You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/283177031

Photocatalytic Chemoselective Aerobic Oxidation of Thiols to Disulfides


Catalyzed by Combustion Synthesized Bismuth Tungstate Nanoparticles in
Aqueous Media

Article  in  Journal of Cluster Science · October 2015


DOI: 10.1007/s10876-015-0928-0

CITATIONS READS

9 80

4 authors, including:

Yagna PRAKASH Bhoi


National Institute of Technology Rourkela
9 PUBLICATIONS   32 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Photocatalysis View project

Photocatalytic degradation View project

All content following this page was uploaded by Yagna PRAKASH Bhoi on 26 October 2015.

The user has requested enhancement of the downloaded file.


J Clust Sci
DOI 10.1007/s10876-015-0928-0

ORIGINAL PAPER

Photocatalytic Chemoselective Aerobic Oxidation


of Thiols to Disulfides Catalyzed by Combustion
Synthesized Bismuth Tungstate Nanoparticles
in Aqueous Media

Yagna Prakash Bhoi1 • Durga Prasad Rout1 •

B. G. Mishra1

Received: 28 June 2015


 Springer Science+Business Media New York 2015

Abstract Bismuth tungstate (Bi2WO6) nanoparticles were synthesized by solution


combustion synthesis method using different N-containing organic compounds as
fuels (urea, glycine, hexamethylenetetramine (HMTA), malonic acid dihydrazide
(MDH)). The Bi2WO6 nanoparticles have been characterized using XRD, UV–Vis–
DRS, photoluminescence, FESEM, HRTEM and sorptometric techniques. Glycine
as fuel was effective for synthesis of phase pure Bi2WO6 nanoparticles. Employing
HMTA and MDH as fuel lead to mixed phase complex oxide system with Bi14
W2O27 as minor (impurity) phase. The nature of the fuel significantly influences the
particle size and morphology. Formation of spherical, flake like and elongated
polyhedral particles were observed in FESEM study depending upon the nature of
the fuel. TEM study of Bi2WO6 prepared using glycine fuel depicted the presence of
spherical particles with size in the range of 5–8 nm. The Bi2WO6 nanoparticles
were used as an efficient photocatalyst for the visible light driven chemoselective
oxidation of substituted thiophenols to disulfide using air as oxidant. The effect of
various reaction parameters such as catalyst amount, reaction media, time, along
with catalyst recyclability were studied in detail. Structurally diverse diphenyl
disulfide moieties were obtained in high yield and excellent selectivity within a time
span of 6 h of visible light irradiation.

Keywords Bismuth tungstate  Chemoselective oxidation  Thiols  Visible light


catalysis

& B. G. Mishra
brajam@nitrkl.ac.in; bgmcat@rediffmail.com
1
Department of Chemistry, National Institute of Technology, Rourkela, Rourkela,
Odisha 769008, India

123
Y. P. Bhoi et al.

Introduction

Chemoselective oxidation of organic compounds to generate specialty and value


added chemicals is an important field of research with potential application in
pharmaceutical and related fine chemical industries [1, 2]. Among selective
oxidative transformations, the oxidative coupling of thiols to disulfides has attracted
a great deal of attention in recent years because of the potential application of
disulfides as anti-oxidants, pharmaceuticals, pesticides, and rubber vulcanization
reagents [3, 4]. The ‘‘Merox process’’ involving the oxidation of thiols to disulfides
using NaOH is a widely used process for removal of thiols in petroleum refinery
industries. However, the Merox process generates significant quantity of caustic
waste as byproduct leading to environmental and disposal problem [4]. During the
oxidation of thiols often formation of over-oxidized products such as thiosulfinates,
thiosulfonates, and sulfonic acids have been observed due to high reactivity of S–S
bonds [3, 4]. Hence, there are many efforts to develop novel catalytic protocols for
selective oxidation of thiols to disulfides. Particularly, notable are the methods
which involve aerial oxidation route using atmospheric oxygen as oxidant because
of the sustainability of the process. Several catalytic methods have been developed
in recent past towards aerobic oxidation of thiols which include VOCl3, molecular
I2, Br2/SiO2, 2,6-dicarboxypyridinium chlorochromate, Al2O3/KF, activated carbon,
PhSeZnCl, Mn(III) Schiff bases [5–10]. However, most of the catalytic methods
reported so far utilize homogeneous catalytic conditions and supported reagents
which suffer from draw backs such as stoichiometric amounts of catalysts, long
reaction times, high cost of the catalyst, formation of toxic byproducts as well as
tedious workup steps. Hence, there is a need to develop heterogeneous catalytic
method which is less expensive, environmentally benign, selective and uses oxygen
as oxidant. Recently, Noel et al. have developed a photocatalytic method for
synthesis of disulfides from thiols by aerobic oxidation route using the Eosin Y as a
visible light active photocatalyst under homogeneous conditions [3]. The Eosin Y
was found to be highly active and selective for the reaction and show better activity
compared to Ruthenium and Iridium based organometallic complexes. The
heterogeneous photocatalytic protocol developed for oxidation of thiols includes
NiFe2O4 nanoparticles [11] and iron phthalocyanine immobilized on graphene oxide
[12]. Although both method provided good yield and selectivity to disulfide, the
oxidation process using nickel ferrite require hydrogen peroxide as oxidant where as
the iron phthalocyanine are highly costly. With an intention to develop novel and
economic photocatalytic protocol for selective aerobic oxidation of thiol to disulfide
under visible light irradiation, in this work we have used solution combustion
synthesized Bi2WO6 nanoparticles as a visible light active photocatalyst.
The bismuth tungstate (Bi2WO6) is the simplest member of the Aurivillius family
of layered perovskites with general formula [Bi2O2][An-1BnO3n?1]. Bismuth
tungstate has been studied widely as visible-light-induced photocatalyst for water
splitting and decomposition of organic contaminants [13–15]. The hydrothermal
method has been extensively used for synthesis of bismuth tungstate nanostructure

