You are on page 1of 14

Res Chem Intermed (2012) 38:323–336

DOI 10.1007/s11164-011-0349-0

Preparation, characterization, and photocatalytic


activity of polyaniline/ZnO nanocomposite

Ali Olad • Rahimeh Nosrati

Received: 8 May 2011 / Accepted: 12 July 2011 / Published online: 27 July 2011
 Springer Science+Business Media B.V. 2011

Abstract Polyaniline (PANI)/zinc oxide (ZnO) nanocomposite was synthesized


by in-situ polymerization. X-ray diffraction patterns, UV–visible spectroscopy,
SEM, and TEM were used to characterize the composition and structure of the
nanocomposite. Nanostructured PANI/ZnO composite was used as photocatalyst in
the photodegradation of methylene blue dye molecules in aqueous solution. The
photocatalytic activity of PANI/ZnO nanocomposite under UV and visible light
irradiation was evaluated and was compared with that of ZnO nanoparticles. ZnO/
PANI core–shell nanocomposite had greater photocatalytic activity than ZnO
nanoparticles and pristine PANI under visible light irradiation. According to these
results, application of PANI as a shell on the surface of ZnO nanoparticles causes
the enhanced photocatalytic activity of the PANI/ZnO nanocomposite. Also UV–
visible spectroscopy studies showed that the absorption peak for PANI/ZnO
nanocomposite has a red shift toward visible wavelengths compared with the ZnO
nanoparticles and pristine PANI. The effect of different operating conditions on the
photocatalytic performance of PANI/ZnO nanocomposite in the photodegradation
of methylene blue dye molecules was investigated in a bath experimental setup.

Keywords Conducting polymers  Polyaniline  ZnO  Photocatalyst

Introduction

In recent years there has been increasing interest to the development of organic/
inorganic hybrid materials [1]. These materials have a good combination of the
different properties of the individual components and have potential applications in

A. Olad (&)  R. Nosrati


Polymer Composite Research Laboratory, Department of Applied Chemistry,
Faculty of Chemistry, University of Tabriz, Tabriz, Iran
e-mail: a.olad@yahoo.com

123
324 A. Olad, R. Nosrati

a variety of fields, for example microelectronics, display technology, catalysis and


photocatalysis, sensors, and batteries and solar cells [2–12].
Inorganic semiconductors have been extensively investigated as hybrids with
organic polymers having synergetic or complementary properties and behavior for
fabrication of a variety of devices [13]. Zinc oxide (ZnO) is one of the most
important semiconductors with a direct wide band gap of *3.37 eV and large
excitation binding energy of 60 meV at room temperature [2]. ZnO has promising
applications in electrical engineering, catalysis, ultraviolet (UV) absorption, and
optical and optoelectronic devices. The shape and size of ZnO particles affect its
applications in these fields [13–15]. Nanostructured ZnO has greatly enhanced
responses and applications because of its high surface area and high surface energy.
It has also been reported that incorporation of ZnO into polymer matrices and
porous materials improves its effectiveness in a variety of applications [15, 16]. A
variety of polymers have, therefore, been widely used as support or host matrixes
for ZnO particles or nanoparticles [5]. However, ZnO nanoparticles tend to
aggregate in the host matrix, because of their high surface energy. Good disper-
sion of ZnO nanoparticles in polymer matrices is very difficult and expensive.
In this regard, control of the surface properties of ZnO is very important and
effective [15].
One of the most important and surface-dependant properties of ZnO particles and
nanoparticles is their photocatalytic behavior under UV electromagnetic irradiation.
This behavior can be used for degradation and removal of environmental pollutants,
especially organic dye contaminants, from water [17]. The use of organic dye
compounds in the paper, textile, food, and other industries causes pollution of the
environment because of their non-biodegradability and high toxicity [18, 19].
Several techniques are used for removal of such hazardous contaminants from
water. Photodegradation of organic pollutants by inorganic semiconductors, for
example TiO2 and ZnO, has attracted much interest in recent years [20]. Although
the big advantage of ZnO as photocatalyst compared with TiO2 is its ability to
absorb a comparatively large fraction of the UV spectrum [21], ZnO nanoparticles
can be activated only by the UV fraction of the sunlight spectrum, and because of
the low fraction (\5%) of UV irradiation (\387 nm) in solar light, use of ZnO as a
photocatalyst under sunlight irradiation is restricted [22]. On the other hand, the use
of UV lamps is expensive and harmful to the human body. Therefore, it is important
environmentally and to health to shift the photoactivation region of ZnO particles
toward visible wavelengths. This can be achieved by a variety of methods, and,
especially, by surface modification of ZnO particles or nanoparticles with various
materials [22].
Chemical and physical methods have been used for surface treatment of ZnO
particles. Application of a polymeric shell to the surface of ZnO core particles or
nanoparticles is an important method for surface modification of ZnO particles [13,
15]. Conducting polymers, a new class of polymeric material, have been used as
shell or surface modifiers for semiconductors such as TiO2, SiO2, and ZnO particles
and nanoparticles. It has been reported that the surface properties of ZnO particles
and nanoparticles can be improved by modification with conducting polyaniline
(PANI) [11, 23, 24]. As a typically p-type organic semiconductor, PANI has

