ArticlePDF Available

Visible-light-driven photocatalytic and photoelectrochemical properties of porous SnSx(x = 1,2) architectures

Authors:

Abstract and Figures

By using a facile and template-free polyol refluxing process, we reported the successful synthesis of porous SnS and SnS2 architectures on a large scale. The as-synthesized samples were characterized by using XRD, SEM, TEM, UV-vis DRS, Raman and N2 adsorption–desorption analyses. Studies revealed that the as-synthesized SnS and SnS2 products mainly consist of porous flower-like microstructures with reasonable BET surface areas of 66 m2 g−1 and 33 m2 g−1, respectively. Photocatalytic properties of trace amounts of samples were investigated by photodegradation of MB and RhB under visible light irradiation. The photoelectrochemical properties of both samples were also studied by configuring the samples as photoelectrochemical (PEC) cells, exhibiting excellent photosensitivity and response with greatly enhanced Ion/off as high as 1.4 × 103, three orders of magnitude higher than previous work. The results indicate the potential applications of the SnSx nanostructures in visible-light-driven photocatalysts, high response photodetectors and other optoelectronic nanodevices.
Content may be subject to copyright.
Visible-light-driven photocatalytic and photoelectrochemical properties of
porous SnS
x
(x ¼ 1,2) architectures
Junfeng Chao,
a
Zhong Xie,
a
XianBao Duan,
c
Yuan Dong,
a
Zhuoran Wang,
a
Jing Xu,
a
Bo Liang,
a
Bin Shan,
c
Jinhua Ye,
b
Di Chen
*
a
and Guozhen Shen
*
a
Received 25th November 2011, Accepted 16th February 2012
DOI: 10.1039/c2ce06586j
By using a facile and template-free polyol refluxing process, we reported the successful synthesis of
porous SnS and SnS
2
architectures on a large scale. The as-synthesized samples were characterized by
using XRD, SEM, TEM, UV-vis DRS, Raman and N
2
adsorption–desorption analyses. Studies
revealed that the as-synthesized SnS and SnS
2
products mainly consist of porous flower-like
microstructures with reasonable BET surface areas of 66 m
2
g
1
and 33 m
2
g
1
, respectively.
Photocatalytic properties of trace amounts of samples were investigated by photodegradation of MB
and RhB under visible light irradiation. The photoelectrochemical properties of both samples were also
studied by configuring the samples as photoelectrochemical (PEC) cells, exhibiting excellent
photosensitivity and response with greatly enhanced I
on/off
as high as 1.4 10
3
, three orders of
magnitude higher than previous work. The results indicate the potential applications of the SnS
x
nanostructures in visible-light-driven photocatalysts, high response photodetectors and other
optoelectronic nanodevices.
Introduction
Porous semiconductor nanomaterials with a large surface area
have been widely investigated due to their important applications
in many fields such as photocatalysts, optoelectric devices,
semiconductor sensors, solar cells and electrode material for
batteries and so on.
1–7
Up to now, many synthetic methods have
been developed to prepare nanomaterials with porous micro-
structures. For example, 3D mesoporous carbon structures were
successfully fabricated from the soft chemical route using
commercially available colloidal silicas as templates.
8
Nanosized
flower-like nickel hydroxide, synthesized by a microwave-assis-
ted hydrothermal method, shows good electrochemical activity
for the electrochemical reduction of O
2
at room temperature.
9
Hierarchical WO
3
architectures including dendrites, spheres and
dumbbells were synthesized from different tungstates with
similar morphologies and show enhanced photocatalytic activi-
ties for the degradation of some organic dyes compared to the
commercial materials.
10
Thus, developing simple and effective
methods to fabricate semiconductors with porous structures is
important for future device applications.
The photocatalytic technique for the degradation of organic
contaminants in solution or in air has attracted extensive interest
to remedy the effects of environmental pollution because of its
simple and exhaustive decomposition process.
10–14
Among all
kinds of semiconductors, TiO
2
is the most investigated one in
view of its low cost, good stability, nontoxicity and so on.
However, with a wide band gap of 3.2 eV, TiO
2
is only active in
the ultraviolet light region and is not responsive to visible light.
Although some dyes could be degraded over TiO
2
by self-
photosensitization under visible light irradiation, the efficiencies
are relatively lower than that under UV light irradiation.
15–18
It is
well known that only about 4% of the solar spectra falls in the
UV region, thus it is appealing to develop efficient visible light
sensitive photocatalysts in view of the better utilization of solar
energy. Until now, many researchers have attempted to synthe-
size new catalysts with narrow band gaps in order to realize
visible-light-driven catalytic reactions.
19–25
Metal sulfides have
been reported to be promising photocatalysts in the visible
region. For instance, ultrathin beta-In
2
S
3
nanobelts prepared by
Qian’s group showed enhanced photocatalytic activity for the
decomposition of methyl orange under visible light irradiation.
26
From a two-step solution route, K. Domen and his coauthors
have synthesized nanostructured CdS with a high photoactivity
for hydrogen evolution when irradiated by visible light.
27
With narrow band gaps, tin sulfides, one of the IV–VI group
semiconductors, include a variety of phases such as SnS, SnS
2
,
Sn
2
S
3
,Sn
3
S
4
and Sn
4
S
5
. Among them, SnS and SnS
2
, with
a
Wuhan National Laboratory for Optoelectronics (WNLO) and College of
Optoelectronic Science and Engineering, Huazhong University of Science
and Technology (HUST), Wuhan, 430074, China. E-mail: gzshen@mail.
hust.edu.cn; dichen@mail.hust.edu.cn; Fax: +86-27-87792225
b
International Center for Materials Nanoarchitectonics (MANA)
National Institute for Materials Science (NIMS), 1-2-1 Sengen,
Tsukuba, Ibaraki, 305-0047, Japan. E-mail: Jjinhua.YE@nims.go.jp;
Fax: (+)81-29-859-2301
c
School of Materials Science and Engineering, Huazhong University of
Science and Technology (HUST), Wuhan, 430074, China
This journal is ª The Royal Society of Chemistry 2012 CrystEngComm, 2012, 14, 3163–3168 | 3163
Dynamic Article Links
C
<
CrystEngComm
Cite this: CrystEngComm, 2012, 14, 3163
www.rsc.org/crystengcomm
PAPER
Downloaded by Huazhong University of Science & Technology on 22 April 2012
Published on 28 February 2012 on http://pubs.rsc.org | doi:10.1039/C2CE06586J
View Online
/ Journal Homepage
/ Table of Contents for this issue
indirect band gaps of 1.3 eV and 2.2 eV, have received more
attention due to their possible applications in optoelectronics,
solar cells, lithium-ion batteries, near-infrared detectors, photo-
catalysts and photoluminescence.
28–35
In this paper, we reported
the synthesis of porous SnS and SnS
2
architectures from a simple
and template-free polyol refluxing route. Photocatalytic prop-
erties of both samples were investigated by photodegradation of
MB and RhB under visible light irradiation. The electronic
structures of the porous SnS and SnS
2
samples were also inves-
tigated by band structure calculations. The photoelectrochemical
properties of both samples were also studied by configuring the
samples as photoelectrochemical (PEC) cells, exhibiting
a reasonable photosensitivity and response with greatly
enhanced I
on/off
ratios.
Experimental
Materials
All reagents, such as SnCl
2
$2H
2
O, SnCl
4
$5H
2
O, thiourea (Tu)
and ethylene glycol were analytical grade and used directly
without further purification.
