Next Article in Journal
Simulation Research on Blood Detection Sensing with Parity-Time Symmetry Structure
Next Article in Special Issue
Pt-Based Multimetal Electrocatalysts and Potential Applications: Recent Advancements in the Synthesis of Nanoparticles by Modified Polyol Methods
Previous Article in Journal
Tunable Bandgaps in Phononic Crystal Microbeams Based on Microstructure, Piezo and Temperature Effects
Previous Article in Special Issue
Enhanced Photoresponsivity of 2H-MoTe2 by Inserting 1T-MoTe2 Interlayer Contact for Photodetector Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Humidity Sensing and Photodetection Based on Tin Disulfide Nanosheets

1
Department of Electronic Engineering, National Changhua University of Education, No. 2, Shi-Da Road, Changhua 50074, Taiwan
2
Department of Electronic Engineering, Ming Chi University of Technology, Taishan, No. 84 Gungjuan Road, New Taipei City 24301, Taiwan
3
Department of Materials Science and Engineering, National Chung Hsing University, Taichung 402, Taiwan
4
Key Laboratory of Optoelectronic Devices and Systems of Ministry of Education and Guangdong Province, College of Physics and Optoelectronic Engineering, Shenzhen University, Shenzhen 518060, China
*
Authors to whom correspondence should be addressed.
Crystals 2021, 11(9), 1028; https://doi.org/10.3390/cryst11091028
Submission received: 27 July 2021 / Revised: 18 August 2021 / Accepted: 20 August 2021 / Published: 26 August 2021

Abstract

:
Tin disulfide has substantial importance for two-dimensional material-based optoelectronics and sensors due to its unique optoelectrical properties. In this report, we fabricate SnS2 nanosheets using the low-pressure thermal sulfurization process, whose crystal structure and surface morphology are confirmed by X-ray diffraction (XRD) and scanning electron microscope (SEM) measurements. From photoconductivity measurement and photocurrent mapping, we observe smaller electrode spacing of SnS2 thin films can enhance photodetection. Then, by the H2O2 oxidation processing, we transform SnS2 to SnO2 to detect humidity. The measured response and recovery time can be optimized to 5.6 and 1.0 s, respectively, shorter than those of commercial DHT11 humidity sensor of 32 and 34 s. At suitable bias, humidity sensor can detect human respiration properly at room temperature. Our results show that SnS2 nanosheets exhibit reasonable performance for emergent photodetector applications and humidity sensing.

1. Introduction

Transition metal dichalcogenides (TMDC) materials [1,2,3] are attracting intense research interest due to their appreciable band gap in optical and electronical areas. Numerous two-dimensional (2D) hexagonal structures have been widely explored for various applications including photocatalysis [4], supercapacitors [5], optoelectronics [6,7] and sensors [8,9,10]. TMDCs including MoS2 [11,12,13], WS2 [14], WSe2 [12,15] and MoTe2 [16] are MX2 stoichiometric compounds, consisting of periodically stacked one transition metal layer and two chalcogenide layers. Among these TMDC semiconductors—although SnS2 is less explored compared with MoS2 or WS2—tin (Sn) and sulfur (S) are cheap and abundant on the earth and their potential applications have increased fast. Tin disulfide (SnS2) is an intrinsic n-type layered semiconductor [14] with a band gap of 2.18–2.44 eV [17,18], and it has a layered hexagonal CdI2 structure with stacked layers with weak van der Waals interaction [19]. Unique characteristics of SnS2 including high on–off ratios, fast photoresponse rate and good stability [20,21,22] makes it a promising candidate for a variety of applications in emergent optical and electronic areas. Humidity sensors are very important in modern monitoring field including environmental monitoring, agricultural production, industrial production, biological, medical and chemical monitoring [23,24]. Humidity sensors based on semiconducting oxides have advantages such as low cost, small size and ease of placing in the operating environment. As one important member of metal oxide semiconductors (SnO2, ZnO, WO3), tin dioxide is a typical n-type wide semiconductor with bandgap around 3.7 eV at 300 K [25,26,27], it has been widely used in various applications such as gas sensors [27,28], solar cells [29] and humidity sensors [30].
In the present work, we synthesized large area SnS2 films by low pressure sulfurized method. We study the crystal structure and surface morphology of SnS2 and SnO2 films by using XRD and SEM measurements, and reveal their optical properties by using the photoconductivity measurement and photocurrent mapping spectroscopy. Optoelectronic response of SnS2 devices is investigated including response time, responsivity and electrode spacing. Furthermore, the SnO2 humidity sensor is fabricated by oxidizing SnS2 film with hydrogen peroxide (H2O2). Our work suggests that SnS2 film has good performance as optical sensors, and SnO2 is promising humidity sensing material for practical and commercial applications.

