You are on page 1of 9

Journal of Photochemistry and Photobiology A: Chemistry 325 (2016) 116–124

Contents lists available at ScienceDirect

Journal of Photochemistry and Photobiology A:


Chemistry
journal homepage: www.elsevier.com/locate/jphotochem

Photodegradation of methyl violet 6B and methylene blue using


tin-oxide nanoparticles (synthesized via a green route)
Archita Bhattacharjee, M. Ahmaruzzaman* , Th. Babita Devi, Jayashree Nath
Department of Chemistry, National Institute of Technology, Silchar 788010, Assam, India

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

Article history:
Received 29 November 2015 Green synthesis of tin dioxide nanoparticles were developed by microwave heating method using 1:1,
Received in revised form 28 March 2016 1:2 and 1:3 volumetric ratio of water and glycerol, wherein glycerol acts as a good complexing as well as
Accepted 30 March 2016 capping agent. This method resulted in the formation of spherical SnO2 nanoparticles with an average
Available online 31 March 2016 diameter 8–30 nm. The synthesized SnO2 NPs were characterized by transmission electron microscopy
(TEM), selected area electron diffraction (SAED) and Fourier transformed infrared spectroscopy (FT-IR).
Keywords: The optical properties were investigated using UV–vis spectroscopy. The photocatalytic activity of
SnO2 nanoparticles synthesized SnO2 NPs was evaluated for the degradation of two different toxic dyes namely, Methyl Violet
Microwave heating
6B and Methylene blue dye under direct sunlight.
Glycerol
ã 2016 Elsevier B.V. All rights reserved.
Photocatalyst
Methyl violet 6B
Methylene blue

1. Introduction poses a threat to water bodies and our ecosystem. Hence, complete
removal of dye from industrial waste water is vital for reducing
Tin dioxide (SnO2) is an n-type semiconductor with a wide band water pollution. In this study, methyl violet 6B (MV6B) and
gap of 3.6 eV [1]. SnO2 with a tetragonal crystal structure, is the methylene blue (MB) dyes are selected. Methyl violet 6B is a water
most intensively explored metal oxide due to its potential soluble dye, used in textile industries, paper dyeing, paints and
applications in catalysis, gas sensors, dye-based solar cells, printing ink. MV6B is also used as a disinfectant and is found very
transparent conducting electrodes, rechargeable lithium batteries, poisonous to animals. MV6B is carcinogenic in nature. Methylene
etc. [2–9]. A variety of methods were employed for the synthesis of blue is also a water soluble dye used as a colorant in textile
SnO2 nanoparticles [10–13]. The majority of the process needs industries. It is toxic and causes anemia, bladder irritation and
longer reaction time, utilization of poisonous chemicals and gastrointestinal problems. Therefore, the removal of such dye
critical reaction conditions. However, large-scale synthesis of effluents from water is very necessary for avoiding adverse health
crystalline SnO2 nanoparticles by a green and facile method is still hazards and also to prevent our ecosystem. In recent years, there
a challenging job. In this paper, we reported a green synthesis of are various techniques developed for the removal of such dye
SnO2 nanoparticles by microwave heating method using glycerol, pollutants from water, however most of them were associated with
wherein glycerol acts as a good complexing as well as capping certain flaws [14]. Herein, we reported the sundriven photo-
agent. catalytic degradation of the dyes namely methyl violet 6B and
Dyes constitute a major class of organic compound having huge methylene blue in presence of synthesized SnO2 nanoparticles,
applications in our daily life. They are used in textile industries, acting as catalyst. This method does not cause secondary pollution
dyeing, printing, cosmetics etc. However, most of the dyes are toxic and therefore proves to be an effective method for the removal of
in nature. The effluents coming out from these industries toxic dyes from waste water. Several metal oxide semiconductors,
contaminate water system thereby causing water pollution. This such as SnO2, TiO2, ZnO, NiO, V2O5, etc have been used as
photocatalyst for the degradation of organic pollutants in water
[4,15–19]. Among them, SnO2 is found to an efficient photocatalyst
due to its large reactivity of surface, immense absorption capacity
* Corresponding author. of light radiation and huge number of active sites [15].
E-mail address: md_a2002@rediffmail.com (M. Ahmaruzzaman).

http://dx.doi.org/10.1016/j.jphotochem.2016.03.032
1010-6030/ ã 2016 Elsevier B.V. All rights reserved.
A. Bhattacharjee et al. / Journal of Photochemistry and Photobiology A: Chemistry 325 (2016) 116–124 117

