You are on page 1of 6

Journal of Colloid and Interface Science 371 (2012) 101106

Contents lists available at SciVerse ScienceDirect

Journal of Colloid and Interface Science


www.elsevier.com/locate/jcis

Photo-catalyzed degradation of hazardous dye methyl orange by use of a


composite catalyst consisting of multi-walled carbon nanotubes and titanium
dioxide
Tawk A. Saleh a, Vinod K. Gupta a,b,
a
b

Chemistry Department, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia
Department of Chemistry, Indian Institute of Technology Roorkee, Roorkee 247 667, India

a r t i c l e

i n f o

Article history:
Received 31 October 2011
Accepted 12 December 2011
Available online 29 December 2011
Keywords:
Carbon nanotube
Titanium oxide
Methyl orange
Photo catalyst

a b s t r a c t
The high rate of electron/hole pair recombination reduces the quantum yield of the processes with TiO2
and represents its major drawback. Adding a co-adsorbent increases the photocatalytic efciency of TiO2.
In order to hybridize the photocatalytic activity of TiO2 with the adsorptivity of carbon nanotube, a composite of multi-walled carbon nanotubes and titanium dioxide (MWCNT/TiO2) has been synthesized. The
composite was characterized by means of X-ray diffraction (XRD), eld emission scanning electron
microscopy (FESEM), Fourier transform infrared absorption spectroscopy (FTIR), and diffuse reectance
UVvis spectroscopy. The catalytic activity of this composite material was investigated by application
of the composite for the degradation of methyl orange. It was observed that the composite exhibits
enhanced photocatalytic activity compared with TiO2. The enhancement in photocatalytic performance
of the MWCNT/TiO2 composite is explained in terms of recombination of photogenerated electronhole
pairs. In addition, MWCNT acts as a dispersing agent preventing TiO2 from agglomerating activity during
the catalytic process, providing a high catalytically active surface area. This work adds to the global discussion of how CNTs can enhance the efciency of catalysts.
2011 Elsevier Inc. All rights reserved.

1. Introduction
Research in water purication is considered a hot area of research due to the shortage in the supply of fresh drinking water
and which is at present a serious concern worldwide. The removal
of pollutants is becoming an important environmental issue owing
to their wide use in industrial applications. Different materials
were used for the removal of dyes by use of adsorption processes
[13]. Bottom ash and deoiled soya were used as adsorbents for
the removal of basic fuchsine and azo dye acid orange. Activated
carbon and activated rice husks were used for the adsorption of
safranin-T from wastewater. Carbon slurry developed from a waste
material has been used for the removal of endosulfan and
methoxychlor and vertigo blue 49 and orange DNA13 from water.
An agricultural waste, wheat husk, was used for the removal of
reactox golden yellow 3 RFN from aqueous solution. However,
the adsorption process has the major drawback of requiring
disposal of the resulting sludge residue. Methods by which the adsorbed pollutants can be degraded are required [4,5]. Photocatalytic purication of water containing pollutants has been
Corresponding author at: Department of Chemistry, Indian Institute of Technology Roorkee, Roorkee 247 667, India. Fax: +91 1332 286202.
E-mail address: vinodfcy@gmail.com (V.K. Gupta).
0021-9797/$ - see front matter 2011 Elsevier Inc. All rights reserved.
doi:10.1016/j.jcis.2011.12.038

extensively used for environmental applications. Titanium dioxide


(TiO2) is commonly used for this purpose because in addition to its
large band gap energy (3.2 eV for anatase titania), it is physically
and chemically stable, non-toxic, cost-effective and water insoluble under most conditions. It is widely used for various applications such as reduction of emissions and degradation of organic
pollutants. However, the efciency of TiO2 is reduced due to electron/hole pair recombination. Adding a co-adsorbent is considered
to be a good strategy for increasing the photocatalytic efciency of
TiO2.
For this purpose, composites of SiO2/TiO2 and glass/TiO2 have
been proposed [6,7]. Carbon-coated TiO2 showed high activity for
the decomposition of polyvinyl alcohol in water under UV irradiation and could be used repeatedly [8]. The inuence of activated
carbons on the photocatalytic degradation of aqueous organic
pollutants by UV-irradiated titania was also reported. Activated
carbontitania slurry under UV irradiation had a benecial effect
on the photocatalytic degradation of methyl orange, 4-chlorophenol, and herbicide but its effect depends on the properties of the
activated carbon [9]. Large surface area and high degree of dispersion of TiO2 powders in the reaction media were reported to favor
efcient photocatalytic performance [1012]. Due to their unique
electronic properties, multi-walled carbon nanotubes (MWCNTs)
have attracted considerable attention. In our recent papers, we