123
Photocatalytic Chemoselective Aerobic Oxidation of Thiols…

using conventional and microwave heating. By varying the solution pH,


hydrothermal treatment temperature, time and solvent, Bi2WO6 materials with
different morphology and size have been obtained [16–18]. By controlling the
synthetic parameters, Bi2WO6 nanosheets, quantum dots, nanodiscs, micro and
nanospheres, hierarchical nanostructures have been synthesized [16, 19–22]. The
visible-light-induced photocatalytic activity of the Bi2WO6 material has been
studied for degradation of organic contaminants such as dyes, chlorinated organic
compounds, and aliphatic and aromatic organic compounds [13–22]. The photo-
generated surface trapped holes and electrons in Bi2WO6 are the prime species
which facilitate photooxidation of the organic substrates [13, 14]. In order to
achieve efficient electron- hole pair separation, Bi2WO6 based hybrid materials with
graphene, Ag2O, TiO2, C3N4 and Fe3O4, have been prepared and studied for their
phtocatalytic activity [18, 23–26]. However, most of these studies focus on the
‘‘nonselective’’ photocatalytic degradation of environmental pollutants. It has been
reported that, the Bi2WO6 possess mild oxidizing power and do not generate the
nonselective hydroxyl radical under visible light illumination [14, 22]. This property
of Bi2WO6 has been exploited recently for efficient and selective oxidation of
glycerol to dihydroxyacetone and benzylalcohol to benzaldehyde [14, 22, 27]. In
order to exploit the photocatalytic potential of Bi2WO6 in selective organic
transformations, in this work we have reported the catalytic application of
combustion synthesized Bi2WO6 nanoparticles for selective aerobic oxidation of
thiols to disulfide under visible light illumination.

Experimental

Combustion Synthesis of Bi2WO6 Nanoparticles

Ammonium tungstate [(NH4)10W12O41] and bismuth carbonate [(BiO)2CO3] were


obtained from Himedia India Ltd. Urea (N2H4CO), hexamethylenetetramine
(C6H12N4) and glycine (C2H5NO2) were procured from Merck India Pvt. Ltd.
Malonicacid dihydrazide was synthesized using the procedure reported elsewhere
[28]. Double distilled water prepared in the laboratory was used in the preparation
process. The Bi2WO6 (BWO) nanoparticles were synthesized by solution combus-
tion synthesis (SCS) using ammonium tungstate and bismuth carbonate as precursor
salt and hexamethylenetetramine (HMTA), malonicacid dihydrazide (MDH), urea
(U) and glycine (G) as fuel. In a typical preparation procedure, required amount of
ammonium tungstate, bismuth carbonate and glycine was dissolved in minimum
amount of double distilled water and the mixture was kept in a furnace, preheated to
400 C. The mixture ignited instantaneously producing a foamy material simulta-
neously releasing a lot of gaseous products. After completion of the combustion
process, the solid material was collected, grinded for 30 min and then calcined at
500 C for 6 h to obtain the Bi2WO6 materials. The fuel to oxidizer ratio (F/O) was
calculated by using the method described by Jain et al. [29]. The F/O ratio was

123
Y. P. Bhoi et al.

varied from 0.5 to 2.0 to study the effect of fuel content on the physicochemical
characteristics of the Bi2WO6 complex oxides. The Bi2WO6 materials prepared
using different fuels are referred to as BWO-F in the subsequent text where F is the
fuel employed (Urea-U, Glyine-G, hexamethylenetetramine-H and malonicacid
dihydrazide-M) for synthesis.