2. Materials and Methods

We grow SnS2 films by low pressure thermal sulfurized method for this study. First, we use E-gun system to evaporate 5/50 nm Ti/Au metallic electrodes on slide glass substrates in a three-zone horizontal tube furnace. Then, we synthesize large area SnS2 films on metallic electrodes as follows. The source materials in a quartz crucible are 1.2 g sulfur powder and 0.4 g SnO2 powder, which are placed in the upstream and center of the tube furnace, respectively. Glass substrates are placed 11 cm upstream from the front end of the tube furnace. The center of the tube furnace is kept at 700 °C for 1 h. During the growth process, the carrier gas (Ar gas) has been constantly supplied in a flow rate of 100 sccm, which has also been maintained until the whole furnace temperature gradually decreases down to room temperature. To fabricate SnO2 humidity sensors, we use 35% hydrogen peroxide (H2O2) at 70 °C for 30 min to oxidize SnS2 film.
We carry out the XRD and SEM measurements to confirm crystal structures and surface morphologies of SnS2 and SnO2 films. Their atomic proportion is also determined by energy dispersive X-ray spectrum (EDS). For photoconductivity and photocurrent measurements, a 130 W tungsten–halogen lamp with a monochromator is used to provide monochromatic light, tunable in a wide spectral range by grating. The AC output signal with a chopper frequency of 9 Hz is recorded by a dual phase lock-in amplifier (SR830) to suppress noise signals. The stable bias voltage is supplied by a source meter (Keithley 2400).
The 2D photocurrent mapping of SnS2 films with different contact spacing are obtained with 520 nm TTL laser illumination; photocurrent signals are recorded by a pico-ammeter (Keithley 6485) with a bias voltage of 20 V. For humidity measurements, we first use carbon dioxide (CO2) to create a dry environment, then we measure sample resistance with respect to relative humidity at room temperature using a source meter (Keithley 2400)

