of dye was varied from 5-25 mgL-1. The change in the chemical
oxygen demand (COD) of dye samples before and after degradation experiments was
measured to investigate the extent of dye mineralization and any possible amount of organic
intermediate formation using back titration method. The titration involving dichromate as an
oxidant and ferrous ammonium sulphate as a standard solution [25].
10

3 Results and Discussion


3.1 Characterization of Photocatalysts
The photocatalysts were characterized using powder X-ray diffraction (PXRD), Transmission
electron microscopy (TEM), UV-Vis spectroscopy and X-ray photoelectron spectroscopy
(XPS).
3.1.1 Catalyst bulk composition
The XRD patterns of pure TiO2 anatase, Degussa P25 and Cu supported TiO2 calcined at 400

C under air for 3h are shown in Fig. 1. The bare TiO2 showed diffraction peaks at 2 values

of 25.4 (101), 37.9 (004), 48.2 (200), 54.0 (105) and 55.1 (211), indicating well crystallized
pure form of anatase. Average crystalline size of catalysts was calculated using Scherer
equation. XRD and TEM results showed that TiO2 anatase particles has an average particle
size of 80 nm and possess a wide particle size distribution while commercially available
Degussa P25 has an average particle size of 25 nm. In TEM images, due to higher atomic
mass of Cu and higher refractive index of Cu2O-CuO (as confirmed by XPS) compared to
TiO2, the presence of dark spots on large bright TiO2 particle confirmed the presence of Cu
species at the TiO2 surface indicated by the white arrow (Fig. 2a). The TEM analysis of
Degussa P25 showed that it contained higher concentration of anatase phase (smaller round
particle) compared to rutile phase (larger and rectangular particles). UV-Vis spectra indicated
the sharp absorption edge at 387.5 nm corresponding to a typical TiO2 anatase band gap of
3.2 eV. The slight absorption in visible region in the case of Cu2O-CuO/TiO2 catalysts are Cu
d-d (650-900 nm) transitions and Cu/TiO2 interfacial charge transfer (380-550 nm) which
must not be confused with band gap excitations as they do not add to the catalyst activity
under visible light irradiation [26, 27]. BET surface area of P25 and anatase catalysts was
comparable and was found to be 49.85 and 45.25 m2g-1, respectively.
11

3.1.2 Catalyst surface composition


XPS analysis of 8 wt. % Cu supported on TiO2 catalyst is presented in Fig. 3. The relative
concentration of each element (Ti, O, C and Cu) in the nearsurface region of sample was
calculated from Ti 2p, O 1s, C 1s and Cu 2p corelevel peak areas. The surface elemental
composition was found to be 20.4 wt. %, 43.3 wt. %, 32.6 wt. %, and 3.86 wt. % for Cu, Ti,
O and C, respectively. The apparent large concentration of Cu can be attributed to the fact
that almost all Cu atoms present on the surface were analyzed whereas, all of Ti and O atoms
owing to large TiO2 particle size (as shown by the TEM analysis) cannot analyzed using XPS
technique. The similar observation has been noted in the case of Au/TiO 2 [28].
The Ti 2p3/2 and Ti 2p1/2 peaks (Fig. 3) were located at the binding energies of 459.1 eV and
464.9 eV respectively consistent with Ti+4 values in TiO2 lattice [29]. Core level spectrum of
O1s (Fig. 3), due to its metallic oxide Ti-O bond, showed its main peak at 529.9 eV. Two
broad peaks at 531 eV and 532.5 eV may be assigned to oxygen atoms of hydroxyl and
carboxyl groups, respectively [30]. The core level spectrum of C1s main peak at 285 eV was
observed corresponding to the adsorbed carbon at the surface of samples and peaks at 286.5
eV and 289.2 eV corresponded to C-O bonds of hydroxyl and carboxylic species,
respectively [31]. The peak fitting and position of Cu 2p XPS spectra indicated that there are
two kinds of Cu species. Cu 2p3/2 peaks at binding energy values of 932.9 eV and 934.2 eV
corresponds to copper in Cu2O and CuO form, respectively [32]. Absence of any emission
peak around 930 eV clearly indicates the absence of copper in metallic form. The presence of
shake up structure clearly indicates that Cu2+ is the dominant copper species. From the XPS
analysis it is clear that catalyst surface is oxidized to some extent and functional groups like
OH and COOH are embedded during calcination at 400 oC.
3.2 Effect of irradiation and catalyst
12