Characterization Techniques

The XRD patterns of the BWO nanomaterials were recorded using a Rigaku,
Ultima-IV Multipurpose X-ray diffraction system using Ni filtered CuKa1
(a = 1.5405 Å) radiation in the range of 20–60 at a scan rate of 2 per minute.
The crystallite size has been calculated from the Fourier line profile analysis of the
broadened XRD profiles following the Warren and Averbach method [30]. The UV–
Vis diffuse reflectance spectra of the BWO samples were recorded using a Jasco
V-650 spectrometer fitted with BaSO4 coated integration sphere. The field emission
scanning electron micrograph (FESEM) was taken using a Nova NanoSEM
microscope model FEI operating at an acceleration voltage of 15 kV. Transmission
electron micrograph (TEM) of the BWO-G material was recorded using JEM-2100
HRTEM equipment using carbon coated copper grids. The photoluminescence (PL)
spectra of the combustion synthesized BWO samples were recorded with a Horiba
Scientific Fluoromax-4 spectrofluorometer with an excitation wavelength of
320 nm. The specific surface area of the combustion synthesized materials was
determined by BET method using N2 adsorption/desorption at 77 K on an
AUTOSORB 1 Quantachorme instrument. Melting points was measured using
LABTRONICS LT-110 model and are uncorrected. 1H NMR spectra were recorded
with Bruker spectrometer at 400 MHz using TMS as internal standard. Reactions
were monitored by thin layer chromatography using 0.2 mm silica gel F-254 plates.

Photocatalytic Oxidation of Thiophenol

The catalytic activity of the BWO nanoparticles was evaluated for the photocat-
alytic selective oxidation of thiophenol to diphenyl disulfide under visible light
irradiation using molecular oxygen from air as oxidant. In a typical catalytic test,
2 mmol of thiophenol dispersed in 05 ml of water was taken in a two neck round
bottom flask fitted with a condenser. 50 mg of Bi2WO6-G catalyst was added to the
reaction mixture and stirred for 15 min under dark condition. Air from a cylinder
was passed through the reaction mixture at a flow rate of 20 ml/min. The reaction
mixture was subsequently irradiated with a 125 W mercury vapor lamp with
k [ 400 nm. Small amount of reaction mixture was periodically collected and the
progress of the reaction was monitored by TLC. The percentage conversion at
different interval of time was analyzed using a Nucon gas chromatograph using a
EC-Wax capillary column and FID detector. After the completion of the reaction,
05 mL of chloroform was added to the reaction mixture and shaken well. The
catalyst particles were filtered and the diphenyl disulfide was recovered from the

123
Photocatalytic Chemoselective Aerobic Oxidation of Thiols…

chloroform solution and recrystallized to afford colourless crystals of the pure


product. The diphenyl disulfide was identified by comparing its physical and
spectral characteristics with the literature reported values.

Physical and Spectra Data of Some Representative Compounds


from Table 2

Diphenyl disulfide (Table 2, Entry 1)

S S

Mp: 72 C, 1H NMR (400 MHz, CDCl3): 7.52 (4H, d, ArH), 7.32 (4H, m, ArH),
7.22 (2H, s, ArH), 13C NMR (400 MHz, CDCl3): 137 (2 C, Ar), 129 (4 C, Ar),
127.46 (4 C, Ar), 127.15 (2 C, Ar),

Fig. 1 X-ray diffraction


patterns of (a) BWO-G, * Bi2WO6
(b) BWO-U, (c) BWO-M, and @ Bi2W2O9
(d) BWO-H materials
# Bi14W2O27

# # #
#
BWO-H d
Intensity (a.u.)

BWO-M c
(008)

@
BWO-U b
*
(131)

(002)

(202)

(133)

* * *
BWO-G a

20 30 40 50 60
2θ (degrees)

123
Y. P. Bhoi et al.

2, 2’-Diamino diphenyl disulfide (Table 2, Entry 6)

H2N

S S

NH2

Mp: 264 C, 1H NMR (400 MHz, CDCl3): 7.1–6.4 (8H, ArH), 4.25 (4H, NH2), 13C
NMR (400 MHz, CDCl3): 148.6, 136.37, 131.67, 118.31, 118.26, 115.27 (6 Ar–C).

Results and Discussion

Characterization of the Combustion Synthesized Bi2WO6 Nanoparticles

The X-ray diffraction patterns of the Bi2WO6 nanomaterials prepared using


different fuels at F/O ratio of 1 are presented in Fig. 1. The BWO-G material exhibit

Fig. 2 XRD patterns of


Bi2WO6 nanoparticles
(131)

synthesized using glycine as fuel


at different F/O ratio. (a) F/
O = 0.5, (b) F/O = 1.0, (c) F/
O = 1.5, and (d) F/O = 2.0
(002)

(202)

(133)

d
F/O = 2.0
Intensity (a.u.)

c
F/O = 1.5

b
F/O = 1.0

F/O = 0.5 a

20 30 40 50 60
2θ (degrees)