3. Results and Discussion

Figure 1a shows the SEM image of SnS2 film on glass substrates. It can be observed that numerous SnS2 nanosheets are distributed in densely interlaced configurations on the substrate. The continuous SnS2 film is composed of flower-like surfaces of nanosheets, which has an advantage for photoresponsivity, similar to SnS2 nanoflake based field effect transistor and high efficiency photodetector [10]. The atomic composition ratio of Sn:S is about 34.6:66.4 (Sn:S = 1:1.98), which is very close to the ideal ratio of SnS2 (Sn:S = 1:2). Figure 1b shows SEM image of SnO2 film, fabricated by oxidizing SnS2 film with hydrogen peroxide (H2O2). We anticipate flower-like surface of SnS2 and SnO2 films can enhance light absorption and thus photo-responsivity, as well as humidity sensing characteristics. We also observe that the SnO2 film remained the similar platelet-shaped nanosheets and from the EDS analysis we found that the atomic composition ratio of Sn:O:S is about 27:70:3. A small amount of sulfur has been observed implies that the SnS2 may not transform into SnO2 completely. The atomic composition ratio of Sn:O is a little bit higher than the ideal ratio of 1:2, which implies that some oxygen atoms could not find their lattice positions but interstitial sites.
Figure 1c shows XRD patterns for SnS2 and SnO2 films. The detected peaks can be attributed to (001), (100), (101), (102), (110), (111), (103), (201), (202), (113) and (203) crystallographic planes of the hexagonal phases SnS2, as shown for hexagonal SnS2 (card no. 23-0677) of the JCPDS database. Our results also show good crystal structure of large area SnS2 films by low pressure thermal sulfurization process. Here, we found that the highest XRD intensity showed (101) preferred orientation, which is similar to the result reported by J. Ma et al. [31] Through anisotropic nucleation at high precursor concentration, three-dimensional aggregation effects may dominate the whole process and develop a flower-like morphology rather than a flat two-dimensional surfaces with a preferred orientation of (001) direction. After oxidization by using H2O2, the SnS2 films have been transformed into SnO2 films. The sulfide atoms have been substituted by oxygen atoms and rearranged in a new crystal type with a preferred orientation of (200). In the XRD patterns of SnO2 film, the first weak peak represents the (001) plane of SnS2 film and the second peak indicates the (200) plane of SnO2 (card no. 41-1445) of the JCPDS database. The coexistence of SnS2 and SnO2 peaks reveals that some residual SnS2 has not been transformed completely. We think residual SnS2 nanosheets may coexist with the SnO2 nanosheets and become a SnO2/SnS2 heterojunction structures, which may contribute better response than pure SnO2 or SnS2 nanosheets as reported by D. Gu et al. [32]
We are also interested in the obvious (200) peak from SnO2 film. It may imply the SnO2 film has been formed in a very disordered phase. The 2θ peak position of the (200) located at 38.24 degree, which is a little bit higher that that indicated in the JCPDS No. 41-1445 (37.88 degree). We think that during the process of oxygen substitution for transformation of SnO2, some oxygen atoms may insert in possible interstitial positions in the crystals and extend the lattice a little bit. Actually, the XRD analysis is done with an X-ray source of Cu Kα radiation (λ = 1.5406 Å), and the lattice constant of a is determined to be 4.7 Å for 2θ = 37.88 degree while a = 4.74 Å for 2θ = 38.24 degree. The difference is less than 1%. In Figure 1d, the XPS spectrum of SnS2 film shows a characteristic peak of Sn 3d5/2 and Sn 3d3/2 bands around 486 and 494 eV, respectively. The signals of S 2p and 2s bands are found at the binding energy around 161 and 225 eV, respectively. Furthermore, a weak O 1s peak locates at 531 eV, which may contribute from the native SnO2 layer on the SnS2 surface [33]. The photo-conductivity spectrum in the inset of Figure 1d also illustrates the proper band gap of SnS2 film around 2.32 eV. After oxidization with H2O2, on the SnO2 film, no sulfur related signals have been found and an obvious O 1s band signal has been observed at binding energy of 531 eV consistent with the results of Love et al. [33]. We anticipate H2O2 oxidization can effectively change SnS2 to SnO2, which has good humidity sensing characteristics.
To measure photoresponse and humidity sensing of our devices, we use the home-built AC opto-electronic setup in Figure 2a and humidity measurement system in Figure 4a, respectively. In Figure 2c,d, we show 2D photocurrent mapping of two SnS2 films with electrode spacing of 600 and 800 μm, respectively. The width of the channel is 2000 um and the film thickness is about 5 um. The only difference for the geometrical parameter of the devices is the channel length. In this a pilot study we did not focus on the comparison of different geometrical parameters. The strong photocurrent is generated only in the gap of voltage biased metal-semiconductor-metal fingers. The photocurrent distribution reveals that some area can generate higher current (about two times) than the average value. This fact implies that the uniformity of film quality needs to be improved for applications. The photoinduced carriers may recombine before they can reach the electrodes due to high density of defects in the crystal surface or interfaces. We also notice the electrode spacing has its influence to photodetection, which provides another parameter to optimize device performance.