102

T.A. Saleh, V.K. Gupta / Journal of Colloid and Interface Science 371 (2012) 101106

Fig. 1. Structure of methyl orange (MO).

have reported various composites of MWCNT with metal oxides


such as MWCNT/alumina, MWCNT/iron oxide as magnetic composites, and MWCNT/tungsten oxide [1317]. In the present work,
a composite of multi-walled carbon nanotube/titanium dioxide
(MWCNT/TiO2) has been synthesized in order to hybridize photocatalytic activity of TiO2 with adsorptivity of MWCNTs. The catalytic activity of the composite was tested by degradation of
methyl orange (Fig. 1), taken as a model compound, under UV
irradiation.
Fig. 2. Pictorial representation of the modication of MWCNTs and preparation of
the MWCNT/TiO2 composite.

2. Experimental
2.1. Reagents and chemicals
MWCNT material was purchased from Cheap Tubes Com. Their
specications are as follows: purity, >95%; outer diameter,
3050 nm; inside diameter, 510 nm and 3050 nm; length, 10
20 lm; average specic surface area, 60 m2/g; Electrical Conductivity, >100 S/cm; Bulk density, 0.28 g/cm3; True density, 2.1 g/
cm3. Methyl orange (MO) was obtained from Fisher Scientic Company (M-216 76166 LOT 726294) and was used as received. The
water employed in all the studies was doubly distilled. The other
reagents obtained from Sigma Aldrich were of analytical purity
and were used as received.
2.2. Synthesis of MWCNT/TiO2 composite
MWCNTs were puried and oxidized before use. After the purication process, MWCNTs were acid treated with HNO3 and H2SO4
using a method reported earlier [15]. Briey, the MWCNTs were reuxed in a H2SO4HNO3 mixture (3:1 (v/v). Then, the mixture was
diluted, ltered, and washed with deionized water to remove the
excess acid. After that, it was dried overnight at 120 C in an oven.
Finally, it was milled/crushed to a powder.
In a typical experiment to prepare MWCNT/TiO2 composites,
MWCNTs were dispersed by sonication for 4 h. At the same time,
the titania precursor, namely titanium tetraisopropoxide solution,
was hydrolyzed and stirred until it became signicantly more viscous. It was then added into the dispersed nanotubes with continuous stirring and reuxed for 6 h. The mixture was washed,
ltered, and dried overnight. The nal step was calcination for
3 h at 350 C, Fig. 2. The composite was characterized as described
below.
2.3. Characterization
The as-synthesized composite was characterized by X-ray diffraction (XRD) (Shimadzu XRD Model 6000), to determine the crystalline phases and average crystallite size. The morphology and
microstructure of the composite were examined by means of eld
emission scanning electron microscope (FESEM, FEI Nova-Nano
SEM-600, the Netherlands). Fourier transform infrared (FTIR),
(MATTSON FTIR-100 Spectrometer) spectra were measured at
room temperature using the KBr Pellet technique. Pellets were prepared by mixing 1 mg of oxidized MWCNTs with 300 mg of KBr in
an agate mortar and then pressing the mixture at a pressure of