123
Photocatalytic Chemoselective Aerobic Oxidation of Thiols…

well-defined and intense diffraction peaks with d values of 3.14, 2.62, 1.92, 1.64 and
1.57 Å (Fig. 1a). These peaks correspond to the orthorhombic phase of Russellite
Bi2WO6 with reflection from (131), (002), (202), (133) and (262) planes,
respectively (JCPDS 79-2381). For BWO-U material, in addition to the character-
istic reflections of orthorhombic Russellite Bi2WO6 phase, a less intense peak is
observed at 2h value of 30.3 corresponding to reflection from (008) plane of
Bi2W2O9 (Fig. 1b). The BWO-U material contains a mixed phase system with
Bi2WO6 as major and Bi2W2O9 as minor phase. When MDH and HMTA were used
as fuel, in addition to the characteristic reflection from the Bi2WO6 phase, less
intense but distinct XRD peaks are observed at 2h value of 27.6, 31.9, 45.9, 54.3,
56.9, 58.7 degrees (Fig. 1c, d). These peaks correspond to the presence of a minor
amount tetragonal Bi14W2O27 phase (JCPDS-39-0061). The XRD study clearly
indicates that the nature of the fuel in the combustion mixture is very crucial for the
generation of phase pure Bi2WO6 material. The phase pure BWO material observed
in case of glycine can be ascribed to the effective complexion capability of glycine
molecule with the constituent ions [31, 32]. In order to study the effect of F/O ratio
on the crystallographic phase formed, the BWO material has been prepared at
different F/O ratio using glycine (Fig. 2). No separate crystalline phase corre-
sponding to either Bi2W2O9 or Bi14W2O27 phase was detected in the XRD study.

0.025 1.0
I BWO-G II BWO-G
BWO-H BWO-H
BWO-M BWO-M
BWO-U BWO-U
0.020 0.8
Fourier Cofficient (AS)

0.015 0.6
PV

0.010 0.4

0.005 0.2

0.000 0.0
0 50 100 150 200 0 50 100 150 200
o o
L(A ) L(A )

Fig. 3 Fourier line profile analysis plots for Bi2WO6 nanomaterials

123
Y. P. Bhoi et al.

However, the XRD peak intensity increases with glycine content in the combustion
mixture possibly due to increase in exothermicity of the combustion reaction
leading to well crystalline materials [31].
The microstructural characteristics of the combustion synthesized BWO mate-
rials is evaluated from the Fourier line shape analysis of the XRD patterns following
the Warren and Averbach method [30] using software BRAEDTH [33]. The
calculated volume-weighted distributions (PV), and size coefficient (As) as function
of Fourier length (L) for different BWO samples is presented in Fig. 3I, II,
respectively. The BWO-U material exhibit a narrow distribution function (PV) in
the range of 3-12 nm whereas the distribution is quite broad for the BWO-M
material (Fig. 3I). The wide distribution function observed for the BWO-M material
indicates that the material is polycrystalline in nature, contain larger grains and the
distribution of the grain size is inhomogeneous. The BWO-G and BWO-H materials
exhibit comparable spread in distribution function (Fig. 3I). The volume weighted
domain size calculated from the Fourier plots are presented in Table 1. All BWO
materials contain nanoparticles with average size less than 12 nm. The crystallite
size increases in the order BWO-U \ BWO-G & BWO-H \ BWO-M. The root
mean square lattice strain calculated from the Fourier analysis for different BWO
materials is presented in Table 1. An inverse correlation is observed between the
crystallite size and the root-mean-square strain.
The optical properties of the combustion synthesized BWO materials have been
studied using UV–Vis–diffuse reflectance and Photoluminescence spectroscopy.
Bi2WO6 is a semiconducting complex oxide which exhibit typical absorption edge
around 450 nm with a band gap value of 2.7 eV [13, 14]. The UV–Vis–DRS spectra
of the combustion synthesized BWO materials are presented in Fig. 4I. All
combustion synthesized BWO samples show strong photoabsorption properties in
UV region which extend up to the visible light region. The absorption edge is in the
range of 440-500 nm depending on the fuel adopted in the synthesis. In case of

Table 1 Physicochemical characteristics and photocatalytic activity of the combustion synthesized


Bi2WO6 nanoparticles
Catalyst Band Surface Crystallite Strain (e2)1/2 Conversionb Selectivity Yieldc
gap (eV) area (m2/g) sizea (nm) (%) (%) (%)

BWO- 2.86 24.2 7.5 ± 0.5 3.67 9 10-3 86.0 100 83.5
G
BWO- 2.82 22.0 9.0 ± 0.6 3.79 9 10-3 76.2 100 73.7
H
BWO- 2.70 20.1 11.5 ± 1.5 3.25 9 10-3 68.4 100 64.2
M
BWO- 2.90 25.8 5.8 ± 1.2 4.67 9 10-3 73.0 100 68.8
U
a
Calculated from the Fourier line profile analysis of the XRD peaks
b
Calculated from the analysis of the reaction mixture using Gas chromatograph after 6 h of reaction time
c
Refers to pure and isolated yield after 6 h of reaction time in aqueous media

123
Photocatalytic Chemoselective Aerobic Oxidation of Thiols…

I II (a) BWO-G BWO-U


II (b)

[FR*(h )]2
F[R] (a.u.)