123
Polyaniline/ZnO nanocomposite 325

attracted much attention because of its simple method of synthesis, environmental


stability, and unique conduction mechanism [3]. The properties and performance of
a variety of PANI/ZnO hybrid structures are strongly dependent on the method of
preparation and the surface treatment technique used for modification of the ZnO
particles or nanoparticles. A layer-by-layer self assembly technique was used by
Kovtyukhova et al. [25] for preparation of PANI/ZnO composite films with
improved I–V characteristics. A diode with good rectifying behavior has been
fabricated by deposition of PANI chains on the surface of a thin film of ZnO [26]. It
has also been reported that a p–n junction is formed between ZnO and conducting
PANI, and that this is important in applications such as polymer-based solar cells,
field-emission transducers, and sensors [3].
Khare et al. [4] have prepared a PANI/ZnO nanocomposite film by a solution-
casting method and have investigated its dielectric properties. They showed that
the dielectric constant and loss factor of PANI were reduced in PANI/ZnO
nanocomposite.
The photocatalytic activity of PANI/ZnO nanocomposite in the degradation of
organic methylene blue dye molecules has been investigated by Ameen et al. [19].
They used visible light irradiation for photoactivation of PANI/ZnO nanocomposite
and for subsequent photocatalytic degradation of the dye. However, the photocat-
alytic degradation efficiency was low.
The objective of this study was the preparation and characterization of core–shell
ZnO/PANI nanostructured hybrid material and investigation of its photocatalytic
activity in the degradation of methylene blue dye in water under visible light
irradiation. High removal efficiency was obtained by using the ZnO/PANI core–
shell system as photocatalyst under visible light irradiation.

Experimental

Reagents and materials

Ammonium persulfate (APS), hydrochloric acid, and methylene blue (MB) were
purchased from Merck Chemicals and were used as received. The structure and
characteristics of MB dye are shown in Table 1. Aniline was purchased from Merck

Table 1 Structure and


Structure
characteristics of methylene
blue dye molecules

IUPAC name 3,7-bis(Dimethylamino)phenothiazin-5-ium


chloride
CAS number 61-73-4
Molecular formula C16H18ClN3S.xH2O(x = 2–3)
Molar mass 319.86 g/mol

123
326 A. Olad, R. Nosrati

Chemicals and was doubly distilled before use. ZnO nanoparticles with an average
particle size of less than 50 nm were purchased from Neutrino Company.

Nanocomposite synthesis

A typical oxidative polymerization method was carried out for polymerization of


aniline in the presence of ZnO nanoparticles. ZnO nanoparticles (0.9313 g) were
dispersed by ultrasound in a 20-mL aqueous solution of aniline monomer (0.01 mol)
and hydrochloric acid (0.01 mol). APS (0.01 mol) was dissolved in a 15 mL
distilled water and was added dropwise to the mixture of ZnO and aniline monomer,
with stirring, in an ice bath. Polymerization proceeded for 5.5 h. The nanocomposite
of PANI with ZnO nanoparticles was obtained as precipitate. This precipitate was
isolated by filtration, washed with distilled water and ethanol several times, and
dried at 50 C.
Pure PANI was also synthesized, by use of an identical method but without use of
ZnO nanoparticles, to compare the structure and photocatalytic behavior of PANI/
ZnO nanocomposite with those of pure PANI.