Fig. 4. presents the effect of TiO2 anatase and Degussa P25 addition on individual dye (RB
49, RR 24 and RY 160) absorption spectrum in an aqueous solution in the presence of UV
radiation for 80 minutes. Reduction in the dye absorbance was an indication of degradation of
main dye chromophores [33]. The amount of dye degradation as a function of irradiation time
is shown in Fig. 5. It can be noted that dye degradation efficiency of TiO 2 anatase is different
than Degussa P25 for same dye type. At Degussa P25, RY 160 dye was degraded up to
69.3% followed by RB 49 dye with 64.2% and RR 24 dye with 53.9% degradation efficiency
as presented in Fig. 5a. However, RB 49 dye was found highly reactive towards TiO 2 anatase
surface indicating maximum degradability with 33.8% (Fig. 5b) while RY 160 showed
degradation up to 28.1% at TiO2 anatase (Fig. 5b). The order of dye degradation efficiency
may be related to their enhanced adsorption as explained latter under section 3.3.
As RB 49 dye showed highest reactivity on TiO2 anatase surface, it was used to further
investigate the effect of amount of Cu loading to obtain the optimized degradation efficiency
as indicated in Fig. 6. A maximum degradation of 76% was recorded at 4 wt. % Cu2OCu2O/TiO2 after 2 h of UV irradiation. The degradation trend was as follows: 4 wt. % Cu2OCuO/TiO2 (76%) > Degussa P25 (69%) > 2 wt. % Cu2O-CuO/TiO2 (64.2 %) > 8 wt. % Cu2OCuO/TiO2 (60%) > TiO2 (32%). Increased dye removal efficiency using 4 wt. % Cu2OCuO/TiO2 can be attributed to the fact that depositing Cu2O-CuO at the surface of TiO2
facilitates electron hole charge separation, hence inhibiting their fast recombination. It has
been demonstrated earlier that metal act as an electron trap via Schottky barrier formation
thus reducing electron-hole recombination rates which ultimately enhance the photocatalytic
activity of TiO2 [18, 28, 34]. The holes present in the valence band may react with the surface
OH groups to produce OH radicals where electrons trapped on Cu2O-CuO react with O2 to
convert into O2 - radical (Fig. 7). Both of these species have been reported to degrade the
13

dyes [11]. The presence of copper in its oxide form under photocatalytic conditions has been
demonstrated recently [35]. Additionally, the visible light absorption centered around 450 nm
clearly seen in UV-Vis spectra of Cu2O-CuO/TiO2 catalysts may also be behind the enhanced
photocatalytic activity (Fig. 2). Degussa P25 containing 20% rutile and 80% anatase TiO2
could enhance photocatalytic activity by forming interparticle electron transfer (IPET) [36].
However, anatase was used as support material as compared to Degussa P25 catalyst due the
fact that former can be easily purchased/synthesized with much higher surface area that
would absorb higher fraction of incoming solar light (5% UV; 5 mWcm-2) per unit surface
area. This would effectively increase the activity even further for industrial scale applications.
The decrease in photocatalytic activity with increasing Cu content from 4 wt. % to 8 wt. %
may arise due to various factors. The decrease in photocatalytic activity with further
increasing metal deposition has been observed by many other workers for which no clear
explanation is yet known [37], [38], [39], [34] and [40]. Upon increasing of metal particle
coverage, large fraction of the semiconductor surface may become unavailable for light
absorption. It can also result in increasing surface defects at metal- semiconductor interface
leading to an increase in electron-hole recombination centers. The change in photocatalytic
activity may not only arise from the geometric effect but also due to the electronic effects. In the latter
case the metal particles can affect the activity away from them on the semiconductor surface [41, 42].