c
b II (c) BWO-H II (d) BWO-M
a

200 300 400 500 600 2.0 2.5 3.0 3.5 2.0 2.5 3.0 3.5
Wavelength (nm) h (eV)

Fig. 4 UV–Vis–DRS spectra of (a) BWO-G, (b) BWO-U, (c) BWO-H, and (d) BWO-M (panel I) and
plots of [F(R)hm]2 as a function of photon energy for the BWO materials (panel II)

BWO-U and BWO-G material, the absorption edges commence near 450 nm where
as for BWO-H and BWO-M materials the absorption edge is red shifted and occur
between 480 and 500 nm. The BWO materials show significant absorption in the
visible region indicating their suitability for application as visible light active
photocatalyst. The intense absorption bands with a stiff edge in the visible light
region indicate that the optical absorption is due to the band gap transition rather
than the transition from impurity levels [17, 18]. For Bi2WO6 material, the valance
band is formed by hybridization of Bi 6s and O 2p orbitals where as the conduction
band is W 5d in nature [34]. The blue shift in absorption edge observed for BWO-G
and BWO-U material can be ascribed to the quantum confinement effect due to the
presence of ultrafine nanoparticles in these samples (Table 1). The band gap of the
BWO materials calculated from intercept of the plot of photon energy (hm) versus
[F(R)hm]2 (Fig. 4II) are presented in Table 1. The band gap value of 2.70 eV
observed for BWO-M material is typical of Bi2WO6 and is in agreement with earlier
literature reported value [15–18, 34]. For BWO-G, H and U materials the band gap
is in the range of 2.8–2.9 eV. The higher band gap value observed for these
materials is due to presence of small nanoparticles which exhibit quantum
confinement effect. This is further validated from the crystallite size obtained from
Fourier analysis which is in the range of 5–9 nm for these samples (Table 1). The
PL spectra of the combustion synthesized BWO materials in the spectral region of
400–600 nm are presented in Fig. 5. The combustion synthesized BWO materials
exhibit broad blue-green emission in the spectral region of 430–560 nm. The intense
peak at 460 nm in the PL spectra can be assigned to intrinsic luminescence of

123
Y. P. Bhoi et al.

Fig. 5 Photoluminescence
spectra of (a) BWO-G,
(b) BWO-H, (c) BWO-U, and
(d) BWO-M materials

Intensity (a.u.)
b

400 450 500 550 600


Wavelength (nm)

Bi2WO6 originating from the charge transfer transition involving the valance band
(hybrid orbitals of Bi 6s and O 2p) and the conductance band (W 5d) [20, 26, 35]. A
weak blue emission at 435 nm is attributed to the emission due to Bi3? ions from
the 3P1 (6s16p1) excited quantum state to the 1S0 (6s2) ground state [36]. The lower
PL intensity observed for BWO-G material indicate that the electron–hole pair
separation is more effective for this material compared to other materials. The green
emission observed at 560 nm in the PL spectra is ascribed to the presence of
metallic defect sites and the oxygen vacancy in the BWO nanocrystals generated
due to the short time span of combustion reaction [20, 26, 36].
The morphological features of the combustion synthesized BWO materials have
been studied using field emission scanning electron microscopy. The FESEM
images of different BWO materials are presented in Fig. 6. The BWO-G and BWO-
M material contains predominantly spherical shape particles of irregular size present
in an agglomerated state (Fig. 6b, d). The presence of some elongated particles is
also noted in the BWO-G materials. Contrary to this observation, the BWO-H and
BWO-U materials contain flake like or plate like particles. For BWO-G material,
with increase in fuel content, the morphology gradually changes from 1D
agglomerated particle to sheet shaped particles (Fig. 6d–f). The presence of
polyhedral sheets with 5–7 lm width and about 100 nm thickness are clearly
observed for BWO-G materials with F/O ratio 2, suggesting the possibility of
anisotropic growth along the XY plane. The transmission electron micrograph of the

123
Photocatalytic Chemoselective Aerobic Oxidation of Thiols…

Fig. 6 Field emission scanning electron micrograph of a BWO-H, b BWO-M, c BWO-U, d BWO-G (F/
O = 1.0), e BWO-G (F/O = 1.5), and f BWO-G (F/O = 2.0) materials

BWO-G material prepared at F/O ratio of 1 is presented in Fig. 7. Near spherical


shape particles with size in the range of 5–7 nm in a well dispersed state are
observed from the TEM image (Fig. 7 inset). The Fourier line shape analysis data
support the TEM observation. The measured surface area of the BWO materials is
presented in Table 1. The combustion synthesized materials are nonporous in nature

123
Y. P. Bhoi et al.

Fig. 7 Transmission electron


micrograph of BWO-G material

without the presence of any structural periodic porosity. The BWO materials exhibit
surface area in the range of 18–26 m2/g.