12 Pa for 3 min. The spectra of the composites were scanned from


4000 to 400 cm1. A nitrogen adsorption isotherm was obtained to
determine the specic surface area of oxidized MWCNT and TiO2
nanocoated carbon nanotubes. The BET (BrunauerEmmettTeller)
specic surface areas were approximately 154 and 190 m2/g,
respectively. Comparing the surface area of oxidized MWCNTs
with pristine nanotubes, one can notice an increase in the surface
area. This suggests that the oxidation of MWCNTs with HNO3 and
H2SO4 is an effective method to remove the amorphous carbon,
carbon black, and carbon nanoparticles introduced by the preparation process and probe inner cavities of CNTs which expose their
internal surface area. Functionalization causes opening up nanotube ends and generation of defects on the sidewall of nanotubes,
and therefore, the access into the cavity of the nanotubes can be
achieved.
The method used to determine the point of zero charge (PZC) of
the nanocomposite was as follows: dilute aqueous solutions of
NaOH and HCl were employed to prepare 12 solutions in the range
from pH 1.012.0. Next, 20 mL aliquots of these solutions were
pipetted into polyethylene vials and allowed to equilibrate for
5 h, and the pH of these was recorded before and after the addition
of 25.0 mg samples of the MWCNT/TiO2 to each solution. Each vial
was capped, shaken, and allowed to equilibrate overnight. The nal
pH was measured, and a plot of initial-vs.-nal pH revealed plateaus indicative of the points of zero charge (PZC).
2.4. Absorbance measurement
The UV-vis absorbance of MO solutions was measured for various
degradation times and various pH, dosage, and initial concentration
of MO. A JASCO V-570 UVvis spectrometer was employed to record
the spectra at wavelengths ranging from 200 to 600 nm. The percentage of degradation was calculated from the following equation:

" 
#
C C
 100
The percentage of degradation

C
where C is the initial concentration of dye and C is the concentration of MO at time t. All C and C values were obtained by the maximum absorption at 480 nm in the absorption spectrum in order to
evaluate the degradation efciency.
2.5. Photocatalytic experiments
To examine the catalytic activity of the composite, photodegradation experiments were carried out using a Pyrex reaction cell

T.A. Saleh, V.K. Gupta / Journal of Colloid and Interface Science 371 (2012) 101106

103

equipped with a cooling system. A 40 W ultraviolet lamp situated


in the middle of the quartz cell was used as the light source. The
photocatalytic activity of MWCNT/TiO2 composites was measured
for MO degradation in aqueous solution under UV irradiation. In
a typical experiment, 50 mM MO aqueous solutions were mixed
with the catalyst under constant stirring. Aliquots were taken
and analyzed, initially and periodically.
3. Results and discussion
3.1. Characterization analysis
The strong diffraction peak for the samples at the angle (2h) of
25.8 can be assigned to the C (0 0 2) reection of the hexagonal
graphite structure. The other diffraction peaks appeared at 2h of
42.7, 43.9, 53.5, and 77.5 can be attributed to the (1 0 0),
(1 0 1), (0 0 4), and (1 1 0) diffractions. The X-ray diffraction pattern
of the oxidized MWCNTs has been reported earlier [16].
As shown in Fig. 3, the XRD pattern of the MWCNT/TiO2 composite conrms the presence of a TiO2-anatase phase. The peak
at 25.8 corresponds to the (0 0 2) reection of MWCNTs. This peak
overlaps with the peak due to the anatase TiO2 (1 0 1) reection
plane at 2h = 25.3. The peak at about 2h = 53.5 corresponds to
the (0 0 4) reection of MWCNTs. The other peaks in the XRD pattern at 37.80 (0 0 4), 48.18 (2 0 0), and 54.09 (1 0 5) clearly represent
the anatase TiO2 phase[18,22]. Therefore, the anatase constitutes
the major crystal form in the MWCNT/TiO2 composite. The average
size of TiO2 nanoparticles on the surface of MWCNTs is estimated
from the Scherrer formula to be 8 nm.
The FESEM image conrms the morphology and microstructure
of MWCNT/TiO2 nanocomposite, as shown in Fig. 4. The high resolution image of the composite is presented in Fig. 5. The inset of the
gure shows clearly the nodes of titania on nanotubes both sides
surfaces and ends of the nanotubes. The presence of titanium oxide
on the surface of the nanotube was also conrmed by the EDX
spectrum. Fig. 6a depicts the quantitative analysis of the nanocomposite. It conrms the presence of carbon, oxygen, and titanium.
The results of the quantitative analysis of the MWCNT/TiO2 nanocomposite are depicted in the inset of Fig. 6a. Fig. 6b presents the
SEM image that shows the size of the selected sample area from
which the EDX was obtained. It should be noted that the sample
area was large in order to obtain a representative analysis.
Fig. 7a depicts the IR spectra of the oxidized MWCNTs. The band
at 1580 cm1 is attributed to C@C double bonds. The bands related
to carbonyl (1650 cm1), carboxyl (1710 cm1), and hydroxyl
(3400 cm1) functional groups are present in the spectrum of
MWCNTs to conrm the nanotubes oxidation forming new func-

Fig. 3. X-ray diffraction pattern of MWCNT/TiO2 composite.