. Chen and Mao reported similar behavior and abrupt decrease in the photocatalytic activity
on increasing Cu deposition on TiO2 [43].
3.3 Effect of dye adsorption
The degradation rate of a dye, in heterogeneous photocatalysis may depend upon its
adsorption capability on catalyst surface [44]. To evaluate this effect, a pre-determined
quantity (5 mg) of RB 49, RR 24 and RY 160 were allowed to adsorb on the surface of 25
14

mgL-1 of each of 4 wt. % Cu2O-CuO/TiO2 and Degussa P25 for 30 minutes under dark
conditions. The catalyst with adsorbed dye was filtered and relative amount of adsorbed dye
was calculated from the absorption measurement of dye solution before catalyst addition and
after catalyst filtration. The amount of RB 49 and RR 24 adsorbed onto 4 wt. % Cu2OCuO/TiO2 catalyst surface was found higher than that of adsorption onto Degussa P25 except
RY 160 dye. RB 49 showed high adsorption at Cu2O-CuO/TiO2 compared to Degussa P25
(Fig. 8). RR 24 dye was found to be least adsorbed dye on both catalysts (Fig. 8). The high
adsorption efficiency of RB 49 and RY 160 dye can be explained on the basis of dyes
individual natural pH which were found to be 6.5, 7.5 and 6.9 for RB 49, RR 24 and RY 160,
respectively. The point of zero charge for TiO2 particles is pHpzc = 6.8 [45]. Typically ZPC of
TiO2 decreases with metal loading due to Schottky barrier formation that results in electron
depletion on the TiO2 and increase in OH groups on the surface. The ZPC of Cu-TiO2 is
typically found in 5.6-6.1 range depending upon the metal loading [46]. At a pH values of a
solution higher than pHpzc, the surface of the catalyst gets negatively charged [47, 48]. RB 49
is strongly adsorbed at its natural pH due to relatively lower negative surface charge while
there is less adsorption in the case of other dyes because surface is relatively more negatively
charged [49].
3.4 Effect of Catalyst Loading
Decomposition of RB 49 was measured at 25, 50, 75, 125 and 150 mg L-1 of 4 wt. % Cu2OCuO/TiO2 and Degussa P25. It was observed that 125 mg L-1 of 4 wt. % Cu2O-CuO/TiO2 and
Degussa P25 was the optimum catalyst dose for efficient degradation of RB 49 dye (Fig. 9).
The complete photodegradation (100%) of RB 49 was found at 125 mg L -1 of 4 wt. % Cu2OCuO/TiO2 after 80 minutes. The increase in catalyst dose provides more active sites for OH /
O2

radicals formation as well as increased dye adsorption ultimately increasing the


15

degradation rate [50]. Decrease in degradation rate at higher catalyst doses above optimum
dose (Fig. 9) might be attributed to decrease in light penetration thus leaving some catalyst
particles out of light reach and the possibility of more TiO2 aggregates formation [51].
Moreover, particleparticle interaction also becomes imperative as the quantity of particles in
solution increases. This may result in an enhanced rate of deactivation of activated particles
by collision with the particles at their ground state and hence decreasing the dye degradation
[52].
3.5 Effect of dye dose and dye degradation kinetics
In order to investigate the effect of increase in initial dye concentration on the rate of its
degradation, various initial dye doses of RB 49 were used. The dye concentration was varied
from 5 to 25 mg in 250 mL of water. The solution was irradiated for 80 minutes in the
presence of optimum catalyst dose (25 mg). Rate constant were calculated for each dye dose.
The rate constant (kd) of RB 49 photodegradation was calculated by using the following
formulas:
Ct = Co exp ( kd t) or

(2)

ln Co/Ct = kd t

(3)

Where, Ct is the absorbance at specific time t and Co is initial dye absorbance. kd is pseudofirst-order rate constant (min1) for degradation reactions and t is the degradation time. The
experimental degradation results are shown in Figure 10 and Table 2. Overall it was observed
that with increasing dye concentration from 5-25 mg L-1 the reaction rate gets slower. A
straight line shown in ln Co/Ct versus irradiation time plots in the presence of 4 wt. % Cu2OCuO/TiO2 (Fig. 10) confirms that the degradation kinetics of RB 49 nearly obey pseudo-firstorder kinetics. Also, high R2 values 0.976 to 0.861 for Cu2O-CuO/TiO2 and 0.993 to 0.917 for
Degussa P25 indicated that the observed values were almost linear with the expected
16

degradation values (Table. 2). It is also apparent from Table 2 that increasing dyes dose from
5 to 25 mgL-1 significantly lower their kd values from 0.9496 min-1 to 0.1103 min-1 on Cu2OCuO/TiO2 and 0.5829 min-1 to 0.0793 min-1 at Degussa P25 in 80 min, respectively. It is
difficult to directly compare rate constants with previous studies as they depend on initial dye
concentration, catalyst type, pH and other conditions. However, in this study such behavior
can be attributed to the fact that rate of degradation is proportional to the number of OH /O2