Characterization

The UV–visible spectra of dye-containing aqueous solutions were recorded by use


of a Shimadzu UV-1700 Pharma Spectrophotometer. The photoactivation wave-
lengths of PANI/ZnO nanocomposite and pure PANI and ZnO were evaluated by
recording their UV–visible spectra. X-ray diffraction patterns for the composite and
its constituents were also recorded, by use of a Siemens D5000, k = 1.54Å
diffractometer, in order to characterize the composition and structure of samples. A
field-emission scanning electron microscopy (FESEM) image was obtained for ZnO
nanoparticles on a Hitachi S-4160 microscope, and a transmission electron
micrograph (TEM) of the PANI/ZnO nanocomposite was obtained by use of a
Leo 906 kv 80-100 (Germany).

Photodegradation experiments

The photocatalytic degradation of MB dye molecules was performed in a Pyrex


batch reactor under a halogen lamp (50 W JCDR-C GU5.3) with UV filter. Dye
solution (10 mg L-1, 50 mL) was prepared in distilled water and known amounts of
PANI/ZnO nanocomposite as photocatalyst were added and the mixture was stirred
in darkness for 10 min to equilibrate. The lamp was switched on to initiate the
photocatalytic degradation reaction. During irradiation, the suspension was stirred
continuously to keep the suspension homogeneous. At 10-min intervals dye solution
samples were taken and were filtered through a 0.22 lm filter (Whatman) and the
dye concentration was analyzed with a UV–visible spectrophotometer at the kmax of
MB dye (665 nm). Figure 1 shows the UV–visible absorption spectrum of MB dye.
A calibration graph based on Beer–Lambert’s law was obtained by plotting the
absorbance against the concentration of dye in solution. The photodegradation
efficiency (R, %) was calculated by use of Eq. 1

123
Polyaniline/ZnO nanocomposite 327

Absorbance (a.u.)

200 300 400 500 600 700 800


Wavelength (nm)

Fig. 1 UV–visible spectrum of methylene blue dye

R% ¼ 100½ðC0  CÞ=C0  ð1Þ


-1
where C0 is the initial concentration of MB dye in solution (10 mg L ), C is the
concentration of MB at each sampling time during the photodegradation process,
and R is the photodegradation efficiency at different times.

Results and discussion

Characterization of PANI/ZnO nanocomposite

X-ray diffraction analysis

Analysis of X-ray diffraction (XRD) patterns was used to investigate the structure of
PANI/ZnO nanocomposite. Figure 2 shows the XRD patterns of PANI/ZnO
nanocomposite, ZnO nanoparticles, and pure PANI. The XRD patterns of pristine
PANI has a weak broad peak at 2h = 25, because of its amorphous structure with
low crystallinity. The XRD patterns of ZnO nanoparticles have characteristic peaks
of ZnO with the crystalline structure of wurtzite. The XRD patterns of PANI/ZnO
nanocomposite include the characteristic peaks of both PANI and ZnO, which
confirms the formation of nanocomposite with lower crystallinity.

UV–visible spectroscopic analysis

To investigate the effect of combining ZnO nanoparticles with PANI on the


spectroscopic absorption of ZnO nanoparticles, the UV–visible absorption spectra
of all components were recorded and analyzed. Figure 3 shows the UV–visible

123
328 A. Olad, R. Nosrati

Intensity (a.u.)

15 25 35 45 55 65 75
2theta (deg.)

Fig. 2 XRD patterns of polyaniline (PANI) (a), PANI/ZnO nanocomposite (b), and ZnO nanoparticles (c)
Absorbance (a.u.)

200 400 600 800 1000


Wavelength (nm)

Fig. 3 UV–visible spectra of ZnO nanoparticles (a), pristine PANI (b), and PANI/ZnO nanocomposite (c)

spectra of ZnO nanoparticles, pristine PANI, and PANI/ZnO nanocomposite.