Fig. 4. SEM image of MWCNT/TiO2.

tional groups on the nanotube surface. The assignment of the carbonyl band to an ester is conrmed by observation of a strong band
in the CAO stretching region at about 1180 cm1 (broad band
11001300 cm1).
Fig. 7b depicts the FTIR spectrum of the MWCNT/TiO2 nanocomposite. In the spectrum, the peak at around 700 cm1 is assigned to
TiAO and TiAOATi bonding of titania. It also shows the characteristics peaks of carbonyl, hydroxyl, and unsaturated carbon bonds in
the composite. The band intensity of the MWCNT/TiO2 in the region of the asymmetric carboxylate stretching mode at about
1400 cm1 is signicantly higher than that of MWCNT. This indicates a (CAOATi) interaction between the titanium atom and the
carboxylate group on the nanotube to form a COOATi moiety, in
which one of the carboxylate oxygen interacts with the titanium
ion [19].
3.2. Results of the photocatalytic activity
3.2.1. Time effect
UVvis spectroscopy was used to monitor the change in photocatalytic degradation of MO under UV light illumination. First, the
material was equilibrated in the dark followed by UV light illumination. Five sets of experiments were performed Fig. 8. The rst set
was conducted with only air oxidation. The second set of experiments was conducted under UV illumination without catalyst, to
determine the contribution of UV illumination to the degradation
of MO in the absence of a catalyst. The third set was conducted under UV irradiation in the presence of MWCNTs to observe the
adsorption-effect of MWCNT. The fourth set was conducted under
UV light illumination with the presence of TiO2 nanoparticles. The
fth set was conducted under UV irradiation in the presence of the
MWCNT/TiO2 nanocomposite. It was observed that when MWCNTs
was added to a solution of MO, only a slight degree of removal occurred. This can be ascribed to adsorption processes. Note that the
degradation of only 65% was observed when TiO2 was applied. This
is because of the rapid recombination between conduction band
(CB) electrons and valence band (VB) holes in pure TiO2. However,
after introducing the MWCNTs by chemically forming MWCNT/
TiO2 composite, the activity is remarkably enhanced and increased,
up to 93%. This can be attributed to the synergistic effects of
adsorption on MWCNT in the MWCNT/TiO2 composite. This
enhancement in the activity of the composite supports the idea
of formation of a common contact interface between the components of the composite, in which MWCNT acts as an efcient

104

T.A. Saleh, V.K. Gupta / Journal of Colloid and Interface Science 371 (2012) 101106

Fig. 5. HRSEM image of the synthesized MWCNT/TiO2.

adsorption trap of the MO, where it undergoes rapid photocatalytic


degradation. It is worth mentioning that a mechanical mixture of
TiO2 and MWCNTs exhibited almost the same activity as that of
TiO2 nanoparticles. As there was no signicant difference, those results are not shown in Fig. 8.
The kinetics of photocatalytic degradation of MO solution was
investigated using a 2 mM initial concentration. A linear relationship between ln(Co/C) and time (lnCo/C = kt) was obtained. This
indicates that the photocatalytic degradation of MO follows pseudo-rst-order kinetics.
3.2.2. Dosage
Photodegradation of 2  105 M of MO aqueous solution was
tested for different dosages of MWCNT/TiO2 composite, dosage
ranges from 0.1 to100 mg/50 ml. The pH of the solutions was 5.5,
and the exposure time of irradiation is 100 min. It was observed
that the decay was enhanced signicantly by increasing the doses
of the composite between 0.1 and 10 mg/50 ml. However, increasing the dosages from 10 to 100 mg/50 ml caused only insignicant
changes in degradation. The optimum removal was obtained when
10 mg/50 ml was used.