Selective Photocatalytic Oxidation of Thiols to Disulfides

The catalytic activity of the BWO nanoparticles has been evaluated for visible light
driven selective oxidation of thiols to disulfides using air as oxidant. Initially, the
oxidation of thiophenol is taken as a model reaction and different BWO materials
are studied for their catalytic activity (Scheme 1). Table 1 show the percentage
conversion and isolated yield of the diphenyl disulfide obtained after 6 h of
irradiation time in aqueous media. The BWO materials have been found to be highly
selective for diphenyl disulfide without the formation of any over oxidized products.
Among the BWO catalysts, the BWO-G material shows highest percentage
conversion with maximum isolated yield of the product. The higher yield observed
in case of BWO-G materials can be ascribed to the efficient separation of the e-–h?
pair as observed from the PL spectra and their subsequent utilization in the
oxidation process. The BWO-G material has been chosen to study the effect of
various reaction parameters on photocatalytic activity. In order to establish that the
reaction is truly photocatalytic in nature, different test reactions are performed under
dark as well as under visible light irradiation in presence and absence of the catalyst.

Bi2WO6, air
SH S S
R nm, 6 h R R

Scheme 1 Selective oxidation of thiols to disulfides catalyzed by Bi2WO6 nanomaterials

123
Photocatalytic Chemoselective Aerobic Oxidation of Thiols…

Fig. 8 Effect of light 100


irradiation time on the isolated BWO-G
yield of the diphenyl disulfide BWO-U
for different BWO catalyst
BWO-M
80 BWO-H

Isolated Yield (%)


60

40

20

0
0 1 2 3 4 5 6 7 8
Irradiation time (h)

The reaction does not take place under dark as well as illuminated condition in
absence of the catalyst. In presence of the catalyst, when the reaction is performed
under dark condition the formation of product is not detected. In order to prove that
the molecular oxygen from the air is the primary oxidant, the reaction is performed
under nitrogen atmosphere. Only trace amount of product was detected in N2
atmosphere (yield 5.2 %) after 6 h of reaction indicating that the molecular oxygen
is the primary oxidant. Figure 8 shows the plot of the isolated yield of diphenyl
disulfide obtained at various irradiation times for different BWO materials. The
yield of the product in general increases with time before attending saturation yield
between 6 and 8 h of reaction time. In case of BWO-G material, the initial rate of
diphenyl disulfide formation is quite higher than other combustion synthesized
materials and equilibrium yield is obtained after 5 h of reaction time. For a reaction
involving 2 mmol of thiophenol, in order to determine optimum catalyst amount,
the catalyst weight in the reaction mixture was varied between 25 and 100 mg. The
yield of diphenyl disulfide increases with catalyst amount up to 50 mg, thereafter,
further increase in catalyst quantity does not significantly impact the yield of the
product (Fig. 9I). The effect of reaction media on the yield of the diphenyl disulfide
for BWO-G catalyst after 6 h of reaction time is presented in Fig. 9II. For nonpolar
solvent like hexane the yield of the product is very less. The photocatalytic reaction
is facilitated in polar solvent. The solubility of thiophenol as well as easy
accessibility of the catalyst surface is the probable reasons for the observed

123
Y. P. Bhoi et al.

100 100
I II

80
Isolated Yield (%)

80
60

40
60

20

0 40
25 50 75 100 H2O CH3CN CH2Cl2 C2H5OH CHCl3 Toluene Hexane

Catalyst weight (mg) Solvent effect

Fig. 9 Effect of (I) catalyst weight and (II) reaction media on the photocatalytic activity of BWO-G
catalyst for the synthesis of diphenyl disulfide after 6 h of irradiation time

enhanced activity in polar solvent. After optimizing the reaction parameters, we


further explored the scope of the developed photocatalytic protocol by using
different substituted thiophenols. Substituted thiophenols containing electron
withdrawing and donating as well as oxidizing functional groups in their structure
reacted smoothly under the optimized condition to afford the corresponding
diphenyl disulfide in high yield and purity (Table 2). The BWO-G catalyst is also
found to be equally active for the conversion of aliphatic thiols to disulfides. The
recyclability of the BWO-G catalyst is studied by conducting the photocatalytic
tests in aqueous media for five consecutive cycles. After completion of each cycle,
the catalyst particles were filtered, dried in hot air oven and then calcined at 400 C
for 2 h to generate the catalyst. No significant different in the yield of the product or
selectivity was observed up to five catalytic cycles indicating the BWO-G material
to be highly recyclable and selective for formation of diphenyl disulfides (Fig. 10I).
The mechanism of photooxidation process over the surface of Bi2WO6 has been
studied earlier [22, 37–39]. Upon absorption of a photon, the conduction band
electrons and valence band holes are generated. The e- and h? could migrate to the
surface to react with the adsorbed reactants, or they could undergo recombination.
The e-–h? recombination is divided into two categories: volume and surface
recombination. The volume recombination is a dominant process in crystalline bulk
semiconducting materials whereas surface recombination occurs due to entrapment
and insufficient driving force for e-–h? pair separation [34, 40]. The average