To understand the photoresponse of SnS2 films, including responsivity speed, defect level influence to carrier lifetime and electrode spacing effect, we carry out frequency dependent photoconductivity measurements. In Figure 3a, we present photocurrent results excited by a 405 nm laser at various chopper frequencies. The observed decrease of photocurrent n a c with increasing frequency obeys the relation (1) [34]:
n a c = k 1 × tan h ( 1 4 f τ 1 ) + k 2 × tan h ( 1 4 f τ 2 )
where n a c is the measured photocurrent, f is the chopper frequency, k 1 and k 2 describe indirect and direct recombination, respectively; τ 1 and τ 2 are carrier lifetime of indirect and direct recombinations. In Table 1, we list fitted k 1 ,   k 2 , τ 1 and τ 2 values. When SnS2 film contact spacing changes from 600 to 800 μm, the large variation of k 2   from 0.04 to 0.36 is the main result, indicating defects are decreased due to the shorter spacing electrode. Our results reveal that SnS2 film of 600 μm electrode spacing can have larger photocurrent than that of 800 μm electrode spacing. Figure 3b shows the relative balance [(Imax − Imin)/Imax] versus frequency up to 10 kHz. The relative balance remains up to 23.4% for the 600 μm electrode spacing SnS2 film at 10 kHz; however, it drops to 9.01% at 30 Hz for the 800 μm electrode spacing SnS2 film and photocurrent becomes too low to be measured at high frequency range. Our observation implies that the SnS2 film with 600 μm contact spacing is capable of monitoring faster optical signals.
In Figure 3c,d, we compare photoresponsivity of SnS2 films of 600 and 800 μm contact spacing with respect to laser power and bias voltage. At weak 405 nm incident laser intensity of 0.2 μW, the responsivities are 2.55 mA/W and 1.55 mA/W for SnS2 films of 600 and 800 μm contact spacing, respectively. The relative photoresponsivity keeps unchanged with increasing incident laser power. The effect of bias voltage on the photoresponsivity shows the same trend in Figure 3d. For both low and high operating voltages up to 20 V, the responsivity values of the SnS2 film with 600 μm contact spacing is always larger and improves faster than that of the SnS2 film with 800 μm contact spacing. Our results imply shorter contact spacing can improve photoresponsivity for SnS2 film photodetection. Although the underlying mechanism needs further study, we believe the reduced grain boundary of shorter electrode spacing can improve both photodetection as well as the following humidity sensing.
In Figure 4b,c, we show the response and recovery time measurements of SnO2 humidity sensors. For SnO2 sensors of 700 and 900 μm contact spacing and the commercial DHT11 humidity sensor, the measured response times are 5.6, 9.0 and 32 s, and for recovery time 1.0, 1.0 and 34 s, respectively. In general, the sensing mechanism occurring in SnO2 humidity sensors has several hypotheses. Since water is a weak electrolyte, adsorbed water molecules can be dissociated on the surface. On the other hand, free electrons can also be released due to surface interactions between the SnO2 and the adsorbed H2O molecules. [35] H. Y. Xu et al. has reported that the exposed {200} crystal faces have more unsaturated metal bounds and dangling bonds, which resulting in more oxygen vacancies for enhancing the absorption of water molecules. [36]
Our results indicate both SnO2 humidity sensors have faster response than that of the DHT11 sensor. In Figure 4d,e, humidity sensitivity versus relative humidity and bias voltage are measured. Humidity sensitivity has been calculated using the formula:
S ( % ) = I R H I D A
where I R H and I D A are measured currents at given and initial relative humidity, respectively [37]. Although the trend of resistance decrease with increasing humidity keeps the same for different humidity sensors in Figure 4f, the maximum sensitivity of bias 30 V at room humidity is observed for SnO2 film with 700 μm contact spacing. Our results SnO2 film with 700 μm contact spacing has better response than film of 900 μm contact spacing, same as the situation of SnO2 film photodetection.
Breath monitoring is one important application for fast response humidity sensor. In Figure 4g, the current-time (I-T) plot of the SnO2 film with 700 and 900 μm contact spacing are at bias voltage of 10 V. Each current peak represents one human exhalation, and we also simulate apnea from 60 to 75 s. The observed respiratory rate is 17 breaths per minute, and we can clearly observe the apnea. Even though our samples have not reached their current top at high humidity, they still get the excellent measurement of human exhalation. On the contrary, the DHT11 humidity sensor with longer response time than SnO2 films, cannot apply to human exhalation. The commercial DHT11 has a plastic case to protect the humidity device and which is put in a mask for this test. We think may be the humidity is not very easy to escape out the plastic case and the mask resulting in a constant high humidity curve in Figure 4g. Our experiments show SnS2 film possesses fast response and recovery time to water molecules generated by human exhalation at room temperature.