Fig. 6. EDX spectrum (a) with low magnication SEM image (b) of the synthesized
MWCNT/TiO2; inset in (a) is the table showing the percentage of each component in
the composite.

3.2.3. pH Effect
The pH is an important factor in inuencing the photocatalytic
process. In order to investigate the photodegradation process of
MO, the experiments were carried out at pH values of 3.0, 5.5,
7.0, and 9.0, respectively. The results obtained showed that the
efciency of the composite on the photodegradation of MO was affected by changing pH. The discussion of this effect is valuable but
complicated since several reaction mechanisms contribute to dye
degradation, such as hydroxyl radical attack, direct oxidation by
positive holes, and direct reduction by electrons. It was observed
that acidic to neutral (pH of 3, 5.5 and 7) solutions were more
effective in degrading MO than basic solutions (pH of 9).

T.A. Saleh, V.K. Gupta / Journal of Colloid and Interface Science 371 (2012) 101106

105

Fig. 9. Schematic diagram of the proposed mechanism of photocatalytic


degradation.

surface, and basic solutions made the surface negatively charged.


Therefore, in acidic to neutral media, a strong adsorption of MO
on the composite particles is preferred as a result of the electrostatic attraction of the positively charged particles toward the
MO. However, in basic solutions (pH = 9), dissociated form MO results in the formation of sodium ions that can react with adsorbed
OH, which reduced the amount of OH radicals [20].
Fig. 7. FTIR spectra of (a) oxidized MWCNTs (b) MWCNT/TiO2 composite.

3.3. Mechanism

Fig. 8. Photodegradation of MO versus irradiation time for several systems; Blank


experiment, UV with no catalyst, UV with: oxidized MWCNTs, TiO2 nanoparticles,
and MWCNT/TiO2 composite (initial concentration of MO is 1  105 M, dosage:
10 mg g/50 mL, pH5.5.

The increase in photocatalytic degradation between acidic pH


and neutral pH can be claried by measuring point-of-zero charge
(PZC) of the composite, which was found to be about 6.6. Thus, the
change in pH can inuence the adsorption of MO molecules onto
the composite surface. The adsorption is considered to be the
important step for the photo-oxidation to take place. In the acidic
to neutral solutions, the catalytic composite had positively charged

The high photocatalytic activity of MWCNT/TiO2 composite can


be understood as illustrated in Fig. 9. The VB electrons (e) of titania, under UV light irradiation, are excited to CB, creating holes (h+)
in the VB. Normally, these charge carriers quickly recombine, and
only a fraction of the electrons and holes participate in the photocatalytic reaction, resulting in low reactivity [21]. However, when
TiO2 was modied by the MWCNT, the catalytic activity of TiO2
was enhanced.
Generally, the photocatalytic performance of a photocatalyst is
governed by several factors such as surface area, adsorption capacity, light-response range, and the recombination time of photogenerated charge carriers [22,23]. Since the change in the surface area
is not signicant, the variation in adsorption capacity, determined
dominantly by the surface area, is not the major reason for the
photoactivity enhancement of MWCNT/TiO2 composites. Reasonably, it can be assumed that the dominant contribution of nanotubes in MWCNT/TiO2 composites is mainly determined by an
increase in recombination time for photogenerated electronhole
pairs. This is supported by the delocalized p-structure nature of
MWCNT, which facilitates the transfer of photoinduced electrons
and can perform as an excellent electron acceptor leading to
holeelectron separation. The injected electron is transferred to
the oxygen to generate the superoxide anion, which is considered
highly active radical that depredates the dye adsorbed on the
nanotube surface as shown in Fig. 9.
To clarify the mechanism, the diffuse reectance UVvis spectrum of the MWCNT/TiO2 composite was examined. Fig. 10 depicts

106

T.A. Saleh, V.K. Gupta / Journal of Colloid and Interface Science 371 (2012) 101106

No. 10-WAT1400-04 as part of the National Science, Technology


and Innovation Plan.
References

Fig. 10. Diffuse reectance UVvis spectra of TiO2 nanoparticles, pure MWCNTs,
MWCNT/TiO2 composite (CNT/TiO2), and mechanical mixture of MWCNTs and TiO2
(CNT + TiO2).