radicals formed at photocatalyst surface, whereas, at a fixed catalyst concentration, the


number of active sites remains the same [53]. Thus, increasing dye molecules blocks more of
these active sites for OH generation. Also, excess amount of dye molecules adsorbed on the
photocatalysts surface reduces the radiation path length with an ultimate reduction in light
intensity for photocatalyst excitation [54]. Similar effect was observed by Neppolian and coworkers during the degradation of textile dyes under solar light [52].
3.6 Extent of dye mineralization and catalyst stability
As the reduction of chemical oxygen demand (COD) reflects the extent of degradation or
mineralization of an organic species, the percentage change in COD was studied for RBD 49
dye samples under optimized conditions (Catalyst concentration = 125 mg L-1; C = 5 mg L-1;
pH = 6.5) after and before the degradation experiment (Table 3). The results indicate that the
photocatalytic process leads, apart from decolorization, to a substantial decrease in the COD
of treated dye solution. The % COD reduction is less than % decolorization which may be
due to the formation of smaller uncolored products. Therefore, it seems that to achieve
complete mineralization of dyes, a longer irradiation time as compared to decolorization time,
is required. It can also be noted that the decrease in COD is more in the case of 4 wt. %
Cu2O-CuO/TiO2 (84%) as compared to Degussa P25 (62%) indicating a faster dye
mineralization rate on the former. Moreover, in order to study the catalyst stability, 4 wt. %
17

Cu2O-CuO/TiO2 catalyst was filtered, dried and reused to test the efficiency of catalyst for
three successive photocatalytic experiments. The results indicated that there was no such
prominent decrease in the degradation activity even after recycling the catalyst three times
(Table 4).

4 Conclusions
The study shows that photocatalytic degradation of commercial dyes in general and RB 49 in
particular could effectively be achieved using 4 wt. % Cu2O-CuO/TiO2. Moreover, the
photocatalysts with 4 wt. % Cu2O-CuO/TiO2 exhibited higher photocatalytic activity
compared to Degussa P25 catalyst. The use of anatase as support material compared to
Degussa P25 catalyst is due the fact that former can be easily purchased/synthesized with
much higher surface area (> 100 m2g-1) which would effectively increase the activity even
further. Such enhanced dye degradation ability of Cu deposited catalyst can be ascribed to the
immediate electron trapping property of Cu present in Cu2O-CuO/TiO2. Under optimized
conditions, the rate of degradation increased by increasing catalyst concentration up to 125
mgL-1 probably due to the excess of OH /O2 - radical formation. However, photocatalytic
activity was found to be inversely proportional to initial dye dose which has been found
related to the decreases amount of OH /O2

radical formation. High photodegradation

efficiency of Cu2O-CuO/TiO2 catalyst for RB 49 dye indicates its potential application for
degradation of other carcinogenic commercial dyes as well.