Pristine PANI strongly absorbs visible light, with a characteristic peak at 440 nm.
The UV–visible spectrum of the PANI/ZnO nanocomposite has an absorption peak
at 445 nm, which is red shifted compared with the absorption peaks of pristine
PANI (440 nm) and of ZnO nanoparticles (375 nm). This may be because of
interactions between PANI chains and ZnO nanoparticles which cause easy charge

123
Polyaniline/ZnO nanocomposite 329

transfer from PANI to ZnO via hydrogen bonding. According to these results, the
PANI/ZnO nanocomposite synthesized in this study has an absorption peak in the
visible light region, in other words it can be photoactivated by visible light
irradiation, which has no effect on health, in contrast with the pristine ZnO
photocatalyst nanoparticles which are photoactivated under harmful UV light
irradiation. This may be the main advantage of the PANI/ZnO nanocomposite
prepared in this study, which has a special structure which is only photoactivated
under visible light irradiation and has higher photocatalytic activity in this region
compared with previously reported PANI/ZnO composite materials [19, 27].

Morphological studies

Figure 4 shows the field-emission scanning electron microscopy (FESEM) micro-


graphs of ZnO nanoparticles and PANI/ZnO nanocomposite. The diameter of
ZnO nanoparticles is approximately 50 nm and they have a crystalline structure
(Fig. 4a). According to these results, the completely spherical structure of PANI/ZnO

Fig. 4 FESEM micrographs of


ZnO nanoparticles and PANI/
ZnO nanocomposites

123
330 A. Olad, R. Nosrati

nanocomposite (Fig. 4b) confirms the loading or precipitation of a smooth PANI


layer on the surface of ZnO nanoparticles and formation of a core–shell structure.
Figure 5 also shows the transmission electron micrograph (TEM) of PANI/ZnO
nanocomposite. A uniform distribution of ZnO nanoparticles in the PANI matrix is
apparent from the TEM results. Figure 5 also shows the spherical structure of PANI/
ZnO nanocomposite, from which, considering the spherical structure of ZnO
nanoparticles, formation of PANI/ZnO nanocomposite with a core–shell structure
can be concluded. In other word, the results of morphological studies demonstrate
the formation of PANI layers around the ZnO nanoparticles with a thicknesses of
approximately 5–25 nm.

Photocatalytic studies

Photocatalytic degradation of MB in aqueous solution using PANI, ZnO nanopar-


ticles, and PANI/ZnO nanocomposite under UV and visible light irradiation were
examined. Use of ZnO particles and nanoparticles as photocatalysts for degradation
of colored contaminants in water under UV irradiation has been well established
[17, 18]. However, application of conducting PANI to the surface of ZnO
nanoparticles causes a red shift in the spectroscopic absorption peak of ZnO
nanoparticles toward the visible light region (Fig. 3). Visible light is healthy,
inexpensive, and simple compared with harmful UV light. The degradation
efficiency (R , %) was calculated by use of Eq. 1.

Effect of light source

The photocatalytic degradation of MB dye molecules in aqueous solution using


PANI/ZnO nanocomposite as photocatalyst under UV and visible light sources was
evaluated. Similar amounts (1500 mg) of PANI/ZnO nanocomposite were brought
into contact with 50 mL MB solutions (10 mg L-1) in a batch experimental setup.
The concentration of MB was then analyzed at regular time intervals (10 min) and

Fig. 5 TEM of PANI/ZnO


nanocomposite

123
Polyaniline/ZnO nanocomposite 331

the normalized concentration (C/C0) of MB was evaluated after different contact


times by dividing the concentration of MB at each contact time by the initial
concentration of MB in solution. Figure 6 shows the variation in the normalized
concentration of MB during contact of 1500 mg of PANI/ZnO nanocomposite
powder under UV (30 W, Philips, Holland) and visible light (halogen lamp, 50 W
JCDR-C GU5.3, with UV filter) irradiation.
According to these results, removal of MB after 1 h exposure to the PANI/ZnO
nanocomposite was 28% under UV irradiation and 82% under visible light
irradiation. This is because of the shifting of the excitation wavelength of PANI/
ZnO nanocomposite toward the visible light region ([400 nm) compared with
pristine ZnO nanoparticles (375 nm) (Fig. 3). Also the excitation wavelength of
PANI coated on the ZnO nanoparticles is in the visible light region (Fig. 3).