4. Conclusions

In conclusion, we have successfully synthesized high quality SnS2 films by low pressure thermal sulfurization process, confirmed by XRD, Raman and SEM measurements. SnS2 films are composed of nanosheets of flower-like surface and large photoresponsivity up to 2.55 and 1.55 mA/W for electrode spacing of 600 and 800 μm films, and the relative balance can remain as 23.4% for the SnS2 film with 600 μm electrode spacing up to 10 kHz. We further transform SnS2 to SnO2 using H2O2 oxidation process. The humidity sensing analysis showed the SnO2 film has shorter response time than DHT11 humidity sensor, and can be even used to monitor human breathing. Our results indicate short contact spacing not only enhances photoresponsivity significantly but also improves the humidity sensing capability for SnS2 and SnO2 films separately. Our results would provide useful insights into the design, fabrication and practical application of photodetector and humidity sensor based on SnS2 film.

Author Contributions

H.-P.H. and D.-Y.L. conceived and designed the experiments. H.-S.H., Y.-C.Y. and C.-F.L. prepared the materials and performed the experiments. W.Z., D.-Y.L., H.-S.H. and Y.-C.Y. analyzed data. W.Z. and D.-Y.L. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Ministry of Science and Technology of Taiwan under project no. MOST. 109-2221-E-018-008.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to express their sincere acknowledgments to H.-K Teng in Nan Kai University of Technology for his kind support in equipment and useful discussion.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Tan, H.; Xu, W.; Sheng, Y.; Lau, C.; Fan, Y.; Chen, Q.; Tweedie, M.; Wang, X.; Zhou, Y.; Warner, J. lateral graphene-contacted vertically stacked WS2/MoS2 hybrid photodetectors with large gain. Adv. Mater. 2017, 29, 1702917. [Google Scholar] [CrossRef]
  2. Fang, H.; Chuang, S.; Chang, T.C.; Takei, K.; Takahashi, T.; Javey, A. High-performance single layered WSe2 p-FETs with chemically doped contacts. Nano Lett. 2012, 12, 3788–3792. [Google Scholar] [CrossRef] [Green Version]
  3. Li, M.; Shi, Y.; Cheng, C.; Lu, L.; Lin, Y.; Tang, M.; Tsai, M.L.; Chu, C.; Wei, K.; He, J.; et al. Epitaxial growth of a monolayer WSe2-MoS2 lateral p-n junction with an atomically sharp interface. Science 2015, 349, 524–528. [Google Scholar] [CrossRef] [Green Version]
  4. Zhang, N.; Zhang, Y.; Xu, Y. Recent progress on graphene-based photocatalysts: Current status and future perspectives. Nanoscale 2012, 4, 5792–5813. [Google Scholar] [CrossRef]
  5. El-Kady, M.; Kaner, R. Scalable fabrication of high-power graphene micro-supercapacitors for flexible and on-chip energy storage. Nat. Commun. 2013, 4, 1475. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, Q.H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J.N.; Strano, M.S. Electronics and optoelectronics of two-dimensional transition metal dichalcogenides. Nat. Nanotechnol. 2012, 7, 699. [Google Scholar] [CrossRef] [PubMed]
  7. Jariwala, D.; Sangwan, V.K.; Lauhon, L.; Marks, T.J.; Hersam, M.C. Emerging device applications for semiconducting two-dimensional transition metal dichalcogenides. ACS Nano 2014, 8, 1102–1120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Choi, W.; Choudhary, N.; Han, G.H.; Park, J.; Akinwande, D.; Lee, Y.H. Recent development of two-dimensional transition metal dichalcogenides and their applications. Mater. Today 2017, 20, 116–130. [Google Scholar] [CrossRef]
  9. Ou, J.Z.; Carey, B.; Ge, W.; Daeneke, T.; Rotbart, A.; Shan, W.; Wang, Y.; Fu, Z.; Chrimes, A.F.; Wlodarski, W. Physisorp-tion-based charge transfer in two-dimensional SnS2 for selective and reversible NO2 gas sensing. ACS Nano 2015, 9, 10313–10323. [Google Scholar] [CrossRef]