the optical absorption spectra of the diffuse reectance UVvis


spectra of the MWCNT/TiO2 composite by comparison with TiO2
nanoparticles, pure MWCNTs, and mechanical mixture of MWCNTs
and TiO2. Pure TiO2 nanoparticles without MWCNT substrate
shows a clear absorption edge at around 380 nm, and no absorption of visible light region above 380 nm is observed
[12,19,24,2549]. For comparison, the diffuse reectance UVvis
spectrum of the mixture of both components was examined, as depicted in Fig. 10. However, the MWCNT/TiO2 composite displayed
unusual UVVis spectra, which broaden to cover almost the entire
UVvis region. This suggests that the nanocomposite combines the
features of MWCNTs and TiO2, as well as generating new properties. This indicates the presence of interaction between MWCNT
and titania, which modify the process of the electron/hole pair formation. The interaction could be by coordination of carbonyl or by
ester bonding between carboxyl group of the nanotube and titania.

4. Conclusions
In this work, MWCNT/TiO2 was synthesized and characterized
via various characterization tools such as X-ray diffraction (XRD),
eld emission scanning electron microscope (FESEM), and Fourier
transform infrared absorption spectroscopy (FTIR). The composite
was tested for MO degradation. The results described in this manuscript show that the use of carbon nanotube as a support for TiO2
is effective for obtaining high degradation rate of methyl orange
dye in aqueous solutions.
Acknowledgment
The authors would like to acknowledge the support provided by
King Abdulaziz City for Science and Technology (KACST) through
the Science & Technology Unit at King Fahd University of Petroleum & Minerals (KFUPM) for funding this work through project

