You are on page 1of 10

Journal of Hazardous Materials 368 (2019) 204–213

Contents lists available at ScienceDirect

Journal of Hazardous Materials


journal homepage: www.elsevier.com/locate/jhazmat

Construction of Ni-doped SnO2-SnS2 heterojunctions with synergistic effect T


for enhanced photodegradation activity
Dayong Chena,c, Shoushuang Huanga, , Ruting Huanga, Qian Zhanga, Thanh-Tung Lea,

Erbo Chenga, Rong Yueb, Zhangjun Hua, , Zhiwen Chena,


⁎ ⁎

a
School of Environmental and Chemical Engineering, Shanghai University, Shanghai 200444, People's Republic of China
b
Department of Physics, College of Sciences, Shanghai University, Shanghai 200444, People's Republic of China
c
School of Chemical and Material Engineering, Chizhou University, Chizhou 247100, People's Republic of China

GRAPHICAL ABSTRACT

ARTICLE INFO ABSTRACT

Keywords: Construction of heterostructures with proper band alignment and effective transport and separation of photo-
Tin disulfide generated charges is highly expected for photocatalysis. In this work, Ni-doped SnO2-SnS2 heterostructures
Oxides (NiSnSO) are simply prepared by thermal oxidation of Ni-doped hierarchical SnS2 microspheres in the air. When
Quantum dots applied for the photodegradation of organic contaminants, these NiSnSO exhibit excellent catalytic performance
Doping
and stability due to the following advantages: (1) Ni doping leads to the enhancement of light harvesting of SnS2
Heterostructure
in the visible light regions; (2) the formed heterojunctions promote the transport and separation of photo-
Photocatalysis
generated electrons from SnS2 to SnO2; (3) Ni-SnO2 quantum dots facilitate the enrichment of reactants, provide
more reactive centers and accelerate product diffusion in the reactive centers; (4) the SnS2 hierarchical mi-
crospheres constituted by nanoplates provide abundant active sites, high structural void porosity and accessible
inner surface to faciliate the catalytic reactions. As a result, the optimized NiSnSO can photodegrade 92.7%
methyl orange within 80 min under the irradiation of simulated sunlight, greatly higher than those of pure SnS2
(29.8%) and Ni-doped SnS2 (52.1%). These results reveal that the combination of heteroatom doping and het-
erostructure fabrication is a very promising strategy to deliver nanomaterials for effectively photocatalytic ap-
plications.


Corresponding authors.
E-mail addresses: sshuang@shu.edu.cn (S. Huang), huzjun@shu.edu.cn (Z. Hu), zwchen@shu.edu.cn (Z. Chen).

https://doi.org/10.1016/j.jhazmat.2019.01.009
Received 14 August 2018; Received in revised form 4 January 2019; Accepted 5 January 2019
Available online 07 January 2019
0304-3894/ © 2019 Elsevier B.V. All rights reserved.
D. Chen et al. Journal of Hazardous Materials 368 (2019) 204–213

Fig. 1. (A) schematic formation process of NiSnSO samples; (B) XRD pattern of SnS2, Ni-SnS2 and calcined Ni-SnS2 samples with different durations: 50, 100, 200 and
400 min (labeled as NiSnSO-50, NiSnSO-100, NiSnSO-200 and NiSnSO-400, respectively); (C) Raman spectrum of SnS2, Ni-SnS2 and NiSnSO-100; (D)
Thermogravimetric curves of the calcined Ni-SnS2 samples under different durations: 50, 100, 200 and 400 min; The inset in Fig. 1 D shows thermogravimetric curves
of SnS2, Ni-SnS2.

1. Introduction carrier mobility, the generation and recombination of excitons, and


therefore the photocatalytic reactivity. Transition metal ion, such as Fe,
In the past period of time, shortage of energy and pollution of the Co, Ni, Mn, Cr, V, Cu and Zn, have been widely used as dopants to
environment are to be worsening due to the development of industry modify the absorption spectrum of semiconductors to improve photo-
and agriculture in the world. Photocatalysis [1–3], biological technol- catalytic performance under visible radiation [32]. Among them, Ni is
ogies [4,5], physical and chemical adsorption [6,7] and catalytic redox considered to be an crucial dopant to improve the photocatalytic ac-
reaction [8] have displayed their unique roles in relieving the ag- tivity of some semiconductors: (i) Ni exhibits the natures of inexpen-
gravations of environmental pollution or energy crisis. Especially, siveness, moderate toxicity, and excellent chemical stability [32]; (ii) Ni
semiconductor photocatalysis has exihibited its great potentials in 3d orbits can create an impurity energy band, and therefore enhance
splitting water [9–14], cutting emssions of CO2 [15–19] and elim- the visible-light response of a semiconductor [33]; (iii) Ni can serve as
inating contaminants [20–22]. Various semiconductor photocatalysts electron trapping centers, which facilitate the separation of photo-
have been explored, such as TiO2 [23,24], ZnO [24,25], CdS [26], BiOX induced charges in a semiconductor [34]; (iv) Ni doping can increase
(X ]I, Br, Cl) [27]. Nevertheless, some disadvantages, such as wide the lifespan and stability of a semiconductor photocatalyst, such as CdS
optical bandgap, low separation efficiency of photoinduced charges [26]; (v) Ni2+ has almost identical ionic radius with Sn4+ (0.69 Å),
and/or the weak oxidation ability of holes, greatly reduce their pho- which makes it easily enter into SnS2 lattice by substitution. Because of
tocatalytic performance and thus limit practical applications [28]. the above merits, doping SnS2 with Ni is highly appreciated. However,
Thus, it is an urgency to develop a novel photocatalyst with excellent the Ni-doped SnS2 nanomaterial is rarely reported. Another tactics to
photocatalytic activity to satisfy the practical application requirements elevate the photocatalytic activity of SnS2 is to combine SnS2 with
in the energy and environment areas. another semiconductor, such as Ag2CO3 [35], AgO [36], BiOCl [37],
Tin disulfide (SnS2), a semiconductor with a bandgap from 2.18 to BiOBr [38], SnS [39] and Bi2S3 [40], to form heterojunction structures.
2.44 eV, has important potential applications in the photodegradation In these cases, suppressing the recombination of charges is the crucial
of contaminations [29,30]. However, the low separation efficiency of factor to elevate the catalytic activity of SnS2. In the photocatalytic
photoinduced charges and low oxidation ability of holes inevitably process, the heterointerface serves as a convenient channel for trans-
restrict its photocatalytic activity [31]. Recently, heteroatom doping ferring the photogenerated charges (electrons or holes). In this way, the
has been demonstrated to be an effective way to improve the charge- separation of photogenerated charges in the heterojunction structure