2. Experimental 60
55 (a) S1

(211)
(110)
2.1. Materials (b) S2

(101)
50
(c) S3
The reagents, stannous chloride dihydrate (SnCl22H2O), 45
glycerol, methyl violet 6B and methylene blue were of analytical 40

Absorbance (a.u.)
grade (AR) and purchased from Merck. The reaction was carried out

(301)
35

(200)

(310)
(220)

(1)12
in LG Microwave oven MG-557B (Input: 230V-50 Hz; Reoutput:

(321)
(002)

(202)
900W; Microwave: 1350W; Frequency: 2450 MHz). Double dis- 30
tilled water was used for the synthesis of SnO2 nanoparticles. 25 (c)
20
2.2. Synthesis of SnO2 nanoparticles 15 (b)
10
In this present work, microwave heating method was employed
for the synthesis of SnO2 nanoparticles. Three different sets of 5 (a)
reactions were carried out by dissolving 0.3505 g of SnCl22H2O in 0
100 ml distilled water. To these three different sets, 1:1, 1:2 and -5
1:3 volumetric ratio of distilled water and glycerol were added 10 20 30 40 50 60 70 80
drop wise with constant stirring. The reaction mixtures were then
Wavelength (nm)
kept in a microwave oven and irradiated with sixty 10 s shots. This
method results in the formation of white precipitate. The obtained Fig. 1. XRD spectra of synthesized SnO2 QDs (a) S1, (b) S2 and (c) S3.
precipitate was centrifuged and washed three to four times with
double distilled water. The white product was finally calcined at sunlight irradiation. Before exposing to direct sunlight, both the
400  C for 2 h and then collected for characterization. The SnO2 solutions were also kept in dark for next 30 min to achieve the
nanoparticles obtained for different ratios of water and glycerol adsorption/desorption equilibrium. The photodegradation of the
mixtures were marked as S1 (1:1), S2 (1:2) and S3 (1:3) dyes were carried out on a sunny day between 10 a.m-2 p.m when
respectively. there were minimum fluctuations in solar intensity. The experi-
ment was carried out in Silchar city in the month of May (outside
2.3. Characterization of SnO2 nanoparticles temperature 35–30  C) on a sunny day when the average solar
radiation was 4.29 kwh/m2/day. At regular interval of time, 5 ml of
SnO2 nanoparticles were characterized by powder X-ray the each suspension was withdrawn and centrifuged immediately.
diffraction (XRD) method using Phillips X’Pert PRO diffractometer The progress of the reaction was monitored by using UV–vis
with Cu-Ka radiation of wavelength 1.5418 Å. The size, morpholo- spectrophotometer. The degradation products were identified
gy and diffracted ring pattern of SnO2 nanoparticles were using LC–MS analysis.
determined by JEM-2100 Transmission Electron Microscope. The
Infrared spectra were recorded in the wave number range from 400 3. Results and discussion
to 4000 cm1 by using Bruker Hyperion 3000 FTIR spectrometer.
UV–vis absorption spectra of the synthesized SnO2 nanoparticles 3.1. FT-IR studies
were recorded using Cary 100 BIO UV–vis spectrophotometer
equipped with 1 cm quartz cell. The degradation products were Fig. S1 (Supplementary information) represented the FTIR
identified by LC–MS (410 Prostar Binary LC with 500 MS IT PDA spectra of synthesized S1, S2 and S3 nanoparticles. The assignment
Detectors). of FTIR bands of S1, S2 and S3 nanoparticles were recapitulated in
Table 1. The peak around 600 cm1 was due to Sn OSn
2.4. Photocatalytic activity of synthesized SnO2nanoparticles stretching mode of surface bridging oxide formed by the
condensation of adjacent hydroxyl groups [10,11,20,2]. FTIR spectra
The photocatalytic activity of synthesized SnO2 nanoparticles were recorded in order to detect the formation of SnO2 and also to
were evaluated by the degradation of two different organic dyes perceive the presence of glycerol as capping agent. The band
namely, methyl violet 6B (MV6B) and methylene blue (MB). To around 3415 cm1 and 1628 cm1 were attributed to O H and
examine the photocatalytic activity, 10 mg of synthesized SnO2 C¼O stretching vibration. This indicated the presence of glycerol
photocatalyst was dispersed in 200 ml of 104 M aqueous solution molecules on the surface of SnO2 which act as a capping agent.
of MV6B and MB dye. Both the solutions were then exposed to Hence, the formation of SnO2 was confirmed from the FTIR spectra.