[1] V.K. Gupta, R. Jain, T.A. Saleh, A. Nayak, S. Malathi, S. Agarwal, Sep. Sci. Technol.
46 (2011) 839846.
[2] V.K. Gupta, R. Jain, M.N. Siddiqui, T.A. Saleh, S. Agarwal, S. Malati, D. Pathak, J.
Chem. Eng. Data 55 (2010) 52255229.
[3] I. Ali, V.K. Gupta, Nat. Protoc. 1 (2006) 26612667.
[4] Y. Kim, C. Kim, I. Choi, S. Rengraj, J. Yi, Environ. Sci. Technol. 38 (2004) 924
931.
[5] S. Wang, T. Terdkiatburana, M.O. Tad, Sep. Purif. Technol. 62 (2008) 6470.
[6] A. Hilmi, J.H.T. Luong, A.L. Nguyen, Chemosphere 38 (1999) 865874.
[7] M. Holgado, A. Cintas, M. Ibisate, C.J. Serna, C. Lopez, F. Meseguer, J. Colloid
Interface Sci. 229 (2000) 611.
[8] M. Inagaki, Y. Hirose, T. Matsunaga, T. Tsumura, Carbon 41 (2003) 26192624.
[9] J. Matos, J. Laine, J.M. Hermann, J. Catal. 200 (2001) 1020.
[10] S. Wang, K. Murata, Stud. Surf. Sci. Catal. 147 (2004) 691696.
[11] S. Wang, S. Zhou, J. Hazard. Mater. 185 (2011) 7785.
[12] Y. Li, L. Xiaodong, L. Junwen, Y. Jing, Water Res. 40 (2006) 11191126.
[13] V.K. Gupta, S. Agarwal, T.A. Saleh, Water Res. 45 (2011) 22072212.
[14] V.K. Gupta, S. Agarwal, T.A. Saleh, J. Hazard. Mater. 185 (2011) 1723.
[15] T.A. Saleh, Appl. Surf. Sci. 257 (2011) 77467751.
[16] T.A. Saleh, V.K. Gupta, J. Colloid Interface Sci. 362 (2011) 337344.
[17] T.A. Saleh, S. Agarwal, V.K. Gupta, Appl. Catal. B: Environ. 106 (2011) 4653.
[18] D. Reyes-Coronado, G. Rodrguez-Gattorno, M.E. Espinosa-Pesqueira, G.
Oskam, Nanotechnology 19 (2008) 145605.
[19] L. Fuks, D. Filipiuk, M. Majdan, J. Mol. Struct. 792793 (2006) 104109.
[20] L. Zhang, F. Lv, W. Zhang, R. Li, H. Zhong, Y. Zhao, Y. Zhang, X. Wang, J. Hazard.
Mater. 171 (2009) 94300.
[21] J.G. Yu, W.G. Wang, B. Cheng, B.L. Su, J. Phys. Chem. C 113 (2009) 6743.
[22] J.G. Yu, T. Ma, G. Liu, B. Cheng, Dalton Trans. 40 (2011) 66356644.
[23] Q. Xiang, J.G. Yu, M. Jaroniec, Nanoscale 3 (2011) 36703678.
[24] A. Akyol, H.C. Yatmaz, M. Bayramoglu, Appl. Catal. B: Environ. 54 (2004) 925.
[25] V.K. Gupta, P.J.M. Carrott, M.M.L. Ribeiro, Suhas, Crit. Rev. Environ. Sci.
Technol. 39 (2009) 83842.
[26] A.K. Jain, V.K. Gupta, B.B. Sahoo, L.P. Singh, Analytical Proceedings including
Analytical Communications 32 (1995) 99101.
[27] S.K. Srivastava, V.K. Gupta, M.K. Dwivedi, S. Jain, Analytical Proceedings
including Analytical Communications 32 (1995) 2123.
[28] A.K. Jain, V.K. Gupta, L.P. Singh, Analytical Proceedings including Analytical
Communications 32 (1995) 263265.
[29] S.K. Srivastava, V.K. Gupta, S. Jain, Electroanalysis 8 (1996) 938940.
[30] A.K. Jain, V.K. Gupta, L.P. Singh, P. Srivastava, J.R. Raisoni, Talanta 65 (2005)
716721.
[31] V.K. Gupta, A. Rastogi, J. Hazardous Materials 154 (2008) 347354.
[32] V.K. Gupta, P. Kumar, Anal. Chim. Acta 389 (1999) 205212.
[33] V.K. Gupta, A. Mittal, L. Krishnan, Jyoti Mittal, Ind. Engg. Chem. Res. 45 (2006)
14461453.
[34] V.K. Gupta, I. Ali, Suhas, D. Mohan, J. Colloid Interface Sci. 265 (2003) 257264.
[35] V.K. Gupta, A. Rastogi, J. Hazardous Materials. 163 (2009) 396402.
[36] V.K. Gupta, A. Rastogi, J. Hazardous Materials. 153 (2008) 759766.
[37] V.K.Gupta, Rajni Mangla, S. Agarwal, Electroanalysis 14 (2002) 11271132.
[38] V.K. Gupta, A. Rastogi, M.K. Dwivedi, D. Mohan, Sep. Sci. Technol. 32 (1997)
28832912.
[39] V.K. Gupta, D. Mohan, S. Sharma, Sep. Sci. Technol. 33 (9) (1998) 13311343.
[40] V.K. Gupta, Arshi Rastogi, J. Colloid Interface Sci. 342 (2010) 533539.
[41] V.K. Gupta, A.K. Jain, P. Kumar, Sens. Actuat. B. 120 (2006) 259265.
[42] V.K. Gupta, Arshi Rastogi, A. Nayak, Adsorption studies on the removal of
hexavalent, J. Colloid Interface Sci. 342 (2010) 135141.
[43] V.K. Gupta, I. Ali, V.K. Saini, Water Research. 41 (2007) 33173326.
[44] V.K. Gupta, M. Al Khayat, A.K. Singh, Manoj. K. Pal, Anal. Chim. Acta 634 (2009)
3643.
[45] V.K. Gupta, R.N. Goyal, R.A. Sharma, Int. J. Electrochem. Sci. 4 (2009) 156172.
[46] A.K. Jain, V.K. Gupta, U. Khurana, L.P. Singh, Electroanalysis 9 (1997) 857860.
[47] V.K. Gupta, A. Rastogi, Coll. Surfaces B. 64 (2) (2008) 170178.
[48] A.K. Jain, V.K. Gupta, L.P. Singh, U. Khurana, Analyst. 122 (1997) 583586.
[49] V.K. Gupta, P. Singh, N. Rahman, J. Colloid Interface Sci. 275 (2004) 398402.

You might also like