Comparison of photocatalysts

The photocatalytic activity of ZnO nanoparticles, pristine PANI, and PANI/ZnO


nanocomposite under visible light irradiation for degradation of MB dye molecules
was evaluated and compared. Figure 7 shows the variations in the normalized
concentration (C/C0) of MB dye in solutions with the same initial concentration of
10 mg L-1, during contact with the same amount (1500 mg L-1) of the different
photocatalysts ZnO nanoparticles, pristine PANI, and PANI/ZnO nanocomposite
under visible light irradiation. The results showed that, ZnO nanoparticles have no
photocatalytic activity under visible light irradiation. This is because of the large
band gap of ZnO, i.e. the high energy required for excitation of electrons from the
valence band to the conduction band, which cannot be provided by visible light
irradiation. Pristine PANI has photocatalytic activity under visible light irradiation

0.8

0.6
C/C0

0.4
under UV
irradiation
0.2
under Visible
light irradiation

0
0 10 20 30 40 50 60
Time (min)

Fig. 6 Normalized concentration of MB in solution (initial concentration: 10 mg L-1) during contact


with PANI/ZnO nanocomposite powder (1500 mg L-1) under UV and visible light irradiation

123
332 A. Olad, R. Nosrati

in the degradation of MB molecules. The degradation efficiency of MB, or


decolorization efficiency of MB solution, under visible light irradiation using
pristine PANI as photocatalyst was evaluated as 39% after 1 h exposure time
(Fig. 7). However, the efficiency of photocatalytic degradation of MB using PANI/
ZnO nanocomposite as photocatalyst under visible light irradiation was much more
than that of pristine PANI at the same contact time. According to these results, 82%
of MB dye molecules were degraded under visible irradiation using PANI/ZnO
nanocomposite after exposure for 1 h (Fig. 7). It can be concluded that coating of
ZnO nanoparticles with a PANI shell makes them photocatalytically active under
visible light irradiation. This may be because of the efficient charge separation of
electron and hole pairs in the excited states and prevention of recombination of
charge pairs for a longer time, in the PANI/ZnO nanocomposite under visible light
irradiation. According to these results, this effect is successfully induced by the
PANI shell in the ZnO/PANI core–shell system.

Effect of photocatalyst loading

The effect of amount of PANI/ZnO nanocomposite as photocatalyst on the


efficiency of photocatalytic decomposition of MB in aqueous solution was
investigated. For this, different amounts (750–2000 mg L-1) of PANI/ZnO
nanocomposite were contacted with 50 mL MB solution in water at the same
initial concentration of 10 mg L-1. The variations in the concentration of MB
during the photodegradation process, as the concentration relative to the initial
concentration of MB (C/C0) were plotted against contact time. Figure 8 shows the
relative concentration of MB during photocatalytic decomposition using different

0.8

0.6
C/C0

0.4

ZnO
0.2
PANI

PANI/ZnO nanocomposite
0
0 10 20 30 40 50 60
Time (min)

Fig. 7 Normalized concentrations of MB in solution against contact time during photocatalytic degradation
by use of the different photocatalysts ZnO nanoparticles, pristine PANI, and PANI/ZnO nanocomposite
under visible light irradiation. [photocatalyst] = 1500 mg L-1, initial dye concentration = 10 mg L-1