3.2. XRD studies


Table 1
Assignment of FTIR bands of synthesized SnO2 nanoparticles.
The XRD patterns of SnO2 nanoparticles (S1, S2 and S3) were
SnO2 NPs Precursor FT-IR bands represented in Fig. 1(a, b and c) respectively. The XRD pattern was
S1 SnCl2 + (Water + Glycerol) (Ratio = 1:1) 3415 cm1 (nOH str.) recorded in order to investigate the crystal structure, purity and
600 cm1 (nSnOSn str.) crystalline nature of synthesized SnO2 nanoparticles. The peak
positions observed at 2u values corresponded to (110), (101), (200),
S2 SnCl2 + (Water + Glycerol) (Ratio = 1:2) 1628 cm1 (nC¼O str.)
3415 cm1 (nOH str.)
(211), (220), (002), (310), (301), (202) and (321) planes, respec-
600 cm1 (nSnOSn str.) tively. The XRD pattern clearly reflected the tetragonal crystal
structure of SnO2 nanoparticles (JCPDS 41-1445) [21,22]. No
S3 SnCl2 + (Water + Glycerol) (Ratio = 1:3) 1628 cm1 (nC¼O str.) characteristic peaks were found for other impurities which
3415 cm1 (nOH str.)
indicated the purity of the synthesized SnO2 nanoparticles. Hence,
600 cm1 (nSnOSn str.)
1628 cm1 (nC¼O str.) from the XRD pattern it was confirmed that SnO2 nanoparticles
118 A. Bhattacharjee et al. / Journal of Photochemistry and Photobiology A: Chemistry 325 (2016) 116–124

with tetragonal crystal structure were formed at a temperature


400  C with different volumetric ratio of water and glycerol. From
the XRD pattern, it was also evident that the peaks became
gradually sharper and the full width half maximum (FWHM) were
reduced with an increase in the ratio of glycerol in the mixture.
This indicated that the particle size and crystallinity of SnO2
nanoparticles increases with an increase in the ratio of glycerol.
The average crystallite size of S1, S2 and S3 NPs calculated using
Debye–Scherrer equation were found to be 9, 14 and 30 nm,
respectively. This shows that the crystalline size of SnO2 nano-
particles increases with an increase in the glycerol volume. Hence,
from the above studies it is evident that the size of SnO2
nanoparticles can be tuned by controlling the volume of glycerol.

Fig. 3. (a) TEM microphotograph of synthesized SnO2 NPs (S2), (b) HRTEM image
(inset of Fig. 3a) synthesized SnO2 NPs (S2), (c) SAED pattern of synthesized SnO2
NPs (S2).

3.3. Electron microscopic analysis

The detailed observations of electron microscopic analysis for


the synthesized SnO2 NPs (S1, S2 and S3) are shown in Figs. 2–4 .
The morphology and the size distribution of S1 NPs formed at
400  C can be depicted from the TEM and HRTEM images. Fig. 2(a,
b) represented the TEM and HRTEM microphotograph of S1 NPs.
Fig. 2(a) showed the formation of spherical SnO2 nanoparticles
with an average diameter of 8–10 nm. The spacing between
adjacent lattice fringes calculated from the HRTEM image (Fig. 2b;
inset of Fig. 2a) was found to be 0.26 nm which corresponds to
(101) lattice plane. The SAED pattern (Fig. 2c) depicted the
Fig. 2. (a) TEM microphotograph of synthesized SnO2 NPs (S1), (b) HRTEM image presence of concentric diffraction rings which indicated the
(inset of a) of synthesized SnO2 NPs (S1), (c) SAED pattern of synthesized SnO2 NPs
(S1).
crystalline nature of SnO2 nanoparticles. The lattice spacings were
A. Bhattacharjee et al. / Journal of Photochemistry and Photobiology A: Chemistry 325 (2016) 116–124 119

The morphology and the size distribution of S3 NPs formed at


400  C can be depicted from the TEM and HRTEM images (Fig. 4a,
b). The TEM image (Fig. 4a) showed the formation of spherical
nanoparticles having an average diameter of 30 nm. The spacing
between lattice fringes calculated from the HRTEM image (Fig. 4b;
inset of Fig. 4a) was found to be 0.174 nm which corresponded to
(211) lattice plane. The SAED pattern (Fig. 4c) showed concentric
diffraction rings which revealed the crystalline nature of SnO2
nanoparticles. The lattice spacing were calculated from the SAED
pattern (Fig. 4c) and it corresponded to (110), (101), (200) and (211)
lattice plane respectively. The lattice plane obtained from the SAED
pattern matches well with the tetragonal crystal structure of SnO2
[21,22].
From the TEM images and SAED pattern, it was evident that the
SnO2 nanoparticles formed at 400  C with different volumetric
ratio of water and glycerol were spherical and crystalline in nature
with an average particle size ranging from 8–30 nm. Hence, it can
be stated that with an increase in the ratio of glycerol in the
mixture, the grain size of the nanocrystallites started to grow,
thereby exhibiting a high degree of crystallinity. At higher volume
of glycerol, the neighboring particles agglomerate thereby
increasing the grain size of SnO2 nanoparticles. The synthesized
SnO2 nanoparticles also possess tetragonal crystal structure which
was in resemblance with the XRD pattern.

