You are on page 1of 13

Sensors & Actuators: B.

Chemical 403 (2024) 135215

Contents lists available at ScienceDirect

Sensors and Actuators: B. Chemical


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

One-pot hydrothermal synthesis of self-assembled MoS2/WS2 nanoflowers


for chemiresistive room-temperature NO2 sensors
Zhiping Liang a, 1, Mingyuan Wang b, 1, Siwei Liu a, Mobashar Hassan a, Xiangzhao Zhang a,
Shuangying Lei b, Guanjun Qiao a, Guiwu Liu a, *
a
School of Materials Science and Engineering, Jiangsu University, Zhenjiang 212013, China
b
SEU-FEI Nano-Pico Center, Key Laboratory of MEMS of Ministry of Education, School of Electrical Science and Engineering, Southeast University, 210096 Nanjing,
China

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

Key words: Transition metal dichalcogenides (TMDs) based heterostructures are becoming hotspots of gas-sensing applica­
Transition metal dichalcogenides tions due to the high electronic mobility, high surface-volume ratio, and abundant adsorption sites. However, the
Self-assembled nanoflowers preparation of TMDs-based heterostructures is usually required multiple steps or higher synthesis temperatures
Heterostructure
(≥ 500 ◦ C). Herein, the flower-like MoS2/WS2 composites were firstly prepared by a facile one-pot hydrothermal
Hydrothermal synthesis
method to effectively detect NO2 at room temperature. The WS2 molecules prefer to grow on the edge of as-
Gas sensing
nucleated MoS2 to form the MoS2/WS2 nanosheets, and then they can grow by layer-by-layer self-assembly to
form the MoS2/WS2 heterostructure. The optimized MoS2/WS2 heterostructured composite exhibits the very high
sensitivity to NO2 at room temperature, involving high response, good reproducibility, excellent selectivity and
long-term stability. Specially, the responses (ΔR/Rg) of MoS2/WS2-0.5 composite to 5 and 100 ppb NO2 are 6.1%
and 23.1%, respectively. Moreover, a plausible gas sensing mechanism was proposed from viewpoints of electron
transfer and gas adsorption by combining multiple experiments and DFT calculations. This work provides a facile
process for synthesis of TMDs-based heterostructures, but also a prospective gas sensing material for room-
temperature NO2 detection.

1. Introduction dichalcogenides (TMDs), are becoming hotspots of electronic, opto­


electronic, energy-saving and gas-sensing applications due to their
Due to growing concerns about the environment and human health, unique properties, involving thickness-dependent bandgap, high elec­
there has been an increasing focus on monitoring hazardous gases in tronic mobility, high surface-volume ratio, and abundant adsorption
modern society [1,2]. NO2 is one of the most common air pollutants that sites [8–11]. WS2 and MoS2 are two typical TMDs with good thermal
can damage the ecosystem by triggering photochemical smog, acid rain stability and room-temperature sensitivity to some harmful gases in
and ground-level O3 depletion [3]. Moreover, the NO2 concentration in ambient environments, so they draw more attentions in the field of gas
ambient environments was regulated by Japan government as the sensing [12]. What’s more, the dangling bonds at the edges of TMDs
threshold limit value (TLV) of ≤ 40–60 ppb because of its harm to possess more active sites, which contribute to the adsorption of the gases
human health [4,5]. Generally, the metal oxide semiconductors (MOSs) and the formation of heterostructures with other materials [13]. How­
are used as the gas and humidity sensing materials. However, their ever, the disadvantages of pristine WS2 and MoS2 cannot be ignored,
higher operating temperatures of 200–400 ◦ C limit their potential for the including the low response and incomplete recovery to target gases at
energy-saving development of gas sensor in future [6,7]. Therefore, the room temperature [14,15]. To address these issues, some feasible stra­
development of room-temperature-operated and highly sensitive gas tegies were proposed by constructing TMDs-based heterostructures to
sensing materials is critical to creating efficient and reliable gas sensors improve the room-temperature gas sensing performance. For instance,
for toxic and harmful gases. Ikram et al. fabricated a large-surface-area WS2/MoS2 heterojunction by
Recently, two-dimensional materials, such as transition metal combining the exfoliation and hydrothermal processes, indicating that

* Corresponding author.
E-mail address: gwliu76@ujs.edu.cn (G. Liu).
1
The authors contribute equally to this paper.

https://doi.org/10.1016/j.snb.2023.135215
Received 30 August 2023; Received in revised form 21 December 2023; Accepted 21 December 2023
Available online 23 December 2023
0925-4005/© 2023 Elsevier B.V. All rights reserved.
Z. Liang et al. Sensors and Actuators: B. Chemical 403 (2024) 135215

the formation of WS2/MoS2 heterostructure can greatly enhance the After the stainless steel autoclaves were furnace-cooled to room tem­
room-temperature gas sensitivity to NO2, with the detection limit of 10 perature, the deposits were removed from the autoclaves, and the main
ppb [16]. Kim et al. fabricated the WS2/WSe2 and MoS2/WSe2 com­ powder products were obtained by repeatedly washing with ethanol and
posites by in-situ chemical vapor deposition (CVD) process, indicating centrifuging, and then dried in a vacuum drying oven at 50 ◦ C for 10 h.
that the TMDs-based heterostructures present the good Finally, the as-dried powders were calcined at 380 ◦ C for 5 h in pure N2
room-temperature gas sensing performance, providing a platform for to obtain highly crystalline composite powder samples. Here, the MoS2/
highly sensitive and selective gas sensor applications [17]. WS2 composites with Mo/W molar ratios of 0.25, 0.5 and 1 are labeled
Indeed, the CVD process was commonly used to prepare the TMDs- as MoS2/WS2-0.25, MoS2/WS2-0.5 and MoS2/WS2-1, respectively. The
based heterostructures [17–20]. Actually, some multi-step processes, preparation method of MoS2 was similar to that of WS2 [24], where only
involving combination of exfoliation and hydrothermal synthesis [21], the Na2WO4⋅2 H2O was replaced by Na2MoO4⋅2 H2O. Fig. 1 shows the
two-step hydrothermal processes [22], and combination of etching, schematic of fabrication of self-assembled MoS2/WS2 nanoflowers and
intercalation and hydrothermal synthesis [23] were adopted to fabricate gas sensor. The detailed characterization methods of materials are
TMD-based heterostructures. However, the preparation of TMDs-based available in Supporting Information.
heterostructures is usually required multiple steps or higher synthesis
temperatures (≥ 500 ◦ C), thereby increasing the complexity of fabrica­
2.2. Preparation and performance characterization of gas sensors
tion process and energy consumption to a certain extent. Therefore, a
facile preparation of TMDs-based heterostructures needs to be explored
The drop casting method was employed to fabricate pristine WS2,
for large-scale gas sensor applications.
MoS2 and MoS2/WS2 composites based gas sensors as follows: 3 mg
In this work, flower-like MoS2/WS2 composite with TMDs hetero­
pristine MoS2 or WS2 or composite powders were added in 0.01 mL
structure were fabricated by a facile one-pot hydrothermal method. We
ethanol and ultrasonically treated for 10 min to get homogeneous sus­
investigated comparatively the room-temperature sensing performance
pensions, then dropped on the sensor substrates consisting of interdigital
of pristine MoS2, WS2 and MoS2/WS2 composites to different concen­
Pt electrode, and finally dried in a vacuum drying oven. The SD101-type
trations of NO2, and subsequently emphasized the reproducibility,
dynamic gas sensor test system was used to record the response and
selectivity and long-term stability of the optimized MoS2/WS2 com­
recovery characteristics of pristine MoS2, WS2 and MoS2/WS2 compos­
posite. The optimized MoS2/WS2 composite exhibits superior room-
ites. This test system consists of the gas cylinder, Al-alloy gas test
temperature (25 ◦ C) gas sensitivity to different concentrations
chamber, four-channel gas mass flow control module and software for
(0.005–50 ppm) of NO2. In particular, the responses of MoS2/WS2-0.5
controlling and testing the gas [24]. Noted that the precision of gas
composite to 5 and 100 ppb NO2 arrive at 6.1% and 23.1%, respectively.
control is ~0.35%. The as-coated sensor substrates were exposed to
The formation and gas sensing mechanisms of the MoS2/WS2 hetero­
0.005–50 ppm NO2, 5 ppm SO2, NH3 and H2S, 100 ppm CO,
structure were elucidated through a combination of multiple experi­
200 ppm H2, or 50 ppb NO2 at various levels of relative humidity (RH),
ments and DFT calculations. This work offers a facile synthesis of dual-
and the real-time resistances (Ra and Rg) in air and target gases were
TMDs heterostructure, and highlights the outstanding room-
recorded. All the operating temperature is 25 ◦ C (room temperature).
temperature gas sensing performance.
The gas responses of sensors were calculated by (Ra-Rg)/Rg for SO2 and
NO2 gases, while those for CO, NH3, H2S, H2 gases, acetone, ethyl
2. Experimental and computational methods
alcohol and water vapor were calculated by (Rg-Ra)/Ra. The response
and recovery times were defined as the ones required for 90% change in
2.1. Synthesis of MoS2/WS2 composites
sensor resistances in the target gas and air, respectively. All sample gases
are dry and are custom-made from the Nanjing Special Gas Co., Ltd, and
MoS2/WS2 composites were prepared by a one-pot hydrothermal
the purities of all these gases are 99.9%. The whole detection process can
route. 0.8 g Na2WO4⋅2 H2O, 0.63 g C2H2O4, and 0.16, 0.33 or 0.66 g
be found in our previous work [24]. Some samples were repeatedly
Na2MoO4⋅2 H2O (with purities of ≥99.5%) were dissolved in deionized
tested 3–5 times to analyze the reproducibility of MoS2/WS2 composites
water (30 mL) in turn, and uninterruptedly stirred by a magnetic bar for
towards various concentrations of NO2. The arithmetical averages were
40 min to obtain homogeneous solutions. Next, the 0.1 M HCl solution
adopted to calculate the mean response & recovery times and responses,
dripped into the above solutions to control the PH value of 1.2–2.
and the deviations were determined by the standard deviation method.
Meanwhile, according to the added amounts of Na2MoO4⋅2 H2O, 0.83,
0.99 and 1.25 g CH4N2S (≥99%) were dissolved in the above solutions
by magnetically stirring for 30 min, respectively. Subsequently, the 2.3. Computational method
resulting solutions were put into 50 mL Teflon-lined stainless steel au­
toclaves, and maintained at 140–230 ◦ C for 4–24 h in a drying oven. In this work, the density functional theory (DFT) based on the Vienna
Ab initio Simulation Package (VASP) [25] was used for calculations of