Synthesis of SnS and SnS
2
architectures
Porous SnS and SnS
2
architectures were successfully prepared
from the simple polygol refluxing process, respectively. In
a typical procedure for the SnS sample, appropriate amounts of
SnCl
2
and Tu were first added into 50 mL ethylene glycol in
a round bottom flask. After stirring for 5 min, the reaction
system was heated to 175
C and kept at that temperature for 1 h
and then cooled to room temperature naturally. Finally, black
precipitates were collected, washed with absolute ethanol and
distilled water and then dried at 60
C. The synthetic process for
porous SnS
2
nanostructures was similar to that of SnS products
except that SnCl
4
was used as the raw material instead of SnCl
2
and the refluxing system was kept at 160
C for 1.5 h.
Material characterization
The crystal structures of the as-prepared samples were confirmed
by the X-ray diffraction pattern (X’Pert PRO, PANalytical B.V.,
the Netherlands) with radiation of a Cu target (Ka, l ¼ 0.15406
nm). The morphologies and sizes of the samples were charac-
terized by scanning electron microscopy (SEM, JSM-6700F) and
transmission electron microscopy (TEM, JEM-2010). UV-vis
diffuse reflectance spectra were recorded on a UV/vis spec-
trometer (UV-2500, Shimadzu) and were converted from reflec-
tion to absorbance by the standard Kubelka–Munk method. The
surface areas were measured on a surface area analyzer (Micro-
meritics, Shimadzu) by nitrogen absorption at 77 K using the
Brunauer–Emmett–Teller (BET) method.
Results and discussion
SnS and SnS
2
are two kinds of important layered structured
semiconductor materials and the corresponding ball-and-stick
models are shown in Fig. 1. Typically, a SnS crystal with
orthorhombic structure consists of double layers perpendicular
to the c-axis in which Sn and S atoms are tightly bound
(Fig. 1, left). While the SnS
2
with hexagonal cadmium iodide
(CdI
2
) structure is composed of sheets of tin atoms sandwiched
between two close-packed sheets of sulfur atoms (Fig. 1, right).
Based on the layered structure properties, research on SnS and
SnS
2
has attracted more and more attention.
XRD was used to study the phase purity of the as-synthesized
SnS and SnS
2
products as shown in Fig. 2. All of the peaks in the
XRD pattern in Fig. 2a can be readily indexed to the pure
orthorhombic SnS phase with space group P6nm (JCPDS card
No: 39-354). The pattern in Fig. 2b can be indexed to the
hexagonal SnS
2
phase with space group P-3m1 (JCPDS card No:
23-677). Clearly, no characteristic peaks from other crystalline
impurities were detected in these XRD patterns, indicating the
formation of pure products.
Raman spectroscopy is an appropriate technique to probe the
detailed structure of materials because the bonding states in the
coordination polyhedra of a material can be deduced directly
from the Raman vibrational spectrum. In orthorhombic SnS, 24
vibrational modes are represented by the following irreducible
representations at the center of Brillouin zone as:
Fig. 1 The ball-and-stick models of SnS (left) and SnS
2
(right) crystals.
Fig. 2 XRD patterns of the as-synthesized porous SnS and SnS
2
architectures.
3164 | CrystEngComm, 2012, 14, 3163–3168 This journal is ª The Royal Society of Chemistry 2012
Downloaded by Huazhong University of Science & Technology on 22 April 2012
Published on 28 February 2012 on http://pubs.rsc.org | doi:10.1039/C2CE06586J
View Online
G ¼ 4A
g
+2B
1g
+4B
2g
+2B
3g
+2A
u
+4B
1u
+2B
2u
+4B
3u
Among them, there are 21 optical phonons, of which 12 are
Raman active modes (4A
g
,2B
1g
,4B
2g
and 2B
3g
), seven are
infrared active modes (3B
1u
,1B
2u
and 3B
3u
) and two are inactive
(2A
u
).
36
Fig. 3 shows the Raman spectra of the as-synthesized
SnS and SnS
2
products, respectively. The Raman modes shown
in Fig. 3a are observed at 69, 95, 165, 192 and 219 cm
1
,
respectively. Among the peaks, the peaks at 69 cm
1
,95cm
1
,
165 cm
1
corresponding to the B
1g
or B
2g
mode, A
g
mode and B
3g
mode, respectively, are in good agreement with the literature.
37
The peaks at 192 cm
1
and 219 cm
1
can be assigned to the A
g
mode. The Raman spectrum of the as-synthesized SnS
2
is shown
in Fig. 3b. A strong Raman peak located at about 313 cm
1
is
observed, which can be assigned to the A
1g
mode according to
the group theory analysis given by Lucovsky et al.
38
According to
the Raman results, it can be further confirmed that pure phase
SnS and SnS
2
products were prepared from the present process.
The morphology and microstructure of the as-prepared
sulfides from the simple polyol refluxing process were charac-
terized by SEM. Fig. 4a shows the low-magnification SEM image
of SnS products, which demonstrates that porous SnS samples
on a large scale were prepared and composed of many nano-
platelets with uniform size. Further observation (Fig. 4b) depicts
that large numbers of platelets interconnect with each other with
a thickness of several tens of nanometres to form a porous
architecture structure. In the present case, a similar porous
architecture structure was also observed in SnS
2
products as
shown in Fig. 4c, when the raw material of SnCl
4
was used in this
refluxing solution at 165
C. The high-magnification FESEM
image of the products in Fig. 4d indicates that the surfaces of the
architectures are highly porous in structure, suggesting that the
samples possess possibly large surface areas. The measurements
show that the BET surface areas of porous SnS and SnS
2
architectures are 66 m
2
g
1
and 33 m
2
g
1
, respectively, which are
favourable for potential applications in waste water treatment
such as photocatalysts.
Further insight into the microstructural details of the porous
SnS and SnS
2
architectures were gained by using TEM and high-
resolution TEM (HRTEM). Fig. 5a shows a TEM image taken
from the porous SnS architectures. It is interesting that the SnS
products are composed of interconnected nanoplatelets, in good
accordance with the SEM observations. The representative
HRTEM image (Fig. 5b) taken from the edge of the micro-
structures exhibits the highly crystalline nature of the SnS
nanoplatelets. The lattice fringe in the image is clearly visible
with a spacing of 0.40 nm, corresponding to the (110) plane of
orthorhombic SnS. Fig. 5c displays the TEM image of the SnS
2
product. Nanoplatelets with a thickness of about several nano-
metres were found to be connected to each other to build the 3D
porous architectures. This is also in good agreement with the
SEM results. The distance between the lattice fringes of the
architectures can be indexed to the (100) planes of hexagonal
SnS
2
.
The optical properties of the products were measured and the
corresponding UV-vis absorption spectra are depicted in Fig. 6a.
A black-colored SnS sample (inset in Fig. 6a) exhibits broad and
strong absorption in the range from 200 to 800 nm. In contrast,
yellow-colored SnS
2
(inset in Fig. 6a) only absorbs light in the
region below 600 nm and it exhibits less intensive absorption
Fig. 3 Raman spectra of the as-synthesized porous SnS (a) and SnS
2
(b)
architectures.
Fig. 4 SEM images of the as-synthesized porous SnS (a, b) and SnS
2
(c, d) architectures from the simply polyol refluxing process.
Fig. 5 TEM and HRTEM images of the porous SnS (a, b) and SnS
2
(c, d) architectures from the simple polyol refluxing process.