123
Polyaniline/ZnO nanocomposite 333

amounts of PANI/ZnO nanocomposite powder under visible light irradiation. Also,


in an exactly similar experiment, the direct photolysis of MB under visible light
irradiation without the use of any photocatalyst was examined. Some dye molecules
may be decomposed under visible light irradiation. Therefore, it is important to
know the extent to which the MB dye molecules are photolyzed if no catalyst was
used and hence estimate the contribution of direct photolysis in the overall
decomposition process. For this, the decomposition of MB under visible light
irradiation was examined without the use of PANI/ZnO nanocomposite (no
photocatalyst) (Fig. 8). It was found that MB dye molecules are not photolyzed
under visible light irradiation. This means that decomposition of MB in the visible
light region is caused solely by the photocatalytic activity of PANI/ZnO (Fig. 8).
On increasing the amount of PANI/ZnO photocatalyst up to 1500 mg L-1, the
efficiency of decolorization of MB was increased. Increasing the amount of
PANI/ZnO photocatalyst in the dye solution increases the number of reactive oxide
sites for dye decomposition generated on the photocatalyst under light irradiation,
because of the increase of the photocatalyst surface subjected to light irradiation.
Maximum degradation of 82% was obtained with a catalyst loading of
1500 mg L-1 after exposure for 60 min. A decrease in the efficiency of degradation
of MB dye was observed on increasing the amount of photocatalyst beyond
1500 mg L-1 (e.g. 2000 mg L-1), however (Fig. 8). This may be because of the
greater turbidity of the dye solution resulting from increasing the amount of the
colored PANI/ZnO photocatalyst in the solution. Increasing the turbidity of
the solution prevents light transmission and results in increased light scattering in
solution.

0.8

0.6
C/C0

0.4
no photocatalyst

0.2 750 mg/l


1500 mg/l
2000 mg/l
0
0 10 20 30 40 50 60
Time (min)

Fig. 8 Normalized concentrations of MB in solution against time during photocatalytic degradation by


use of different amounts of PANI/ZnO nanocomposite powder as photocatalyst under visible light
irradiation

123
334 A. Olad, R. Nosrati

Photocatalytic reaction mechanisms

On the basis of the experimental results obtained in this study, and in accordance
with previous reports on photocatalytic reactions of ZnO and PANI/ZnO
nanocomposite [21, 22, 27], the mechanism of photocatalytic activity of PANI/
ZnO under visible light irradiation can be predicted. Under visible light irradiation
ZnO itself is not excited and has no photocatalytic activity (Fig. 7). Visible light
absorption and photocatalytic behavior was observed when the surface of ZnO
nanoparticles was modified by PANI photosensitizer. Visible light irradiation causes
the PANI molecules to be excited and transfer electron to the conduction band of
ZnO nanoparticles (Eq. 2). Electrons in the conduction band of ZnO can reduce
molecular oxygen and produce the superoxide anion radical (Eq. 3). This radical
may form hydrogen peroxide or organic peroxides in the presence of oxygen and
organic molecules (Eqs. 4, 5, 6). Hydrogen peroxide can be generated in another
path (Eq. 7). Hydrogen peroxide can form hydroxyl radicals which are powerful
oxidizing agents (Eqs. 8, 9). The radicals produced are capable of attacking organic
pollutants with formation of intermediates which react with hydroxyl radicals to
produce the photocatalytic products (P) (Eq. 10).
PANI=ZnO þ visible light ! PANIþ =ZnO þ e
CB ð2Þ
e
CB þ O2 ! O
2 ð3Þ
H2 O þ O 
2 ! OOH þ OH

ð4Þ

2OOH ! O2 þ H2 O2 ð5Þ
O
2 þ MB ! MB  OO 
ð6Þ
OOH þ H2 O þ e
CB ! H2 O2 þ OH

ð7Þ
H2 O2 þ e
CB

! OH þ OH 
ð8Þ
H2 O2 þ O
2

! OH þ OH þ O2 
ð9Þ
OH =O
2 =PANI

þ MB ! intermediates ! P ð10Þ
Under UV irradiation, ZnO nanoparticles can be excited, producing photogen-
erated electron–hole pairs and achieving photocatalytic activity. Generated holes in
the valence band of ZnO can be transferred to the PANI molecules and can reach the
surface of photocatalyst where they can make contact with the dye molecules, to
cause to their degradation. Photogenerated electrons on the conduction band of ZnO
can also react with oxygen and produce superoxide anion radicals which cause
degradation of MB dye molecules. Therefore, PANI and ZnO together have a
synergetic effect on the photocatalytic behavior of each other. PANI causes a shift
in the wavelength of the light needed for excitation of ZnO nanoparticles toward the
visible light region. Also PANI causes some delay in recombination of electron–
hole pairs generated in ZnO nanoparticles and so causes an increase in the efficiency
of photocatalytic activity. In other words, by increasing the lifetime of the electron–
hole pairs, the photocatalytic efficiency of the ZnO–PANI system is also increased.
The photocatalytic degradation pathway of methylene blue under UV or visible light