Fig. 1. Schematic of fabrication of self-assembled MoS2/WS2 nanoflowers and gas sensor.

2
Z. Liang et al. Sensors and Actuators: B. Chemical 403 (2024) 135215

adsorption energy, electrons transfer and binding energy. The electron 3. Results and discussion
correlation was described by generalized gradient approximation with
the Perdew-Burke-Ernzerhof functional [26]. The projector augmented 3.1. The growth of MoS2/WS2 composite
wave method was employed to treat the ion–electron interaction [27]. A
vacuum space of over 15 Å was adopted to avoid the interactions be­ To reveal the growth mechanism of the MoS2/WS2 composite, the
tween two periodic layers. Moreover, the DFT-D3 was applied due to the SEM and corresponding EDS can be used to analyze the microstructures
van der Waals (vdW) interactions [28]. The cut-off energy of 500 eV and and chemical compositions of the related products under different hy­
the Brillouin Zone of 3 × 3 × 1 grid were set to calculate the electronic drothermal temperatures and reaction time. When the molar ratio of Mo
properties and total energies. The adsorption energy (ΔE) was calculated to W is fixed to 0.5 in raw materials, the needle-like WO3 and agglom­
from Eq. (1) [29,30]: erated MoS2 nanoparticles can be observed under 140 ◦ C× 12 h condi­
tion (Fig. 2a and S1a). When the temperature rises to 170 ◦ C, the needle-
ΔE = Etotal − Esurface − Egas (1)
like WO3 crystals grow up and transform into rod-like WO3, and
meanwhile the MoS2 nanoparticles transform into pleated MoS2 nano­
where Etotal is the total energy of surface with different adsorption sites,
sheets (Fig. 2b and S1c). Indeed, the WS2 still fails to nucleate under
Esurface is the energy of substrate slab, and the Egas is the energy of
170 ◦ C× 12 h condition and the partial WO3 can turn into WS2 as the
adsorbent in vacuum. We considered different adsorption sites and gas
temperature climbs to 200 ◦ C (Fig. S1b). So, the MoS2/WS2 nanosheets
orientations (including transverse and longitudinal orientations, as well
start to form with coexisting of a little WO3 while employing 200 ◦ C
as the positions of the atoms of different gas molecules relative to the
(Figs. 2c and S1d). When the temperature increases further to 230 ◦ C,
adsorption sites) during our calculations.
the MoS2/WS2 nanosheets can form and develop gradually into MoS2/
WS2 nanoflowers with the increase of reaction time (Fig. 2d–e). Actu­
ally, a little WO3 still remains under 230 ◦ C× 4 h condition ((Fig. S1e).
Moreover, the petals of MoS2/WS2 nanoflowers gradually become small
with the molar ratio of Mo to W increasing from 0.25 to 1 (Fig. 2f–h),

Fig. 2. SEM images of MoS2/WS2-0.5 composite after hydrothermal processes at: (a) 140 ◦ C× 12 h, (b) 170 ◦ C× 12 h, (c) 200 ◦ C× 12 h; (d) 230 ◦ C× 4 h, (e)
230 ◦ C× 12 h, and (f) 230 ◦ C× 24 h; SEM images of MoS2/WS2-0.25 MoS2/WS2-1 composites after hydrothermal process at 230 ◦ C× 24 h; XRD patterns of pristine
MoS2, WS2, and MoS2/WS2 composites.

3
Z. Liang et al. Sensors and Actuators: B. Chemical 403 (2024) 135215

and they are greatly smaller than those of pristine MoS2 and WS2 3.2. TEM and chemical analyses
(Fig. S2a and b).
As shown in Fig. 2i, the characteristic peaks of WS2 (JCPDS 08–0237) Fig. 4a shows the typical TEM image of MoS2/WS2-0.5 composite
can be observed in XRD patterns, and the distinct peaks of MoS2 cor­ fabricated at 230 ◦ C × 24 h, indicating the nanoflower-like morphology
responding to the (006) and (201) planes (JCPDS 37–1492) can also be with diameter of 1 µm. Clear MoS2/WS2 interface and symbiotic struc­
observed [31], indicating successful synthesis of MoS2/WS2 composites ture are produced, which can be demonstrated by the HRETEM images
by one-pot hydrothermal process [32]. Moreover, the corresponding and the corresponding Fast Fourier Transform (FFT) images [33], indi­
EDS spectra indicate that the atomic ratios of Mo/W in the three cating the formation of MoS2/WS2 heterostructure (Fig. 4b and d–f).
MoS2/WS2 composites are in good agreement with the normal molar Specifically, the lattice spacings of 6.18 nm, 0.273 nm, and 6.15 nm are
ratios, respectively (Fig. S2c and d). assigned to the (002), (100) planes of WS2, and (002) plane of MoS2,
According to the above results, the schematic of evolution of respectively [34–36]. Moreover, the clear diffraction rings are observed
morphology and phase in the MoS2/WS2 system with the hydrothermal in the SAED pattern, corresponding to the (002), (100), (101), (103)
temperature can be roughly deduced (Fig. 3a). At 140 ◦ C, the needle-like planes of WS2, and (100), (101), (103) and (201) planes of MoS2
WO3 and MoS2 nanoparticles coexist, and then they turn into nanorods (Fig. 4c). Actually, the WS2 is mainly distributed at the edge of petals
and nanosheets with the temperature increasing to 170 ◦ C, respectively. instead of the center of the petals, and excess WS2 may grow on the MoS2
The self-assembled MoS2/WS2 nanosheets start to form when the tem­ surface [37,38]. The EDS elemental mappings of MoS2/WS2-0.5 com­
perature is raised to 200 ◦ C, and then evolve into nanoflowers with the posite also demonstrate that the Mo element is mainly distributed at the
temperature further increasing to 230 ◦ C. center of petals and the W element is uniformly distributed on the whole
To further understand the grow mechanism of MoS2/WS2 composite, flower (Fig. 4g), implying that MoS2 can act as the nucleation center of
the binding energies of WS2 molecules deposited on the MoS2 and WS2 WS2 to form the MoS2/WS2 heterostructure in the hydrothermal process.
edges were calculated by DFT. When the temperature rises to 200 ◦ C Raman spectra can effectively characterize the molecular structure of
(while MoS2 has already nucleated), the WS2 molecules can be deposited materials, especially the TMDs. The two vibrational bands at 351 and
on the edge of MoS2 firstly due to too high relative binding energy on the 417 cm− 1 can be observed in pristine WS2, and the other two at 384 and
surface (8.47 eV) (Fig. S3), and then grow laterally to form the MoS2/ 408 cm− 1 are assigned to the pristine MoS2, corresponding to the similar
WS2 nanosheets due to the lower relative binding energy (− 1.2 eV) vibration modes: in-plane optical (E12g ) and out-of-plane vibration (A1g ),
(Fig. 3b). When the lateral growth of WS2 arrives at a certain point, the respectively (Fig. 5a) [39–41]. The characteristic peaks of WS2 and MoS2
WS2 molecules can be deposit on the surfaces of MoS2/WS2 nanosheets can be observed in the MoS2/WS2 composites, and they have a slight red
via van der waals force due to over high specific surface energy at the or blue shift compared with the pristine WS2 and MoS2, implying the
edge and insufficient driving force (Fig. 3c). As a result, the MoS2/WS2 formation of heterostructure. Moreover, with the increase of Mo con­
nanosheets can grow by layer-by-layer self-assembly to form the flower- centration in the MoS2/WS2 composites, the intensity of WS2 charac­
like MoS2/WS2 heterostructured composite. teristic peaks gradually decreases and the offset level of these peaks