18

References
[1] T. Robinson, G. McMullan, R. Marchant, P. Nigam, Remediation of dyes in textile
effluent: a critical review on current treatment technologies with a proposed alternative,
Bioresource Technology, 77 (2001) 247-255.
[2] A.G. Prado, L.B. Bolzon, C.P. Pedroso, A.O. Moura, L.L. Costa, Nb 2O5 as efficient and
recyclable photocatalyst for indigo carmine degradation, Applied Catalysis B:
Environmental, 82 (2008) 219-224.
[3] A. Bafana, S.S. Devi, T. Chakrabarti, Azo dyes: past, present and the future,
Environmental Reviews, 19 (2011) 350-371.
[4] A. Houas, H. Lachheb, M. Ksibi, E. Elaloui, C. Guillard, J.-M. Herrmann, Photocatalytic
degradation pathway of methylene blue in water, Applied Catalysis B: Environmental, 31
(2001) 145-157.
[5] H. Zollinger, Color chemistry: syntheses, properties, and applications of organic dyes and
pigments, Wiley-Vch, 2003.
[6] K. Shakir, A.F. Elkafrawy, H.F. Ghoneimy, S.G. Elrab Beheir, M. Refaat, Removal of
rhodamine B (a basic dye) and thoron (an acidic dye) from dilute aqueous solutions and
wastewater simulants by ion flotation, Water Research, 44 (2010) 1449-1461.
[7] A.K. Verma, R.R. Dash, P. Bhunia, A review on chemical coagulation/flocculation
technologies for removal of colour from textile wastewaters, Journal of Environmental
Management, 93 (2012) 154-168.
[8] M. Rafatullah, O. Sulaiman, R. Hashim, A. Ahmad, Adsorption of methylene blue on
low-cost adsorbents: a review, Journal of Hazardous Materials, 177 (2010) 70-80.
[9] S.H. Lin, M.L. Chen, Treatment of textile wastewater by chemical methods for reuse,
Water Research, 31 (1997) 868-876.
[10] P. Pizarro, C. Guillard, N. Perol, J.-M. Herrmann, Photocatalytic degradation of
imazapyr in water: Comparison of activities of different supported and unsupported TiO 2based catalysts, Catalysis Today, 101 (2005) 211-218.
[11] A. Ajmal, I. Majeed, R.N. Malik, H. Idriss, M.A. Nadeem, Principles and mechanisms of
photocatalytic dye degradation on TiO2 based photocatalysts: a comparative overview, RSC
Advances, 4 (2014) 37003-37026.
[12] A.H. Alwash, A.Z. Abdullah, N. Ismail, Elucidation of Reaction Behaviors in
Sonocatalytic Decolorization of Amaranth Dye in Water Using Zeolite Y Co-Incorporated
with Fe and TiO2, ACES, 3 (April 2013).
[13] O. Carp, C.L. Huisman, A. Reller, Photoinduced reactivity of titanium dioxide, Progress
in Solid State Chemistry, 32 (2004) 33-177.
[14] S.P. Patil, V.S. Shrivastava, G.H. Sonawane, S.H. Sonawane, Synthesis of novel Bi2O3
montmorillonite nanocomposite with enhanced photocatalytic performance in dye
degradation, Journal of Environmental Chemical Engineering, 3 (2015) 2597-2603.
[15] K. Pirkanniemi, M. Sillanp, Heterogeneous water phase catalysis as an environmental
application: a review, Chemosphere, 48 (2002) 1047-1060.
[16] S. Kaur, V. Singh, TiO2 mediated photocatalytic degradation studies of Reactive Red
198 by UV irradiation, Journal of Hazardous Materials, 141 (2007) 230-236.
[17] W. Liu, S. Chen, W. Zhao, S. Zhang, Study on the photocatalytic degradation of
trichlorfon in suspension of titanium dioxide, Desalination, 249 (2009) 1288-1293.
[18] M. Nadeem, K. Connelly, H. Idriss, The photoreaction of TiO 2 and Au/TiO2 single
crystal and powder with organic adsorbates, International Journal of Nanotechnology, 9
(2012) 121-162.
19

