You are on page 1of 9

Journal of Alloys and Compounds 807 (2019) 151650

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


journal homepage: http://www.elsevier.com/locate/jalcom

Optimization NH3 sensing performance manifested by gas sensor


based on Pr-SnS2/ZnS hierarchical nanoflowers
Qixian Zhang, Shuyi Ma*, Rong Zhang, Kaiming Zhu, Yue Tie, Shitu Pei
Key Laboratory of Atomic and Molecular Physic & Functional Materials of Gansu Province, College of Physics and Electronic Engineering, Northwest Normal
University, Lanzhou, 730070, China

a r t i c l e i n f o a b s t r a c t

Article history: Pure and Pr doped 3D SnS2/ZnS hierarchical nanoflowers were prepared by two-step hydrothermal
Received 8 April 2019 method. The as-prepared samples were characterized by various techniques. For the Pr-SnS2/ZnS sam-
Received in revised form ples, XPS results confirmed that Pr3þ/Pr4þ ions were successfully doped into the samples. Pr doping led to
15 July 2019
lattice distortion, inhibiting grain growth and reducing grain size. Compared with sensors based on SnS2/
Accepted 30 July 2019
Available online 30 July 2019
ZnS nanoflowers, Pr doped SnS2/ZnS samples exhibited better gas sensing performance to 50 ppm
ammonia gas at 160  C, including higher response (14.03) and ultrafast responding and recovering time
(6 s/13 s). The excellent gas sensing performance may be attributed to more free electrons generated by
Keywords:
Pr-SnS2/ZnS
Pr doping, SnS2/ZnS heterojunction and the high specific surface area of such unique hierarchical flower-
Ammonia like structure. In addition, the gas sensing mechanism of Pr-SnS2/ZnS sensor was also analyzed.
Hydrothermal © 2019 Elsevier B.V. All rights reserved.
Nanoflowers
Sensing performance

1. Introduction sensitive NH3 gas sensor have been reported [10]. In the recent
years, multicomponent heterojunction semiconductor has caused
Ammonia (NH3) is a commonly used chemical agent in our lives. extensive attention. The construction of heterojunction facilitates
In addition, it is harmful to human body and can lead to vomiting, the transmission of electrons and formation of an interfacial
diarrhea, pulmonaryedema [1]. According to the OSHA standard depletion region. It will eventually lead to enhanced sensor per-
issued by the Occupational Safety and Health Administration, the formance [11e13]. Meanwhile, rare earth element doping will
threshold for ammonia in the workplace is 50 ppm [2]. Thus, it is improve the sensing performance of gas sensing materials. For
particularly urgent to monitor low concentrations of ammonia gas example, Tie et al. [14] have reported that the formaldehyde gas
in the environment. In the past few decades, most sensitive ma- sensing of ZTO/SnO2 gas sensor has been significantly improved by
terials based on metal oxides semiconductor have been studied, Y-doped ZTO/SnO2. Chen et al. [15]. have proved that the ethanol
including ZnO [3], SnO2 [4] and WO3 [5] et al. Although they have sensing performance has been enhanced by Pr and Ce co-doped
promoted the development of gas sensors, they still have high ZnSn(OH)6.
optimal operating temperatures and poor selectivity. In this paper, we successfully synthesized Pr doped 3D SnS2/ZnS
Tin sulfide (SnS2) is an IV-VI compound, n-type semiconductor nanoflowers by using two-step hydrothermal method. In particular,
material with a band gap of approximately 2.1 eV. It belongs to a the structure, morphology and specific surface area were investi-
typical layered hexagonal CdI2 type crystal. Its atoms are combined gated. Moreover, the gas sensing performance of pure SnS2/ZnS and
into an “SeSneS” structure by means of covalent bond, and a Pr-doped SnS2/ZnS materials were studied. Compared with pure
layered structure is formed in a hexagonal close-packed manner sample, Pr-doped 3D SnS2/ZnS nanoflowers exhibited better
[6]. Due to its excellent physical and chemical properties, it is used sensing performance to ammonia.
in lithium ion batteries [7], photocatalysis [8], and gas sensor [9]. In
our previous work, 3D SnS2 hierarchical micro-flowers for ultra-
2. Experiments

2.1. Chemical reagent


* Corresponding author.
E-mail address: zhangqixian1992@163.com (S. Ma). All chemicals reagents in our experiments were analytical grade

https://doi.org/10.1016/j.jallcom.2019.151650
0925-8388/© 2019 Elsevier B.V. All rights reserved.
2 Q. Zhang et al. / Journal of Alloys and Compounds 807 (2019) 151650

and no need for purification, including Sodium stannate (Na2S- the total resistance change is defined as the response/recovery time
nO3$3H2O), Zinc acetate ((CH3COO)2Zn$2H2O), ammonia aqueous [16].
solution (NH3$H2O, 25%), Thioacetamide (CH3CSNH2), Sodium
alginate ((C6H7NaO6)n), and Acetic acid glacial (CH3COOH). 3. Results and discussion

2.2. Preparation of ZnSn(OH)6 hollow nanospheres 3.1. Samples structure characterization

Firstly, 5 mmol Na2SnO3$3H2O was fully dissolved in 15 ml of Fig. 2(a) displays the XRD pattern for the ZnSn(OH)6 samples. It
deionized water, and 0.02 g sodium alginate was dissolved in 5 ml can be seen from the figure, all the diffraction peaks are well cor-
of deionized water. Subsequently, 5 mmol Zn (CH3COO)2$2H2O was responding to the standard spectrum (JCPDS NO. 73e2384). There
fully dissolved in 10 ml of deionized water, and 5 ml of ammonia are no observed other impurities peaks, which indicates the high
aqueous solution was added. Then, after the two above solutions purity of the samples. Fig. 2(b) displays two sets of diffraction peaks
were fully mixed, reacted at 160  C for 10 h. To be cooled to room corresponding from SnS2 (JCPDS NO. 89e2358) and ZnS (JCPDS No.
temperature, white precipitates were obtained by centrifugation 65e0309), respectively. Nevertheless, due to the low content of Pr
and washed with deionized water and ethanol for several times, element, diffraction peaks of Pr element are not detected in Pr-
dried at 80  C for 10 h. SnS2/ZnS samples. In Fig. 2(c), it shows a large enlarged view of the
(100) and (011) diffraction peaks in the SnS2/ZnS and Pr-SnS2/ZnS
2.3. Preparation of Pr-SnS2/ZnS nanoflowers XRD pattern. It displays that the (100) and (011) peaks of Pr-SnS2/
ZnS compared with pure SnS2/ZnS have obviously right shift of
 
In this process, 0.286 g ZnSn(OH)6 samples were completely 0:06 and 0:07 , respectively. It may be due to the large ionic radius
dispersed in 20 ml of deionized water with magnetic stirring. of Pr (0.099 nm)/Pr4þ (0.085 nm) than that of Sn4þ (0.071 nm)

Meanwhile, 5 mmol thioacetamide (TAA) and a certain amount of [15,17]. When the Pr3þ/Pr4þ ions substitute for Sn4þ ions, due to the
PrN3O9$6H2O (0 wt % and 3 wt %) were dissolved in 10 ml of different valence states of them, more S vacancies are generated
deionized water, and 5 ml acetic acid glacial was added. The above based on the charge compensation effect, leading lattice distortion
solutions were mixed together with stirring until well mixed. Then, [17]. Besides, the crystallite size D of the SnS2/ZnS and Pr-SnS2/ZnS
the mixture kept at 160  C for 8 h. After cooling to room tempera- samples are calculated through the XRD data and the Scherrer
ture, yellow precipitates were collected by cleaned with deionized formula (Eq. (1) [18,19]), corresponding results are displayed in
water and ethanol for several times. It was dried at 80  C for 12 h. Table 1. Compared with pure SnS2/ZnS, the crystallite size of Pr-
SnS2/ZnS decreased. However, it is known that the width of the
2.4. Preparation and measurement of sensors diffraction peak depends on the size of the particles, ambient
temperature and lattice strain. The Williamson-Hall approach
The sensor was prepared as follows: (1) The SnS2/ZnS and claims that the grain size and lattice strain are two principal factors
Pr  SnS2/ZnS samples were mixed with deionized water in weight affecting the FWHM of X-ray diffraction peaks at a constant tem-
ratio of 4:1 and form a slurry; (2) As shown in Fig. 1(a), the slurry is perature [20]. Lattice strain is a measure of the distribution of lat-
uniformly coated on the surface of the ceramic tube with a thick- tice constants arising from crystal imperfections, such as lattice
ness. (3) As shown in Fig. 1(b), ceramic tube is soldered with the disruption, in which crystallite size reduction can induces a large
pedestal, the sensor was kept at 200  C for 48 h to enhance its amount of strain. Thus, the grain size DW-H and εWH of pure and Pr
stability. In order to study the gas sensing performance, the cor- doped SnS2/ZnS samples are identified by the Williamson-Hall
responding test circuit schematic is shown in Fig. 1(c). The gas approach (Eq. (2) [20e23]). Where D is the Scherrer's crystallite
sensing measurement system has been in Ref. [16]. In the experi- size, k ¼ 0.89, linc ¼ 1:5406 Å is the X-ray incident wavelength and
ment, the response was defined as S]Ra/Rg (Ra is the resistance of qhkl is the diffraction angle, bhkl is full width at half maximum of
gas sensor in air and Rg is the resistance of gas sensor in a test gas). intensity, DW-H and εWH is the size of the nanocrystals and lattice
When the test gas is injected/extracted, the time for getting 90% of strain in the Williamson-Hall method, respectively.

D ¼ klinc =ðcosqhkl bhkl Þ (1)

bhkl cosqhkl ¼ 4 εWH sinqhkl þ klinc =DWH (2)


Fig. 2 (d-e) show the Williamson-Hall plots of pure and Pr doped
SnS2/ZnS samples. DW-H can be calculated from the y-axis intercept.
The DW-H and εWH of the pure and Pr doped SnS2/ZnS samples are
22.663372 and 19.756974 nm, 0.0004125 and 0.0002075,
respectively. From the results of Scherrer's and Williamson-Hall
outputs, implying on the same trends of the grain size in two ap-
proaches, which has proved that Pr doping can lead to lattice
distortion and suppress the growth of crystallite [14,24].

3.2. XPS analysis

As shown in Fig. 3, the Pr-SnS2/ZnS hierarchical nanoflowers are


analyzed by XPS. In Fig. 3(a), the survey spectrum of the sample
displays that it is composed of Sn, Zn, S, Pr, C and O elements, where
C and O elements may come from XPS system and O2 adsorption on
Fig. 1. (a) The structure diagram of gas sensor configuration, (b) Sensor diagram, (c) the surface of materials [15]. The peaks located at 495.3 eV and
measurement circuit. 486.8 eV are well corresponding to Sn 3d5/2 and Sn 3d3/2 for Sn4þ
Q. Zhang et al. / Journal of Alloys and Compounds 807 (2019) 151650 3

Fig. 2. XRD patterns of the synthesized samples: (a) ZnSn(OH)6; (b) SnS2/ZnS and Pr-SnS2/ZnS; (c) Comparison the shift of (100) and (011) peaks of SnS2/ZnS and Pr-SnS2/ZnS;
Williamson-Hall plots (d) pure SnS2/ZnS, (e) Pr doped SnS2/ZnS.

Table 1 3.3. Morphological analysis


Parameters of the (100) and (011) diffraction peaks of undoped SnS2/ZnS and Pr-
SnS2/ZnS.
The morphology of the ZnSn(OH)6 samples was firstly tested by
Materials Planes (h k l) 2q ( ) FWHM ( ) D (nm) SEM. Fig. 4(a) shows that ZnSn(OH)6 samples exhibit regular
0 wt% Pr-SnS2/ZnS 100 28.243 0.378 21.457 spherical architecture and good dispersity. These nanospheres have
011 32.137 0.399 20.515 an average size of about 800 nm. Fig. 4(b) clearly displays the sur-
3 wt% Pr  SnS2/ZnS 100 28.304 0.473 17.150 face of these nanospheres have rough surface and some apparent
011 32.211 0.411 19.920
pores. It indicates that these nanospheres may have hollow nano-
structure. As shown in Fig. 4(c), there is a bright center and a much
darker edge in TEM image, demonstrating these nanospheres are
states in Fig. 3(b) [25], respectively. In Fig. 3(c), the peaks located at hollow.
1022.9 eV and 1045.9 eV are corresponding to Zn 2p3/2 and Zn 2p1/2 Fig. 4(d) and (e) show that the typical SEM images of pure SnS2/
for Zn2þ states [26], respectively. In Fig. 3(d), the peaks centered at ZnS nanoflowers samples. There are a lot of nanoflowers self-
161.62 eV and 162.90 eV are well corresponding to S 2p3/2 and S assembled by nanosheets. And there are no other morphologies,
2p3/2 for S2 states [17], respectively. As seen from Fig. 3(e), the indicating a high yield of SnS2/ZnS nanoflowers. The typical TEM
spectrum can be resolved into 5 peaks at s (929.5 eV), s’ (949.8 eV), image (Fig. 4(f)) displays a similar result to SEM images. And the 3D
m (933.5 eV), m’ (953.9 eV) and t’ (955.8 eV) [27], respectively. The SnS2/ZnS hierarchical nanoflowers are composed of uniform
peaks marked with s and s’ are corresponding to Pr4þ states, those nanosheets, the thickness of which is about 30 nm. As shown in
marked with m and m’ are corresponding to Pr3þ states, and the Fig. 4 (g-i), compared with the pure SnS2/ZnS, the morphology and
peak marked with t’ only exists in 3d3/2 part [28,29]. The corre- size of 3 wt % Pr-SnS2/ZnS samples have no obvious change.The
sponding element atomic concentration is shown in Table 2. The individual elemental mappings of Zn, Sn, S and Pr are shown in
XPS results show that Pr element exist in Pr-SnS2/ZnS with Pr3þ and Fig. 4(jen), indicating the homogenous distribution for the
Pr4þ. mentioned elements on the surface of Pr doped SnS2/ZnS
4 Q. Zhang et al. / Journal of Alloys and Compounds 807 (2019) 151650

Fig. 3. XPS spectra of the Pr-SnS2/ZnS hierarchical nanoflowers: (a) a survey spectrum, (b) Sn 3d, (c) Zn 2p, (d) S 2p, (e) Pr 3d.

Table 2 target gas and improve gas sensing performance of the sensing
The atomic concentration of the Pr doped SnS2/ZnS. materials.
Element C 1s O 1s S 2p Zn 2p3 Sn 3d5 Pr 3d

At/% 69.15 9.48 10.90 1.24 9.17 0.06 3.4. Gas sensing performance

The optimal operating temperature and selectivity are critical


nanoflowers. for semiconductor gas sensors. In this paper, the optimal operating
As shown in Fig. 5, the isotherms are identified as type IV ac- temperature was firstly tested. Moreover, the gas sensing perfor-
cording to the IUPAC classification, which indicates pure SnS2/ZnS mance of pure SnS2/ZnS and Pr-SnS2/ZnS are studied for getting
and Pr-SnS2/ZnS samples belong to mesoporous materials [29]. The compared. In Fig. 6(a), the responses of the gas sensors of pure
specific surface area of the synthesized samples is 21.4228 m2/g for SnS2/ZnS and Pr-SnS2/ZnS to 50 ppm ammonia are tested at the
pure SnS2/ZnS, 24.8033 m2/g for 3 wt % Pr doped SnS2/ZnS nano- range of 100  Ce240  C. It is obviously observed that the response
flowers. Obviously, the Pr-SnS2/ZnS nanoflowers samples provide of two gas sensors first increases and then decreases, reaching their
larger specific surface area which can promote the adsorption of maximum response at 160  C. It demonstrates that the optimal
operating temperature of two sensors is 160  C. The relationship
Q. Zhang et al. / Journal of Alloys and Compounds 807 (2019) 151650 5

Fig. 4. SEM and TEM images of samples: (a) and (b) SEM of ZnSn(OH)6, (c) TEM of ZnSn(OH)6 hollow nanospheres; (d) and (e) SEM of SnS2/ZnS, (f) TEM of pure SnS2/ZnS
nanoflowers; (g) and (h) SEM of 3 wt % Pr-SnS2/ZnS nanoflowers, (i) TEM of Pr-SnS2/ZnS, (jen) SEM images of Pr-SnS2/ZnS and its elemental mapping images of Zn, Sn, S and Pr,
respectively.