3.4. Optical properties

UV–vis spectra were recorded in order to explore the optical


properties of synthesized SnO2 NPs. Figs. S2a, S3a and S4a
(Supplementary information) represented the absorption spectra
of S1, S2 and S3 NPs respectively. From Fig. S2(a), it was evident
that broad peaks of negligible intensity started to originate around
250–260 nm and 220 nm after calcination at 400  C. From Figs. S3
(a) and S4(a) it was apparent that the intensity of these broad peaks
started to increase with an increase in the volumetric ratio of
glycerol.
Absorption spectra were recorded also to obtain the band gad
energy of synthesized SnO2 NPs. The band gap energy (Eg) of the
synthesized SnO2 nanoparticles can be calculated from the
absorption spectra using Tauc plot. For semiconductor nano-
particles, following equation has been used to relate absorption
coefficient with incident photon energy:
a(n)hn = K (hn  Eg)n (1)
where Eg is the band gap energy, hn is the incident photon energy,
K is a constant, a(n) is absorption coefficient which can be defined
by the Beer–Lambert’s law as follows: a(n) = 2:303 Ar/cl, where A
Fig. 4. (a) TEM microphotograph of synthesized SnO2 NPs (S3), (b) HRTEM image is the absorbance, r is the density of the SnO2 nanoparticles, c is the
(inset of a) synthesized SnO2 NPs (S3), (c) SAED pattern of synthesized SnO2 NPs concentration and l is the path length. The exponent ‘n’ in Eq. (1)
(S3).
depends on the type of the transition and ‘n’ have values 1/2, 2, 3/2,
3 for allowed direct, allowed indirect, forbidden direct and
calculated from Fig. 2(c) and found to be 0.34, 0.28, 0.24, 0.178 and
forbidden indirect transitions, respectively. In case of SnO2
0.16 nm which corresponded to the lattice plane (110), (101), (200),
nanoparticles the value of n is 1/2 for allowed direct transition.
(211) and (220), respectively. The lattice plane obtained from the
Therefore, by plotting (ahn)2 versus hn and extrapolating the curve
SAED pattern matches well with the tetragonal crystal structure of
to zero absorption coefficient band gap energy (Eg) can be
SnO2 [21,22].
determined using Eq. (1) [23,22]. Figs. S2(b), S3(b) and S4(b)
Fig. 3(a, c) represented the TEM image and SAED pattern of
represented graphically the plot of (ahn)2 versus incident photon
S2 nanoparticles. The microstructure of the nanoparticles was
energy (hn) for S1, S2 and S3 NPs formed at 400  C. By extrapolating
investigated by the HRTEM image (Fig. 3b; inset of Fig. 3a). Fig. 3(a)
the curve to zero absorption coefficient, band gap energy was
showed the formation of spherical SnO2 nanoparticles having an
calculated using Eq. (1) and it was found to be 4.17, 4.00 and 3.78 eV
average size of 13–17 nm. From the HRTEM image (Fig. 3b), the
for S1, S2 and S3 NPs, respectively. In case of semiconductors, band
lattice spacing was found to be 0.23 nm which corresponded to
gap energy depends on the particle size. Hence, there is a decrease
(200) plane. The lattice spacings were calculated from the SAED
in the band gap energy with an increase in grain size of the
pattern (Fig. 3c) and this corresponded to (110), (101), (200) and
synthesized SnO2 NPs. Table 2 represented the summary of results
(211) lattice plane respectively. The SAED pattern clearly reflected
obtained from absorption spectra of synthesized SnO2 NPs.
the tetragonal crystal structure of SnO2 [21,22].
120 A. Bhattacharjee et al. / Journal of Photochemistry and Photobiology A: Chemistry 325 (2016) 116–124