[19] M. Pelaez, N.T. Nolan, S.C. Pillai, M.K. Seery, P. Falaras, A.G. Kontos, P.S. Dunlop,
J.W. Hamilton, J.A. Byrne, K. O'Shea, A review on the visible light active titanium dioxide
photocatalysts for environmental applications, Applied Catalysis B: Environmental, 125
(2012) 331-349.
[20] R. Tung, Schottky-barrier formation at single-crystal metal-semiconductor interfaces, in:
Electronic Structure of Metal-Semiconductor Contacts, Springer, 1990, pp. 169-172.
[21] K. Chiang, R. Amal, T. Tran, Photocatalytic degradation of cyanide using titanium
dioxide modified with copper oxide, Advances in Environmental Research, 6 (2002) 471485.
[22] T.-C. Ou, F.-W. Chang, L.S. Roselin, Production of hydrogen via partial oxidation of
methanol over bimetallic AuCu/TiO2 catalysts, Journal of Molecular Catalysis A: Chemical,
293 (2008) 8-16.
[23] Z. Li, J. Liu, D. Wang, Y. Gao, J. Shen, Cu 2 O/Cu/TiO2 nanotube Ohmic heterojunction
arrays with enhanced photocatalytic hydrogen production activity, International Journal of
Hydrogen Energy, 37 (2012) 6431-6437.
[24] E. ayan, Optimization and modeling of decolorization and COD reduction of reactive
dye solutions by ultrasound-assisted adsorption, Chemical Engineering Journal, 119 (2006)
175-181.
[25] K.H. Mancy, Instrumental Analysis for Water Polution Control, Ann Arbor Science
Publishers, Inc., , Michigan, USA, 1971.
[26] H. Irie, K. Kamiya, T. Shibanuma, S. Miura, D.A. Tryk, T. Yokoyama, K. Hashimoto,
Visible light-sensitive Cu(II)-grafted TiO2 photocatalysts: activities and X-ray absorption fine
structure analyses, The Journal of Physical Chemistry C, 113 (2009) 10761-10766.
[27] X. Qiu, M. Miyauchi, H. Yu, H. Irie, K. Hashimoto, Visible-Light-Driven Cu (II)(Sr1 y
Nay)(Ti1 x Mox) O3 Photocatalysts Based on Conduction Band Control and Surface Ion
Modification, Journal of the American Chemical Society, 132 (2010) 15259-15267.
[28] M.A. Nadeem, M. Murdoch, G.I.N. Waterhouse, J.B. Metson, M.A. Keane, J. Llorca, H.
Idriss, Photoreaction of ethanol on Au/TiO2 anatase: Comparing the micro to nanoparticle
size activities of the support for hydrogen production, Journal of Photochemistry and
Photobiology A: Chemistry, 216 (2010) 250-255.
[29] B. Erdem, R.A. Hunsicker, G.W. Simmons, E.D. Sudol, V.L. Dimonie, M.S. El-Aasser,
XPS and FTIR surface characterization of TiO2 particles used in polymer encapsulation,
Langmuir, 17 (2001) 2664-2669.
[30] I.A. Khan, A. Badshah, M.A. Nadeem, N. Haider, M.A. Nadeem, A copper based metalorganic framework as single source for the synthesis of electrode materials for highperformance supercapacitors and glucose sensing applications, International Journal of
Hydrogen Energy, 39 (2014) 19609-19620.
[31] N. Farhangi, R.R. Chowdhury, Y. Medina-Gonzalez, M.B. Ray, P.A. Charpentier,
Visible light active Fe doped TiO2 nanowires grown on graphene using supercritical CO2,
Applied Catalysis B: Environmental, 110 (2011) 25-32.
[32] M.F. Al-Kuhaili, Characterization of copper oxide thin films deposited by the thermal
evaporation of cuprous oxide (Cu2O), Vacuum, 82 (2008) 623-629.
[33] M. Behnajady, N. Modirshahla, R. Hamzavi, Kinetic study on photocatalytic
degradation of CI Acid Yellow 23 by ZnO photocatalyst, Journal of Hazardous Materials,
133 (2006) 226-232.
[34] M. Murdoch, G. Waterhouse, M. Nadeem, J. Metson, M. Keane, R. Howe, J. Llorca, H.
Idriss, The effect of gold loading and particle size on photocatalytic hydrogen production
from ethanol over Au/TiO2 nanoparticles, Nature chemistry, 3 (2011) 489-492.
20

[35] L. Sinatra, A.P. LaGrow, W. Peng, A.R. Kirmani, A. Amassian, H. Idriss, O.M. Bakr, A
Au/Cu2OTiO2 system for photo-catalytic hydrogen production. A pn-junction effect or a
simple case of in situ reduction?, Journal of Catalysis, 322 (2015) 109-117.
[36] R.-a. Doong, S.-m. Chang, C.-w. Tsai, Enhanced photoactivity of Cu-deposited titanate
nanotubes for removal of bisphenol A, Applied Catalysis B: Environmental, (2012).
[37] W.-T. Chen, V. Jovic, D. Sun-Waterhouse, H. Idriss, G.I.N. Waterhouse, The role of
CuO in promoting photocatalytic hydrogen production over TiO 2, International Journal of
Hydrogen Energy, 38 (2013) 15036-15048.
[38] G.I. Waterhouse, M.A. Nadeem, J.B. Metson, J. Llorca, Photocatalytic hydrogen
production from ethanol over Au/TiO2 anatase and rutile nanoparticles. Effect of Au loading
and particle size, Prepr. Pap.-Am. Chem. Soc., Div. Fuel Chem, 55 (2010) 1.
[39] I. Majeed, M.A. Nadeem, M. Al-Oufi, M.A. Nadeem, G.I.N. Waterhouse, A. Badshah,
J.B. Metson, H. Idriss, On the role of metal particle size and surface coverage for photocatalytic hydrogen production: A case study of the Au/CdS system, Applied Catalysis B:
Environmental, 182 (2016) 266-276.
[40] M. Bowker, Sustainable hydrogen production by the application of ambient temperature
photocatalysis, Green Chemistry, 13 (2011) 2235-2246.
[41] M. Nadeem, G. Waterhouse, H. Idriss, A study of ethanol reactions on O 2-treated
Au/TiO2. Effect of support and metal loading on reaction selectivity, Surface Science, (2015).
[42] M.A. Nadeem, I. Majeed, G.I.N. Waterhouse, H. Idriss, Study of ethanol reactions on H 2
reduced Au/TiO2 anatase and rutile: effect of metal loading on reaction selectivity, Catalysis,
Structure & Reactivity, 1 (2015) 61-70.
[43] X. Chen, S.S. Mao, Titanium dioxide nanomaterials: synthesis, properties,
modifications, and applications, Chemical Reviews, 107 (2007) 2891-2959.
[44] J. Wiszniowski, D. Robert, J. Surmacz-Gorska, K. Miksch, J.-V. Weber, Photocatalytic
decomposition of humic acids on TiO2: Part I: Discussion of adsorption and mechanism,
Journal of Photochemistry and Photobiology A: Chemistry, 152 (2002) 267-273.
[45] I.A. Alaton, I.A. Balcioglu, D.W. Bahnemann, Advanced oxidation of a reactive dyebath
effluent: comparison of O3, H2O2/UV-C and TiO2/UV-A processes, Water Research, 36
(2002) 1143-1154.
[46] I.J. Puentes-Crdenas, G.M. Chvez-Camarillo, C.M. Flores-Ortiz, M.d.C. CristianiUrbina, A.R. Netzahuatl-Muoz, J.C. Salcedo-Reyes, A.M. Pedroza-Rodrguez, E. CristianiUrbina, Adsorptive Removal of Acid Blue 80 Dye from Aqueous Solutions by Cu-TiO2,
Journal of Nanomaterials, 2016 (2016).
[47] M. Behnajady, N. Modirshahla, R. Hamzavi, Kinetic study on photocatalytic
degradation of CI Acid Yellow 23 by ZnO photocatalyst, Journal of Hazardous Materials,
133 (2006) 226-232.
[48] S.K. Kansal, M. Singh, D. Sud, Studies on TiO 2/ZnO photocatalysed degradation of
lignin, Journal of Hazardous Materials, 153 (2008) 412-417.
[49] C. Raillard, V. Hquet, P. Le Cloirec, J. Legrand, Comparison of different TiO 2
photocatalysts for the gas phase oxidation of volatile organic compounds, Water Science &
Technology, 50 (2004) 241-250.
[50] K. Byrappa, A. Subramani, S. Ananda, K.L. Rai, R. Dinesh, M. Yoshimura,
Photocatalytic degradation of rhodamine B dye using hydrothermally synthesized ZnO,
Bulletin of Materials Science, 29 (2006) 433-438.
[51] S. Kavitha, P. Palanisamy, Photocatalytic and sonophotocatalytic degradation of
Reactive Red 120 using dye sensitized TiO2 under visible light, International Journal of Civil
and Environmental Engineering, 3 (2011) 1-6.
21

[52] B. Neppolian, H. Choi, S. Sakthivel, B. Arabindoo, V. Murugesan, Solar light induced


and TiO2 assisted degradation of textile dye reactive blue 4, Chemosphere, 46 (2002) 11731181.
[53] W.Z. Tang, H. An, UV/TiO2 photocatalytic oxidation of commercial dyes in aqueous
solutions, Chemosphere, 31 (1995) 4157-4170.
[54] A. Subramani, K. Byrappa, S. Ananda, K.L. Rai, C. Ranganathaiah, M. Yoshimura,
Photocatalytic degradation of indigo carmine dye using TiO 2 impregnated activated carbon,
Bulletin of Materials Science, 30 (2007) 37-41.