205
D. Chen et al. Journal of Hazardous Materials 368 (2019) 204–213

Fig. 2. SEM images of (A) SnS2, (B) Ni-SnS2 and calcined Ni-SnS2 samples under different durations: (C) 50, (D) 100, (E) 200 and (F) 400 min. Magnified SEM image
(G, H) of the NiSnSO-100. EDX spectrum (I) of the as-obtained NiSnSO-100 sample.

can be accelerated under the built-in electric fields [41–45]. Other than are seldom explored, especially for 3D SnS2 microspheres and 0D SnO2
the charge separation, oxidation ability of the holes in the valence band QDs. The present methods for the preparation of SnS2/SnO2, such as
also plays a crucial role in determining the photocatalytic performance physical anchoring [53] hydrothermal method [54,55] are incapable to
of a semiconductor [46,47]. Compared with SnS2, tin dioxide (SnO2) easily adjust their microstructures. Therefore, it is necessary to explore
has much higher oxidation ability of holes in its valence band due to a facile and cost-effective method to fabricate 3D SnS2 and 0D SnO2
low valence band edge potential [31]. On the basis of the above in- QDs heterostructures with controllable microstructures. In this work,
vestigations, combining SnS2 with SnO2 to form heterostructured na- based on the above considerations, Ni-doped 0D-3D heterojunctions of
nocomposites has the chance to exert their respective advantages in SnO2 QDs-SnS2 hierarchical microspheres (NiSnSO) were synthesized
their photocatalytic applications. by a facile thermal oxidation procedure. The photocatalytic abilities of
Besides their chemical constituents, the performances of micro/na- NiSnSO were comparatively evaluated by phtodegradation of organic
nomaterials also greatly rely on their microstructures. As a tiny particle contaminants. Especially, the influence of calcination duration on the
with a size in 1–10 nm, zero-dimensional (0D) quantum dots (QDs) microstructure, optical properties and the photocatalytic performance
display many unique advantages, such as large specific surface area, of the as-synthesized heterostructure was also investigated in detail.
effective charge-transfer and tunable optoelectronics [48]. Due to these
advantages, QDs present excellent performance in the domain of pho-
tocatalysis [49,50]. However, QDs are vulnerable to self-aggregation, 2. Experimental
and abundant surface defects make them unstable compared with their
bulk materials. One of the most efficient routes to solve these problems 2.1. Synthesis of flower-like Ni-doped SnS2 microspheres
is to load QDs onto three-dimensional (3D) microstructures to form a
0D/3D nanocomposite. Three-3D micro/nanomaterials constructed In a typical process to prepare Ni-doped SnS2 microspheres, 0.7 g of
from two-dimensional (2D) possess the characteristics of large porosity, SnCl4·5H2O and 0.5 g L-cysteine was dissolved in 60 mL distilled water
high exposures of low-energy surfaces, high surface/volume ratio, high and stirred for 30 min to obtain a homogeneous solution. Then, 0.8 mL
percentages of surface atoms, and high accessibility of inner surfaces, of 0.05 M NiCl2·6H2O solution was added into the above solution. After
which are less-pronounced in low-dimensional nanomaterials [51,52]. being treated with ultrasound for a few minutes, the mixture was
These characteristics endow 3D micro/nanomaterials with increased transferred to a Teflon lined autoclave and then heated in an oven at
active centers, facilitated mass transport and diffusion. Own to these 160 °C for 24 h. The product was collected and washed with distilled
excellent features, interactions between 0D/3D moieties can make 0D water or absolute ethyl alcohol several times. Finally, the collected
QDs more dispersive and stable, and the accelerated charge transfer can precipitate was subjected to vacuum-drying at 60 °C for 12 h. The pure
effectively quench the photoluminescence of QDs and subsequently SnS2 sample was synthesized by a similar method with the Ni-SnS2
suppress the recombination of photoexcited charges. However, the sample, but without adding NiCl2·6H2O solution.
heterostructures composed of 3D microstructures and 0D quantum dots

206
D. Chen et al. Journal of Hazardous Materials 368 (2019) 204–213

Fig. 3. TEM images of (A) SnS2, (B) Ni-SnS2 and the calcined Ni-SnS2 samples under different calcination durations: (C) 50, (D) 100 min, the inset in Fig. 3D shows an
individually magnified nanoflower of NiSnSO-100. Magnified TEM image (E) of NiSnSO-100, the inset in Fig. 3E displays the magnified image of the blue circle.
HRTEM image (F) of NiSnSO-100 (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article).

2.2. Synthesis of Ni-doped SnO2-SnS2 heterostructures

To prepare the Ni-doped SnS2-SnO2 heterostructures, the as-pre-


pared Ni-SnS2 was placed in a quartz boat and calcined at 350 °C in a
furnace under the air atmosphere. The calcination durations of 50, 100,
200 and 400 min were applied for the samples, respectively. These
samples are labeled as NiSnSO-50, NiSnSO-100, NiSnSO-200, and
NiSnSO-400, respectively.

2.3. Materials characterization

X-ray diffraction (XRD) patterns of the samples were obtained from


a diffractometer (Rigaku D/max-2500) with 2θ ranging from 10° to 85°.
Fourier transform infrared (FT-IR) and Raman spectra were recorded on
a Bruker FT-IR instrument and DXR Raman microscope, respectively.
The weight loss of the samples was analyzed by an HCT-3 thermal
analyzer in the air. The temperature rising rate is 10 °C/min. The mi-
crostructures of the samples were observed on a JEOL electron micro-
Fig. 4. FT-IR spectra of SnS2, Ni-SnS2, and the calcined Ni-SnS2 samples under scope (JEM-2100 F). Optical absorption and band gap of the samples
different calcination durations: 50, 100, 200 and 400 min. were investigated by a Hitachi U4100 UV Spectrometer. The photo-
luminescence (PL) spectra were investigated on an RF-5301PC fluor-
escence spectrometer. Elements and their chemical states of the samples

207
D. Chen et al. Journal of Hazardous Materials 368 (2019) 204–213

Fig. 5. XPS spectra of the as-synthesized NiSnSO-100 sample: (A) Survey, (B) Sn 3d, (C) Ni 2p, (D) S 2p, (E) O 1 s, and atomic percent (F) of Sn, S, O and Ni.