Table 2 and 3.9. The progress of the photodegradation reaction was also
Summary of results obtained from absorption spectra of synthesized SnO2 NPs.
monitored by recording the changes in the absorption spectra of
SnO2 NPs Precursor Band gap energy (eV) MV6B and MB solution at regular interval of time.
S1 SnCl2 + (Water + Glycerol) (Ratio = 1:1) 4.17 Fig. 5(a) and (b) represented the absorption spectra for the
S2 SnCl2 + (Water + Glycerol) (Ratio = 1:2) 4.00 photocatalytic degradation of MV6B and MB dye respectively using
S3 SnCl2 + (Water + Glycerol) (Ratio = 1:3) 3.78 SnO2 NPs as photocatalyst under direct sunlight. The UV–vis
spectra of the MV6B and MB showed a strong absorption band at
580 and 663 nm respectively. It was evident that addition of SnO2
3.5. Role of glycerol in the synthesis of SnO2 NPs
nanoparticles lead to a decrease in the absorption band with time.
The intensity of the bands gradually decreased with an increase in
The glycerol content in the reaction system was decisive for the
irradiation time in both cases. The absorption band at 580 and
particle size. The TEM image of the sample showed that the
663 nm for MV6B and MB dye disappeared completely within
particles agglomerate when the glycerol content was increased to
270 min and 240 min respectively. The color of both the solution
some extent. The formation mechanism of SnO2 particles in
also faded away which indicated complete destruction of the
glycerol was shown in Scheme 1. In glycerol solution, SnCl2 was
chromophoric structure of dye. The photodegradation reaction
surrounded and protected by glycerol molecules and Cl was
follows pseudo first order kinetics and the rate constant of the
replaced by glycerol molecules [24]. When the system was
reaction can be obtained from the linear plot of lnAt versus
irradiated with microwaves, the water molecules might attack
irradiation time, t [25]. The value of rate constant (k) was found to
the Sn OCH2CH(OH)CH2OH bonds as a result of replacing the
be 0.7  102 min1 and 1.07  102 min1 for MV6B and MB dye,
glycerol molecules to form Sn(OH)2 and finally decomposed into
respectively (Figs. S5a and S6a ; Supplementary information). It
SnO and then was further oxidized to SnO2. The plausible
was also evident that 96.2% and 96% of MV6B and MB degraded
mechanism for the formation of nanoparticles can be visualized
photochemically within 270 min and 240 min, respectively using
as (Scheme 1):
SnO2 NPs as photocatalyst (Figs. S5b and S6b ; Supplementary
information).
3.6. Photocatalytic activity of synthesized SnO2 NPs
Simultaneously, to confirm the photocatalytic activity of
synthesized SnO2 NPs, a control experiment was carried out
The photocatalytic activity of synthesized SnO2 NPs was
where, MV6B and MB dye solutions were irradiated with sunlight
evaluated by carrying out the degradation of Methyl violet 6B
in the absence of SnO2 NPs. Both the dye does undergo negligible
(MV6B) and Methylene Blue (MB) dye under solar irradiation. The
degradation (about 5–6%) within 240 min. Similarly, when both the
degradation of dyes depends on the band gap energy of the
dye solution (MV6B and MB) were treated with SnO2 NPs in the
photocatalyst. Under solar irradiation, photocatalyst possessing
absence of sunlight i.e., in dark, the dyes undergo negligible
lower band gap energy is more preferable. Therefore, the SnO2
degradation. This confirmed that the dyes namely MV6B and MB
nanoparticles possessing band gap energy of 3.78 eV (S3) was
undergo photocatalysis under solar radiation.
utilized as photocatalyst for the degradation of dyes namely MV6B
and MB. The degradation of MV6B and MB passes through various
3.7. Mechanism for the degradation of MV6B and MB dyes
intermediate stages. For example, on sunlight irradiation, MB
undergoes degradation to leuco MB in the first step, which is prior
The semiconductor photocatalyst (SnO2 nanoparticles) having
to mineralization. Hence, on solar irradiation, various chemical
wide band gap energy cannot absorb the visible light directly from
reactions take place which are explained vividly in section 3.7,3.8
solar irradiation. However, light absorption in the visible range can

Scheme 1. Formation of SnO2 NPs.


A. Bhattacharjee et al. / Journal of Photochemistry and Photobiology A: Chemistry 325 (2016) 116–124 121

0.5
(a) Methyl violet 6B (MV6B) (b) Methylene Blue (MB)
1.2
0 min
0.4 15 min
30 min 0 min
45 min
1.0
10 min
60 min
90 min 20 min
120 min
0.3 0.8

Absorbance (a.u.)
150 min 30 min
Absorbance (a.u.)

180 min
40 min
210 min
240 min 50 min
270 min 0.6
0.2 60 min
90 min
0.4 120 min
180 min
0.1
240 min
0.2

0.0
0.0
400 500 600 700 400 500 600 700
Wavelength (nm)
Wavelength (nm)
Fig. 5. (a) Photodegradation of methyl violet 6B (MV6B) dye by solar irradiation using synthesized dye-sensitized SnO2 photocatalyst, (b) Photodegradation of methylene blue
(MB) dye by solar irradiation using synthesized dye-sensitized SnO2 photocatalyst.

be extended by photosensitization [26]. Since, the dye molecules solar irradiation [27,28]. Hence, SnO2 nanoparticles act as dye-
were sensitive to visible light, hence dye molecules enhance the sensitized photocatalyst for the visible light degradation of both
visible light absorption of sensitized photocatalysts by injecting the dyes. The probable mechanism for the degradation of the dyes
electrons from the lowest unoccupied molecular orbital (LUMO) of viz., MV6B and MB using dye-sensitized SnO2 photocatalyst can be
excited dye into the conduction band (CB) of the photocatalyst depicted as follows [27,28]:
[26]. The electrons of the dye molecules get excited into singlet and Dye + hn ! 1Dye + 3Dye.
triplet state which was further followed by the electron injection
Dye* or 3Dye + SnO2 ! Dye+ + SnO2 (eCB)
from the excited dye molecule to the conduction band of the
photocatalyst. This lead to the formation of cationic dye radicals
(dye+). The electron injected to the conduction band of the dye-
SnO2 (eCB) + O2 ! O2 + SnO2
sensitized photocatalyst (i.e., SnO2 nanoparticles) reacts with the
oxygen adsorbed from air to form oxidizing species (O2, HOO
and OH radicals) which oxidize the dye pollutants. The
semiconductor photocatalyst act as electron-transfer mediator O2 + H+ ! OOH+
and carry out the photooxidation of the dyes successfully under

Scheme 2. Probable mechanistic pathway for the degradation of the dyes using dye-sensitized SnO2 photocatalyst under direct sunlight.
122 A. Bhattacharjee et al. / Journal of Photochemistry and Photobiology A: Chemistry 325 (2016) 116–124

OOH+ + O2 + H+ ! O2 +H2O2 N(CH 3)2

H2O2 + O2 ! OH + OH + O2

Dye+ + OH ! Degradation products H 3C

N NH2

H 3C m/z= 344.4
Dye+ + O2 ! Degradation products

N(CH3)2 N(CH3)2
Dye+ + O2 ! Degradation products

Scheme 2 visualizes the degradation pathway obeyed by the


dye molecules (viz., MV6B and MB) on irradiation with sunlight in
presence of dye-sensitized SnO2 photocatalyst.

3.8. Identification of degradation products of methyl violet 6B dye H3C


H2N NH2 N NH2
m/z= 316.4 H
MS fragmentation pattern of the degradation product obtained m/z= 330.4
during the degradation of MV6B dye was shown in Fig. S7(a)
(Supplementary information). The degradation products formed H N H

during the photodegradation of MV6B under direct sunlight were H N CH3

identified by the mass fragmentation pattern obtained from LC–MS


and the mechanism for the formation of the degradation products
were schematically depicted in Scheme 3.
The OH radical formed during the degradation process
initiated the degradation of MV6B dye. The OH radical act as an
oxidizing agent and attack the methyl group to form aldehyde. The H2N NH 2
aldehyde group oxidizes to carboxylic acid group which decar- H2N NH2 m/z= 288.3
m/z= 302.3
boxylates into CO2. All the methyl groups present in the ring were
demethylated by OH radical thereby leading to the formation of p-
rosaniline which was also identified in the mass fragmentation H N H
pattern (m/z = 288). The OH radicals further attack the demethy-
lated compound and lead to the decomposition of the dye into H2N
various fragments as detected in the mass fragmentation pattern m/z= 93

(m/z = 213, 137, 93, 94). OH

3.9. Identification of degradation products of methylene blue dye


H2N NH2

The degradation products generated during the degradation of m/z= 288 H2N NH2

MB dye was analyzed by LC–MS and from the mass spectra the
fragmented ions were identified. Fig. S7(b) (Supplementary
information) represented the MS spectra of the degradation
O
O
products obtained during the degradation of MB dye. The
identified fragmented products analyzed from the mass spectra OH
were represented in Scheme 4 and the pathway for the formation
H 2N
of the degradation products was also depicted in Scheme 4. H 2N NH2
m/z= 137 m/z=213
On sunlight irradiation, MB undergoes degradation to leuco MB
in the first step, detected in the MS fragmentation pattern at m/
z = 284, which is prior to mineralization. During the photo-
degradation process, the photo-generated holes (h+) and the
OH radicals act as oxidizing agent. The OH radical generated
during photodegradation process attack the C–S+ = C functional H2N HO NH2

group of MB to form C–S(¼O)–C thereby making its oxidizing state m/z= 94 m/z= 93

to pass from 2 to 0, which was identified from the MS


fragmentation pattern obtained at m/z = 301. The sulphoxide group Scheme 3. Photodegradation pathway of MV6B dye and identification of the
degradation products.
was again attacked by the OH radical to produce sulphone (not
detected in MS) wherein the oxidation state of sulphur increases
from 0 to +5 and also lead to the dissociation of the ring. The
sulphone was then attacked by third OH radical to produce to form phenolic compound (detected in MS fragmentation
sulphonic acid (oxidation state = +6) which was identified from MS pattern).
pattern showing m/z value at 201. The attack of OH radical again The methyl group present in the ring was also attacked by OH
lead to the release of SO42 ions. This reacts again with OH radical radical to form corresponding aldehyde which on further oxidation
A. Bhattacharjee et al. / Journal of Photochemistry and Photobiology A: Chemistry 325 (2016) 116–124 123

CH 3
H 3C
N S N
m/z= 284 CH 3
H3C

H
N

CH 3
H3C
N S N
CH3
H 3C
O
m/z= 301

H 2N

H3C CH 3
N SO3H N
H3C CH 3
m/z= 201 m/z= 136

HO
O HO
O
O

HC CH 3
CH
HOC
N SO3 H N
N SO3 H N
H3C CH3
m/z= 215 CH3
H 3C m/z= 151
m/z= 137

HO
HO
O
H H
COH
N SO3H N
H
CH 3 N
HC m/z= 201 m/z= 123 CH 3
N SO3H

H 3C O
m/z= 187 HO
HO

H
H
H N
H N CO
N SO3H CH H
N SO3H
HOC
m/z= 137 O
H m/z= 173 O

O
HO

N
H
OH
m/z= 109
m/z= 94
Scheme 4. Photodegradation pathway of MB dye and identification of degradation products.
124 A. Bhattacharjee et al. / Journal of Photochemistry and Photobiology A: Chemistry 325 (2016) 116–124