This journal is ª The Royal Society of Chemistry 2012 CrystEngComm, 2012, 14, 3163–3168 | 3165
Downloaded by Huazhong University of Science & Technology on 22 April 2012
Published on 28 February 2012 on http://pubs.rsc.org | doi:10.1039/C2CE06586J
View Online
compared to the SnS sample. The optical band gaps of the two
samples could be evaluated from the following equation:
39–41
ahn ¼ A(hn Eg)
n/2
Here a, n, A, Eg and n are the absorption coefficient, incident
light frequency, constant, optical band gap, and an integer
(normally equal to 1, 2, 4 and 6), respectively. The n values for
the two observed optical transitions for the two samples should
be indirect. The a and n values at the steep edges of the
absorption spectra were used to construct the plots of (ahn)
2/n
against photon energy. As seen in Fig. 6b, the curves for the two
samples show a linear region with n ¼ 4 for the indirect band gap.
The optical band gaps of the samples could be determined by the
intersections of the extrapolated linear portions of the plots with
the energy axis. These values were calculated to be about 1.35 eV
for SnS and 2.18 eV for SnS
2
, which is in good agreement with
the previous reports.
42,32
To investigate the photocurrent properties of the as-synthe-
sized SnS and SnS
2
products, photoelectrochemical (PEC) cells
were prepared by using the as-synthesized products as the cor-
responding photoelectrode, respectively. The photoelectrodes of
the photoelectrochemical devices were prepared by the well-
known ‘doctor-blading’ method widely used for dye-sensitized
solar cell (DSSC) fabrication. The corresponding SnS or SnS
2
product was first deposited onto a FTO substrate. The coun-
terelectrode was lightly platinized by coating with a drop of an
isopropanol solution of H
2
PtCl
6
, followed by firing at 400
C for
20 min. A solar simulator of AM 1.5G (100 mW cm
2
) was used
as the illumination source. In principle, when the external circuit
was closed, the illuminated PEC devices should generate
a photocurrent. Fig. 7 demonstrates the typical time-dependent
current change curves of the PEC devices made of SnS and SnS
2
architectures, respectively. From Fig. 7a, we can see that, upon
illumination, the current density rapidly increased from 0.052 mA
cm
2
to a stable value of 70.5 mAcm
2
. The significant
enhancement of the current density induced by electron-hole
generation is around 1.4 10
3
. Similar results were also obtained
for the SnS
2
architectures, as can be seen in Fig. 7b. Both values
are about three orders of magnitude higher than previous work
on the same materials.
43
Besides, there are current transients in
both the light-on state and the light-off state, which can be
attributed to a typical sign of surface recombination processes
according to previous reports.
44–46
It was found that the photocurrent response can be repro-
ducibly switched from the ‘ON’ state to the ‘OFF’ state by
periodically turning the solar simulator on and off. When
switched from the dark to light condition, the photocurrent
response of the SnS-based PEC device changes abruptly and then
stably with time (in Fig. 7a), while the response curve of the SnS
2
-
based PEC device has a sharp-edged peak at first and then drops
slowly (in Fig. 7b). The response time and decay time of both
devices were around 0.5 s, indicating rapid photoresponse
characteristics. On the other hand, the relatively low dark current
also guarantees the outstanding optical switch performance for
the device. When the solar simulator illuminates, photons with
energy higher than the SnS
x
band gap are absorbed by the SnS
x
nanomaterials and a lot of electron–hole pairs are generated. The
holes accept electrons from the I
3
/I
redox electrolyte, and the
electrons move from the porous SnS
x
surface to the anode
contact. Along with the photocurrent flows, electrons pass
through the external closed circuit. Owing to the porous surface
area, the practical contact area of the device is much larger than
the flat light area, so the number of electron–hole pairs created
can be tremendously larger than that of the geometric device
area.
The photocatalytic activities of the as-synthesized porous SnS
and SnS
2
architectures were further evaluated for the degrada-
tion of some organic dyes such as methylene blue (MB) and
rhodamine B (RhB) under visible light irradiation at room
temperature. Since both samples possess large BET surface areas,
trace amounts of the SnS (10 mg) and the SnS
2
(20 mg) samples
were used in the present cases, respectively. First, an appropriate
Fig. 6 (a) UV-vis absorption spectra and (b) (ahn)
1/2
vs. hn plots for the
porous SnS and SnS
2
architectures.
Fig. 7 JT curves of the porous SnS (a) and SnS
2
(b) architectures under
one sun illumination (AM 1.5G).
3166 | CrystEngComm, 2012, 14, 3163–3168 This journal is ª The Royal Society of Chemistry 2012
Downloaded by Huazhong University of Science & Technology on 22 April 2012
Published on 28 February 2012 on http://pubs.rsc.org | doi:10.1039/C2CE06586J
View Online
catalyst was suspended into 100 mL MB or RhB aqueous solu-
tion in a pyrex reactor, respectively (the dye concentration was
8mgL
1
in both cases). The suspension was stirred in the dark
for a certain time to reach an adsorption–desorption equilibrium,
then the light was turned on. A 500 W Xe-lamp equipped with
a cutoff filter (l > 420 nm) and a water filter was used as the light
source. At given irradiation time intervals, appropriate reaction
suspension was collected and filtrated. The absorption spectra of
the filtrates were measured on a Shimadzu UV-2550 UV-vis
spectrometer and are shown in Fig. 8. Clearly, during the dark,
porous SnS architectures with a larger BET surface area of 66 m
2
g
1
showed higher adsorption abilities to MB and RhB than the
SnS
2
product. As shown in Fig. 8(a–d), porous SnS architectures
can adsorb about 75% of MB within 30 min. The dye concen-
tration adsorbed onto the surface of the catalyst further increases
with the prolonging of time (Fig. 8b). After 2 h, an adsorption–
desorption equilibrium in the solution was reached, since the
intensities of the main peaks located in the visible region and in
the UV region are similar to those for 2.5 h. When the light was
turned on, the main peaks decreased continuously with increased
irradiation time, indicating that the MB solution was decom-
posed completely in the present system. Fig. 8(c,d) shows the
great adsorption and degradation activities of the SnS architec-
tures when trace catalyst was dispersed into the RhB solution.
Due to the good adsorption and degradation effect of the SnS
architectures exhibited in the organic dye solution, we can
conclude that the obtained sample might be used in the waste-
water treatment in industry.
In addition, the photocatalytic abilities of the porous SnS
2
architectures were also investigated and Fig. 8(e,f) show the
photodegradation process of MB and RhB solutions in
the presence of SnS
2
catalysts, respectively. It was obvious that
the porous SnS
2
products with less BET surface area show
weaker adsorption and photocatalytic abilities compared to the
porous SnS architectures. However, in our experiments, we
found that trace amounts (10 mg or 20 mg) of the two samples
can accomplish good adsorption and degradation to the MB and
RhB when irradiated by visible light, which is better or compa-
rable to previous reports.
33,47–50
The photocatalytic performances of sulfide photocatalysts
strongly depend on their electronic structures. In the present
work, the electronic structures of SnS and SnS
2
were further
investigated by plane-wave DFT calculations, respectively, and
Fig. 9 shows the energy structures and density of states (DOS)
calculated. As shown in Fig. 8a left, the conduction band
minimum (CBM) of SnS crystal is located in the area between the
G point and the Z point while the valence band maximum (VBM)
is located at the Y point of Brillouin zone, indicating that SnS is
indeed an indirect semiconductor. Similarly, the calculated energy
band of SnS
2
is shown in Fig. 9b left, it can be clearly seen that the
conduction band minimum (CBM) of SnS
2
crystal is located at the
M point of the Brillouin zone and the valence band maximum
(VBM) is located in the region between the K point and the G
point, indicating that SnS
2
is also an indirect semiconductor. For
an indirect semiconductor, the lifetime of the photoexcited elec-
tron–hole pair is comparably longer than that in the direct
samples, which is beneficial for the photocatalytic reactions.
The band gap of a given semiconductor is determined mainly by
the potential of the valence band top and the potential of the
Fig. 8 (a,b) UV-vis spectral changes of MB aqueous solution and the
magnification of the lit portion,(c,d) UV-vis spectral changes of RhB
aqueous solution and the magnification in the presence of SnS architec-
tures and (e) MB aqueous solution and (f) RhB aqueous solution in the
presence of SnS
2
architectures under visible light irradiation. Reaction
conditions: 100 mL, 8 mg L
1
, 500 W Xe-lamp, l > 420 nm.
Fig. 9 The calculated (left) energy bands and (right) density of states
(DOS) of as-synthesized porous (a) SnS and (b) SnS
2
samples.
This journal is ª The Royal Society of Chemistry 2012 CrystEngComm, 2012, 14, 3163–3168 | 3167
Downloaded by Huazhong University of Science & Technology on 22 April 2012
Published on 28 February 2012 on http://pubs.rsc.org | doi:10.1039/C2CE06586J
View Online
conduction band bottom. The total DOS displays (Fig. 9a and 9b
right) that the bands of SnS and SnS
2
are classified into two parts.
The bottom of the conduction band is mainly constituted by the
Sn 5p orbital, while the conduction band top is composed mainly
of the S 3p orbital. The band structure indicates that charge
transfer upon photoexcitation occurs from the S 3p orbital to the
empty Sn 5p orbital. The calculated band gaps of SnS and SnS
2
are
1.5 and 2.14 eV, respectively, which are slightly larger than the
experimental results based on the UV-vis absorption spectra.
Conclusions
In summary, we reported the synthesis of porous SnS and SnS
2
architectures from the simple polyol refluxing process on a large
scale. SEM investigations revealed that both the SnS and the
SnS
2
products are composed of many nanoplatelets inter-
connecting with each other to form the porous architecture. The
photoelectrochemical characteristics recorded under AM 1.5
illumination exhibited the increased photoresponse properties of
SnS and SnS
2
based devices. Greatly enhanced I
on/off
with a value
of about 1.4 10
3
, three orders of magnitude higher than
previous work, was obtained. Furthermore, during the photo-
catalytic degradation process, black SnS architectures with
a larger BET surface area and the narrower band gap of 1.35 eV
show much stronger adsorption and photodegradation abilities
for MB and RhB solutions compared to yellow SnS
2
products.
Our results indicate the potential applications of the SnS
x
nanostructures in visible-light-driven photocatalysts, high
response photodetectors and other optoelectronic nanodevices.
Acknowledgements
This work was supported by the National Natural Science
Foundation (21001046, 51002059), the 973 Program of China
(No.2011CB933300, 2011CBA00703), the Basic Scientific
Research Funds for Central Colleges (2010QN045), the Research
Fund for the Doctoral Program of Higher Education
(20090142120059, 20100142120053) and the Director Fund of
WNLO. We thank the Analytical and Testing Center of Huaz-
hong University Science & Technology, and the Center of Micro-
Fabrication and Characterization (CMFC) of WNLO for the
samples measurements.
Notes and references
1 C. McManamon, J. D. Holmes and M. A. Morris, J. Hazard. Mater.,
2011, 193, 120.
2 T. Kokubu, Y. Oaki and E. Hosono, Adv. Funct. Mater., 2011, 21,
3673.
3 Y. S. Li, J. Xu, J. F. Chao, D. Chen, S. X. Ouyang, J. H. Ye and
G. Z. Shen, J. Mater. Chem., 2011, 21, 12852.
4 K. Varoon, X. Y. Zhang and B. Elyassi, Science, 2011, 334, 6052.
5 Z. X. Yang, G. D. Du, Z. P. Guo, X. B. Yu, Z. X. Chen, T. L. Guo
and R. Zeng, Nanoscale, 2011, 3, 4440.
6 Z. X. Y. G. Zhu, J. Xu, H. T. Huang, D. Chen and G. Z. Shen,
CrystEngComm, 2011, 13, 6393.
7 T. N. Huan, T. Ganesh, K. S. Kim, S. Kim, S. H. Han and H. Chung,
Biosens. Bioelectron., 2011, 27, 183.
8 H. I. Lee, G. d. Stucky, J. H. Kim, C. Pak, H. Chang and J. M. Kim,
Adv. Mater., 2011, 23, 2357.
9 L. P. Xu, Y. S. Ding, C. H. Chen, L. L. Zhao, C. Rimkus, R. Joesten
and S. L. Suib, Chem. Mater., 2008, 20, 308.
10 D. Chen and J. H. Ye, Adv. Funct. Mater., 2008, 18, 1922.
11 D. Chen, Z. Liu, S. X. Ouyang and J. H. Ye, J. Phys. Chem. C, 2011,
115, 15778.
12 M. A. Fox and M. T. Dulay, Chem. Rev., 1993, 93, 341.
13 D. Chen, S. X. Ouyang and J. H. Ye, Nanoscale Res. Lett., 2009, 4, 274.
14 J. Xu, Y. G. Zhu, H. T. Huang, Z. Xie, D. Chen and G. Z. Shen,
CrystEngComm, 2011, 13, 2629.
15 H. M. Sung-Suh, J. R. Choi, H. J. Hah, S. M. Koo and Y. C. Bae,
J. Photochem. Photobiol., A, 2004, 163, 37.
16 T. Wu, G. Liu, J. Zhao, H. Hidaka and N. Serpone, J. Phys. Chem. B,
1998, 102, 5845.
17 F. Zhang, J. Zhao, L. Zang, T. Shen, H. Hidaka, E. Pelizzetti and
N. Serpone, J. Mol. Catal. A: Chem., 1997, 120, 173.
18 J. M. Wu andT. W. Zhang, J. Photochem. Photobiol., A, 2004, 162, 171.
19 F. Yang, Y. takahashi and N. Sakai, J. Phys. Chem. C., 2011, 115
,
18270.
20 D. Chen, S. X. Ouyang and J. H. Ye, Nanoscale Res. Lett., 2009, 4, 274.
21 J. Xu, W. Meng, Y. Zhang, L. Lei and C. S. Guo, Appl. Catal., B,
2011, 107, 355.
22 G. Q. Li, N. Yang, X. L. Yang, W. L. Wang and W. F. Zhang,
J. Phys. Chem. C, 2011, 115, 13734.
23 P. Schmitt, N. Brem, S. Schunk and C. Feldmann, Adv. Funct. Mater.,
2011, 21, 3037.
24 X. C. Song, Y. F. Zheng, R. Ma, Y. Y. Zhang and H. Y. Yin,
J. Hazard. Mater., 2011, 192, 186.
25 D. Chen and J. H. Ye, J. Phys. Chem. Solids, 2007, 68, 2317.
26 W. M Du, J. Zhu, S. X. Li and X. F. Qian, Cryst. Growth Des., 2008,
8, 2130.
27 N. Z. Bao, L. M. Shen, T. Takata and K. Domen, Chem. Mater.,
2008, 20, 110.
28 D. Chen, G. Z. Shen, K. B. Tang, S. J. Lei, H. G. Zheng and
Y. T. Qian, J. Cryst. Growth, 2004, 260, 469.
29 M. Muller, R. Zentel, T. Maka, S. G. Romanov and C. M. S. Torres,
Adv. Mater., 2000, 12, 1499.
30 S. S. Hegde, A. G. Kunjomana, K. A. Chandrasekharan, K. Ramesh
and M. Prashantha, Phys. B, 2011, 406, 1143.
31 Y. Lei, S. Song, W. Fan, Y. Xing and H. Zhang, J. Phys. Chem. C,
2009, 113, 1280.
32 Y. C. Zhang, Z. N. Du, K. W. Li and M. Zhang, Sep. Purif. Technol.,
2011, 81, 101.
33 W. Du, D. Deng, Z. Han, W. Xiao, C. Bian and X. Qian, Cryst. Eng.
Comm, 2011, 13, 2071.
34 Y. C. Zhang, J. Li, M. Zhang and D. D. Dionysiou, Environ. Sci.
Technol., 2011, 45, 9324.
35 S. Sohila, M. Rajalakshmi, C. Ghosh, A. K. Arora and
C. Muthamizhchelvan, J. Alloys Compd., 2011, 509, 5843.
36 P. M. Nikolic, P. Lj Milikovic, B. Mihajlovic and Lavrencic, J. Phys.
C: Solid State Phys., 1977, 10, L289.
37 S. Sohila, M. Rajalakshmi, C. Ghosh, A. K. Arora and
C. Muthamizhchelvan, J. Alloys Compd.
, 2011, 509, 5843.
38 G. Lucovsky and J. C. Mikkelsen Jr, Phys. Rev. B: Solid State, 1976,
14, 1663.
39 H. Fu, L. Zhang, W. Yao and Y. Zhu, Appl. Catal., B, 2006, 66, 100.
40 J. Tang, Z. Zou and J. Ye, J. Phys. Chem. B, 2003, 107, 14265.
41 G. Li, N. Yang, W. Wang and W. F. Zhang, Electrochim. Acta, 2010,
55, 7235.
42 Y. Xu, N. Al-Salim, C. W. Bumby and R. D. Tilley, J. Am. Chem.
Soc., 2009, 131, 15990.
43 S. G. Hickey, C. Waurisch, B. Rellinghaus and A. Eychmuller, J. Am.
Chem. Soc., 2008, 130, 14978.
44 Z. R. Wang, H. Wang, B. Liu, W. Z. Qiu, J. Zhang, S. H. Ran,
H. T. Huang, J. Xu, H. W. Han, D. Chen and G. Z. Shen, ACS
Nano, 2011, 5, 8412.
45 H. Ye, H. S. Park, V. A. Akhavan, B. W. Goodfellow,
M. G. Panthani, B. A. Korgel and A. J. Bard, J. Phys. Chem. C,
2011, 115, 234.
46 S. G. Hickey, C. Waurisch, B. Rellinghaus and A. Eychmuller, J. Am.
Chem. Soc., 2008, 130, 14978.
47 Y. C. Zhang, Z. N. Du, S. Y. Li and M. Zhang, Appl. Catal., B, 2010,
95, 153.
48 C. Yang, W. Wang, Z. Shan and F. Huang, J. Solid State Chem.,
2009, 182, 807.
49 P. Tang, H. Chen, F. Cao, G. Pan, K. Wang, M. Xu and Y. Tong,
Mater. Lett., 2011, 65, 450.
50 Y. Lei, S. Song, W. Fan, Y. Xing and H. Zhang, J. Phys. Chem. C,
2009, 113, 1280.
3168 | CrystEngComm, 2012, 14, 3163–3168 This journal is ª The Royal Society of Chemistry 2012
Downloaded by Huazhong University of Science & Technology on 22 April 2012
Published on 28 February 2012 on http://pubs.rsc.org | doi:10.1039/C2CE06586J
View Online
... Because of their exceptional properties, they have inclusive applications in photocatalysis since they possess immense surface area among other types of nanomaterials. In addition, they can be exploited in optoelectric applications, sensors, and batteries' electrodes (Chao et al. 2012). These semiconductors show a band gap ranging between 0.9 and 1.9 eV and are appropriate absorbers for sunlight utilized in photovoltaic devices (Gerein and Haber 2006) and photocatalysis (Zhang et al. 2022). ...
Article
Full-text available
The development of efficient photocatalysts for the photodegradation of organic dyes is highly desired. In this work, BixSn6-2xSy (0.33 ≤ x ≤ 2.95) photocatalysts were synthesized using the nanoemulsion technique with morphological structures that changed from nanowhiskers to quantum dots (QDs). The optical properties of these materials were examined by UV-visible absorbance spectroscopy and photoluminescence, while Mott-Schottky analysis was utilized to study their electronic properties. BixSn6-2xSy materials possess appreciable absorption in the UV-visible light range with a direct band gap that increases from 1.23 to 1.46 eV. Both crystal structure and composition greatly affect the photocatalytic activity of BixSn6-2xSy semiconductors. Among the various synthesized photocatalysts, BiSn4S4.5 can efficiently photodegrade methylene blue dye within 10 min under UV-visible light. The photocatalytic activity is positively affected by the change of crystal structure from orthorhombic to cubic symmetry. Based on the Mott-Schottky plots, the flat band potential (Efb) and the semiconductor behavior of the fabricated BixSn6-2xSy nanomaterials were determined. The obtained Efb values for SnS, Bi0.3Sn5.34S5.8, BiSn4S5.5, and Bi2.14Sn1.71S4.7 are -0.18 V, -0.42 V, -0.53 V, and -0.51 V (vs. Ag/AgCl), respectively. The Efb value is clearly shifted towards more negative potential values from -0.18 to -0.53 V with the increase of Bi molar ratio (x). However, Bi2.95Sn0.1S4.5 semiconductor was found to be of n-type character, having a positive Efb value of +0.66 V (vs. Ag/AgCl). Photocurrent and EIS responses confirm the high stability and photocatalytic activity of BiSn4S5.5 by attaining lowest charge transfer resistance. The modified electronic properties of the BixSn6-2xSy semiconductors significantly improve their photocatalytic activity, rendering them to be promising absorbers for sunlight harvesting applications.
... Tin sulfide (SnS) is an interesting IV-VI semiconductor with an orthorhombic structure, which has received significant attention over the few past decades due to its desirable properties, such as high optical absorption coefficient, narrow band gap and high native freecarrier concentration [1]. Such properties of SnS makes it an attractive candidate for various applications such as, photo-catalysis [2], gas sensors [3], photo-electrochemical [4], photodetector [5], solar cells [6] [7], Schottky diodes [8]. In addition, SnS is widely used as an alternative solar absorber layer to conventional thin film absorbers such as Copper Gallium Selenide (CGS), Copper Zinc Tin Sulfide (CZTS), Copper Indium Gallium Selenide (CIGS), Cadmium Telluride (CdTe), and Copper Indium Gallium Selenide (CIGS) due to its several advantages, properties including biological non-toxicity [9] [10]. ...