were detected on an X-ray photoelectron spectrometer with a mono- methyl orange (MO) solution (10 mg/L), which was then treated ul-
chromated Al Kα radiation. trasonically for 10 min to make the catalyst evenly disperse in the MO
aqueous solutions. The resulting suspension was subjected to magnetic
2.4. Electrochemical measurements stirring for 30 min in dark to balance the adsorption-desorption of MO
on photocatalysts. Subsequently, the suspensions were illuminated by a
The transient photocurrent measurement was performed in a three- 400 W metal halide lamp. During the illumination, a certain volume of
electrode system using a CHI600 A electrochemical workstation. In the the suspension (2 mL) was withdrawn in a 20 min interval, and filtered
three-electrode system, indium-tin oxide (ITO) glass coated with a by a sterile syringe filter to remove photocatalysts from the suspension.
sample, platinum wire and Ag/AgCl (saturated KCl) were used as the The degradation processes of MO were studied by monitoring the de-
working, counter and reference electrodes, respectively. TBAPF6 acet- colorization on an ultraviolet-visible spectrophotometer. The test wa-
onitrile solution (0.1 mol/L) was applied as electrolyte. In the testing velengths used to monitor the decolorization of MO range from 200 to
process, the three-electrode system was illuminated by a Xe lamp at an 700 nm.
interval of 20 s.

2.5. Photocatalytic activity

In a typical experiment, 10 mg catalysts were dispersed in 50 mL of

208
D. Chen et al. Journal of Hazardous Materials 368 (2019) 204–213

Fig. 6. UV–vis diffuse reflectance


spectra (A), (Ahv)2 vs hv plots (B) of the
as-synthesized SnS2, Ni-SnS2, NiSnSO-
50, NiSnSO-100, NiSnSO-200 and
NiSnSO-400 samples; The inset in Fig.
R4B shows the optical band gaps of the
corresponding samples; PL spectrum
(C) of the SnS2, Ni-SnS2, and NiSnSO-
100 samples; Transient photocurrent
responses (D) of the SnS2 and NiSnSO-
100 samples.

Fig. 7. Photodegradation results (A) of


MO by the SnS2, Ni-SnS2, NiSnSO-50,
NiSnSO-100, NiSnSO-200 and NiSnSO-
400; time-dependent UV–vis adsorp-
tion spectra (B) of MO solution photo-
degraded by NiSnSO-100; rate con-
stants (C) of MO photodegradation by
the SnS2, Ni-SnS2, NiSnSO-50, NiSnSO-
100, NiSnSO-200 and NiSnSO-400; cy-
cling runs (D) for photodegradation of
MO on NiSnSO-100.

3. Results and discussion Fig. 1A. 3D flower-like Ni-doped SnS2 hierarchical microspheres were
firstly prepared through a facile hydrothermal process at 160 °C for
3.1. Structure and morphology 24 h. Then, SnO2 QDs in-situ grew on the surface of Ni-doped SnS2
microspheres by a facile thermal oxidation procedure in the air. The
The strategy for the preparation of NiSnSO samples was displayed in loading amount of Ni-SnO2 QDs on Ni-doped SnS2 hierarchical

209
D. Chen et al. Journal of Hazardous Materials 368 (2019) 204–213

Energy-dispersive X-ray spectroscopy (EDX) demonstrates that the


NiSnSO-100 is consisted of Sn, O, S and Ni elements (Fig. 2H and 2I).
All above results indicate that the element of Ni has been successfully
doped into the crystal lattice of SnS2, and that SnS2 has been partially
transformed to SnO2 after being calcined for 100 min in the air.
Transmission electron microscopy (TEM) images in Fig. 3A and B
show deep black, indicating that the flower-like SnS2 microspheres are
closely packed by nanoplates. Fig. 3C and D show that the NiSnSO-50
and NiSnSO-100 have similar shapes with that of pure Ni-SnS2. The
inset in Fig. 3D displays an individual NiSnSO-100 with larger magni-
fication, which provides more evidence of their flower-like features.
From Fig. 3E, it can be seen that lots of ultrasmall Ni-SnO2 QDs are
homogeneously dispersed on the surface of
Ni-SnS2 microspheres after being calcined for 100 min. The inset in
Fig. 3E indicates that the sizes of Ni-SnO2 QDs are about 5 nm. The
Fig. 8. Diagram of the energy band structure and electron transfer in the Ni-
doped SnO2-SnS2 composites. high-resolution TEM image of NiSnSO-100 (Fig. 3F) reveals that the Ni-
SnS2 microspheres and Ni-SnO2 QDs are well crystallized. The inter-
planar distances of 0.335, 0.264 and 0.237 nm can be assigned to the
microspheres can be tuned by adjusting the calcination durations. (100), (101) and (200) planes of the rutile SnO2, respectively. The
Fig. 1B shows the X-ray diffraction patterns (XRD) of the as-synthesized spacings of 0.316 and 0.278 nm correspond to the (110) and (101)
samples. SnS2 and Ni-SnS2 show the identical diffraction peaks of lattice planes of the hexagonal SnS2, respectively. These results verify
hexagonal SnS2 (JCPDS card no. 23-0677) at 15.0°, 28.2°, 32.1°, 41.9°, the coexistence of Ni-SnO2 and Ni-SnS2 in the NiSnSO-100 nano-
50.0° and 52.5°, indicating that the crystallographic structure of Ni- composite.
SnS2 is not influenced by the doping of Ni element. When Ni-SnS2 was In order to understand the calcination procedures, the FT-IR spec-
calcined for 50 min in the air, it gives three new diffraction peaks at troscopy measurement was conducted. In Fig. 4, two peaks are found at
26.6°, 33.9° and 51.8°, which can be assigned to (110), (101), (211) 3430.8 and 2926.9 cm−1, which can be assigned to stretching vibration
planes of rutile SnO2 (JCPDS card no. 41-1445), respectively, indicating of the OeH bond and stretching vibrations of CeH in CH2, respectively.
the successful in-situ growth of SnO2 on Ni-SnS2. By comparing the XRD The peak at 1646.2 cm−1 is assigned to stretching vibration of C]O.
patterns of the four calcined samples, the diffraction peaks belong to While, the peak located at 1096.3 cm-1 is associated with CeN
hexagonal SnS2, such as (001), (100), (101), (102), tended to be weaker stretching vibrations coupled with out-of-plane NH2 [56]. The absorp-
and weaker with the increased calcination durations from 50, 100, 200 tion peaks observed at 663.7 and 559.8 cm−1 are associated with the
to 400 min. This phenomenon can be attributed to the decreased Sn-S and S-Sn-S in SnS2, respectively [57]. The comparison of the FT-IR
amount of Ni-SnS2 due to the thermal oxidation of Ni-SnS2 to Ni-doped spectra of the NiSnSO-50 and NiSnSO-100 with the Ni-SnS2 reveals that
SnO2 (Ni-SnO2). In Fig. S1, the diffraction peaks belong to SnS2 dis- the intensities of Sn-S and S-Sn-S peaks of the NiSnSO-50 and NiSnSO-
appeared with further increasing the calcination duration to 800 min, 100 are obviously decreased. It means that the Sn-S and S-Sn-S bonds
indicating the complete transformation of Ni-SnS2 to Ni-SnO2. The are breaked by the thermal oxidation treatments. Fig. 4 shows that the
diffraction peaks of SnO2 are considerably broadened in the calcined absorption intensity of the characteristic peak for S-Sn-S bond decreases
Ni-SnS2 samples, suggesting that the Ni-SnO2 particles feature an ul- with increasing calcination durations, which almost disappears at 200
trasmall size. Raman spectra of the samples are displayed in Fig. 1C. and 400 min. The above evidence suggests that the calcination treat-
One can see that all the Raman spectra display A1g mode of SnS2 at ment is an effective way to promote the microstructure evolution from
311 cm−1, confirming the existence of SnS2. In the Raman spectrum of Ni-SnS2 microspheres to Ni-doped SnS2-SnO2 heterostructured nano-
the Ni-SnS2, the peaks related to NiS or NiS2 were not detected, which composites.
indicates that the Ni is successfully doped into SnS2. In contrast to that X-ray photoelectron spectroscopy (XPS) was further used to analyze
of Ni-SnS2, the Raman spectrum of NiSnSO-100 displays three weak the valence state and the surface composition of the NiSnSO-100
peaks at 228.0, 481.9 and 623.4 cm−1, corresponding to sub-bridging sample. In the survey spectrum, the primary chemical compositions of
oxygen-vacancy, Eg modes and A1g modes of rutile SnO2, respectively Sn, Ni, S and O can be determined (Fig. 5A). In Fig. 5B, the Sn 3d
[52]. In Fig. 1D and the inset in Fig. 1D, the weight loss of the Ni-SnS2 spectrum displays two peaks with the binding energy at 495.2 and
and calcined Ni-SnS2 samples declines with the increments of calcina- 486.7 eV, which can be attributed to the photoelectron emissions of Sn
tion durations from 50, 100, 200 to 400 min, indicating the transfor- 3d3/2 and Sn 3d5/2, respectively. The characteristic peaks located at
mation of Ni-SnS2 to Ni-SnO2, which is well consistent with the results 856.1 and 874.4 eV can be indexed to Ni 2p3/2 and Ni 2p1/2, re-
of XRD, Raman and FT-IR. spectively [58]. The peak positions of Ni 2p3/2 and Ni 2p1/2 are quite
The morphology details of the sample were then investigated by different from those of NiS,
scanning electron microscopy (SEM). Fig. 2A and 2B reveal that the NiS2, for which the peak positions range from 852.1 to 851.1 eV and
SnS2 and Ni-SnS2 are uniform flower-like hierarchical microspheres from 869.7 to 873.8 eV, respectively [59–62]. It hints that the Ni is
with a size in the range of 1–2 μm, implying that the doping of Ni successfully doped into the SnS2 and SnO2. The peaks centered at 162.2
element into Sn sites does not alter the morphology and size of SnS2. and 161.1 eV in Fig. 5D correspond to S 2p3/2 and S 2p1/2 photo-
Additionally, it can be seen that the morphologies of NiSnSO-50 and electron emissions, respectively, which can be considered as char-
NiSnSO-100 are similar with that of Ni-SnS2. This result suggests that acteristic of S2− in the Ni-SnS2. The peaks at 531.1 and 532.0 eV in
the original morphonology of Ni-SnS2 can be remained after it under- Fig. 5E belong to the O element of Sn-O and O-H of H2O that absorbed
went annealing treatment for 50 and 100 min in the air. Fig. 2E and 2 F on the surface of the NiSnSO-100, respectively. This spectrum of O 1 s
indicat that the Ni-SnS2 microspheres are destroyed after being calci- reveals the existence of Ni-SnO2 in the calcinated Ni-SnS2 under
nated for 200 and 400 min. From the magnified SEM image of the 100 min. According to the above results, we can conclude that the
NiSnSO-100 in Fig. 2G, it can be seen that lots of ultrasmall nano- element of Ni is chemically doped into SnS2 and SnO2. From Fig. 5F, it
particles are loaded on Ni-SnS2 microspheres. These ultrasmall particles is estimated that the percent of Ni element is about 2 at %, and that the
can be determined to be Ni-SnO2 nanoparticles, which are derived from atomic ratio of O relative to S is about 2.2.
the transformation of Ni-SnS2 during the thermal oxidation process.

210
D. Chen et al. Journal of Hazardous Materials 368 (2019) 204–213

3.2. Photocatalytic activity be seen that the concentration of MO solution is nearly invariable after
being adsorbed on NiSnSO-100 for 80 min in dark, which means that
Fig. 5A shows the UV–Vis spectrum of the as-prepared samples. the influence of adsorption can be negligible in the photocatalysis
Compared with bare SnS2, the Ni-SnS2 samples exhibit much improved process. Based on Fig. 7A, the rate constant (k) are calculated. In Fig. 7
light absorption in visible light region (400–800 nm), which means that C, the k value of the NiSnSO-100 is 7.32 and 3.48 times as high as the
Ni doping can obviously enhance light absorption of SnS2 in visible SnS2 and Ni-SnS2 samples, respectively. Fig. 7 D shows NiSnSO-100 has
light region. After Ni-SnO2/SnS2 heterojunction being formed, we can excellent recyclability for degradation of MO. As shown in Fig. S6,
see that the absorption intensity of the NiSnSO samples increases in the NiSnSO-100 also displays excellent photocatalytic activity for photo-
ultraviolet light region (200–400 nm) with the increasing annealing degrading rhodamine B (Rh B), which is obviously superior to the bare
time in air. This means that the loading of SnO2 can improve the light SnS2. The above results distinctly reveal that the Ni doping combined
absorption of SnS2 and Ni-SnS2 in ultraviolet light region. These results with the followed calcination procedure greatly affects the photo-
indicate that the combining of Ni doping and the formation of hetero- catalytic activities of SnS2.
junction can widen the light absorption in the range of 200–800 nm, Fig. 8 displays the possible photo response mechanism. The Ni
and they are of great promise to be efficient visible-light-driven pho- doping narrows the bandgap of SnS2 from 2.33 to 1.93 eV. This nar-
tocatalysts. The enhanced optical absorption can be ascribed to the rowed band gaps increase the light harvesting and extend light response
impurity level of Ni in the band of SnS2, which enhances the absorption range. The Ni dopant, serving as electron trapping centers, results in
of light owing to the transition from the Ni 3d level to the conduction much lower recombination rate of photogenerated charges for Ni-SnS2
band level of SnS2 [26,33]. Fig. 6B depicts the plot of (Ahv)2 against hv. than the pristine SnS2. In addition, thermal oxidation reaction facil-
The calculated band gap values of the SnS2, Ni-SnS2, and the calcined itates the in-situ growth of Ni-SnO2 QDs on Ni-SnS2, constructing het-
Ni-SnS2 samples under different calcination durations: 50, 100, 200 and erostructures between Ni-SnS2 and Ni-SnO2 QDs. These hetero-
400 min are 2.33, 1.93, 1.95, 1.98, 2.18 and 2.28 eV, respectively. Fig. structures offer an ideal interface for charge migration from Ni-SnS2 to
S2 and Fig. S3 reveal that the calculated band gap value is 3.28 eV for Ni-SnO2 QDs. Once illuminated by solar light, the electrons in Ni-SnS2
the Ni-SnS2 calcined for 800 min (Ni-SnO2). This value is much smaller are excited, and then migrate from the VB to CB of Ni-SnS2. Subse-
than that of the SnO2 QDs (3.89 eV) prepared under the same condition quently, these electrons (e−) can further migrate from the CB of Ni-SnS2
with Ni-SnO2. The XPS survey spectrum (Fig. S4 A) displays the Sn, O, to that of Ni-SnO2 QDs due to the formation of the heterojunction be-
Ni spectra in the Ni-SnO2. The peaks at 856.2 and 874.5 are attributed tween Ni-SnS2 and Ni-SnO2 QDs. Meanwhile, the photoinduced holes
to the Ni 2p3/2 and Ni 2p1/2, respectively [58]. These peaks further (h+) migrate reversely from the VB of Ni-SnO2 QDs to that of Ni-SnS2
verify the existence of Ni element in the Ni-SnO2 QDs. The band gap of through the heterointerface. This synergistic effect between Ni-SnS2
the Ni-SnO2 is determined to be 3.28 eV (Fig. S3), smaller than that of and Ni-SnO2 QDs with modified band gap structure promotes photo-
bare SnO2 (3.89 eV). The smaller band gap of the Ni-SO2 is beneficial to excited charge transition, decreases recombination of photoexcited
improve its photocatalytic performance. Actually, similar results can carriers, and therefore prolongs the photoexcited charge lifetime. Ad-
also be found in the previous studies as shown in Table S1. For example, ditionally, benzoquinone (BQ), tert-butyl alcohol (TBA) and Na2C2O4
Bannu et al. reported that the band gap of SnO2 thin films became are applied as scavengers to capture %O2ˉ, %OH and h+, respectively. In
smaller from 3.70 to 3.30 eV by Zn2+ doping [63]. In present work, the Fig. S7, after being added with BQ, TBA and Na2C2O4, NiSnSO-100
smaller band gap of Ni-doped SnO2 may be attributed to the following photodegrades 17.2%, 35.5% and 23.9% MO, respectively. These de-
reasons [64]: (i) Ni2+ owns two valence electrons, Sn4+ owns four gradation efficiencies are much lower than that on NiSnSO-100 without
valence electrons. Thus, the substitution of Ni2+ for Sn4+ results in an adding scavengers at the same conditions, verifying that the %O2ˉ, h+
electron deficiency in the bonding orbitals, forming an impurity band and ·OH influence significantly the photodegradation reaction. The
between the conduction band (CB) and valence band (VB) of SnO2; (ii) possible mechanism of MO photodegradation can be described as fol-
Due to the similar ionic radius of Ni2+ with Sn4+, the substitution of lowing: (i) the heterointerface between the Ni-SnS2 and Ni-SnO2 facil-
Ni2+ for Sn4+ distorts the crystal structure of SnO2 and results in a itate the transportation of e− from Ni-SnS2 to Ni-SnO2 (eqn (1) and (2));
bandtailing akin to the Urbach tail, causing a redshift of absorption (ii) superoxide anion radicals (%O2ˉ) are produced due to the scavenging
edge in SnO2. Fig. 6C displays the photoluminescence spectra (PL) of e− in Ni-SnO2 by O2 (eqn (3)); (iii) MO was degraded by %O2ˉ ; (iv) at
spectra of the as-prepared SnS2, Ni-SnS2, and NiSnSO-100. Compared the same time, ·OH was produced by the route of %O2ˉ → H2O2 → %OH
with the pure SnS2 sample, the Ni-SnS2 sample shows a weaker in- (eqn (5) and (6)) [39]; (v) MO was also degraded by %OH and h+ ((eqn
tensity of emission peak at 472 nm. This is because that the Ni dopant (7) and (8)). The SnO2 QDs, featuring ultrasmall nanoparticles, promote
serves as electron trapping centers, resulting in the enhancement in the the enrichment of reactants and provide more reactive centers by the
non-radiative recombination process [34]. The PL intensity of the confinement effect. Therefore, the quantum dots structure of SnO2 ac-
NiSnSO-100 is significantly quenched in comparison with that of Ni- celerates the photodegradation rate of MO.
SnS2, demonstrating that recombination of photo-induced carriers is
strongly inhibited due to the formation of the heterostructure. As shown
in Fig. 6D, the bare SnS2 sample displays a low photocurrent density 4. Conclusions
due to its high charge recombination rate. In contrast, the photocurrent
density of the NiSnSO-100 is significantly improved, indicating that the In summary, we have explored a combined strategy of Ni doping
charge recombination rate is reduced. This result also suggests that the and thermal oxidation to prepare Ni-doped SnO2-SnS2 heterostructures.
doping of Ni into SnS2 and the formation of heterojunctions can effi- The correlations between the calcination duration and photocatalytic
ciently separate the photoinduced charges. activity of the samples were systematically investigated. It was found
The photodegradation of MO was applied to estimate the photo- that the 3D Ni-SnS2 microspheres and 0D Ni-SnO2 QDs heterostructured
catalytic performances of the fabricated samples. Fig. 7A shows that the nanocomposites calcinated for 100 min exhibited a photodegraded rate
NiSnSO-100 exhibits the highest photocatalytic activity, photo- of 92.7% for the degradation of MO after exposure to solar light for
degrading 92.7% of MO, which is 3.11, 1.77, 1.05, 1.10 and 1.20 times 80 min, which was greatly higher than that of the pristine SnS2 mi-
as high as the SnS2, Ni-SnS2 and the Ni-SnS2 calcined for 50, 200 and crospheres (29.8%). These findings revealed that the combination of
400 min, respectively. Fig. 7 B displays that the absorbance of MO so- heteroatom doping and heterostructure fabrication is one of the pro-
lution photodegraded by NiSnSO-100 dramatically decreases with mising approaches to adjust the microstructure and to optimize the
prolonging reaction time from 20 to 80 min, which demonstrates the photocatalytic performance of nanocomposites.
rapid photodegradation of MO by NiSnSO-100. From the Fig. S5, it can

211
D. Chen et al. Journal of Hazardous Materials 368 (2019) 204–213

Acknowledgements [21] B. Yuan, Y. Long, L. Wu, K. Liang, H. Wen, S. Luo, H. Huo, H. Yang, J. Ma, TiO2@h-
CeO2: a composite yolk–shell microsphere with enhanced photodegradation ac-
tivity, Catal. Sci. Technol. 6 (2016) 6396–6405.
This work was supported by both the National Natural Science [22] Q. Zhang, R. Dong, Y. Wu, W. Gao, Z. He, B. Ren, Light-driven Au-WO3@C janus
Foundation of China (21601120, 11375111) and the Science and micromotors for rapid photodegradation of dye pollutants, ACS Appl. Mater.
Technology Commission of Shanghai Municipality (17ZR1410500). Interfaces 9 (2017) 4674–4683.
[23] M. Gao, L. Zhu, W.L. Ong, J. Wang, G.W. Ho, Structural design of TiO2-based
This work was also supported by the Natural Science Research Project photocatalyst for H2 production and degradation applications, Catal. Sci. Technol. 5
of Anhui Education Department (KJ2016A510), the Excellent Youth (2015) 4703–4726.
Talents Project of Anhui Education Department (gxyq2017104) and the [24] S.G. Kumar, K.S.R.K. Rao, Zinc oxide based photocatalysis: tailoring surface-bulk
structure and related interfacial charge carrier dynamics for better environmental
Educational Quality Project of Anhui Province (2015jyxm398). applications, RSC Adv. 5 (2015) 3306–3351.
[25] S. Maiti, S. Pal, K.K. Chattopadhyay, Recent advances in low temperature, solution
Appendix A. Supplementary data processed morphology tailored ZnO nanoarchitectures for electron emission and
photocatalysis applications, CrystEngComm. 17 (2015) 9264–9295.
[26] M. Luo, Y. Liu, J. Hu, H. Liu, J. Li, One-pot synthesis of CdS and Ni-doped CdS
Supplementary material related to this article can be found, in the hollow spheres with enhanced photocatalytic activity and durability, ACS Appl.
online version, at doi:https://doi.org/10.1016/j.jhazmat.2019.01.009. Mater. Interfaces 4 (2012) 1813–1821.
[27] L. Ye, Y. Su, X. Jin, H. Xie, C. Zhang, Recent advances in BiOX (X = Cl, Br and I)
photocatalysts: synthesis, modification, facet effects and mechanisms, Environ. Sci.
References Nano 1 (2014) 90–112.
[28] M. Ge, C. Cao, J. Huang, S. Li, Z. Chen, K.Q. Zhang, S.S. Al-Deyab, Y. Lai, A review
[1] K.M. Cho, K.H. Kim, K. Park, C. Kim, S. Kim, A. Al-Saggaf, I. Gereige, H.T. Jung, of one-dimensional TiO2 nanostructured materials for environmental and energy
Amine-functionalized graphene/CdS composite for photocatalytic reduction of CO2, applications, J. Mater. Chem. A Mater. Energy Sustain. 4 (2016) 6772–6801.
ACS Catal. 7 (2017) 7064–7069. [29] J. Qu, D. Chen, N. Li, Q. Xu, H. Li, J. He, J. Lu, Coral-inspired nanoscale design of
[2] S.N. Habisreutinger, L. Schmidt-Mende, J.K. Stolarczyk, Photocatalytic reduction of porous SnS2 for photocatalytic reduction and removal of aqueous Cr (VI), Appl.
CO2 on TiO2 and other semiconductors, Angew. Chem. Int. Ed. 52 (2013) Catal. B: Environ. 207 (2017) 404–411.
7372–7408. [30] X. Li, J. Zhu, H. Li, Comparative study on the mechanism in photocatalytic de-
[3] A.V. Akimov, J.T. Muckerman, O.V. Prezhdo, Nonadiabatic dynamics of positive gradation of different-type organic dyes on SnS2 and CdS, Appl. Catal. B: Environ.
charge during photocatalytic water splitting on GaN(10-10) surface: charge loca- 123 (2012) 174–181.
lization governs splitting efficiency, J. Am. Chem. Soc. 135 (2013) 8682–8691. [31] X. Zhou, T. Zhou, J. Hu, J. Li, Controlled strategy to synthesize SnO2 decorated SnS2
[4] S. Ye, G. Zeng, H. Wu, C. Zhang, J. Dai, J. Liang, J. Yu, X. Ren, H. Yi, M. Cheng, nanosheets with enhanced visible light photocatalytic activity, CrystEngComm 14
C. Zhang, Biological technologies for the remediation of co-contaminated soil, Crit. (2012) 5627–5633.
Rev. Biotechnol. 37 (2017) 1062–1076. [32] M.P.B. Vega, M. Hinojosa-Reyes, A. Hernández-Ramírez, J.L.G. Mar, V. Rodríguez-
[5] W.C. Li, M.H. Wong, Interaction of Cd/Zn hyperaccumulating plant (Sedum al- González, L. Hinojosa-Reyes, Visible light photocatalytic activity of sol–gel Ni-
fredii) and rhizosphere bacteria on metal uptake and removal of phenanthrene, J. doped TiO2 on p-arsanilic acid degradation, J. Solgel Sci. Technol. 85 (2018)
Hazard. Mater. 209 (2012) 421–433. 723–731.
[6] Z. Zeng, S. Ye, H. Wu, R. Xiao, G. Zeng, J. Liang, C. Zhang, J. Yu, Y. Fang, B. Song, [33] J. Chen, F. Xin, S. Qin, X. Yin, Photocatalytically reducing CO2 to methyl formate in
Research on the sustainable efficacy of g-MoS2 decorated biochar nanocomposites methanol over ZnS and Ni-doped ZnS photocatalysts, Chem. Eng. J. 230 (2013)
for removing tetracycline hydrochloride from antibiotic-polluted aqueous solution, 506–512.
Sci. Total Environ. 648 (2018) 206–217. [34] N.S. Gultom, H. Abdullah, D.-H. Kuo, Enhanced photocatalytic hydrogen produc-
[7] T. Madrakian, A. Afkhami, M. Ahmadi, H. Bagheri, Removal of some cationic dyes tion of noble-metal free Ni-doped Zn(O,S) in ethanol solution, Int. J. Hydrogen
from aqueous solutions using magnetic-modified multi-walled carbon nanotubes, J. Energy 42 (2017) 25891–25902.
Hazard. Mater. 196 (2011) 109–114. [35] J. Luo, X. Zhou, J. Zhang, Z. Du, Fabrication and characterization of Ag2CO3/SnS2
[8] S. Ye, G. Zeng, H. Wu, C. Zhang, J. Liang, J. Dai, Z. Liu, W. Xiong, J. Wan, P. Xu, composites with enhanced visible-light photocatalytic activity for the degradation
M. Cheng, Co-occurrence and interactions of pollutants, and their impacts on soil of organic pollutants, RSC Adv. 5 (2015) 86705–86712.
remediation-a review, Crit. Rev. Environ. Sci. Technol. 47 (2017) 1528–1553. [36] L. Deng, Z. Zhu, L. Liu, H. Liu, Synthesis of Ag2O and Ag co-modified flower-like
[9] G. Cui, W. Wang, M. Ma, J. Xie, X. Shi, N. Deng, J. Xin, B. Tang, IR-driven pho- SnS2 composites with enhanced photocatalytic activity under solar light irradiation,
tocatalytic water splitting with WO2-NaxWO3 hybrid conductor material, Nano Lett. Solid State Sci. 63 (2017) 76–83.
15 (2015) 7199–7203. [37] H. Meng, T. Wang, X. Yu, Y. Zhu, Y. Zhang, BiOCl/SnS2 hollow spheres for the
[10] S.J.A. Moniz, S.A. Shevlin, D.J. Martin, Z.X. Guo, J. Tang, Visible-light driven photocatalytic degradation of waste water, RSC Adv. 5 (2015) 107088–107097.
heterojunction photocatalysts for water splitting – a critical review, Energy Environ. [38] F. Qiu, W. Li, F. Wang, H. Li, X. Liu, J. Sun, In-situ synthesis of novel Z-scheme
Sci. 8 (2015) 731–759. SnS2/BiOBr photocatalysts with superior photocatalytic efficiency under visible
[11] Y. Feng, X.Y. Yu, U. Paik, N-doped graphene layers encapsulated NiFe alloy nano- light, J. Colloid Interface Sci. 493 (2017) 1–9.
particles derived from MOFs with superior electrochemical performance for oxygen [39] K. Yao, J. Li, S. Shan, Q. Jia, One-step synthesis of urchinlike SnS/SnS2 hetero-
evolution reaction, Sci. Rep. 6 (2016) 34004. structures with superior visible-light photocatalytic performance, Catal. Commun.
[12] W. Hu, L. Lin, R. Zhang, C. Yang, J. Yang, Highly efficient photocatalytic water 101 (2017) 51–56.
splitting over edge-modified phosphorene nanoribbons, J. Am. Chem. Soc. 139 [40] C. Yu, K. Wang, P. Yang, S. Yang, C. Lu, Y. Song, S. Dong, J. Sun, J. Sun, One-pot
(2017) 15429–15436. facile synthesis of Bi2S3/SnS2/Bi2O3 ternary heterojunction as advanced double Z-
[13] L. Mu, Y. Zhao, A. Li, S. Wang, Z. Wang, J. Yang, Y. Wang, T. Liu, R. Chen, J. Zhu, scheme photocatalytic system for efficient dye removal under sunlight irradiation,
F. Fan, R. Li, C. Li, Enhancing charge separation on high symmetry SrTiO3 exposed Appl. Surf. Sci. 420 (2017) 233–242.
with anisotropic facets for photocatalytic water splitting, Energy Environ. Sci. 9 [41] X. Yan, K. Ye, T. Zhang, C. Xue, D. Zhang, C. Ma, J. Wei, G. Yang, Formation of
(2016) 2463–2469. three-dimensionally ordered macroporous TiO2@nanosheet SnS2 heterojunctions
[14] P. Zhang, J. Zhang, J. Gong, Tantalum-based semiconductors for solar water for exceptional visible-light driven photocatalytic activity, New J. Chem. 41 (2017)
splitting, Chem. Soc. Rev. 43 (2014) 4395–4422. 8482–8489.
[15] F. Sastre, A.V. Puga, L. Liu, A. Corma, H. Garcia, Complete photocatalytic reduction [42] Y.C. Zhang, J. Li, H.Y. Xu, One-step in situ solvothermal synthesis of SnS2/TiO2
of CO2 to methane by H2 under solar light irradiation, J. Am. Chem. Soc. 136 (2014) nanocomposites with high performance in visible light-driven photocatalytic re-
6798–6801. duction of aqueous Cr(VI), Appl. Catal. B: Environ. 123-124 (2012) 18–26.
[16] I. Shown, H.C. Hsu, Y.C. Chang, C.H. Lin, P.K. Roy, A. Ganguly, C.H. Wang, [43] M. Sun, Q. Yan, T. Yan, M. Li, D. Wei, Z. Wang, Q. Wei, B. Du, Facile fabrication of
J.K. Chang, C.I. Wu, L.C. Chen, K.H. Chen, Highly efficient visible light photo- 3D flower-like heterostructured g-C3N4/SnS2 composite with efficient photo-
catalytic reduction of CO2 to hydrocarbon fuels by Cu-nanoparticle decorated catalytic activity under visible light, RSC Adv. 4 (2014) 31019–31027.
graphene oxide, Nano Lett. 14 (2014) 6097–6103. [44] F. Deng, L. Zhao, X. Pei, X. Luo, S. Luo, Facile in situ hydrothermal synthesis of g-
[17] L.B. Hoch, P. Szymanski, K.K. Ghuman, L. He, K. Liao, Q. Qiao, L.M. Reyes, Y. Zhu, C3N4/SnS2 composites with excellent visible-light photocatalytic activity, Mater.
M.A. El-Sayed, C.V. Singh, G.A. Ozin, Carrier dynamics and the role of surface Chem. Phys. 189 (2017) 169–175.
defects: designing a photocatalyst for gas-phase CO2 reduction, Proc. Natl. Acad. [45] M. Kovacic, H. Kusic, M. Fanetti, U.L. Stangar, M. Valant, D.D. Dionysiou,
Sci. U. S. A. 113 (2016) E8011–E8020. A.L. Bozic, TiO2-SnS2 nanocomposites: solar-active photocatalytic materials for
[18] X. Li, J. Yu, M. Jaroniec, Hierarchical photocatalysts, Chem. Soc. Rev. 45 (2016) water treatment, Environ. Sci. Pollut. Res. 24 (2017) 19965–19979.
2603–2636. [46] M.T. Niu, L.F. Cui, P. Huang, Y.L. Yu, Y.S. Wang, Hydrothermal synthesis, structural
[19] R. Kuriki, K. Sekizawa, O. Ishitani, K. Maeda, Visible-light-driven CO2 reduction characteristics, and enhanced photocatalysis of SnO2/Fe2O3 semiconductor nano-
with carbon nitride: enhancing the activity of ruthenium catalysts, Angew. Chem. heterostructure, ACS Nano 4 (2010) 681–688.
Int. Ed. Engl. 54 (2015) 2406–2409. [47] T. Gao, T. Wang, Sonochemical synthesis of SnO2 nanobelt/CdS nanoparticle core/
[20] A. Ghosh, P. Guha, A.K. Samantara, B.K. Jena, R. Bar, S.K. Ray, P.V. Satyam, Simple shell heterostructures, Chem. Commun. (2004) 2558–2559.
growth of faceted Au-ZnO hetero-nanostructures on silicon substrates (nanowires [48] M.Y. Ye, Z.H. Zhao, Z.F. Hu, L.Q. Liu, H.M. Ji, Z.R. Shen, T.Y. Ma, 0D/2D hetero-
and triangular nanoflakes): a shape and defect driven enhanced photocatalytic junctions of vanadate quantum dots/graphitic carbon nitride nanosheets for en-
performance under visible light, ACS Appl. Mater. Interfaces 7 (2015) 9486–9496. hanced visible-light-driven photocatalysis, Angew. Chem. Int. Ed. 56 (2017)
8407–8411.

212
D. Chen et al. Journal of Hazardous Materials 368 (2019) 204–213

[49] Y.J. Gao, X.B. Li, H.L. Wu, S.L. Meng, X.B. Fan, M.Y. Huang, Q. Guo, C.H. Tung, 5002.
L.Z. Wu, Exceptional catalytic nature of quantum dots for photocatalytic hydrogen [57] G. Hatui, G. Chandra Nayak, G. Udayabhanu, Y.K. Mishra, D.D. Pathak, Template-
evolution without external cocatalysts, Adv. Funct. Mater. 28 (2018) 1801769. free single pot synthesis of SnS2@Cu2O/reduced graphene oxide (rGO) nanoflowers
[50] S. Yu, X.B. Fan, X. Wang, J. Li, Q. Zhang, A. Xia, S. Wei, L.Z. Wu, Y. Zhou, for high performance supercapacitors, New J. Chem. 41 (2017) 2702–2716.
G.R. Patzke, Efficient photocatalytic hydrogen evolution with ligand engineered all- [58] J. Shen, S. Guo, C. Chen, L. Sun, S. Wen, Y. Chen, S. Ruan, Synthesis of Ni-doped α-
inorganic InP and InP/ZnS colloidal quantum dots, Nat. Commun. 9 (2018) 4009. MoO3 nanolamella and their improved gas sensing properties, Sens. Actuators B
[51] J. Shi, X. Wang, S. Zhang, L. Xiao, Y. Huan, Y. Gong, Z. Zhang, Y. Li, X. Zhou, Chem. 252 (2017) 757–763.
M. Hong, Q. Fang, Q. Zhang, X. Liu, L. Gu, Z. Liu, Y. Zhang, Two-dimensional [59] T. Tian, L. Huang, L. Ai, J. Jiang, Surface anion-rich NiS2 hollow microspheres
metallic tantalum disulfide as a hydrogen evolution catalyst, Nat. Commun. 8 derived from metal–organic frameworks as a robust electrocatalyst for the hydrogen
(2017) 958. evolution reaction, J. Mater. Chem. A 5 (2017) 20985–20992.
[52] D. Voiry, J. Yang, M. Chhowalla, Recent strategies for improving the catalytic ac- [60] J. Zhao, B. Guan, B. Hu, Z. Xu, D. Wang, H. Zhang, Vulcanizing time controlled
tivity of 2D TMD nanosheets toward the hydrogen evolution reaction, Adv. Mater. synthesis of NiS microflowers and its application in asymmetric supercapacitors,
28 (2016) 6197–6206. Electrochim. Acta 230 (2017) 428–437.
[53] L. Ma, L. Xu, X. Xu, L. Zhang, X. Zhou, Fabrication of SnO2/SnS2 hybrids by an- [61] P. Thangasamy, V. Maruthapandian, V. Saraswathy, M. Sathish, Supercritical fluid
choring ultrafine SnO2 nanocrystals on SnS2 nanosheets and their photocatalytic processing for the synthesis of NiS2 nanostructures as efficient electrocatalysts for
properties, Ceram. Int. 42 (2016) 5068–5074. electrochemical oxygen evolution reactions, Catal. Sci. Technol. 7 (2017)
[54] X. Jiao, X. Li, X. Jin, Y. Sun, J. Xu, L. Liang, H. Ju, J. Zhu, Y. Pan, W. Yan, Y. Lin, 3591–3597.
Y. Xie, Partially oxidized SnS2 atomic layers achieving efficient visible-light-driven [62] T. Liu, C. Jiang, B. Cheng, W. You, J. Yu, Hierarchical NiS/N-doped carbon com-
CO2 reduction, J. Am. Chem. Soc. 139 (2017) 18044–18051. posite hollow spheres with excellent supercapacitor performance, J. Mater. Chem. A
[55] Y.C. Zhang, L. Yao, G. Zhang, D.D. Dionysiou, J. Li, X. Du, One-step hydrothermal 5 (2017) 21257–21265.
synthesis of high-performance visible-light-driven SnS2/SnO2 nanoheterojunction [63] M.S. Bannur, A. Antony, K.I. Maddani, P. Poornesh, A. Rao, K.S. Choudhari,
photocatalyst for the reduction of aqueous Cr(VI), Appl. Catal. B: Environ. 144 Tailoring the nonlinear optical susceptibility χ(3), photoluminescence and optical
(2014) 730–738. band gap of nanostructured SnO2 thin films by Zn doping for photonic device ap-
[56] Z. Wang, Y. Dong, H. Li, Z. Zhao, H. Bin Wu, C. Hao, S. Liu, J. Qiu, X.W. Lou, plications, Phys. E 103 (2018) 348–353.
Enhancing lithium–sulphur battery performance by strongly binding the discharge [64] D. Mocatta, G. Cohen, J. Schattner, O. Millo, E. Rabani, U. Banin, Heavily doped
products on amino-functionalized reduced graphene oxide, Nat. Commun. 5 (2014) semiconductor nanocrystal quantum dots, Science 332 (2011) 77–81.

213

You might also like