Fig. 3. (a) Schematic of growth of MoS2/WS2 composite at different temperatures; (b) The relative binding energies of WS2 molecules deposited on the MoS2 and
WS2 edges; (c) Schematic of layer-by-layer growth of WS2 on the MoS2 and WS2 surfaces via van der waals force.

4
Z. Liang et al. Sensors and Actuators: B. Chemical 403 (2024) 135215

Fig. 4. (a) TEM, (b) HRTEM, (c) SAED and (g) EDS elemental mapping of MoS2/WS2-0.5 composite; (d–f) HRTEM and FFT images of the select regions “d–f” in (b),
respectively.

increases, indicating the enhanced resonance between WS2 and MoS2. It characteristic peaks can be observed due to the presence of W-S bonds in
is worth noting that the slight red shift of A1g of MoS2 only appears in the the WS2-containing materials (Fig. 5c) [43]. The peaks of W6+ can be
MoS2/WS2-1 composite, which may contribute to a small amount of observed due to the oxidation of pristine WS2 and MoS2/WS2-0.5 com­
covering of WS2 on the surface of MoS2/WS2 nanosheets [40]. The blue posite. Obviously, the two peaks of W4f in the MoS2/WS2-0.5 composite
shift of A1g of MoS2 can be observed in the other two composites may be slightly shift towards higher binding energy compared with the pristine
due to the partial encapsulation of MoS2 by the WS2. WS2, indicating that the electrons can flow away from the WS2. The two
The photoluminescence (PL) emission spectra of pristine MoS2, WS2 characteristic peaks at 323.40 and 229.29 eV are observed in the
and MoS2/WS2 composites were collected in the 450–800 nm wave­ MoS2/WS2-0.5 composite, corresponding to Mo 3d3/2 and Mo 3d5/2,
length region at the excitation wavelength of 360 nm to study the sep­ respectively (Fig. 5d) [31,44]. The two peaks shift towards lower
aration rate of electron-hole pairs (Fig. 5b). It is clear that the relative binding energy compared with the pristine MoS2, indicating that the
intensities of the emission spectra of the two composites with higher Mo electrons can flow to the MoS2. The marked deviation of binding en­
concentrations are greatly lower than those of pristine WS2 and MoS2, ergies of W 4 f and Mo 3d indicates strong interactions between MoS2
indicating that the photoexcited holes and electrons in the two com­ and WS2 in the MoS2/WS2-0.5 composite. In particular, the shifts to­
posites are difficultly recombined [20,42]. The remarkable quenching wards the higher and lower binding energy are originated from the
effect indicates that the unique MoS2/WS2 heterostructure can effec­ decrease and increase of the electron density, implying that WS2 and
tively suppress the recombination of electrons and holes, which is MoS2 act as the electron donor and acceptor in the composite as a whole,
favorable to the efficient interfacial electrons transfer between WS2 and respectively [44,45]. In addition, the specific surface area of
MoS2. However, the relative intensity of MoS2/WS2-0.25 composite is MoS2/WS2-0.5 composite is calculated to be 11.59 m2/g based on the N2
higher than those of pristine WS2 and MoS2, which may be attributed to adsorption-desorption isotherm curves (Fig. S4c), which is much larger
the narrower bandgap. than that of WS2 nanosheets (<8.4 m2/g) [46]. The pore sizes of the
To further understand the chemical states and electrons transfer MoS2/WS2-0.5 composite are mainly between 2 to 30 nm (Fig. S4d),
between MoS2 and WS2, the XPS analysis was carried out. The survey showing the mesoporous characteristic.
XPS spectrum of MoS2/WS2-0.5 composite demonstrates that it is
composed of Mo, S, and W elements (Fig. S4a). For the W 4 f, the two

5
Z. Liang et al. Sensors and Actuators: B. Chemical 403 (2024) 135215

Fig. 5. (a) Raman and (b) PL spectra of the MoS2/WS2 composites, pristine MoS2 and WS2; XPS spectra of (c) W4f and (d) Mo 3d of MoS2/WS2-0.5 composite and
pristine WS2 or MoS2.

3.3. Gas sensing performance MoS2/WS2-0.5 composite increases almost linearly with the concentra­
tion of NO2 increasing from 5 to 100 ppb, with R2 of 0.987 (Fig. 6c and
The room-temperature gas sensing performance of pristine MoS2, d), indicating the good potential in practical applications. Specifically,
WS2, and the three MoS2/WS2 composite samples were recorded by the the average responses of MoS2/WS2-0.5 composite to 100, 50, 10 and 5
dynamic gas sensor test system (Fig. 6 and S5). From the response and ppb are 23.1%, 16.1%, 7.3% and 6.1%, respectively.
recovery curves, the resistances of pristine MoS2 and MoS2/WS2 com­ We also investigated the responses of MoS2/WS2-0.5 at various RH
posites decrease at different levels when these gas sensors contact levels to 50 ppb NO2. The response of MoS2/WS2-0.5 composite de­
various concentrations of NO2, showing the characteristic of p-type creases gradually from 16.1% to 13.2%, 10.3%, 6.7%, and 3.1% with the
semiconductor (Fig. 6a). The responses and response time of pristine RH increasing from 0, to 25%, 33%, 50% and 65% (Fig. 6e inset). Under
MoS2 and MoS2/WS2 composites increase and decrease with the increase the RH of ≤ 33%, the MoS2/WS2-0.5 composite shows the good recov­
of NO2 concentration, respectively (Fig. 6a and b). Indeed, the room- ery, with the response of over 60% initial value. When the RH is lower
temperature responses of pristine MoS2 and WS2 are going to zero (≤25%), a small quantity of H2O molecules are adsorbed on the MoS2/
when the NO2 concentration is less than 1 or 2 ppm (Fig. 6a and S5a). WS2 composite surface and can react with the adsorbed oxygen species
The outstanding gas sensing performance is achieved in the MoS2/WS2- via Eq. (2):
0.5 composite, with responses of 31.4%, 140.1%, 185.1% and 368.2% to
1, 5, 10 and 50 ppm NO2, respectively. These responses are ~8–15 and H2O-gas + O-ads → 2-OHads + 2e- (2)
13–23 times higher than those of pristine MoS2 and WS2, respectively which can provide the less electrons to the MoS2/WS2 composite,
(Fig. 6b). According to the Langmuir model [47,48], all the MoS2/WS2 resulting in the lower resistance of material and the shorter response/
composites show good linearity with R2 of over 0.98 (Fig. 6b). Moreover, recovery times [24,49,50]. When the NO2 molecules enter into the gas
the MoS2/WS2-0.5 composite presents the shortest response and recov­ chamber, the NO2 molecules are adsorbed on the MoS2/WS2 composite
ery times to for 5 ppm NO2 (Fig. S5b and c), showing the higher sensi­ surface, and draw electrons from the sensing material and react with
tivity than the other four materials. -OH via Eq. (3) [51,52]:
Low-concentration detection of hazardous gases at room tempera­
ture contributes to human health protection and energy conservation, so NO2 + -OHads + e- → HNO-3 (3)
the room-temperature sensing performance of MoS2/WS2-0.5 composite
When the RH increases, lots of H2O molecules are adsorbed on the
to ppb-level NO2 was deliberately investigated here. The response of
MoS2/WS2 composite surface and occupy more adsorption sites, and