Preprint
Full-text available
In this manuscript, we report the results of optical properties of SnS thin films, deposited on FTO coated glass substrates at room temperature by thermal evaporation technique. In addition, the effect of film thickness on the optical behavior of FTO/SnS is analyzed and obtained results are compared with data of SnS films grown on glass and ITO substrates. Our study indicates that the properties of SnS film are independent of the substrate material. Further, the influence of the film thickness on the other optical parameters including, linear and third order nonlinear optical constants and dispersion parameters have also been investigated using the transmission, reflection, and absorption spectra. It is found that the optical band gap decreases from 2.07 to 1.30 eV with increase in SnS film thickness, whereas the refractive index increases with increasing thickness. Additionally, the oscillator energy, and the dispersion energy are estimated using WempleDiDomenico approach. The dispersion energies are in the range of 7.20 to 4.59 eV, while the oscillator energies of the thin films are in the range of 5.49 to 2.24 eV. Moreover, the nonlinear refractive index, and optical susceptibility are calculated by using the empirical relation of Tichy and Ticha. The volume of data suggests optical properties of SnS thin films are strongly dependent on film thickness.
... eV, high carrier mobility (up to 230 cm 2 V -1 per second), and a high excitonic binding energy (170 meV). SnS 2 has already proven a broad range of commercial applications, including water splitting, batteries, solar cells, photocatalysis, gas sensors, field-effect transistors, hydrogen generation, and photodetectors [27][28][29][30][31][32][33][34][35][36]. On the other hand, SnS 2 , when used as a CE for DSSCs, may retain excellent stability while also producing high PCE. ...
Article
2D Layered transition metal dichalcogenides (TMDs) have great attention concerning their unique electrical and optical properties. Counter electrodes (CEs) are crucial to creating low-cost and high power conversion efficiency (PCE) dye-sensitized solar cells (DSSCs);thereby, the development of high-performance CEs utilizing non platinum (Pt) materials is essential. This study reports on new highly electrochemically active 2D/2D cobalt substituted SnS2/rGO (reduced graphene oxide) heterostructure that had been successfully synthesized by a two step hydrothermal method with different weight percentages (1, 3 and 5 wt. %) of rGO. The fabricated 2D/2D heterostructure acts as a Pt-free counter electrode in a dye sensitized solar cell. The crystalline structure, morphology, and chemical state were confirmed by X-ray diffraction (XRD), RAMAN, High resolution scanning electron microscopy (HRSEM), High resolution transmission electron microscopy (HRTEM) and X-ray photoelectron spectroscopy (XPS). Among all these, Co-SnS2/rGO (5 wt. %) counter electrode delivers the highest power conversion efficiency of 5.54 % with a short circuit current density (JSC) of 11.85 mA cm− 2, open circuit voltage (VOC) of 0.77 V and fill factor (FF) of 60.13. High surface area (59.202 → 69.148 m2/g), facilitating more active sites via increasing rGO content was analyzed through Brunauer-Emmett-Teller (BET) analysis. Lower peak-to-peak separation (Epp) and charge transfer resistance (RCT) are responsible for the observed improvement in the electrolyte and counter electrode interface. Notably, rGO with Co-SnS2 creates a synergetic effect through fast electron transporting capability. Furthermore, the Co-SnS2/rGO (5 wt. %) CE shows excellent electrochemical stability over 100 cycles in an iodine-based electrolyte. The superior electrocatalytic activity and photovoltaic output highlight the promise of Co-SnS2/rGO (5 wt. %) for DSSCs CE. This efficiency is comparable to the expensive Pt CE (6.0 %).
... Furthermore, MB dye can soak up the incident UV/solar radiation brings to the photo-induced electrons that have carried to the exciting level of the dye thanks to the intramolecular π-π* transition (HOMO-LUMO energy levels), and dyes can be also oxidized [41]. Consequently, according to the obtained outcomes the solution-processed SnS and Sn 1-x Zn x S (especially x = 1%) act for a favorable economic, the most rapid and efficient photocatalytic materials in addition to the available ones based on SnS nanoparticles or nanostructures [36][37][38][39][40][41][42][43][44][45][46][47][48][49][50] up to date (see Table 5). Our promising photocatalyst materials can be straightly utilized under UV/Solar radiation-based distillation or purification of dye contaminated water comes from textile industries, which can stop the calamity in the living sequence and the first step about supplying safe drinkable water to the human beings living in town and countryside communities. ...
Article
Zinc (Zn) doped/substituted Tin(II) sulfide (SnS) is one of the most important semiconductors for optoelectronic, photocatalytic, and gas-sensing applications. Herein, an ultra-fast, highly efficient and room temperature operated photocatalyst degrading powerful organic dyes under direct exposure to the UV–Vis/Sun-light and photo sensor based on Sn1-xZnxS (x = 0.00–0.20) nanostructured thin films were fabricated for the first time by using facile solution process. Compared to the SnS the crystallite/grain size, absorption, active surface area, absorption/reflectivity intensity, photoluminescence, and electrical conductivity of the Sn1-xZnxS thin films were improved considerably at a certain substitution level of Zn (at. %1) according to XRD, SEM, XPS, Raman, UV–Vis spectroscopy, Photoluminescence (PL), and electrical analysis. Amongst all, the Sn0.99Zn0.01S exhibits the highest photocatalytic efficiency (100% for 25 min) for degradation of MB at pH = 11 under sunlight at room temperature, which is higher than the most reported efficient photocatalyst based on SnS thin films whilst it shows 97% photocatalytic efficiency under the UV–Vis-light (at pH = 4, for 210 min) according to photocatalytic measurements. The observed highest values have been attributed to the synergic effect of pH and the highest crystallite size of the photocatalyst. Moreover, the Sn0.99Zn0.01S film sample also has the highest photosensitivity response (225%) with a decay time of 1.06 s amongst all the SnS and Zn-substituted SnS thin films due to having higher crystallite size and hence having more active crystallite site for the other Zn-substitution levels. These outcomes make the Zn-substituted SnS thin films promising and highly efficient multifunctional materials for photocatalytic, photosensitivity, and optoelectronic applications.
... Furthermore, MB dye can soak up the incident UV/solar radiation brings to the photo-induced electrons that have carried to the exciting level of the dye thanks to the intramolecular π-π* transition (HOMO-LUMO energy levels), and dyes can be also oxidized [41]. Consequently, according to the obtained outcomes the solution-processed SnS and Sn 1-x Zn x S (especially x = 1%) act for a favorable economic, the most rapid and efficient photocatalytic materials in addition to the available ones based on SnS nanoparticles or nanostructures [36][37][38][39][40][41][42][43][44][45][46][47][48][49][50] up to date (see Table 5). Our promising photocatalyst materials can be straightly utilized under UV/Solar radiation-based distillation or purification of dye contaminated water comes from textile industries, which can stop the calamity in the living sequence and the first step about supplying safe drinkable water to the human beings living in town and countryside communities. ...