123
Photocatalytic Chemoselective Aerobic Oxidation of Thiols…

Table 2 BWO-G catalyzed selective oxidation of thiols to disulfide after 6 h of reaction time

Sl Substrate product Selectivity Yield

No. (%) (%)

1 SH
100 83.5
S S

2 Cl
100 87.8
SH Cl S S Cl

3 Br SH
100 82.4
Br S S Br

4 H3C SH
100 81.5
H3C S S CH3

5 H3CO SH H3CO S S OCH3


100 85.0

6 SH
H2N
100 73.5
S S
NH2

NH2

7 SH S 100 87.0
S

8 100 83.6
SH S S

diffusion time for the photogenerated charge carrier from the bulk to the surface is
governed by the relation s = r2/p2D where r is the grain radius and D is the
diffusion coefficient of the carrier [17]. The small grain radius of the BWO quantum
dots in the present study is quite favorable for diffusion of the charge carriers from
bulk to the surface for utilization in photocatalytic process. The surface electron
from the conduction band of Bi2WO6 can react with O2 present in the system,
reducing it to superoxide radical anion O 2 . The presence of superoxide radical
anions in the aqueous aerated solution of Bi2WO6 has been detected earlier by ESR
spectral analysis using spin trapping agent [22]. The formation of hydroxyl radical
(OH) as a possible oxidizing species has been ruled out in case of Bi2WO6 in many
studies [22, 37, 38]. Using the band gap value, the valence band and conduction
band levels of the Bi2WO6-G material are calculated to be ?0.43 and ?3.29 eV,
which is consistent with the literature reported values [37]. In the phocatalytic
oxidation of thiophenol since the photogenerated holes (h?), electrons (e-),
superoxide radical anions ðO 2 Þ and the hydroxyl radicals (OH) are the possible

123
Y. P. Bhoi et al.

Fig. 10 I Recyclability of the BWO-G catalyst. II Isolated yield of diphenyl disulfide in presence of
various radical scavengers after 6 h in aqueous media (the photocatalytic test was conducted under
identical reaction condition by adding 2 mmol of the scavenger to the reaction mixture before
illumination)

oxidizing species, the corresponding scavengers are used to investigate their role in
the photooxidation process (Fig. 10II) [22, 37]. The yield of disulfide decreases
drastically in presence ammonium oxalate which is a hole scavengers. Similarly,

Fig. 11 Probable mechanism for the selective oxidation of thiol to disulfide catalyzed by BWO-G
material under visible light irradiation

123
Photocatalytic Chemoselective Aerobic Oxidation of Thiols…

considerable decrease in yield is observed in presence of benzoquinone which is a


scavenger for O
2 radical anion. The radical scavenger experiments confirmed that
the superoxide radical anions and the photogenerated holes are the predominant
active species responsible for the oxidation of thiophenol to disulfide. Based on
these results a possible mechanism for the formation of disulfide is proposed
(Fig. 11).

Conclusions

In this work, we have used solution combustion synthesis to prepare Bi2WO6


nanomaterials and studied their photocatalytic application for chemoselective
aerobic oxidation of thiols to disulfides. Phase pure Bi2WO6 nanoparticles were
obtained using glycine as fuel whereas formation of Bi2W2O9 and Bi14W2O27 as
impurity phases was observed for urea, MDH and hexamethylenetetraamine.
Bi2WO6 nanoparticles with spherical, flake like and elongated polyhedral shape
were obtained by varying the fuel and its content in the combustion mixture. The
BWO material synthesized using glycine contain uniformly distributed nanoparti-
cles with size in the range of 5–7 nm, exhibit higher band gap due to quantum
confinement and better e-–h? pair separation compared to other samples. The
combustion synthesized BWO-G material was used as an efficient photocatalyst for
rapid, convenient and environmentally benign selective oxidation of thiols to
disulfides in aqueous media under visible light irradiation. The BWO-G catalyst was
found to be highly active and 100 % selective to the formation of disulfides for a
variety of substrates containing electron withdrawing and donating functional
group. The photocatalytic protocol developed in this work is advantageous in terms
of its mild nature, preclusion of toxic solvent and chemicals, recyclable and
stable photocatalyst, good yield and excellent selectivity for disulfide formation.

Acknowledgments YPB would like to thank NIT, Rourkela for a research fellowship.

References

1. Z. Guo, B. Liu, Q. Zhang, W. Deng, Y. Wang, and Y. Yang (2014). Chem. Soc. Rev. 43, 3480.
2. F. Cavani (2010). Catal. Today 157, 8.
3. A. Talla, B. Driessen, N. J. W. Straathof, L.-G. Milroy, L. Brunsveld, V. Hessel, and T. Noel (2015).
Adv. Synth. Catal. doi:10.1002/adsc.201401010.
4. S. Dharmarathna, C. K. Kingondu, L. Pahalagedara, C.-H. Kuo, Y. Zhang, and S. L. Suib (2014).
Appl. Catal. B 147, 124.
5. H. Golchoubian and F. Hosseinpoor (2007). Catal. Commun. 8, 697.
6. M. Kirihara, K. Okubo, T. Uchiyama, Y. Kato, Y. Ochiai, S. Matsushita, A. Hatano, and K.
Kanamori (2004). Chem. Pharm. Bull. 52, 625.
7. M. Kirihara, Y. Asai, S. Ogawa, T. Noguchi, A. Hatano, and Y. Hirai (2007). Synthesis 2007, 3286.
8. M. Chakraborty, P. C. Mandal, and S. Mukhopadhyay (2012). Polyhedron 45, 213.
9. M. Hayashi, K. Okunaga, S. Nishida, K. Kawamura, and K. Eda (2010). Tetrahedron Lett. 51, 6734.
10. C. Tidei, M. Piroddi, F. Galli, and C. Santi (2012). Tetrahedron Lett. 53, 232.
11. A. M. Kulkarni, U. V. Desai, K. S. Pandit, M. A. Kulkarni, and P. P. Wadgaonkar (2014). RSC Adv.
4, 36702.
12. P. Kumar, G. Singh, D. Tripathi, and S. L. Jain (2014). RSC Adv. 4, 50331.

123
Y. P. Bhoi et al.

13. H. Tong, S. Ouyang, Y. Bi, N. Umezawa, M. Oshikiri, and J. Ye (2012). Adv. Mater. 24, 229.
14. N. Zhang, R. Ciriminna, M. Pagliaro, and Y.-J. Xu (2014). Chem. Soc. Rev. 43, 5276.
15. L. Zhou, W. Wang, and L. Zhang (2007). J. Mol. Catal. A 268, 195.
16. W.-L. W. Lee, S.-T. Huang, J.-L. Chang, J.-Y. Chen, M.-C. Cheng, and C.-C. Chen (2012). J. Mol.
Catal. A 80, 361–362.
17. S. Sun, W. Wang, D. Jiang, Y. Sun, and Y. Xie (2013). ChemSusChem 6, 1873.
18. S. Murcia-López, M. C. Hidalgo, and J. A. Navı́o (2012). Appl. Catal. A 423, 34.
19. Y. Tian, et al. (2011). J. Alloys Compd. 509, 724.
20. D. Kim, S. Kim, and M. Kang (2009). Bull. Korean Chem. Soc 30, 630.
21. F. Amano, K. Nogami, M. Tanaka, and B. Ohtani (2010). Langmuir 26, 7174.
22. Y. Zhang, N. Zhang, Z. Tang, and Y. Xu (2013). Chem. Sci. 4, 1820.
23. X. Xu, X. Shen, G. Zhu, L. Jing, X. Liu, and K. Chen (2012). Chem. Eng. J. 200–202, 521.
24. L. Chen, H. Hua, Q. Yang, and C. Hu (2015). Appl. Surf. Sci. 327, 62.
25. Y. Wang, X. Bai, C. Pan, J. He, and Y. Zhu (2012). J. Mater. Chem. 22, 11568.
26. Z. Sun, J. Guo, S. Zhu, L. Mao, J. Ma, and D. Zhang (2014). Nanoscale 6, 2186.
27. Y. Zhang and Y.-J. Xu (2014). RSC Adv. 4, 2904.
28. K. C. Patil, M. S. Hegde, T. Rattan, and S. T. Aruna Chemistry of Nanocrystalline Oxide Materials,
Combustion Synthesis, Properties and Applications (World Scientific Publishing Co. Pte. Ltd,
London, 2008).
29. S. R. Jain, K. C. Adiga, and V. R. Verneker (1981). Combust. Flame 40, 71.
30. B. E. Warren and B. L. Averbach (1950). J. Appl. Phys. 21, 595.
31. S. Samantaray and B. G. Mishra (2011). J. Mol. Catal. A 339, 92.
32. S. Samantaray, B. G. Mishra, D. K. Pradhan, and G. Hota (2011). Ceram. Int. 37, 3101.
33. D. Balzar and H. Ledbetter (1993). J. Appl. Crystallogr. 26, 97.
34. H. Fu, L. Zhang, W. Yao, and Y. Zhu (2006). Appl. Catal. B 66, 100.
35. F. Duan, Y. Zheng, and M. Chen (2011). Appl. Surf. Sci. 257, 1972.
36. Q. Xiao, J. Zhang, C. Xiao, and X. Tan (2008). Catal. Commun. 9, 1247.
37. H. Li, W. Hou, X. Tao, and N. Du (2015). Appl. Catal. B 172–173, 27.
38. J. Zhai, H. Yu, H. Li, L. Sun, K. Zhang, and H. Yang (2015). Appl. Surf. Sci. 344, 101.
39. Z. Cui, D. Zeng, T. Tang, J. Liu, and C. Xie (2010). Catal. Commun. 11, 1054.
40. H. Fu, C. Pan, W. Yao, and Y. Zhu (2005). J. Phys. Chem. B 109, 22432.

123
View publication stats

You might also like