  10. Jia, X.; Tang, C.; Pan, R.; Long, Y.-Z.; Gu, C.; Li, J. Thickness-dependently enhanced photodetection performance of vertically grown SnS2 nanoflakes with large size and high production. ACS Appl. Mater. Interfaces 2018, 10, 18073–18081. [Google Scholar] [CrossRef]
  11. Svatek, S.A.; Antolin, E.; Lin, D.-Y.; Frisenda, R.; Reuter, C.; Molina-Mendoza, A.J.; Muñoz, M.; Agrait, N.; Ko, T.-S.; de Lara, D.P.; et al. Gate tunable photovoltaic effect in MoS2 vertical p–n homostructures. J. Mater. Chem. C 2016, 5, 854–861. [Google Scholar] [CrossRef] [Green Version]
  12. Li, H.; Wu, J.; Yin, Z.; Zhang, H. Preparation and applications of mechanically exfoliated single-layer and multilayer MoS2 and WSe2 nanosheets. Acc. Chem. Res. 2014, 47, 1067–1075. [Google Scholar] [CrossRef]
  13. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef]
  14. Perea-Lopez, N.; Elías, A.L.; Berkdemir, A.; Castro-Beltran, A.; Gutiérrez, H.R.; Feng, S.; Lv, R.; Hayashi, T.; López-Urías, F.; Ghosh, S.; et al. Photosensor device based on few-layered WS2 Films. Adv. Funct. Mater. 2013, 23, 5511–5517. [Google Scholar] [CrossRef]
  15. Huang, J.-K.; Pu, J.; Hsu, C.-L.; Chiu, M.-H.; Juang, Z.-Y.; Chang, Y.-H.; Chang, W.-H.; Iwasa, Y.; Takenobu, T.; Li, L.-J. Large-area synthesis of highly crystalline WSe2 Monolayers and device applications. ACS Nano 2014, 8, 923–930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Sung, J.H.; Heo, H.; Si, S.; Kim, Y.H.; Noh, H.R.; Song, K.; Kim, J.; Lee, C.-S.; Seo, S.-Y.; Kim, D.-H.; et al. Coplanar semiconductor–metal circuitry defined on few-layer MoTe2 via polymorphic heteroepitaxy. Nat. Nanotechnol. 2017, 12, 1064–1070. [Google Scholar] [CrossRef] [PubMed]
  17. Burton, L.A.; Whittles, T.; Hesp, D.; Linhart, W.; Skelton, J.M.; Hou, B.; Webster, R.F.; O’Dowd, G.; Reece, C.; Cherns, D.; et al. Electronic and optical properties of single crystal SnS2: An earth-abundant disulfide photocatalyst. J. Mater. Chem. A 2015, 4, 1312–1318. [Google Scholar] [CrossRef] [Green Version]
  18. Huang, Y.; Sutter, E.; Sadowski, J.; Cotlet, M.; Monti, O.L.; Racke, D.A.; Neupane, M.R.; Wickramaratne, D.; Lake, R.; Parkinson, B.A.; et al. Tin disulfide—An emerging layered metal dichalcogenide semiconductor: Materials properties and device characteristics. ACS Nano 2014, 8, 10743–10755. [Google Scholar] [CrossRef]
  19. Ham, G.; Shin, S.; Choi, H.; Park, J.; Lee, J.; Jeon, H. Engineering the crystallinity of tin disulfide deposited at low temperatures. RSC Adv. 2016, 6, 54069–54075. [Google Scholar] [CrossRef]
  20. Chang, R.-J.; Tan, H.; Wang, X.; Porter, B.; Chen, T.; Sheng, Y.; Zhou, Y.; Huang, H.; Bhaskaran, H.; Warner, J.H. High-performance all 2D-layered tin disulfide: Graphene photodetecting transistors with thickness-controlled interface dynamics. ACS Appl. Mater. Interfaces 2018, 10, 13002–13010. [Google Scholar] [CrossRef] [PubMed]
  21. Su, G.; Hadjiev, V.; Loya, P.E.; Zhang, J.; Lei, S.; Maharjan, S.; Dong, P.; Ajayan, P.M.; Lou, J.; Peng, H. Chemical vapor deposition of thin crystals of layered semiconductor SnS2 for fast photodetection application. Nano Lett. 2015, 15, 506–513. [Google Scholar] [CrossRef]
  22. De, D.; Manongdo, J.; See, S.; Zhang, V.; Guloy, A.; Peng, H. High on/off ratio field effect transistors based on exfoliated crystalline SnS2 nano-membranes. Nanotechnology 2012, 24, 025202. [Google Scholar] [CrossRef]
  23. Song, X.; Qi, Q.; Zhang, T.; Wang, C. A humidity sensor based on KCl-doped SnO2 nanofibers. Sens. Actuators B Chem. 2009, 138, 368–373. [Google Scholar] [CrossRef]