Article
Zinc (Zn) doped/substituted Tin(II) sulfide (SnS) is one of the most important semiconductors for optoelectronic, photocatalytic, and gas-sensing applications. Herein, an ultra-fast, highly efficient and room temperature operated photocatalyst degrading powerful organic dyes under direct exposure to the UV–Vis/Sun-light and photo sensor based on Sn1-xZnxS (x = 0.00–0.20) nanostructured thin films were fabricated for the first time by using facile solution process. Compared to the SnS the crystallite/grain size, absorption, active surface area, absorption/reflectivity intensity, photoluminescence, and electrical conductivity of the Sn1-xZnxS thin films were improved considerably at a certain substitution level of Zn (at. %1) according to XRD, SEM, XPS, Raman, UV–Vis spectroscopy, Photoluminescence (PL), and electrical analysis. Amongst all, the Sn0.99Zn0.01S exhibits the highest photocatalytic efficiency (100% for 25 min) for degradation of MB at pH = 11 under sunlight at room temperature, which is higher than the most reported efficient photocatalyst based on SnS thin films whilst it shows 97% photocatalytic efficiency under the UV–Vis-light (at pH = 4, for 210 min) according to photocatalytic measurements. The observed highest values have been attributed to the synergic effect of pH and the highest crystallite size of the photocatalyst. Moreover, the Sn0.99Zn0.01S film sample also has the highest photosensitivity response (225%) with a decay time of 1.06 s amongst all the SnS and Zn-substituted SnS thin films due to having higher crystallite size and hence having more active crystallite site for the other Zn-substitution levels. These outcomes make the Zn-substituted SnS thin films promising and highly efficient multifunctional materials for photocatalytic, photosensitivity, and optoelectronic applications.
... However, according to the Raman spectrometer results of the samples derived from precursor B in Fig. 4b, the peak of the SnS 2 phase is clearly visible for the samples obtained from the solutions with high sulfur content. The characteristic Raman peak of the SnS 2 phase is at ~307 cm − 1 [29,30]. As seen in Fig. 4b, the films obtained with precursor B with a Sn/S molar ratio of 1/2 consist of two phases, SnS and SnS 2 . ...
Article
This study represents the investigation of earth-abundant and non-toxic SnS absorber and SnS2 buffer layer materials by using a green solution process. Some physical properties of prepared thin films were investigated for different starting chemicals and molar ratios of Sn/S. Here, by using different starting chemicals such as tin (II) 2-ethylhexanoate and tin (II) chloride dihydrate, we observed that the formation of p-SnS and n-SnS2 phases can be simply controlled. All the SnS films obtained by the tin (II) 2-ethylhexanoate showed p-type conductivity. Moreover, it has been observed that an n-type SnS2 thin film can be produced by using tin (II) chloride dihydrate as the starting chemical instead of tin (II) 2-ethylhexanoate. According to the Hall effect measurement results, the resistivity of the p-type films obtained with both starting chemicals is in the order of 10⁶ Ωcm. The resistivity of films with n-type properties drops to the order of 10¹ Ωcm. The I–V characteristic of the n-SnS2/p-SnS bilayer exhibits the p-n heterojunction diode characteristic. This study paves the way for solution processed SnS and SnS2 thin films as a potential absorber and buffer layer to be used in photovoltaic devices.
Article
SnS2 is an emerging transition metal double halogenated hydrocarbon semiconductor. SnS2 has the characteristics of adjustable band gap, large specific surface area and high carrier mobility, which has the photoelectric...
Article
The nanoparticles (Nps) of tin disulfide (SnS2) are synthesized by sonochemical route. The Nps are characterized by dispersive analysis of X-ray energy (EDAX) and X-ray photoelectron spectroscopy (XPS) to get the chemical composition. The diffraction of X-ray (XRD) is used for determination of phase and crystal structure. The as-synthesized Nps are polycrystalline and possess hexagonal structure. The surface morphology of the as-synthesized Nps is examined by electron microscopy in scanning (SEM) and high-resolution transmission modes. The residual sample after the thermal analysis is characterized by EDAX, XPS, XRD, SEM and Fourier transformed infra-red spectroscopy. The obtained results of the post-thermal analyzed and as-synthesized SnS2 Nps samples are compared. The thermal analysis of the Nps is carried out by recording the thermogravimetric and differential thermogravimetric curves. These simultaneous thermo-curves are recorded in the temperature range of ambient to 850 K in inert nitrogen atmosphere for three heating rates of 5, 10, 15 and 20 K·min⁻¹. The thermal curves data are analyzed by the iso-conversional Kissinger-Akahira-Sunose, Flynn-Wall-Ozawa, and Friedman methods, and the thermodynamic parameters; activation energy (Ea), change in activation entropy (ΔS*), change in activation enthalpy (ΔH*) and change in activation Gibb's free energy (ΔG*) are determined. All the obtained outcomes are discussed in detail.
Article
The symmetric and asymmetric perylenediimide (PDI)-based organic ligands Ph-PDI-Ph (1), Pyr-PDI-Pyr (2), and Pyr-PDI-Ph (3) were synthesized. Au/(Pyr-PDI-Ph)/p-Si (D1), and Au/(Ph-PDI-Ph)/p-Si (D2) devices were fabricated using thermal evaporation and their photodiode performances were examined under the irradiation of 100 mW.cm⁻² at −2 V. In addition to the analysis of photodiode performances, the electrical properties depending on temperature and frequency of these structures were calculated using three different methods, which are the thermionic emission, Norde function, and Cheung&Cheung function. Furthermore, the molecular geometry, electronic absorption spectra, HOMO-LUMO and MEP surface analysis of the molecules mentioned were computed using the DFT/B3LYP method with 6-31 G (d,p) basis set and DFT calculations supported both the non-planar structures or propeller structures and compatibility of the theoretical results with the experimental data. The results supported that the fabricated D1 and D2 devices can form a basis for practical applications in optoelectronic device applications, especially photodiodes and photodetectors.
Conference Paper
Full-text available
Here, we follow a simple, cost effective, and environmentally friendly one step hydrothermal method for the synthesis of nanohybrid material reduced graphene oxide-tin disulfide (rGO-SnS 2 ). Herein, the spectroscopic and structural investigation of SnS 2 anchored reduced graphene oxide (rGO) such as UV-vis DRS spectroscopy, photoluminescence (PL) spectroscopy, fourier transform infrared spectroscopy (FTIR), and X-ray diffraction (XRD) are studied. A broad peak observes at 365 nm in UV-vis DRS spectra. Optical band gap energy is calculated using Tauc’s plot and the estimated values are 1.96 and 3.91 eV, respectively. Functional groups such as Sn-S, Sn-O, C = C,-OH, C-H, C-O are identified using FTIR which attributes to the successful preparation of nanohybrid material. Structural analysis is done using XRD. XRD result confirms the hexagonal phase of the nanohybrid material and the peak at 2θ = 15.19° confirms the characteristics of SnS 2 and a small peak at 2θ = 26.03° indicates the existence of carbon. Interplanar spacing is calculated from Bragg’s law. The value is estimated to be 5.82 Å. Average crystallite size of the material is estimated using Scherrer’s formula and the value is found to be 28.16 nm. Strain of the material is calculated from Williamson-Hall plot and the value is estimated to be 0.00086.
Article
Biomimetic solid solutions of manganese and cobalt carbonates (MnCO3–CoCO3) are used as the precursors for the synthesis of the corresponding metal oxides. The different structures of the metal oxides with porous morphologies, such as pure manganese monoxide (MnO), the solid solution of manganese and cobalt monoxide (MnO–CoO), and the nanocomposite of MnO and metallic cobalt (MnO/Co), are obtained by the appropriate thermal treatments under a reduction atmosphere. The contents of cobalt species can be controlled by the initial cobalt concentration in the MnCO3–CoCO3 precursors. The resultant MnO/Co nanocomposites show the enhanced charge–discharge cycle stability and rate performance as an anode material of lithium-ion battery.
Article
Photocatalytic remote oxidation was examined under visible light by using WO3-based photocatalysts: WO3, WO3 modified with Pt nanoparticles (Pt–WO3), and WO3 modified with Cu(II) ions (Cu(II)–WO3). Alkyl chains on a glass plate separated from the photocatalyst by 7.5 μm thick air were removed in 40 min under visible light (≥420 nm, 100 mW cm–2). Among the photocatalysts examined, Pt–WO3 exhibits the highest remote oxidation activity. Its activity is higher than that of TiO2 modified with Pt nanoparticles in the wavelength range 360–460 nm. It is likely that Pt–WO3 emits •OH under visible light, which diffuses in air and attacks alkyl chains, whereas the major path of the remote oxidation induced by UV light involves emission of H2O2 and its photocleavage into •OH. Remote storage of oxidative energy in Ni(OH)2 is also possible by using visible light-irradiated Pt–WO3.
Article
A simple room-temperature mineralization method has been developed to synthesize SrWO4 micro/nanostructures with diverse morphologies, including nanoparticle, dumbbell, shuttle, sphere, spherical flower, bundle of straw, and dendrite. The products were characterized by X-ray diffraction, scanning electron microscope, UV–vis absorption, and Raman spectroscopy. The influence of pH on the size and morphology of the as-obtained products was investigated in detail. Taking the synthesized SrWO4 spheres as an example, we also investigated the photocatalytic properties of as-synthesized products for photocatalytically splitting water into H2 and O2 under UV light irradiation.
Article
N-doped antimonic acid was prepared by heating the pristine antimonic acid with urea at 400 °C for 2 h. The pristine and N-doped antimonic acids show higher visible light activities for Rhodamine B (RhB) degradation in comparison with the commercial product P25. The pristine antimonic acid, which shows no obvious visible light absorption, exhibits the highest visible light activity for RhB degradation due to RhB self-sensitized degradation. The N-doped antimonic acid, which could absorb the visible light, shows higher visible light activity for the phenol degradation than the pristine antimonic acid. The order of •OH radical formation over the pristine and N-doped antimonic acids is consistent with that of phenol photocatalytic degradation, but opposite to that of RhB degradation. We demonstrate that the extent of degradation abilities under visible light irradiation depends on whether the reagent has the self-sensitized degradation or not.
Article
Tin sulfide nanoparticles were synthesized through wet chemical route. Structure and phase purity were confirmed by powder XRD. Morphology and size were identified from TEM and AFM. The room temperature photoluminescence spectrum shows the band edge emission at 1.57 eV. The direct and indirect band gaps are estimated from UV–vis-NIR absorption spectrum as 1.78 and 1.2 eV, respectively. Blue shift of 0.48 eV observed for direct transition and 0.2 eV for indirect transition as compared to bulk band gap is due to quantum confinement effect. The Raman spectrum of SnS nanoparticles shows all the predicted Raman modes which show shift towards lower wave number side in comparison with those of the SnS single crystal. This is attributed to phonon confinement.
Article
Heterogeneously dispersed semiconductor surfaces provide both a fixed environment to influence the chemical reactivity of a wide range of adsorbates and a means to initiate light-induced redox reactivity in these weakly associated molecules. Upon photoexcitation of several semiconductors nonhomogeneously suspended in either aqueous or nonaqueous solutions or in gaseous mixtures, simultaneous oxidation and reduction reactions occur. This conversion often accomplishes either a specific, selective oxidation or a complete oxidative degradation of an organic substrate present. The paper discusses the following: survey of reactivity (functional group transformations and environmental decontamination); mechanism of photocatalysis (photoelectrochemistry, carrier trapping, inhibition of electron hole recombination by oxygen, involvement of the hydroxy radical, adsorption effects, Langmuir-Hinshelwood kinetics, pH effects, temperature effects, and sensitization); and semiconductor pretreatment and dispersion (photocatalytically active semiconductors, photocatalyst preparation, and surface perturbation). 215 refs.
Article
With greatly enhanced surface-to-volume ratios, one-dimensional (1-D) nanostructures are believed to be able to deliver better performance as chemical sensors. In this paper, by using a hydrothermal-annealing process, we reported the synthesis of porous In2O3 nanobelts with a high aspect ratio. By annealing hydrothermally-synthesized single crystalline ultralong InOOH nanobelts, single crystalline porous In2O3 nanobelts with an aspect ratio larger than 100 are obtained on a large scale. The gas sensing properties of the as-prepared porous In2O3 nanobelts were investigated, and they can detect different chemicals (methanol, ethanol, and acetone) down to ppb level. The results also showed that the nanobelts exhibited excellent gas sensing performance in terms of high sensitivity, low detection limit, fast response and recovery times, and sensing selectivity.
Article
Pure porous WO3 products consisting of 10–20 nm nanoparticles have been synthesized by controlling the hydrolysis of WCl6 in ethanol solution with the assistance of polystyrene (PS) microspheres at room temperature. The products were characterized by X-ray powder diffraction (XRD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), Brunauer–Emmett–Teller (BET) surface area and diffuse reflectance spectroscopy. The photocatalytic and gas-sensing properties of the porous WO3 samples were studied in detail. With large surface areas, these porous WO3 products exhibit both strong adsorption abilities and high degradation activities for methylene blue (MB) under visible light irradiation. And they also exhibit high sensitivity to organic gases (acetone, methanol, ethanol and formaldehyde), especially good selectivity to acetone at low concentration (0.5–5 ppm).
Article
Single-crystalline, hexagonal tin disulfide (SnS2) nanoplatelets were successfully synthesized through an improved solvothermal process with SnCl4·5H2O and carbon disulfide (CS2) as precursors. The crystal phase, morphology, and crystal lattice of the prepared products were characterized by powder X-ray diffraction (XRD), transmission electron microscopy (TEM), and high-resolution transmission electron microscopy (HRTEM), respectively. Results reveal that the synthesized SnS2 nanoplatelets are in the hexagonal structure with 20–60 nm in diameter and 7–10 nm in thickness. The structural and HRTEM analysis indicates that the formation of SnS2 nanoplatelets with hexagonal morphology result from the accelerating growth of six energetically equivalent high-index crystalline planes {110} and the retarded growth of {001} crystalline planes. The photocatalytic degradation properties of hexagonal SnS2 nanoplatelets driven by solar light were further investigated with Rhodamine B (RhB) as simulating pollutant. Results show that these SnS2 nanoplatelets have a good performance of photocatalytic degradation on RhB, and the decolorizing rate can reach 97.7% after being irradiated for 70 min by solar light. Better photocatalytic properties indicate that hexagonal SnS2 nanoplatelets are a type of promising photocatalyst driven by solar light and have potentially applied prospects in wastewater treatment and environmental protection.
Article
Using a zinc-oleate complex as an efficient self-templated precursor, we report the synthesis of ZnOnanowires and olive-shaped nanorods with growth directions along the (0001) planes via a mild hydrothermal route. As-obtained ZnO products were characterized by using X-ray powder diffraction, scanning electronic microscopy, transmission electron microscopy, UV-vis absorption spectroscopy and photoluminescence spectroscopy. Compared to ZnOnanorods, long nanowires show enhanced photoactivity for methylene blue degradation under UV light irradiation. The surface wettability of the samples was also measured and the hydrophobicity of ZnO long nanowires and the hydrophilicity of ZnO olive-shaped nanorods were observed. The gas sensitivity for acetone and ethanol of both ZnOnanorods and nanowires was measured.