give rise to the corresponding acid which further undergoes References


decarboxylation (detected in MS fragmentation pattern).
The dimethyl phenyl amino group (detected in MS; m/z = 136) [1] S.C. Yeow, W.L. Ong, A.S.W. Wong, G.W. Ho, Sens. Actuators B 143 (2009) 295–
301.
was also attacked by OH radical producing an aldehyde which was [2] S.K. Kansal, M. Singh, D. Sud, J. Hazard. Mater. 141 (2007) 581–590.
detected in the MS fragmentation pattern. The aldehyde was [3] S. Wu, H. Cao, S. Yin, X. Liu, X. Zhang, J. Phys. Chem. C 113 (2009) 17893–17898.
further oxidized into acid which decarboxylates into CO2. The [4] N. Srivastava, M. Mukhopadhyay, Ind. Eng. Chem. Res. 53 (2014) 13971–13979.
[5] N.L.V. Carreno, H.V. Fajardo, A.P. Maciel, A. Valentini, F.M. Pontes, L.F.D. Probst,
amino group present in the aromatic ring can be substituted by the E.R. Leite, E. Longo, J. Mol. Catal. A: Chem. 207 (2004) 91–96.
OH radical to form phenolic compound (detected in MS [6] F. Li, J. Xu, X. Yu, L. Chen, J. Zhu, Z. Yang, X. Xin, Sens. Actuators B 81 (2002) 165–
fragmentation pattern). The amino group released can form 169.
[7] G. Shang, J. Wu, M. Huang, J. Lin, Z. Lan, Y. Huang, L. Fan, J. Phys. Chem. C 116
ammonium ions which can be further oxidized to nitrate. (2012) 20140–20145.
[8] Y.S. He, J.C. Campbell, R.C. Murphy, M.F. Arendt, J.S. Swinnea, J. Mater. Res. 8
4. Conclusion (1993) 3131–3134.
[9] J. Li, Y. Zhao, N. Wang, L. Guan, Chem. Commun. 47 (2011) 5238–5240.
[10] J. Zhang, L. Gao, J. Solid State Chem. 177 (2004) 1425–1430.
In this article, a facile and green synthesis of SnO2 NPs was [11] K.C. Song, Y. Kang, Mater. Lett. 42 (2002) 283–289.
developed by microwave heating method using different volu- [12] J.-J. Zhu, J.-M. Zhu, X.-H. Liao, J.-L. Fang, M.-G. Zhou, H.-Y. Chen, Mater. Lett. 53
metric ratio of water and glycerol. Glycerol acts as a good (2002) 12–19.
[13] H.-C. Chiu, C.S. Yeh, J. Phys. Chem. C 111 (2007) 7256–7259.
complexing as well as capping agent in the synthesis of SnO2 NPs. [14] J. Pal, M.K. Deb, D.K. Deshmukh, B.K. Sen, Appl. Nanosci. 4 (2014) 61–65.
The TEM, HRTEM images and SAED pattern showed the formation [15] G. Sangami, N. Dharmaraj, Spectrochim. Acta A Mol. Biomol. Spectrosc. 97
of well crystalline, spherical SnO2 NPs with an average particle size (2012) 847–852.
[16] B. Qiu, M. Xing, J. Zhang, J. Am. Chem. Soc. 136 (2014) 5852–5855.
of 8–30 nm. From the TEM images, it was evident that with an [17] Y.K. Mishra, G. Modi, V. Cretu, V. Postica, O. Lupan, T. Reimer, I. Paulowicz, V.
increase in the volumetric ratio of glycerol, the average grain size of Hrkac, W. Benecke, L. Kienle, R. Adelung, ACS Appl. Mater. Interfaces 7 (2015)
SnO2 NPs increases. At higher volume of glycerol, the neighboring 14303–14316.
[18] Q. Dong, S. Yin, C. Guo, X. Wu, N. Kumada, T. Takei, A. Miura, Y. Yonesaki, T. Sato,
particles agglomerate thereby increasing the grain size of SnO2 Appl. Catal. B 147 (2014) 741–747.
nanoparticles. The SAED pattern reflects the tetragonal crystal [19] M. Shanmugam, A. Alsalme, A. Alghamdi, R. Jayavel, ACS Appl. Mater.
structure of SnO2. The XRD pattern also confirms the tetragonal Interfaces 7 (2015) 14905–14911.
[20] F. Gu, S.F. Wang, C.F. Song, M.K. Lü, Y.X. Qi, G.J. Zhou, D. Xu, D.R. Yuan, Chem.
rutile structure of SnO2 NPs. The synthesized SnO2 NPs showed Phys. Lett. 372 (2003) 451–454.
excellent dye-sensitized visible light photocatalytic activity for the [21] G.E. Patil, D.D. Kajale, V.B. Gaikwad, G.H. Jain, Int. Nano Lett. 2 (2012) 17–21.
degradation of methyl violet 6B and methylene blue dyes. [22] A. Bhattacharjee, M. Ahmaruzzaman, T. Sinha, Spectrochim. Acta A Mol.
Biomol. Spectrosc. 136 (2015) 751–760.
[23] F. Gu, S.F. Wang, M.K. Lü, G.J. Zhou, D. Xu, D.R. Yuan, J. Phys. Chem. B 108 (2004)
Acknowledgements 8119–8123.
[24] G. Zhang, M. Liu, J. Mater. Sci. 34 (1999) 3213–3219.
We, the authors, express our heartfelt thanks and gratitude to [25] A. Bhattacharjee, M. Ahmaruzzaman, Mater. Lett. 139 (2015) 418–421.
[26] L. Wang, J. Shang, W. Hao, S. Jiang, S. Huang, T. Wang, Z. Sun, Y. Du, S. Dou, T. Xie,
the Director, NIT Silchar and TEQIP-II for providing lab facilities and D. Wang, J. Wang, Sci. Rep. 4 (2014) 7384-1–7384-8.
scholarship. Our special thanks are extended to NEHU, IIT Bombay [27] H.-F. Zhai, A.-D. Li, J.-Z. Kong, X.-F. Li, J. Zhao, B.-L. Guo, J. Yin, Z.-S. Li, D. Wu, J.
for providing TEM, IR data. Solid State Chem. 202 (2013) 6–14.
[28] G. Liu, T. Wu, J. Zhao, Environ. Sci. Technol. 33 (1999) 2081–2087.

Appendix A. Supplementary data

Supplementary data associated with this article can


be found, in the online version, at http://dx.doi.org/10.1016/j.
jphotochem.2016.03.032.

You might also like