123
Polyaniline/ZnO nanocomposite 335

irradiation because of the generation of reactive hydroxy and hydroperoxy radicals


has been proposed [28–30]. Hydroxy and hydroperoxy radicals produced by the
action of photocatalyst are capable of oxidizing MB molecules at the surface layer
of photocatalyst. MB molecules tend to be adsorbed by the surface of the catalyst.
This process can be enhanced by controlling the surface charge of the photocatalyst.
MB is a cationic dye and a negatively charged photocatalyst surface accelerates
adsorption of the MB and, consequently, the photodegradation process [28, 29]. The
initial step of MB degradation is cleavage of the functional group C–S?=C (Eq. 11).
Attack by OH• radicals then causes production of the sulfone and dissociation of the
benzene rings (Eq. 12). The sulfone group is attacked by another OH• radical,
giving a sulfonic acid (Eq. 13). Attack by a fourth OH• radical causes release of
SO42- ions (Eq. 14). Further attack by OH• radicals causes substitution of amino
groups and formation of phenolics. The NH•2 radical released generates ammonia
and ammonium ions (Eq. 15, 16, 17). The other two symmetrical dimethylphe-
nylamino groups in MB undergo progressive oxidative degradation of one methyl
group by attack of OH• radicals, producing an alcohol, then an aldehyde, which is
spontaneously oxidized into acid, and this decarboxylates into CO2 as a result of the
photo-Kolbe reaction [29].
R  Sþ ¼ R0 þ OH ! R  Sð¼ OÞ  R0 þ Hþ ð11Þ

NH2  C6 H3 ðRÞ  Sð¼ OÞ  C6 H4  R þ OH
! NH2  C6 H3 ðRÞ  SO2 þ C6 H5  R ð12Þ
SO2  C6 H4  R þ OH ! R  C6 H4  SO3 H ð13Þ
R  C6 H4  SO3 H þ OH ! R  C6 H4 þ SO2
4 þ 2H
þ
ð14Þ

R  C6 H4  NH2 þ OH ! R  C6 H4  OH þ NH2 ð15Þ
NH2 
þ H ! NH3 ð16Þ
NH3 þ Hþ ! NHþ
4 ð17Þ

Conclusion

A nanostructured core–shell ZnO/PANI composite system has been successfully


synthesized by in-situ polymerization. Photocatalytic degradation of MB dye by use
of the PANI/ZnO nanocomposite was achieved under visible light irradiation. The
nanostructured ZnO/PANI core–shell nanocomposite system prepared in this study
has greater photocatalytic activity than previously reported PANI/ZnO composite
materials. When the optimum amount of nanocomposite (1500 mg L-1) was used
as photocatalyst, MB at an initial concentration of 10 mg L-1 was 82%
photodegraded under visible light irradiation, which is higher than the degradation
efficiency obtained by use of ZnO nanoparticles and pristine PANI as photocat-
alysts. PANI/ZnO nanocomposite had less photocatalytic activity under UV light
irradiation. It was concluded PANI had a synergetic effect on the photocatalytic
activity of ZnO under visible light irradiation.

123
336 A. Olad, R. Nosrati

Acknowledgments The financial support of this research by the University of Tabriz is gratefully
acknowledged.