6
Z. Liang et al. Sensors and Actuators: B. Chemical 403 (2024) 135215

Fig. 6. (a,c) Dynamic response and recovery curves, and (b,d) the corresponding fitting curves of pristine MoS2, WS2, and/or MoS2/WS2 composites to 1–50 ppm or
5–100 ppb NO2; (e) Humidity-dependence to 50 ppb NO2 (The inset is the fitting curves of the composite response and baseline resistance under various RH) and (f)
reproducibility of MoS2/WS2-0.5 composite; (g) long-term stability and (h) Selectivity of pristine MoS2, WS2 and MoS2/WS2-0.5 composite; (i) Comparison of room-
temperature responses of MoS2- and WS2-containing materials to ppb-level NO2.

meanwhile more NO2 can react with the -OH, as a result of that the 38.5%, respectively (Fig. 6 g). However, no significant decrease can be
adsorption of NO2 on the composite surface is reduced and the in­ observed in the response of MoS2/WS2-0.5 composite to ppb-level NO2.
teractions between NO2 and MoS2/WS2-0.5 composite are weakened. Under long-term exposure to air, O2 can oxidize slowly the WS2-con­
Hence, the response/recovery times increase, and the response de­ taining materials and partially occupy the sulfur vacancies [53], leading
creases with incomplete recovery. Moreover, the standard deviation of to a certain decrease in the signal of EPR and the response of composite
response of MoS2/WS2 composite at various levels of RH ranges from (Fig. S5i). In particular, a little decrease of 3.4% in the response appears
0.6% to 3.1%, indicating good reproducibility (Fig. S5d). to 5 ppm NO2, indicating that the MoS2/WS2-0.5 composite has excel­
Good reproducibility, long-term stability and selectivity are the basic lent long-term stability (Fig. 6 g).
requirements for gas sensors in order to precisely monitor the target gas The room-temperature responses of pristine MoS2, WS2 and MoS2/
and prevent the interference from other gases during the whole sensing WS2-0.5 composite to 5 ppm NO2 and other hazardous gases in same or
process. Figs. 6 f and S5d–f show the room-temperature reproducibility higher concentrations (5–100 ppm) were comparatively investigated to
of MoS2/WS2-0.5 composite to 5 ppb, 100 ppb and 5 ppm NO2 and of identify the selectivity of the MoS2/WS2-0.5 composite. Obviously, a
pristine WS2 and MoS2 to 5 ppm NO2. Compared with the pristine WS2 higher response of 140.1% to 5 ppm NO2 can be obtained in the MoS2/
and MoS2, no obvious variations are generated in the response and re­ WS2-0.5 composite, which is 14 times higher than those of pristine MoS2
covery process of MoS2/WS2-0.5 composite during 5 cycles and WS2, and meanwhile the MoS2/WS2-0.5 composite presents much
(Fig. S5d–f), with respective responses of 140.1% ± 0.57% (0.41%), lower responses to other gases (H2, Co, H2S, NH3, SO2, acetone, ethanol
23.1% ± 0.09% (0.39%) and 6.1% ± 0.04% (0.65%) (Fig. 6 f), indi­ and water vapor), indicating excellent selectivity (Fig. 6 h).
cating excellent reproducibility. After exposure in air for 30 days, the The responses of related WS2- and MoS2-containing materials are
responses of pristine WS2 and MoS2 present decrease by 31.8% and collected in this work to compare their gas sensing properties to ppb-

7
Z. Liang et al. Sensors and Actuators: B. Chemical 403 (2024) 135215

Fig. 7. (a) The dynamic response change curves (d(Ra/Rg)/dt) of different materials (including MoS2/WS2 composites, pristine WS2 and MoS2) to 5 ppm NO2 at
25 ◦ C; (b) The dynamic response change curves (d(Ra/Rg)/dt) of MoS2/WS2-0.5 composite to 5–100 ppb NO2 at 25 ◦ C.

Fig. 8. (a) The dynamic curve of MoS2/WS2-0.5 composite baseline resistance in air around room temperature (20–40 ◦ C); (b) The fitting curves of MoS2/WS2-0.5
composite baseline resistance in air around room temperature (20–40 ◦ C) and of MoS2/WS2-0.5 composite resistance to 5–100 ppb NO2 at 25 ◦ C; (c) The dynamic
curve of MoS2/WS2-0.5 composite baseline resistance to the concentration of NO2 under 25% RH at 25 ◦ C; (d) The fitting curves of MoS2/WS2-0.5 composite baseline
resistance under various RH and of MoS2/WS2-0.5 composite resistance to ppb-level NO2 (20–100 ppb) under 25% RH at 25 ◦ C.

8
Z. Liang et al. Sensors and Actuators: B. Chemical 403 (2024) 135215

level NO2 (Fig. 6i) [22–24,54–60]. Obviously, the responses of can detect over 50 ppb NO2 in view of the changes of temperature
MoS2/WS2-0.5 composite are much higher than those of MoS2- and WS2- (variation of ~10 ◦ C in a day) and RH on practical application.
containing materials to same concentrations of NO2 at room tempera­
ture, respectively. Moreover, room-temperature responses of some WS2-
3.4. Gas sensing mechanism
and MoS2- containing materials to NO2 are also exhibited in Table S1,
indicating that the as-fabricated MoS2/WS2 composite has excellent NO2
For further reveal gas sensing mechanism of MoS2/WS2 composite,
sensing performance.
the DFT was used to calculate the related adsorption energies and
Obviously, the rate of response change (d(Ra/Rg)/dt, i.e., the relative
electrons transfer, and the related structure models are shown in Figs. S7
adsorption rate) sharply decreases firstly and then flattens out with the
and S8. The excellent room-temperature sensing performance of MoS2/
time increasing. The MoS2/WS2-0.5 composite has higher d(Ra/Rg)/dt
WS2-0.5 composite to NO2 can be mainly ascribed to the following:
than other MoS2/WS2 composites, pristine WS2 and MoS2 to 5 ppm NO2
at 25 ◦ C (Fig. 7a). Moreover, the d(Ra/Rg)/dt of MoS2/WS2-0.5 com­
(1) The special morphology and microstructure. Nanoflower-like
posite slightly climbs with the NO2 concentration increasing from 5 to
MoS2/WS2 composite provides more adsorption sites to the
100 ppb NO2 (Fig. 7b), and the d(Ra/Rg)/dt of MoS2/WS2-0.5 composite
target gas due to the relatively large specific surface area [63–65]
to 5 ppm NO2 is far higher than those to ppb-level NO2. Therefore, the
(Fig. S4c). From the EPR spectrum of MoS2/WS2-0.5 composite
relative adsorption rate is positively related to the number of active sites
(Fig. 9a), the clear paramagnetic transition at g= 2.003 is
and the target gas concentration, which is good agreement with the
observed due to the formation of unpaired electrons originated
previous reports [61,62].
from the sulfur vacancies [66]. The sulfur vacancies can be usu­
Considering the practical application of MoS2/WS2-0.5 composite to
ally produced on the WS2 or MoS2 in hydrothermal process [67,
ppb-level NO2 at room temperature, the effects of working temperature
68]. Compared with the pristine WS2, MoS2 and the air-aged
(around 25 ◦ C) and RH cannot be ignored. Obviously, the baseline
MoS2/WS2-0.5 composite for 30 days, the MoS2/WS2-0.5 com­
resistance of MoS2/WS2-0.5 composite in air decreases lineally with the
posite presents the larger peak area and higher peak intensity
increase of temperature, and the composite resistance to NO2 decreases
(Figs. S5i and S9a). Actually, the larger peak area and higher peak
almost lineally with the concentration of ppb-level NO2 at 25 ◦ C (Fig. 8a
intensity indicates higher sulfur vacancy concentration. The
and b). Moreover, the effect of temperature on the composite response is
presence of abundant sulfur vacancies can provide more multiple
calculated as 3.41 ppb/◦ C (Fig. 8b), namely that the change of 1 ◦ C
active sites to adsorb gas molecules and thus increase the
temperature on the response of MoS2/WS2-0.5 composite is equivalent
adsorption ability of MoS2/WS2 composites to NO2. According to
to that of the adsorption of 3.41 ppb NO2. Similarly, the baseline
the DFT calculations, the MoS2/WS2 composite with S vacancy
resistance of composite increases lineally with the increase of RH, and
has the highest surface adsorption energy (− 0.34 eV) to NO2
the composite resistance to NO2 decreases lineally with the concentra­
(Fig. 9e).
tion of ppb-level NO2 increasing under 25% RH (Figs. 6e, 8c and d). The
Actually, the edge of MoS2/WS2 composite also increases the
− 8.47 ppb/%RH is obtained from these data (Fig. 8d), indicating that
adsorption ability to gases. The MoS2/WS2-0.5 composite has
the effect of 1% RH change on the composite response is equivalent to
more active W sites at the edge due to the presence of more S
that of desorption of 8.47 ppb NO2. In short, the MoS2/WS2 composite
vacancies than pristine MoS2 and WS2 (Fig. S9a), and its W site

Fig. 9. (a) EPR spectrum of MoS2/WS2-0.5 composite and the corresponding structure model; (b) Energy banding diagram of WS2 and MoS2; (c) Schematic of band
structure model for MoS2/WS2 heterostructure in air and NO2; (d) Charge density difference and planar-average-charge density of MoS2/WS2 heterostructure with S
vacancy to NO2; (e) The charges transfer and the adsorption energies at the surfaces of MoS2, WS2, MoS2/WS2 and MoS2/WS2 with S vacancy; (f) Schematic of
electrons transfer models of MoS2/WS2 composite in air and NO2.

9
Z. Liang et al. Sensors and Actuators: B. Chemical 403 (2024) 135215

has the highest adsorption energy (− 3.93 eV) (Fig. S9b), result­ decrease of the conductivity of p-type MoS2/WS2 composite (i.e., the
ing in the maximum adsorption energy at the edge. Moreover, the increase of the resistance) [53,73]. However, the adsorption energies of
NO2 can gain the more electrons (0.97 e) from the W site at the the reducing gases to WS2 or MoS2 are lower than that of the NO2 [74,
edge of MoS2/WS2 composite to form the W-O bonds, and 75], so the responses of MoS2/WS2 composite to other reducing gases
meanwhile the dangling S atoms at the edge of MoS2/WS2 com­ are lower than that to NO2.
posite can provide the more electrons (1.16 e) to NO2 to form S-O
bonds (Figs. S7 and S9b), indicating the higher electron transfer 4. Conclusion
ability at the edge. However, the edge area of MoS2/WS2
composite-0.5 is much smaller than its surface area, so the Nanoflower-like MoS2/WS2 composites were synthesized by a facile
adsorption ability and electrons transfer ability at the surface of one-pot hydrothermal method. With increasing the hydrothermal tem­
MoS2/WS2-0.5 composite mainly determine the NO2 sensing perature from 140 to 230 ◦ C, the needle-like WO3 and MoS2 nano­
performance. particles turn into nanorods and nanosheets firstly, respectively, then
(2) The effect of MoS2/WS2 heterostructure. Based on the UV-Vis they are self-assembled into MoS2/WS2 nanosheets, and finally evolve
absorption spectra of pristine MoS2, WS2, and MoS2/WS2-0.5 into nanoflowers. The WS2 molecules prefer to grow on the edges of as-
composites, their bandgaps are calculated to be 1.30, 1.03 and nucleated MoS2 to form the MoS2/WS2 nanosheets, and then they can
1.10 eV, respectively (Fig. S10a and S10b). The valence band grow by layer-by-layer self-assembly to form the MoS2/WS2 hetero­
(VB) edges of pristine MoS2 and WS2 measured by valence-band structure. The MoS2/WS2-0.5 composite presents very high sensitivity to
XPS (VB-XPS) spectra are determined to be 0.31 and 0.14 eV different concentrations of NO2 at room temperature. Specifically, the
below their Fermi levels, respectively (Fig. S10c) [69]. Moreover, responses of MoS2/WS2-0.5 composite to 5 ppb, 100 ppb and 5 ppm NO2
the work function (Φ) of MoS2 and WS2 can be determined to be are 6.1%, 23.1% and 140.1%, respectively. Moreover, the MoS2/WS2-
5.2 and 4.9 eV according to these reports [70,71]. So, the energy 0.5 composite exhibits excellent reproducibility, selectivity, and long-
band structure can be described in Fig. 9b. term stability to NO2. The excellent NO2 sensing performance of
MoS2/WS2-0.5 composite can be mainly attributed to the flower-like
The electrons can transfer from WS2 to MoS2 due to the lower work morphology, abundant sulfur vacancies and self-assembled hetero­
function, resulting in the separation of holes and electrons, and the structure. This work provides a feasible strategy to fabricate self-
formation of charge depletion and accumulation layers in air (Fig. 9c, assembled dual-TMDs heterostructures for high-performance gas
d and f). So, the built-in electrical fields are established [72], which can sensing materials.
provide higher potential barrier and then decrease the conductivity of
MoS2/WS2 composite (Fig. 9c). When in air, the O2 molecules are CRediT authorship contribution statement
absorbed on the composite surface, and then draw electron to form O-2.
The relevant reactions in air are described by Eqs. 4 and 5. When in NO2, Zhiping Liang: Investigation, Methodology, Visualization, Writing
partial NO2 molecules can prefer to be adsorbed on the edges of − original draft. Mingyuan Wang: Formal analysis, Software, Writing −
MoS2/WS2 nanoflowers due to the higher adsorption energy at the edge original draft, Visualization. Siwei Liu: Investigation. Mobashar Has­
(Fig. 9e), while most of the NO2 molecules are adsorbed on the san: Methodology, Funding acquisition. Xiangzhao Zhang: Conceptu­
MoS2/WS2 composite surface due to the larger surface area. And the alization, Visualization. Shuangying Lei: :Visualization. Guanjun
small amount of NO2 can react with the O−2 on the composite surface to Qiao: Resource, Funding acquisition. Guiwu Liu: Formal analysis,
form NO−2 or NO−3 . The relevant reactions in NO2 are described by Eqs. Resource, Supervision, Writing − review & editing.
(6)–(9). Moreover, the NO2 can draw the more electrons (0.2 e) from
WS2 or MoS2 in the MoS2/WS2 composite with S vacancy (Fig. 9e), Declaration of Competing Interest
which can break the balance of build-in electrical field. The decrease of
electrons can narrow down the height of potential barrier, and increase The authors declare that they have no known competing financial
the free-moving holes in the p-type MoS2/WS2 composite, leading to the interests or personal relationships that could have appeared to influence
enhancement of the conductivity of MoS2/WS2 composite. The the work reported in this paper.
MoS2/WS2 composite with S vacancies can provide the most electrons to
NO2 whatever at the surface and at the edge (Fig. 9e and S9b), leading to Data availability
the biggest drop in resistance (Fig. 6a and c). Therefore, the
MoS2/WS2-0.5 composite exhibits the optimal NO2 sensing performance No data was used for the research described in the article.
at room temperature.
In air: Acknowledgments
O2 (g)→O2 (ad) (4)
This work was supported by National Natural Science Foundation
O2 (ad) + e− →O−2 (ad) (5) (51950410596) of China, and the Key Research and Development Plan
(BE2019094), Qing Lan Project ([2016]15), Innovation/Entrepreneur­
In NO2:
ship Program (JSSCTD202146) of Jiangsu Province. We thank the Hefei
NO2 (g)→NO2 (ad) (6) Advanced Computing Center for the facility support on the calculation
in this paper.
NO2 (ad) + e− →NO−2 (ad) (7)
Appendix A. Supporting information
NO2 (ad) + O−2 →NO−2 (ad) + O2 (8)
Supplementary data associated with this article can be found in the
2NO2 (ad) + O−2 + e− →2NO−3 (ad) (9) online version at doi:10.1016/j.snb.2023.135215.
For reducing gases (CO, H2S, NH3 and H2), the mechanism of gas
sensing can be mainly explained in terms of electron transfer and gas References
adsorption. In a dry atmosphere, they can be adsorbed on the composite
[1] S.W. Liu, M.S. Wang, C.X. Ge, S.Y. Lei, S. Hussain, M.S. Wang, G.J. Qiao, G.W. Liu,
surfaces and provide some electrons to the composite, resulting in the Enhanced room-temperature NO2 sensing performance of SnO2/Ti3C2 composite

10
Z. Liang et al. Sensors and Actuators: B. Chemical 403 (2024) 135215

with double heterojunctions by controlling co-exposed {221} and {110} facets of [29] M. Wang, R. Song, Q. Zhang, C. Li, Z. Xu, G. Liu, N. Wan, S. Lei, Synergy effect of
SnO2, Sens. Actuators B Chem. 365 (2022), 131919. Cu-Ru dual atoms anchored to N-doped phosphorene for nitrogen reduction
[2] W. Chen, H.Z. Tang, H.M. Zhao, Urban air quality evaluations under two versions reaction, Fuel 321 (2022), 124101.
of the national ambient air quality standards of China, Atmos. Pollut. Res. 7 (1) [30] H. Yang, Y. Wu, G. Li, Q. Lin, Q. Hu, Q. Zhang, J. Liu, C. He, Scalable production of
(2016) 49–57. efficient single-atom copper decorated carbon membranes for CO2 electroreduction
[3] Y. Liu, H. Ding, S.T. Chang, R. Lu, H. Zhong, N. Zhao, T. Lin, Y. Bao, L. Yap, W. Xu, to methanol, J. Am. Chem. Soc. 141 (2019) 12717–12723.
M. Wang, Y. Li, S. Qin, Y. Zhao, X. Geng, S. Wang, E. Chen, Z. Yu, T. Chan, S. Liu, [31] Y. Rao, J. Wang, P. Liang, H. Zheng, M. Wu, J. Chen, F. Shi, K. Yan, J. Liu, K. Bian,
Exposure to air pollution and scarlet fever resurgence in China: a six-year C. Zhang, K. Zhu, Heterostructured WS2/MoS2@ carbon hollow microspheres
surveillance study, Nat. Commun. 11 (1) (2020) 4229. anchored on graphene for high-performance Li/Na storage, Chem. Eng. J. 443
[4] A. Maity, S.B. Majumder, NO2 sensing and selectivity characteristics of tungsten (2022), 136080.
oxide thin films, Sens. Actuators B Chem. 206 (2015) 423–429. [32] Y. Zhao, J. Liu, X. Zhang, C. Wang, X. Zhao, J. Li, H. Jin, Convenient synthesis of
[5] Q. Li, Y. Cen, J.Y. Huang, X.J. Li, H. Zhang, Y.F. Geng, B.I. Yakobson, Y. Du, X. WS2–MoS2 heterostructures with enhanced photocatalytic performance, J. Phys.
Q. Tian, Zinc oxide–black phosphorus composites for ultrasensitive nitrogen Chem. C. 123 (45) (2019), 273633− 27368.
dioxide sensing, Nanoscale Horiz. 3 (5) (2018) 525–531. [33] G.J. Lai, L.M. Lyu, Y.S. Huang, G.C. Lee, M.P. Lu, T.P. Perng, M.Y. Lu, L.J. Chen,
[6] S.M. Majhi, A. Mirzaei, H.W. Kim, S.S. Kim, T.W. Kim, Recent advances in energy- Few-layer WS2–MoS2 in-plane heterostructures for efficient photocatalytic
saving chemiresistive gas sensors: A review, Nano Energy 79 (2021), 105369. hydrogen evolution, Nano Energy 81 (2021), 105608.
[7] A. Mirzaei, M.H. Lee, H. Safaeian, T.U. Kim, J.Y. Kim, H.W. Kim, S.S. Kim, Room [34] B. Groven, A.N. Mehta, H. Bender, J. Meersschaut, T. Nuytten, P. Verdonck,
temperature chemiresistive gas sensors based on 2D MXenes, Sensors 23 (2023) T. Conard, Q. Smets, T. Schram, B. Schoenaers, A. Stesmans, V. Afanasʼev,
8829. W. Vandervorst, M. Heyns, M. Caymax, I. Radu, A. Delabie, Two-dimensional
[8] M. Hassan, S. Liu, Z. Liang, S. Hussain, J. Liu, G. Liu, G. Qiao, Revisiting traditional crystal grain size tuning in WS2 atomic layer deposition: An insight in the
and modern trends in versatile 2D nanomaterials: Synthetic strategies, structural nucleation mechanism, Chem. Mater. 30 (21) (2018) 7648–7663.
stability, and gas-sensing fundamentals, J. Adv. Ceram. 12 (2023) (2023) [35] L. Cao, X. Liang, X. Ou, X. Yang, Y. Li, C. Yang, Z. Lin, M. Liu, Heterointerface
2149–2247. engineering of hierarchical Bi2S3/MoS2 with self-generated rich phase boundaries
[9] X. Hu, Y.J. Liu, J.W. Li, G.X. Wang, J.X. Chen, G.B. Zhong, H.B. Zhan, Z.H. Wen, for superior sodium storage performance, Adv. Funct. Mater. 30 (16) (2020),
Self-assembling of conductive interlayer-expanded WS2 nanosheets into 3D hollow 1910732.
hierarchical microflower bud hybrids for fast and stable sodium storage, Adv. [36] W. Ding, L. Hu, J. Dai, X. Tang, R. Wei, Z. Sheng, C. Liang, D. Shao, W. Song,
Funct. Mater. 30 (5) (2020), 1907677. Q. Liu, M. Chen, X. Zhu, S. Chou, X. Zhu, Q. Chen, Y. Sun, S.X. Dou, Highly
[10] S.R. Cho, D.H. Kim, M. Jeon, P. Rani, M. Gyeon, Y. Kim, M. Jo, S. Song, J.Y. Park, ambient-stable 1T-MoS2 and 1T-WS2 by hydrothermal synthesis under high
J. Kim, I.D. Kim, K. Kang, Overlaying monolayer metal-organic framework on magnetic fields, ACS Nano 13 (2) (2019), 16943− 1702.
PtSe2-Based gas sensor for tuning selectivity, Adv. Funct. Mater. 32 (47) (2022), [37] L. Wang, Q. Yue, C. Pei, H. Fan, J. Dai, X. Huang, H. Li, W. Huang, Scrolling bilayer
2207265. WS2/MoS2 heterostructures for high-performance photo-detection, Nano, Research
[11] T.S. Kim, K.P. Dhakal, E. Park, G. Noh, H.J. Chai, Y. Kim, S. Oh, M. Kang, J. Park, 13 (2020) 9593–9966.
J. Kim, S. Kim, H.Y. Jeong, S. Bang, J.Y. Kwak, J. Kim, K. Kang, Gas-Phase alkali [38] S. Tongay, W. Fan, J. Kang, J. Park, U. Koldemir, J. Suh, D.S. Narang, K. Liu, J. Ji,
metal-assisted MOCVD growth of 2D transition metal dichalcogenides for large- J. Li, R. Sinclair, J. Wu, Tuning interlayer coupling in large-area heterostructures
scale precise nucleation control, Small 18 (20) (2022), 2106368. with CVD-grown MoS2 and WS2 monolayers, Nano Lett. 14 (6) (2014)
[12] Q. Guang, B. Huang, J. Yu, M. Bonyani, M. Moaddeli, M. Kanani, A. Mirzaei, H. 31853–33190.
W. Kim, S.S. Kim, X. Li, PtS-decorated WS2 microflakes based sensors for selective [39] G.C. Wang, L. Li, W.H. Fan, R.Y. Wang, S.S. Zhou, J.T. Lü, L. Gan, T.Y. Zhai,
ammonia detection at room temperature, Sens. Actuators B Chem. 394 (2023), Interlayer coupling Induced infrared response in WS2/MoS2 heterostructures
134399. enhanced by surface plasmon resonance, Adv. Funct. Mater. 28 (2018) 1800339.
[13] W. Wu, Q. Zhang, X. Zhou, L. Li, J. Su, F. Wang, T. Zhai, Self-powered photovoltaic [40] J. Zhang, J. Wang, P. Chen, Y. Sun, S. Wu, Z. Jia, X. Lu, H. Yu, W. Chen, J. Zhu,
photodetector established on lateral monolayer MoS2-WS2 heterostructures, Nano G. Xie, R. Yang, D. Shi, X. Xu, J. Xiang, K. Liu, G. Zhang, Observation of strong
Energy 51 (2018) 45–53. interlayer coupling in MoS2/WS2 heterostructures, Adv. Mater. 28 (22) (2016),
[14] E. Lee, Y.S. Yoon, D.J. Kim, Two-dimensional transition metal dichalcogenides and 19503− 1956.
metal oxide hybrids for gas sensing, ACS Sens 3 (10) (2018) 2045–2060. [41] A.A. Puretzky, L. Liang, X. Li, K. Xiao, K. Wang, M. Mahjouri-Samani, L. Basile, J.
[15] W. Yang, L. Gan, H.Q. Li, T.Y. Zhai, Two-dimensional layered nanomaterials for C. Idrobo, B.G. Sumpter, V. Meunier, D.B. Geohegan, Low-frequency raman
gas-sensing applications, Inorg. Chem. Front. 3 (4) (2016) 433–451. fingerprints of two-dimensional metal dichalcogenide layer stacking
[16] M. Ikram, L.J. Liu, Y. Liu, L.F. Ma, H. Lv, M. Ullah, L. He, H.Y. Wu, R.H. Wang, K. configurations, ACS Nano 9 (6) (2015) 6333–6342.
Y. Shi, Fabrication and characterization of a high-surface area MoS2@WS2 [42] S. Luo, S. Dong, C. Lu, C. Yu, Y. Ou, L. Luo, J. Sun, J. Sun, Rational and green
heterojunction for the ultra-sensitive NO2 detection at room temperature, J. Mater. synthesis of novel two-dimensional WS2/MoS2 heterojunction via direct exfoliation
Chem. A 7 (24) (2019) 14602–14612. in ethanol-water targeting advanced visible-light-responsive photocatalytic
[17] Y. Kim, S. Lee, J.G. Song, K.Y. Ko, W.J. Woo, S.W. Lee, M. Park, H. Lee, Z. Lee, performance, J. Colloid Interface Sci. 513 (2018), 3893− 399.
H. Choi, W.H. Kim, J. Park, H. Kim, 2D transition metal dichalcogenide [43] R. Scarfiello, E. Mazzotta, D. Altamura, C. Nobile, R. Mastria, S. Rella, C. Giannini,
heterostructures for p- and n-type photovoltaic self-powered gas sensor, Adv. P.D. Cozzoli, A. Rizz, C. Malitesta, An insight into chemistry and structure of
Funct. Mater. 30 (43) (2020), 2003360. colloidal 2D-WS2 nanoflakes: Combined XPS and XRD study, Nanomaterials 11
[18] Y. Gong, J. Lin, X. Wang, G. Shi, S. Lei, Z. Lin, X. Zou, G. Ye, R. Vajtai, B. (2021) 1969.
I. Yakobson, H. Terrones, M. Terrones, B.K. Tay, J. Lou, S.T. Pantelides, Z. Liu, [44] H. Kuang, Z. He, M. Li, R. Huang, Y. Zhang, X. Xu, L. Wang, Y. Chen, S. Zhao,
W. Zhou, P.M. Ajayan, Vertical andin-plane heterostructures from WS2/MoS2 Enhancing co-catalysis of MoS2 for persulfate activation in Fe3+-based advanced
monolayers, Nat. Mater. 13 (12) (2014) 1135–1142. oxidation processes via defect engineering, Chem. Eng. J. 417 (2021), 127987.
[19] J. Zhu, W. Li, R. Huang, L. Ma, H. Sun, J.H. Choi, L. Zhang, One-pot selective [45] J. Tao, M. Wang, X. Zhang, L. Lu, H. Tang, Q. Liu, S. Lei, G. Qiao, G. Liu, A novel
epitaxial growth of large WS2/MoS2 lateral and vertical heterostructures, J. Am. CoP@AAH cocatalyst leads to excellent stability and enhanced photocatalytic H2
Chem. Soc. 142 (38) (2020), 162763− 16284. evolution of CdS by structurally separating the photogenerated carriers, Appl.
[20] V.T. Vu, T.L. Phan, T.T.H. Vu, M.H. Park, V.D. Do, V.Q. Bui, K. Kim, Y.H. Lee, W. Catal. B Environ. 320 (2023), 122004.
J. Yu, Synthesis of a selectively Nb-doped WS2-MoS2 lateral heterostructure for a [46] Y.T. Han, Y. Liu, C. Su, X.W. Chen, B.L. Li, W.K. Jiang, M. Zeng, N.T. Hu, Y.J. Su, Z.
high-detectivity PN photodiode, ACS Nano 16 (8) (2022), 120733− 12082. H. Zhou, Z.G. Zhu, Z. Yang, Hierarchical WS2-WO3 nanohybrids with P-N
[21] Y. Xia, L. Xu, S.F. He, L.X. Zhou, M.J. Wang, J. Wang, S. Komarneni, UV-activated heterojunctions for NO2 detection, ACS Appl. Nano Mater. 4 (2) (2021)
WS2/SnO2 2D/0D heterostructures for fast and reversible NO2 gas sensing at room 1626–1634.
temperature, Sens. Actuators B Chem. 364 (2022), 131903. [47] C.Y. Lee, M.S. Strano, Understanding the dynamics of signal transduction for
[22] Y.S. Xu, J.Y. Xie, Y.F. Zhang, F.H. Tian, C. Yang, W. Zheng, X.H. Liu, J. Zhang, adsorption of gases and vapors on carbon nanotube sensors, Langmuir 21 (2005)
N. Pinna, Edge-enriched WS2 nanosheets on carbon nanofibers boosts NO2 5192.
detection at room temperature, J. Hazard. Mater. 411 (2021), 125120. [48] S. Park, S. Jeon, H. Kim, J. Philips, D. Oh, J. Ahn, M. Kim, C. Park, S. Hong, J. Kim,
[23] Q. Zhao, W. Zhou, M. Zhang, Y. Wang, Z. Duan, C. Tan, B. Liu, F. Ouyang, Z. Yuan, W. Jung, I.D. Kim, Imparting metal oxides with high sensitivity toward light-
H. Tai, Y. Jiang, Edge-enriched Mo2TiC2Tx/MoS2 Heterostructure with Coupling activated NO2 detection via tailored interfacial chemistry, Adv. Funct. Mater. 33
Interface for Selective NO2 Monitoring, Adv. Funct. Mater. 32 (2022) 2203528. (17) (2023), 2214008.
[24] Z.P. Liang, X.Z. Zhang, J. Yang, Y. Cheng, H.H. Hou, S. Hussain, J.L. Liu, G.J. Qiao, [49] W.J. Yan, M.A. Worsley, T. Pham, A. Zettl, C. Carraro, R. Maboudian, Effects of
G.W. Liu, Facile fabrication of nanoflower-like WO3/WS2 heterojunction for highly ambient humidity and temperature on the NO2 sensing characteristics of WS2/
sensitive NO2 detection at room temperature, J. Hazard. Mater. 443 (2023), graphene aerogel, Appl. Surf. Sci. 450 (2018) 372–379.
130316. [50] J.H. Bang, Y.J. Kwon, J.H. Lee, A. Mirzaei, H.Y. Lee, H. Choi, S.S. Kim, Y.K. Jeong,
[25] G. Kresse, J. Furthmuller, Efficient iterative schemes for ab initio total-energy H.W. Kim, Proton-beam engineered surface-point defects for highly sensitive and
calculations using a plane-wave basis set, Phys. Rev. B 54 (1996) 11169–11186. reliable NO2 sensing under humid environments, J. Hazard. Mater. 416 (2021),
[26] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made 125841.
simple, Phys. Rev. Lett. 77 (1996) 3865–3868. [51] Z. Wu, H. Wang, Q. Ding, K. Tao, W. Shi, C. Liu, J. Chen, J. Wu, A self-powered,
[27] G. Kresse, D. Joubert, From ultrasoft pseudopotentials to the projector augmented- rechargeable, and wearable hydrogel patch for wireless gas detection with
wave method, Phys. Rev. B 59 (1999) 1758–1775. extraordinary performance, Adv. Funct. Mater. 33 (2023) 2300046.
[28] S. Grimme, J. Antony, S. Ehrlich, H. Krieg, A consistent and accurate ab Initio [52] X. Chen, J. Hu, P. Chen, M. Yin, F. Meng, Y. Zhang, UV-light-assisted NO2 gas
parametrization of density functional dispersion correction (DFT-D) for the 94 sensor based on WS2/PbS heterostructures with full recoverability and reliable
elements H-Pu, J. Chem. Phys. 132 (2010), 154104. anti-humidity ability, Sens. Actuators B Chem. 339 (2021), 129902.

11
Z. Liang et al. Sensors and Actuators: B. Chemical 403 (2024) 135215

[53] Z. Qin, C. Ouyang, J. Zhang, L. Wan, S. Wang, C. Xie, D. Zeng, 2D WS2 nanosheets Zhiping Liang is a PHD. student at the School of Materials
with TiO2 quantum dots decoration for high-performance ammonia gas sensing at Science and Engineering, Jiangsu University. Her research
room temperature, Sens. Actuators B Chem. 253 (2017) 1034–1042. mainly focuses on the two-dimensional (2D) gas-sensing ma­
[54] Y. Xia, S. He, J.L. Wang, Zhou, J. Wang, S. Komarneni, MXene/WS2 hybrids for terials applied at room temperature.
visible-light-activated NO2 sensing at room temperature, Chem. Commun. 57
(2021) 9136–9139.
[55] W.T. Koo, J.H. Cha, J.W. Jung, S.J. Choi, J.S. Jang, D.H. Kim, I.D. Kim, Few-
layered WS2 nanoplates confined in Co, N-doped hollow carbon nanocages:
Abundant WS2 edges for highly sensitive gas sensors, Adv. Funct. Mater. 28 (2018)
1802575.
[56] J.H. Cha, S.J. Choi, S. Yua, I.D. Kim, 2D WS2-edge functionalized multi-channel
carbon nanofibers: effect of WS2 edge-abundant structure on room temperature
NO2 sensing, J. Mater. Chem. A 5 (2017) 8725–8732.
[57] A. Alagh, F.E. Annanouch, P. Umek, C. Bittencourt, A.S. Castillo, E. Haye, J.
F. Colomer, E. Llobet, CVD growth of self-assembled 2D and 1D WS2 nanomaterials
for the ultrasensitive detection of NO2, Sens. Actuators B Chem. 326 (2021),
Mingyuan Wang is a PHD. student at the School of Electrical
128813.
Science and Engineering, Southeast University. He is currently
[58] T.T. Xu, Y.Y. Liu, Y.Y. Pei, Y.P. Chen, Z.M. Jiang, Z.F. Shi, J.M. Xu, D. Wu, Y.
working on the synthesis, characterization, application and
T. Tian, X.J. Li, The ultra-high NO2 response of ultra-thin WS2 nanosheets
DFT calculation of 2D materials.
synthesized by hydrothermal and calcination processes, Sens. Actuators B Chem.
259 (2018) 789–796.
[59] S.Y. Cho, S.J. Kim, Y. Lee, J.S. Kim, W.B. Jung, H.W. Yoo, J. Kim, H.T. Jung, Highly
enhanced gas adsorption properties in vertically aligned MoS2 layers, ACS Nano 9
(2015) 9314–9321.
[60] Y. Ding, X. Guo, D. Kuang, X. Hu, Y. Zhou, Y. He, Z. Zang, Hollow Cu2O
nanospheres loaded with MoS2/reduced graphene oxide nanosheets for ppb-level
NO2 detection at room temperature, J. Hazard. Mater. 416 (2021), 126218.
[61] U. Kumar, S.M. Huang, Z.Y. Deng, C.X. Yang, W.M. Haung, C.H. Wu, Comparative
DFT dual gas adsorption model of ZnO and Ag/ZnO with experimental applications
as gas detection at ppb level, Nanotechnology 33 (2021), 105502.
[62] C.H. Wu, G.J. Jiang, C.C. Chiu, P. Chong, C.C. Jeng, R.J. Wu, J.H. Chen, Fast gas
concentration sensing by analyzing the rate of resistance change, Sens. Actuators B Siwei Liu is a postdoctoral fellow at the School of Materials
Chem. 209 (2015) 906–910. Science and Engineering, Jiangsu University. She is currently
[63] Y. Zhang, S. Han, M.Y. Wang, S.W. Liu, G.W. Liu, X.F. Meng, Z.W. Xu, M.S. Wang, working on the synthesis of MXene-based nano-materials and
G.J. Qiao, Electrospun Cu-doped In2O3 hollow nanofibers with enhanced H2S gas their application in gas sensors.
sensing performance, J. Adv. Ceram. 11 (2022) 427–442.
[64] S. Hussain, J.N.O. Amu-Darko, M. Wang, A.A. Alothman, M. Ouladsmane, S.
A. Aldossari, M.S. Khan, G. Qiao, G. Liu, CuO-decorated MOF derived ZnO
polyhedral nanostructures for exceptional H2S gas detection, Chemosphere 317
(2023), 137827.
[65] C. Zhang, Y. Huan, Y. Li, Y. Luo, M. Debliquy, Low concentration isopropanol gas
sensing properties of Ag nanoparticles decorated In2O3 hollow spheres, J. Adv.
Ceram. 11 (2022) 379–391.
[66] F. Chen, H. Guo, L. Zhao, X. Guo, S. Li, Y. Chu, X. Wang, Y. Zhu, Defect engineering
can enhance the electrochemical performance of WS2 for thermal batteries,
J. Electrochem. Soc. 168 (2021), 103507.
[67] M. Wang, P. Zhou, L. Bian, K. Cui, J. Bai, M. Hao, J. Liang, B. Fang, F. Wang, Cost-
effective synthesis of highly-dispersed WS2 nanosheets assembled on tourmaline Mobashar Hassan is a postdoctoral fellow at the School of
for improved photocatalytic activity, Appl. Surf. Sci. 639 (2023), 158195. Materials Science and Engineering, Jiangsu University. He
[68] H. Zhang, K. Wang, H. Wang, H. Lin, Y. Zheng, In-plane defect engineering on mainly engaged in the development of new and efficient elec­
MoS2 through a novel two-phase hydrothermal synthesis, Catal. Today 404 (2022) trode nanomaterials and their application in supercapacitors,
269–278. fuel cells, batteries, and gas sensors.
[69] X.H. Wang, L. Lu, B. Wang, Z. Xu, Z.Y. Xin, S.C. Yan, Z.R. Geng, Z.G. Zou,
Frustrated lewis pairs accelerating CO2 reduction on oxyhydroxide photocatalysts
with surface lattice hydroxyls as a solid-state proton donor, Adv. Funct. Mater. 43
(2018) 1804191.
[70] X. Liu, K. Chen, X. Li, Q. Xu, J. Weng, J. Xu, High-performance photovoltaic effect
with electrically balanced charge carriers in black phosphorus and WS2
heterojunction, Adv. Mater. Interfaces 5 (2018) 1800671.
[71] S.Y. Cho, H.J. Koh, H.W. Yoo, J.S. Kim, H.T. Jung, Tunable volatile-organic-
compound sensor by using Au nanoparticle incorporation on MoS2, ACS Sens 2
(2017) 183–189.
[72] M. Zhang, K. Liu, X. Zhang, B. Wang, X. Xu, X. Du, C. Yang, K. Zhang, Interfacial
energy barrier tuning of hierarchical Bi2O3/WO3 heterojunctions for advanced Xiangzhao Zhang is an associate Professor at the School of
triethylamine sensor, J. Adv. Ceram. 11 (2022) 1860–1872. Materials Science and Engineering, Jiangsu University. His
[73] J. Kim, I. Sakaguchi, S. Hishita, T. Ohsawa, T.T. Suzuki, N. Saito, Ru-implanted research includes First principles calculation, inorganic mate­
WS2 nanosheet gas sensors to enhance sensing performance towards CO gas in self- rials and devices, gas-sensing materials and sensors.
heating mode, Sens. Actuators B Chem. 370 (2022), 132454.
[74] Z. Cui, K. Yang, Y. Shen, Z. Yuan, Y. Dong, P. Yuan, E. Li, Toxic gas molecules
adsorbed on intrinsic and defective WS2: gas sensing and detection, Appl. Surf. Sci.
613 (2023), 155978.
[75] S. Zhao, J. Xue, W. Kang, Gas adsorption on MoS2 monolayer from first-principles
calculations, Chem. Phys. Lett. 595 (2014) 35–42.

12
Z. Liang et al. Sensors and Actuators: B. Chemical 403 (2024) 135215

Shuangying Lei is an associate Professor at the School of Guiwu Liu is a Professor of the School of Materials Science and
Electrical Science and Engineering, Southeast University. Her Engineering, Jiangsu University. His current research includes
current research includes nano materials and devices based on wetting and joining of ceramics, inorganic materials and de­
experiment and DFT calculation. vices, such as gas-sensing, thermoelectric, light absorbing
materials and sensors.

Guanjun Qiao is a Professor and Dean of the School of Mate­


rials Science and Engineering, Jiangsu University. His current
research interest involves inorganic materials and devices,
such as design and optimization of advanced ceramics, and gas
sensors.

13

You might also like