22

Figure captions
Figure 1. Chemical structure of (a) Reactive Blue 49, (b) Reactive Red 24 and (c) Reactive
Yellow 160.
Figure 2. TEM images of photocatalysts (a) 8 wt. % Cu2O-CuO/TiO2 calcinated at 400 C (b)
Pure Degussa P-25: XRD patterns and UV-Vis spectra of different catalysts as indicated.
Figure 3. XPS survey scan and core level spectra of Ti 2p, O1s, C1s and Cu 2p for 4 wt. %
Cu/TiO2.
Fig. 4. Changes in UV-Vis absorption spectrum of three aqueous dye solutions of RB 49, RR
24, RY 160 after 30 min of UV irradiation on 2 % Cu2O-CuO/TiO2 (a, c, f) and Degussa P-25
(b, d, f).
Fig. 5. Photodegrading efficiency (%) of RB 49, RY 160 and RR 24 dye as a function of time
up to 120 min (C = 5 mg L-1; Catalyst concentration = 25 mg L-1; pH = 6.5) (a) at Degussa
P25 (b) on the surface of 2 wt. % Cu2O-CuO/TiO2 -TiO2.
Fig. 6. Degradation efficiency of RB 49 dye on bare, 2, 4, 8 wt.% Cu2O-CuO/TiO2 and TiO2
Degussa P25 after 120 min UV irradiation (C = 5 mg L-1 ; Catalyst concentration = 25 mg L1
; pH = 6.5).
Fig. 7. Dye degradation mechanism on Cu2O-CuO/TiO2 photocatalyst.
Fig. 8. Adsorption efficiency of RB 49, RR 24 and RY 160 dye at Degussa P25 and 4 wt. %
Cu2O-CuO/TiO2 after 30 min under dark (Catalyst concentration = 25 mg L -1; C = 5 mg L-1;
pH = 6.5).
Fig. 9. Effect of varying catalyst concentration on the rate of RB 49 dye photodegradation
under 80 min of UV irradiation (C = 5 mg L-1 ; pH = 6.5; t = 80 min).
Fig. 10. Pseudo first order Ln (C/Ct) plots of varying RB 49 dye concentrations versus
irradiation time representing the effect of increasing dye concentration on degradation rate.
(left) Cu2O-CuO/TiO2 and (right) Degussa P-25.

23

Figure 1.

24

Figure 2.

25

Figure 3.

26

Fig. 4.

27

Fig. 5.

28

Fig. 6.

29

Fig. 7.

30

Fig. 8.

31

Fig. 9.

32

Fig. 10.

33

Tables
Table. 1. General characteristics of selected dyes.
Dye Name

C.I Reactive Blue 49

C.I Reactive Red 24

C.I Reactive Yellow 160

Commercial name

Synocron Blue P-3R

Synocron Red P3R

Tubatin Yellow ME-4GL

Molecular formula

C32H23ClN7Na3O11S3

C26H17ClN7Na3O10S3

C25H22ClN9Na2O12S3

C.I number

621526

18208

129898

Molecular weight

882.19

788.07

818.13

max (nm)

585

545

420

34

Table 2. Effect of increasing dye dose on degradation rate constant kd and R2 value for
Reactive Blue 49 dye degradation at Cu2O-CuO/TiO2 and Degussa P25 catalysts.
Rate constant values for Reactive Blue 49 dye degradation from literature under
comparable conditions to this study.
Initial Dye
kd (min-1)
Concentration
Cu2O-CuO /TiO2
(mg L-1)
5
0.9496
10
0.5475
15
0.2686
20
0.1945
25
0.1103

R2
P25

Cu2O-CuO /TiO2

P25

0.5829
0.3019
0.1797
0.1409
0.0793

0.861
0.975
0.976
0.946
0.965

0.993
0.990
0.917
0.982
0.961

35

Table. 3. A comparison between color and chemical oxygen demand (COD) of RB 49 dye
solution before and after 80 min irradiation using Degussa P25 and 4 wt. % Cu2O-CuO/TiO2
(Catalyst concentration = 125 mg L-1; C = 5 mg L-1 ; pH = 6.5).
Chemical oxygen demand (COD)
Color removal (%)
Before (mg O2 L-1) After (mg O2 L-1) Removal (%)

Catalysts
Degussa P25

75 5

143

54

62

4 wt. % Cu2O-CuO/TiO2

100

143

23

84

36

Table. 4. A comparison between color and chemical oxygen demand (COD) of RB 49


dye solution before and after 80 min irradiation using 4 wt. % Cu2O-CuO/TiO2 for
three successive experimental runs (Catalyst concentration = 125 mg L -1; C = 5 mg L-1;
pH = 6.5).
Chemical oxygen demand (COD)
Color removal
Catalyst
(%)
Before (mg O2 L-1) After (mg O2 L-1) Removal (%)
(4 wt. % Cu2O-CuO/TiO2)
Experimental run
Results

R1

R2 R3

R1

R2

R3

R1

R2 R3 R1 R2 R3

100 100 100 143

143

143

23

22 23 84

84 84

37

Vous aimerez peut-être aussi