References

1. J.H. Chen, C.Y. Cheng, W.Y. Chiu, C.F. Lee, N.Y. Liang, Eur. Polym. J. 44, 3271 (2008)
2. B.K. Sharma, A.K. Gupta, N. Khare, S.K. Dhawan, H.C. Gupta, Synth. Met. 159, 391 (2009)
3. Y. Li, J. Gong, M. McCune, G. He, Y. Deng, Synth. Met. 160, 499 (2010)
4. B.K. Sharma, A.K. Gupta, N. Khare, S.K. Dhawan, H.C. Gupta, J. Alloy Compd. 477, 370 (2009)
5. S. Ameen, M.S. Akhtar, S.G. Ansari, O.B. Yang, H.S. Shin, Superlattice Microst. 46, 872 (2009)
6. S.P. Sharma, M.V.S. Suryanarayana, A.K. Nigam, A.S. Chauhan, L.N.S. Tomar, Catal. Commun. 10,
905 (2009)
7. A.H. Gemeay, R.G. El-Sharkawy, I.A. Mansour, A.B. Zaki, J. Colloid Interf. Sci. 308, 385 (2007)
8. R. Qiu, L. Song, Y. Mo, D. Zhang, E. Brewer, React. Kinet. Catal. 94, 183 (2008)
9. D.S. Dhawale, D.P. Dubal, A.M. More, T.P. Gujar, C.D. Lokhande, Sens. Actuators B Chem. 147,
488 (2010)
10. A.A. Khan, M. Khalid, J. Appl. Polym. Sci. 117, 1601 (2010)
11. X. Chen, Z. Zhou, W. Lv, T. Huang, S. Hu, Mater. Chem. Phys. 115, 258 (2009)
12. S. Sonawane, B. Neppolian, B. Teo, F. Grieser, M. Ashokkumar, J. Phys. Chem. C 114, 5148 (2010)
13. Y. Yang, Y. Chu, Y. Zhang, F. Yang, J. Liu, J. Solid State Chem. 179, 470 (2006)
14. Y. He, Powder Technol. 147, 59 (2004)
15. X. Peng, Y. Chen, F. Li, W. Zhou, Y. Hu, Appl. Surf. Sci. 255, 7158 (2009)
16. M. Chang, X. Cao, H. Zeng, J. Phys. Chem. C 113, 15544 (2009)
17. W. Shen, Z. Li, H. Wang, Y. Liu, Q. Guo, Y. Zhang, J. Hazard. Mater. 152, 172 (2008)
18. S. Chakrabarti, B.K. Dutta, J. Hazard. Mater. B 112, 269 (2004)
19. S. Ameen, M.S. Akhtar, Y.S. Kim, O.B. Yang, H.S. Shin, Colloid Polym. Sci. 289, 415 (2011)
20. C. Wang, X. Wang, B.Q. Xu, J. Zhao, B. Mai, P. Peng, G. Sheng, J. Fu, J. Photoch, A. Photobio,
Chem. 168, 47 (2004)
21. M.A. Behnajady, N. Modirshahla, R. Hamzavi, J. Hazard. Mater. B 133, 226 (2006)
22. R. Qiu, D. Zhang, Y. Mo, L. Song, E. Brewer, X. Huang, Y. Xiong, J. Hazard. Mater. 156, 80 (2008)
23. S. Yang, Y. Ishikawa, H. Itoh, Q. Feng, J. Colloid Interface Sci. 356, 734 (2011)
24. X. Liu, H. Wu, F. Ren, G. Qiu, M. Tang, Mater. Chem. Phys. 109, 5 (2008)
25. N.I. Kovtyukhova, A.D. Gorchinskiy, C. Waraksa, Mater. Sci. Eng. B 69–70, 424 (2000)
26. S. Mridha, D. Basak, Appl. Phys. Lett. 92, 142111 (2008)
27. H. Zhang, R. Zong, Y. Zhu, J. Phys. Chem. C 113, 4605 (2009)
28. J. Tschirch, R. Dillert, D. Bahnemann, B. Proft, A. Biedermann, B. Goer, Res. Chem. Intermed. 34,
381 (2008)
29. F. Ping-feng, Z. Zhuo, P. Peng, D. Xue-gang, The Chinese Journal of Process Engineering 8, 65
(2008)
30. J. Kim, W. Choi, H. Park, Res. Chem. Intermed. 36, 127 (2010)

123

You might also like