  24. Bharatula, L.D.; Erande, M.B.; Mulla, I.S.; Rout, C.S.; Late, D.J. SnS2 nanoflakes for efficient humidity and alcohol sensing at room temperature. RSC Adv. 2016, 6, 105421–105427. [Google Scholar] [CrossRef]
  25. Wang, L.; Wang, S.; Wang, Y.; Zhang, H.; Kang, Y.; Huang, W. Synthesis of hierarchical SnO2 nanostructures assembled with nanosheets and their improved gas sensing properties. Sens. Actuators B Chem. 2013, 188, 85–93. [Google Scholar] [CrossRef]
  26. Zhang, D.-F.; Sun, L.-D.; Yin, J.-L.; Yan, C.-H. Low-temperature fabrication of highly crystalline SnO2 nanorods. Adv. Mater. 2003, 15, 1022–1025. [Google Scholar] [CrossRef]
  27. Xu, J.; Li, Y.; Huang, H.; Zhu, Y.; Wang, Z.; Xie, Z.; Wang, X.; Chen, D.; Shen, G. Synthesis, characterizations and improved gas-sensing performance of SnO2 nanospike arrays. J. Mater. Chem. 2011, 21, 19086–19092. [Google Scholar] [CrossRef]
  28. Zhang, J.; Guo, J.; Xu, H.; Cao, B. Reactive-template fabrication of porous SnO2 nanotubes and their remarkable gas-sensing performance. ACS Appl. Mater. Interfaces 2013, 5, 7893–7898. [Google Scholar] [CrossRef] [PubMed]
  29. Yang, R.; Zhao, W.; Zheng, J.; Zhang, X.; Li, X. One-step synthesis of carbon-coated tin dioxide nanoparticles for high lithium storage. J. Phys. Chem. C 2010, 114, 20272–20276. [Google Scholar] [CrossRef]
  30. Tomer, V.K.; Duhan, S. In-situ synthesis of SnO2/SBA-15 hybrid nanocomposite as highly efficient humidity sensor. Sens. Actuators B Chem. 2015, 212, 517–525. [Google Scholar] [CrossRef]
  31. Ma, J.; Lei, D.; Duan, X.; Li, Q.; Wang, T.; Cao, A.; Mao, Y.; Zheng, W. Designable fabrication of flower-like SnS2 aggregates with excellent performance in lithium-ion batteries. RSC Adv. 2012, 2, 3615–3617. [Google Scholar] [CrossRef]
  32. Gu, D.; Li, X.; Zhao, Y.; Wang, J. Enhanced NO2 sensing of SnO2/SnS2 heterojunction based sensor. Sens. Actuators B Chem. 2017, 244, 67–76. [Google Scholar] [CrossRef]
  33. Wan, W.; Li, Y.; Ren, X.; Zhao, Y.; Gao, F.; Zhao, H. 2D SnO2 nanosheets: Synthesis, characterization, structures, and excellent sensing performance to ethylene glycol. Nanomaterials 2018, 8, 112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Aly, K.; Abousehly, A.; Othman, A. Photoelectrical properties of (Sb15As30Se55)100−xTex (0 ≤ x ≤ 12.5 at.%) thin films. J. Non-Cryst. Solids 2008, 354, 909–915. [Google Scholar] [CrossRef]
  35. Heiland, G.; Kohl, D.; Seiyama, T. (Eds.) Chemical Sensor Technology, 1st ed.; Kodansha: Tokyo, Japan, 1988; Volume 1, p. 15. [Google Scholar]
  36. Xu, H.Y.; Chen, Z.R.; Liu, C.Y.; Ye, Q.; Yang, X.P.; Wang, J.Q.; Cao, B.Q. Preparation of {200} crystal faced SnO2 nanorods with extremely high gas sensitivity at lower temperature. Rare Met. 2021, 40, 2004–2016. [Google Scholar] [CrossRef]
  37. Li, T.; Li, L.; Sun, H.; Xu, Y.; Wang, X.W.; Luo, H.; Liu, Z.; Zhang, T. Porous ionic membrane based flexible humidity sensor and its multifunctional applications. Adv. Sci. 2017, 4, 1600404. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (Color online) (a) SEM image of SnS2 film and (b) SnO2 film oxidized by hydrogen peroxide (H2O2); (c) X-ray diffraction (XRD) patterns of SnS2 (red) and SnO2 (blue) films and SnS2 and SnO2 of the JCPDS database (black); (d) XPS spectra of SnS2 (black) and SnO2 (brown) films. The inset is the photoconductivity measurement of SnS2 film with band gap of 2.32 eV indicated by the dashed line.
Figure 1. (Color online) (a) SEM image of SnS2 film and (b) SnO2 film oxidized by hydrogen peroxide (H2O2); (c) X-ray diffraction (XRD) patterns of SnS2 (red) and SnO2 (blue) films and SnS2 and SnO2 of the JCPDS database (black); (d) XPS spectra of SnS2 (black) and SnO2 (brown) films. The inset is the photoconductivity measurement of SnS2 film with band gap of 2.32 eV indicated by the dashed line.