Fig. 5. N2 adsorption-desorption isotherm of samples: (a) pure SnS2/ZnS, (b) 3 wt % Pr-SnS2/ZnS.


6 Q. Zhang et al. / Journal of Alloys and Compounds 807 (2019) 151650

Fig. 6. (a) The response of the synthesized samples to 50 ppm ammonia under different operating temperatures; (b) The response of the synthesized samples to 50 ppm various test
gases at 160  C.

between temperature and response is also studied. Firstly, when Pr-SnS2/ZnS, respectively. Therefore, the Pr-SnS2/ZnS sensor pos-
the operating temperature is too slow, the chemical reaction rate is sesses a faster response and recovery speed for detecting ammonia
slow due to the temperature, resulting in a lower response. Sub- than pure SnS2/ZnS sensor.
sequently, with the operating temperature increasing, the Therefore, the sensor of Pr-SnS2/ZnS has better gas sensing
ammonia gas obtains a certain energy to promote the reaction, so performance to ammonia than pure SnS2/ZnS at 160  C. Compared
the response is correspondingly increasing. However, when the with pure 3D SnS2/ZnS hierarchical nanoflowers, Pr doped SnS2/
operating temperature is too high, the ammonia molecules will ZnS effectively improves the gas sensing performance to detect low
escape, resulting in a downturn response. Therefore, the semi- concentration of ammonia.
conductor gas sensor can effectively monitor ammonia at the op- In addition, as depicted in Table 3, we have made a comparison
timum operating temperature. As shown in Fig. 6(b), the response concerning ammonia sensing properties of Pr doped SnS2/ZnS
of the sensors is tested at 160  C to 50 ppm of various testing gases. nanoflowers and other SnS2 based materials. The Pr doped SnS2/
It can be clearly observed that both sensors of SnS2/ZnS and Pr- ZnS sensor possesses a relatively higher response value and faster
SnS2/ZnS have excellent gas sensing performance to ammonia gas, response/recovery time in detecting low concentration of ammonia
and the response of Pr-SnS2/ZnS to ammonia gas is higher than [2,6,30e33]. Therefore, the sensor based on Pr-SnS2/ZnS materials
SnS2/ZnS. Therefore, the sensor based on Pr-SnS2/ZnS samples has a has potential applications in monitoring low concentrations of
better selectivity than SnS2/ZnS to ammonia gas. ammonia.
In Fig. 7(a), the dynamic response variation of SnS2/ZnS and Pr-
SnS2/ZnS sensors to 5e500 ppm concentrations of ammonia at
160  C is tested. When the ammonia gas is injected, the response of 3.5. Sensing mechanism
two sensors firstly increases. When the ammonia gas took away,
the response of two sensors decreases. Moreover, when the Pr As well known, SnS2 and ZnS are both n-type semiconductors,
doped SnS2/ZnS sensors are exposed to 5 ppm ammonia at 160  C, it the gas sensing mechanism of gas sensor is based on the gas/
can still deliver a response value of 3.06, which indicates the semiconductor surface interaction [16,34,35] (as shown in
detection minimum limit of such gas sensor is 5 ppm as well as high Fig. 9(a)).
sensitivity to ammonia. In Fig. 7(b), a relationship image about the In the air atmosphere, the oxygen molecules adsorb on the
response and gas concentration, the response rises with increasing surfaces of the materials and obtain electrons from the conduction
of ammonia concentration. If the concentration is too high, the gas band, forming the chemisorbed oxygen species (such as O 
2, O ,
2
O ) through equations (1)e(4):
concentration will reach a saturated state with a certain response
[16]. The response of Pr-SnS2/ZnS to 50 ppm ammonia is 14.03,
O2ðgasÞ /O2ðadsÞ (3)
which is 1.7 times higher than the SnS2/ZnS sample. Usually, in the
range of relatively low concentration, the response is satisfied with
a functional relationship which already reported in Ref. [16]. O2ðadsÞ þ e /O
2ðadsÞ (4)
Therefore, as shown in Fig. 7(c), in the range of 5e100 ppm
ammonia, the linear fitting relationship of SnS2/ZnS is Ln (S- O  
2ðadsÞ þ e / 2OðadsÞ (5)
1) ¼ 1.004 Ln (C) - 1.6934, correlation coefficient R2 ¼ 0.97562.
Meanwhile, the linear fitting relationship of Pr-SnS2/ZnS is Ln (S-
1) ¼ 0.81825 Ln (C) - 0.78269, correlation coefficient R2 ¼ 0.97922. O  2
ðadsÞ þ e /OðadsÞ (6)
It can be seen that the correlation coefficient of Pr-SnS2/ZnS is
closest to 1, which proves that the Pr-SnS2/ZnS gas sensor can In this process, oxygen is adsorbed on the semiconductor metal
accurately predict the response of ammonia in the range of oxide surface which forms a potential barrier at the grain boundary.
5e100 ppm. Compared with SnS2/ZnS, the sensor of Pr-SnS2/ZnS is And an electron depletion layer is formed on the surface of the
more suitable for monitoring ammonia in the range of low material and increase the potential barrier at the grain boundaries.
concentration. The layer restricts the flow of electrons, corresponding resulting a
Fig. 8 shows the resistance change of SnS2/ZnS and Pr-SnS2/ZnS relatively higher resistance [34,36].
to 50 ppm ammonia at 160  C. It can be seen that the response/ When the sensors are exposed to ammonia gas atmosphere, it
recovery time are 11/22 s and 6/13 s for the sensors of SnS2/ZnS and involves the electrochemical oxidation reaction of ammonia (Eq. (7)
and (8)) [34,37,38]:
Q. Zhang et al. / Journal of Alloys and Compounds 807 (2019) 151650 7

Fig. 7. (a) The dynamic response change of the synthesized samples toward 5e500 ppm ammonia at 160  C; (b) The response of the synthesized samples to different ammonia
concentrations ranging from 5 to 2000 ppm; (c) Plot of ln (S-1) and ln C for all samples toward 5e100 ppm ammonia at 160  C.

Fig. 8. The resistance change of the sensors to 50 ppm ammonia at operating temperature of 160  C: (a) SnS2/ZnS, (b) Pr-SnS2/ZnS.

Table 3
A comparision of the performance of ammonia sensors based on various SnS2 structures.

Sensing materials Concentration Topt ( C) Response Response/Recovery Time (s) Ref.

Flower-shaped SnS2 100 ppm 200 7.4 50.6/624 [2]


2D SnS2 500 ppm RT 4.2 16/450 [6]
SnO2/SnS2 nanotubes 100 ppm RT 2.48 21/110 [30]
SnO2eSnS2 hybrids 500 ppm RT 2.3 11/NA [31]
SnS2 nanosheets 50 ppm RT 2.04 16/NA [32]
Sn/SnO2/N doped carbon nanocomposite 300 ppm 65 172.70 60/55 [33]
Pr-SnS2/ZnS nanoflowers 50 ppm 160 14.03 6/13 This work

RT denotes room temperature, NA denotes no specific data are available.


8 Q. Zhang et al. / Journal of Alloys and Compounds 807 (2019) 151650

Fig. 9. Schematic illustration of the sensing mechanism.

Pr 3þ þ O2 / Pr 4þ þ O (9)
4NH3 þ O
2 ðadsÞ ¼ 2N2 þ 6H2 O þ 6e

(7) 2

Pr 4þ þ e / Pr 3þ (10)
2 
2 = 3 NH3 þ O / 1=3 N2 þ H2 O þ 2 e (8) Finally, the excellent gas sensing performance of Pr doped SnS2/
In this process, the electron concentration of the system ZnS sensor are also related to hierarchical flower-like structure. The
increased, leading to reduction of the sensor resistance and distinctive three-dimensional hierarchical flower-like nano-struc-
increased conductivity [39e41]. Moreover, the corresponding tures are composed of numerous thin nanosheets, which increases
response value of the material can be obtained by using R ¼ Ra/Rg. effective specific surface area and provides more convenient
To comprehensively understand the enhancement gas sensing channels for adsorption and desorption of oxygen and target gases.
property of Pr doped SnS2/ZnS nanoflowers to ammonia, we pre-
sent qualitative explanation of the possible mechanisms: 4. Conclusion
Firstly, the SnS2/ZnS heterojunction is beneficial to improve gas
sensing performance. The energy band structures of SnS2 and ZnS In summary, the Pr doped 3D SnS2/SnS2 hierarchical nano-
are shown in Fig. 9(b). The bandgap energies of SnS2 and ZnS are flowers are successfully prepared by using ZnSn(OH)6 as precursor.
2.11 and 3.7 eV, respectively [42,43]. It is easily found that the Due to the right shift of the diffraction peak and the presence of the
conduction band (CB) of ZnS is higher than that of SnS2, the elec- Pr element in the XPS patterns, it proved that Pr is doped into SnS2/
trons on the CB of ZnS can transfer toward the CB of SnS2 [44]. ZnS. Pr doping led to the lattice distortion inhibit grain growth and
Meanwhile, the SnS2/ZnS n-n heterojunction could reduce the reduce grain size. The specific surface area of SnS2/ZnS and Pr-SnS2/
recombination probability of electron-hole pairs and improve the ZnS samples is 21.4228 m2/g and 24.8033 m2/g, respectively. The
electrons transmission efficiency [45]. Consequently, the absorbed sensor based on Pr doped SnS2/ZnS hierarchical nanoflowers
oxygen molecules will capture the additional electron which can effectively improved gas sensing performance to ammonia
further increase and decrease the sensor's resistance in air or including higher response and quicker responding and recovering
ammonia, respectively. Therefore, the SnS2/ZnS heterojunction speed. The response of Pr-SnS2/ZnS to 50 ppm ammonia is 14.03 at
plays an important modulating role in electrical conductivity for Pr- 160  C, which is 1.7 times higher than the SnS2/ZnS sample. At the
SnS2/ZnS sensor. same time, the Pr doped SnS2/ZnS sample has a faster response/
Secondly, the improvement of gas sensing performance is recovery time (6s/13s). Apart from the increase in specific surface
closely related to the doping of Pr3þ and Pr4þ ions. On the one hand, area, the SnS2/ZnS heterojunction and Pr3þ/Pr4þ doping improve
the growth of SnS2/ZnS crystallites can be restrained by Pr doping, the electrons transmission efficiency, which is the main reason for
which leads to crystallite sizes of Pr doped SnS2/ZnS samples the improvement of gas sensitivity.
smaller than in SnS2/ZnS ones. It has been reported in the relevant
literature that reducing the crystallite size can significantly Acknowledgments
enhance gas sensing [46]. On the other hand, Pr3þ may give an
electron to O2 adsorbed on the surface of Pr doped SnS2/ZnS to form This work was supported by the National Natural Science
O 4þ
2 by transforming into Pr . In contrast, the Pr

species may Foundation of China (Grant No. 11864034 and 51562035), the Sci-
capture electrons to form Pr3þ, it may include the following reac- entific Research Project of Gansu Province (Grant No. 18JR3RA089
tion of Eq. (9) and (10) [47]. Moreover, when Pr3þ/Pr4þ ions were and 17JR5RA072).
doped into SnS2 lattice, they will substitute Sn site to form the
electron-donor defects. In order to maintain electronic neutrality, References
Pr ions will liberate small amount of free electrons [25]. Therefore,
Pr doping can increase the concentration of free electrons, which [1] Z.P. Tshabalala, K. Shingange, F.R. Cummings, O.M. Ntwaeaborwa,
produce more O and O2 on the sample's surface through Eq. (5) G.H. Mhlongo, D.E. Motaung, Ultra-sensitive and selective NH3 room tem-
perature gas sensing induced by manganese-doped titanium dioxide nano-
and (6). When ammonia molecules reacting with those O and O2 , particles, J. Colloid Interface Sci. 504 (2017) 371e386.
there will be more free electrons released back to the conduction [2] Y. Xiong, W. Xu, D. Ding, W. Lu, L. Zhu, Z. Zhu, Y. Wang, Q. Xue, Ultra-sensitive
band on the basis of Eq. (7) and (8). The process will increase the NH3 sensor based on flower-shaped SnS2 nanostructures with sub-ppm
detection ability, J. Hazard Mater. 341 (2018) 159e167.
concentration of free electron. Hence, the more transport of free [3] J. Xu, Y. Chen, D. Chen, J. Shen, Hydrothermal synthesis and gas sensing
electrons results in excellent gas sensing performance. characters of ZnO nanorods, Sens. Actuators B Chem. 113 (1) (2006) 526e531.
Q. Zhang et al. / Journal of Alloys and Compounds 807 (2019) 151650 9

[4] Y. Wang, M.U. Qiuying, G. Wang, Z. Zhou, Sensing characterization to NH3 of [25] Q. Ge, S.Y. Ma, Y.B. Xu, X.L. Xu, H. Chen, Z. Qiang, H.M. Yang, L. Ma, Q.Z. Zeng,
nanocrystalline Sb-doped SnO2 synthesized by a nonaqueous sol-gel route, Preparation, characterization and gas sensing properties of Pr-doped ZnO/
Sens. Actuators B Chem. 145 (2) (2010) 847e853. SnO2 nanoflowers, Mater. Lett. 191 (2017) 5e9.
[5] J. Zhang, X. Fu, H. Hao, W. Gan, Facile synthesis 3D flower-like Ag@WO3 [26] H.M. Yang, S.Y. Ma, G.J. Yang, Q. Chen, Q.Z. Zeng, Q. Ge, L. Ma, Y. Tie, Synthesis
nanostructures and applications in solar-light photocatalysis, J. Alloy. Comp. of La2O3 doped Zn2SnO4 hollow fibers by electrospinning method and appli-
757 (2018) 134e141. cation in detecting of acetone, Appl. Surf. Sci. 425 (2017) 585e593.
[6] Z. Qin, K. Xu, H. Yue, H. Wang, J. Zhang, C. Ouyang, C. Xie, D. Zeng, Enhanced [27] H. Ogasawara, A. Kotani, R. Potze, G.A. Sawatzky, B.T. Thole, Praseodymium
room-temperature NH3 gas sensing by 2D SnS2 with sulfur vacancies syn- 3d- and 4d-core photoemission spectra of Pr2O3, Phys. Rev. B 44 (11) (1991)
thesized by chemical exfoliation, Sens. Actuators B Chem. 262 (2018) 5465.
771e779. [28] C.K.N, L.P. Haack, W. Chun, H.W.J, G.W. Graham, Single-phase PrOyZrO2
[7] JungWook Seo, JungTak Jang, SeungWon Park, Chunjoong Kim, materials and their oxygen storage Capacity: a comparison with single-phase
Byungwoo Park, Two-dimensional SnS2 nanoplates with extraordinary high CeO2ZrO2, PrOyCeO2, and PrOyCeO2ZrO2 materials, J. Phys. Chem. B 103
discharge capacity for lithium ion batteries, Adv. Mater. 20 (22) (2010) (18) (1999) 3634e3639.
4269e4273. [29] W.Q. Li, S.Y. Ma, Y.F. Li, X.B. Li, C.Y. Wang, X.H. Yang, L. Cheng, Y.Z. Mao, J. Luo,
[8] Y. Lei, S. Song, W. Fan, X. Yan, H. Zhang, Facile synthesis and assemblies of D.J. Gengzang, G.X. Wan, X.L. Xu, Preparation of Pr-doped SnO2 hollow
flowerlike SnS2 and in3þ-doped SnS2: hierarchical structures and their nanofibers by electrospinning method and their gas sensing properties,
enhanced photocatalytic property, J. Phys. Chem. C 113 (4) (2009) 1280e1285. J. Alloy. Comp. 605 (2014) 80e88.
[9] S.G. Leonardi, W. Wlodarski, Y. Li, N. Donato, A. Bonavita, G. Neri, Ammonia [30] R. Li, K. Jiang, S. Chen, Z. Lou, T. Huang, D. Chen, G. Shen, SnO2/SnS2 nanotubes
sensing properties of two-dimensional tin disulphide/tin oxides (SnS2/SnO2-x) for flexible room-temperature NH3 gas sensors, RSC Adv. 7 (83) (2017)
mixed phases, J. Alloy. Comp. 781 (2019) 440e449. 52503e52509.
[10] Q.X. Zhang, S.Y. Ma, G.J. Yang, R. Zhang, X.T. Wang, Q. Chen, L. Ma, S.T. Pei, [31] K. Xu, N. Li, D. Zeng, S. Tian, S. Zhang, D. Hu, C. Xie, Interface bonds determined
K.M. Zhu, W.Q. Wang, Y. Tie, 3D SnS2 hierarchical micro-flowers synthesized gas-sensing of SnO2-SnS2 hybrids to ammonia at room temperature, ACS Appl.
by ZnSn(OH)6 for ultra-sensitive NH3 sensor, Mater. Lett. 236 (2019) 600e603. Mater. Interfaces 7 (21) (2015) 11359e11368.
[11] Y.-S. Tsai, T.-W. Chou, C.Y. Xu, W. Chang Huang, C.F. Lin, Y.S. Wu, Y.-S. Lin, [32] H. Wang, K. Xu, D. Zeng, Room temperature sensing performance of
H. Chen, ZnO/ZnS core-shell nanostructures for hydrogen gas sensing per- graphene-like SnS2 towards ammonia, in: IEEE Sensors, 2015, 2015, pp. 1e4.
formances, Ceram. Int. (2019). https://doi.org/10.1016/j.ceramint.2019.05. [33] A.M. Al-Enizi, M. Naushad, A.a.H. Al-Muhtaseb, Ruksana, S.M. Alshehri,
345. Z.A. Alothman, T. Ahamad, Synthesis and characterization of highly selective
[12] L. Wang, H. Fu, Q. Jin, H. Jin, H. Haick, S. Wang, K. Yu, S. Deng, Y. Wang, Directly and sensitive Sn/SnO2/N-doped carbon nanocomposite (Sn/SnO2@NGC) for
transforming SnS2 nanosheets to hierarchical SnO2 nanotubes: towards sen- sensing toxic NH3 gas, Chem. Eng. J. 345 (2018) 58e66.
sitive and selective sensing of acetone at relatively low operating tempera- [34] A. Dey, Semiconductor metal oxide gas sensors: a review, Mater. Sci. Eng., B
tures, Sens. Actuators B Chem. 292 (2019) 148e155. 229 (2018) 206e217.
[13] Y. Han, Y. Liu, C. Su, S. Wang, H. Li, M. Zeng, N. Hu, Y. Su, Z. Zhou, H. Wei, [35] W.Q. Wang, S.Y. Ma, X.L. Xu, K.M. Zhu, L. Ma, Q.X. Zhang, Y. Tie, Optimization
Z. Yang, Interface engineered WS2/ZnS heterostructures for sensitive and of formaldehyde detection performance based on Ni2þ sensitized mono-
reversible NO2 room temperature sensing, Sens. Actuators B Chem. 296 (2019) disperse amorphous zinc tin oxide microcubes, Mater. Lett. 239 (2019)
126666. 207e211.
[14] L. Ma, S.Y. Ma, Z. Qiang, X.L. Xu, Q. Chen, H.M. Yang, H. Chen, Q. Ge, Q.Z. Zeng, [36] S. Kanan, O. El-Kadri, I. Abu-Yousef, M. Kanan, Semiconducting metal oxide
B.Q. Wang, Preparation of Co-doped LaFeO3 nanofibers with enhanced acetic based sensors for selective gas pollutant detection, Sensors 9 (10) (2009)
acid sensing properties, Mater. Lett. 200 (2017) 47e50. 8158e8196.
[15] Q. Chen, S.Y. Ma, H.Y. Jiao, B.Q. Wang, G.H. Zhang, D.J. Gengzang, L.W. Liu, [37] R.P. Patil, C. Hiragond, G.H. Jain, P.K. Khanna, V.B. Gaikwad, P.V. More, La
H.M. Yang, Sodium alginate assisted hydrothermal method to prepare pra- doped BaTiO3 nanostructures for room temperature sensing of NO2/NH3:
seodymium and cerium co-doped ZnSn(OH)6 hollow microspheres and syn- focus on La concentration and sensing mechanism, Vacuum 166 (2019)
ergistically enhanced ethanol sensing performance, Sens. Actuators B Chem. 37e44.
252 (2017) 295e305. [38] L. Wang, W. Meng, Z. He, W. Meng, Y. Li, L. Dai, Enhanced selective perfor-
[16] Q. Chen, S.Y. Ma, X.L. Xu, H.Y. Jiao, G.H. Zhang, L.W. Liu, P.Y. Wang, mance of mixed potential ammonia gas sensor by Au nanoparticles decorated
D.J. Gengzang, H.H. Yao, Optimization ethanol detection performance man- CeVO4 sensing electrode, Sens. Actuators B Chem. 272 (2018) 219e228.
ifested by gas sensor based on In2O3/ZnS rough microspheres, Sens. Actuators [39] M. Tak acs, C. Dücso } , Z. L
abadi, A.E. Pap, Effect of hexagonal WO3 morphology
B Chem. 264 (2018) 263e278. on NH3 sensing, Procedia Eng. 87 (2014) 1011e1014.
[17] Q. Zhao, H. Zhang, Y. Liu, M. Zhu, M. Zhang, Magnetic and optical properties of [40] T.T. Wang, S.Y. Ma, L. Cheng, X.H. Jiang, M. Zhang, W.Q. Li, W.X. Jin, Facile
two-dimensional SnS2 nanosheets doped with Ho ions, Appl. Surf. Sci. 481 fabrication of multishelled SnO2 hollow microspheres for gas sensing appli-
(2019) 1370e1376. cation, Mater. Lett. 164 (2016) 56e59.
[18] P.M. Kibasomba, S. Dhlamini, M. Maaza, C.-P. Liu, M.M. Rashad, D.A. Rayan, [41] Y. Xiong, W. Xu, D. Ding, W. Lu, L. Zhu, Z. Zhu, Y. Wang, Q. Xue, Ultra-sensitive
B.W. Mwakikunga, Strain and grain size of TiO2 nanoparticles from TEM, NH3 sensor based on flower-shaped SnS2 nanostructures with sub-ppm
Raman spectroscopy and XRD: the revisiting of the Williamson-Hall plot detection ability, J. Hazard Mater. 341 (2018) 159e167.
method, Results Phys. 9 (2018) 628e635. [42] K. Zhang, L. Jin, Y. Yang, K. Guo, F. Hu, Novel method of constructing CdS/ZnS
[19] A.W. Burton, K. Ong, T. Rea, I.Y. Chan, On the estimation of average crystallite heterojunction for high performance and stable photocatalytic activity,
size of zeolites from the Scherrer equation: a critical evaluation of its appli- J. Photochem. Photobiol. A Chem. 380 (2019) 111859.
cation to zeolites with one-dimensional pore systems, Microporous Meso- [43] Y. Huo, Y. Yang, K. Dai, J. Zhang, Construction of 2D/2D porous graphitic C3N4/
porous Mater. 117 (1e2) (2009) 75e90. SnS2 composite as a direct Z-scheme system for efficient visible photocatalytic
[20] Y. Shahmoradi, D. Souri, Growth of silver nanoparticles within the tellur- activity, Appl. Surf. Sci. 481 (2019) 1260e1269.
ovanadate amorphous matrix: optical band gap and band tailing properties, [44] S. Harish, Prachi, J. Archana, M. Navaneethan, M. Shimomura, H. Ikeda,
beside the Williamson-Hall estimation of crystallite size and lattice strain, Y. Hayakawa, Synergistic interaction of 2D layered MoS2/ZnS nanocomposite
Ceram. Int. 45 (6) (2019) 7857e7864. for highly efficient photocatalytic activity under visible light irradiation, Appl.
[21] M. Ghasemi Hajiabadi, M. Zamanian, D. Souri, Williamson-Hall analysis in Surf. Sci. 488 (2019) 36e45.
evaluation of lattice strain and the density of lattice dislocation for nanometer [45] X.-T. Wang, Y. Li, X.-Q. Zhang, J.-F. Li, Y.-N. Luo, C.-W. Wang, Fabrication of a
scaled ZnSe and ZnSe:Cu particles, Ceram. Int. 45 (11) (2019) 14084e14089. magnetically separable Cu2ZnSnS4/ZnFe2O4 p-n heterostructured nano-
[22] A. Kalita, M.P.C. Kalita, Williamson-Hall analysis and optical properties of photocatalyst for synergistic enhancement of photocatalytic activity
small sized ZnO nanocrystals, Phys. E Low-dimens. Syst. Nanostruct. 92 (2017) combining with photo-Fenton reaction, Appl. Surf. Sci. 479 (2019) 86e95.
36e40. [46] N. Yamazoe, New approaches for improving semiconductor gas sensors, Sens.
[23] K. Venkateswarlu, M. Sandhyarani, T.A. Nellaippan, N. Rameshbabu, Estima- Actuators B Chem. 5 (1) (1991) 7e19.
tion of crystallite size, lattice strain and dislocation density of nanocrystalline [47] Y. Hanifehpour, B. Soltani, A.R. Amani-Ghadim, B. Hedayati, B. Khomami,
carbonate substituted hydroxyapatite by X-ray peak variance analysis, Pro- S.W. Joo, Praseodymium-doped ZnS nanomaterials: hydrothermal synthesis
cedia Mater. Sci. 5 (2014) 212e221. and characterization with enhanced visible light photocatalytic activity, J. Ind.
[24] T.T. Wang, S.Y. Ma, L. Cheng, J. Luo, X.H. Jiang, W.X. Jin, Preparation of Yb- Eng. Chem. 34 (2016) 41e50.
doped SnO2 hollow nanofibers with an enhanced ethanolegas sensing per-
formance by electrospinning, Sens. Actuators B Chem. 216 (2015) 212e220.

You might also like