Crystals 11 01028 g001
Figure 2. (Color online) (a) Scheme of photocurrent measurement: (b) and (c) 2D photocurrent mapping of SnS2 films with interdigitated electrodes of 600 and 800 μm spacing, respectively. Device optical image is in the upper right corner.
Figure 2. (Color online) (a) Scheme of photocurrent measurement: (b) and (c) 2D photocurrent mapping of SnS2 films with interdigitated electrodes of 600 and 800 μm spacing, respectively. Device optical image is in the upper right corner.
Crystals 11 01028 g002aCrystals 11 01028 g002b
Figure 3. (Color online) (a) Frequency response of photoconductivity and time constant fitting of SnS2 film with different finger spacing; (b) the relative balance [(Imax − Imin)/Imax] versus frequency of SnS2 film with different finger spacing. (c) Laser power and (d) bias voltage-dependent photoresponsivity of SnS2 film with different finger spacing.
Figure 3. (Color online) (a) Frequency response of photoconductivity and time constant fitting of SnS2 film with different finger spacing; (b) the relative balance [(Imax − Imin)/Imax] versus frequency of SnS2 film with different finger spacing. (c) Laser power and (d) bias voltage-dependent photoresponsivity of SnS2 film with different finger spacing.
Crystals 11 01028 g003
Figure 4. (Color online) (a) Scheme of humidity sensing system; (b,c) humidity response and recovery time of SnO2 film with 700 and 900 μm electrode spacing, respectively; the bias voltage of 10 V has been applied in this test and the two humidity levels used for this test were 15% and 65%, respectively; (d) relativity humidity and (e) bias voltage-dependent sensitivity of SnO2 film with different finger spacing; (f) relativity humidity dependent resistance of SnO2 film with different finger spacing; (g) repeatability measurement of three sensor devices upon exposure to human exhalation.
Figure 4. (Color online) (a) Scheme of humidity sensing system; (b,c) humidity response and recovery time of SnO2 film with 700 and 900 μm electrode spacing, respectively; the bias voltage of 10 V has been applied in this test and the two humidity levels used for this test were 15% and 65%, respectively; (d) relativity humidity and (e) bias voltage-dependent sensitivity of SnO2 film with different finger spacing; (f) relativity humidity dependent resistance of SnO2 film with different finger spacing; (g) repeatability measurement of three sensor devices upon exposure to human exhalation.
Crystals 11 01028 g004
Table 1. Fitting results of τ1, τ2, k1 and k2 for SnS2 film with 600 and 800 μm electrode spacing.
Table 1. Fitting results of τ1, τ2, k1 and k2 for SnS2 film with 600 and 800 μm electrode spacing.
Sample600 μm800 μm
k10.640.96
k20.360.04
τ1 (s)4.54 × 10−35.26 × 10−3
τ2 (s)2.41 × 10−49.38 × 10−5
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lin, D.-Y.; Hsu, H.-P.; Hu, H.-S.; Yang, Y.-C.; Lin, C.-F.; Zhou, W. Humidity Sensing and Photodetection Based on Tin Disulfide Nanosheets. Crystals 2021, 11, 1028. https://doi.org/10.3390/cryst11091028

AMA Style

Lin D-Y, Hsu H-P, Hu H-S, Yang Y-C, Lin C-F, Zhou W. Humidity Sensing and Photodetection Based on Tin Disulfide Nanosheets. Crystals. 2021; 11(9):1028. https://doi.org/10.3390/cryst11091028

Chicago/Turabian Style

Lin, Der-Yuh, Hung-Pin Hsu, Han-Sheng Hu, Yu-Cheng Yang, Chia-Feng Lin, and Wei Zhou. 2021. "Humidity Sensing and Photodetection Based on Tin Disulfide Nanosheets" Crystals 11, no. 9: 1028. https://doi.org/10.3390/cryst11091028

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop