Next Article in Journal
Acoustic Denoising Using Artificial Intelligence for Wood-Boring Pests Semanotus bifasciatus Larvae Early Monitoring
Previous Article in Journal
A Deep Learning Approach for Gait Event Detection from a Single Shank-Worn IMU: Validation in Healthy and Neurological Cohorts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Monitoring Gases Content in Modern Agriculture: A Density Functional Theory Study of the Adsorption Behavior and Sensing Properties of CO2 on MoS2 Doped GeSe Monolayer

1
College of Engineering and Technology, Southwest University, Chongqing 400716, China
2
School of Computer Science & Technology, Beijing Institute of Technology, Beijing 100081, China
*
Author to whom correspondence should be addressed.
Sensors 2022, 22(10), 3860; https://doi.org/10.3390/s22103860
Submission received: 3 May 2022 / Revised: 14 May 2022 / Accepted: 18 May 2022 / Published: 19 May 2022
(This article belongs to the Section Nanosensors)

Abstract

:
The reasonable allocation and control of CO2 concentration in a greenhouse are very important for the optimal growth of crops. In this study, based on density functional theory (DFT), an MoS2–GeSe monolayer was proposed to unravel the issues of the lower selectivity, poorer sensitivity and non-recyclability of traditional nanomaterial gas sensors. The incorporation of MoS2 units greatly enhanced the sensitivity of the pure GeSe monolayer to CO2 and the high binding energy also demonstrated the thermal stability of the doped structures. The ideal adsorption energy, charge transfer and recovery time ensured that the MoS2–GeSe monolayer had a good adsorption and desorption ability. This paper aimed to solve the matter of recycling sensors within agriculture. This research could provide the theoretical basis for the establishment of a potentially new generation of gas sensors for the monitoring of crop growth.

1. Introduction

In modern agriculture, greenhouses are often used to artificially change the growth environment of crops in order to cope with random and changeable weather conditions and ensure the diversity of seasonal agricultural products [1,2,3]. In the growth process of plants, carbon is an important organic matter, mainly in the form of CO2 [4,5]. CO2 is a necessary material for plants, which plays a very important role in photosynthesis [6,7,8]. It directly affects the normal growth of plants and the abundance of nutrients within plants. In addition, the concentration of CO2 can change the biological and abiotic conditions in soil to some extent [9,10]. Increases in the emissions of potent greenhouse gases from soil under increasing atmospheric CO2 indirectly affect the growth of crops. Moreover, the demand for CO2 is different in different growth stages of crops. For example, in this study, demand for CO2 reached the maximum value of 1000–1200 ppm before sunrise and then decreased to about 100 ppm 2.5–3 h after sunrise, which was only about 30% of the atmospheric concentration. It did not begin to rise again until 2 h after sunrise and returned to the atmospheric level around 4 p.m. The concentration of carbon dioxide that is required for vegetables to grow is generally 1000~1500 ppm, so CO2 deficiency in a greenhouse is quite serious and has become an important factor that affects the yield of greenhouse vegetables. Therefore, the reasonable allocation and control of CO2 concentration in greenhouses are critical for the optimal growth of crops. Zhang et al. have conducted a lot of research on the study of density functional theory (DFT) [11,12,13,14,15,16]. They studied the microscopic interactions between two-dimensional (2D) materials and gas molecules to explore the application of these materials as gas adsorbents and sensors, such as the use of an Au-doped graphene monolayer as a gas sensor [13,16]. Other relevant work has proved that the simulation calculation is very consistent with experimental results. Based on DFT, we theoretically designed a new type of CO2 concentration detection sensor, which can be indirectly observed and used to control the growth of plants.
Compared to traditional gas sensors, the use of nanomaterial sensors for the detection of gas type and gas content has the advantages of small size, fast response, high sensitivity and low costs. Tao et al. explored the use of nanomaterial sensors, both in theory and experimentally [17,18,19,20,21]. Thanks to their large surface areas and adequate adsorption sites, 2D materials are increasingly being used as gas adsorbents and sensors and were studied by Gui et al. [22,23,24,25]. After graphene, GeSe monolayers have attracted a lot of interest as 2D materials with high optical responsiveness, high mobility and high flexibility. Zhou et al. also systematically explored the physical and chemical properties of a GeSe monolayer [26,27,28] and Dai et al. successfully prepared GeSe thin films [29]. Recently, several researchers have explored the use of GeSe monolayers as gas sensors or adsorbents. Liu et al. studied the gas adsorption characteristics of an intrinsic GeSe monolayer [30], but due to its weak adsorption ability, it could not provide enough of a theoretical basis for gas sensors. Tao et al. explored the adsorption characteristics of a modified GeSe monolayer for SF6 decomposition gas by doping metal oxides [31,32,33]. However, despite being an important gas that is applied in various fields, CO2 lacks theoretical exploration using GeSe-based monolayers. Li et al. found that the adsorption ability of Ce–CaO to CO2 is too strong to be desorbed, which is not conducive to the recyclability of sensors [34,35,36,37]. The adsorption ability of Al–BN monolayers and graphene to CO2 is also too small, which is the reason that they cannot obtain enough of a response [38,39]. Gui et al. proposed that metal oxide doping can effectively improve the gas selectivity and sensitivity of 2D materials [40,41], which offered the essential theoretical basis for our study.
Here, we innovatively propose the use of a MoS2–GeSe monolayer to detect CO2 for the first time. Based on the first principle, parameters such as binding energy (Eb), adsorption energy (Eads), charge transfer (ΔQ), band gap (Eg), density of states (DOS), work functions (WFs), charge density difference (CDD), deformation charge density (DCD) and electron localization function (ELF) were analyzed. Doping with MoS2 units caused the GeSe monolayer to have a higher sensitivity to CO2 and an excellent thermal stability (Eb = −5.994 eV). For the better simulation of the practical applications of gas sensors, we analyzed the recovery time at different temperatures. Appropriate Eads (−0.955 eV) and ΔQ (0.266 e) values caused the MoS2–GeSe monolayer to maintain an excellent response and also produce great recovery characteristics (1.25 s in 398 K). This could provide solid theoretical support for the application of resistive chemical sensors in agriculture.

2. Methods

All calculations were performed using the DMOL3 and CASTEP frameworks [42,43]. The generalized gradient approximation (GGA) and the Perdew–Burke–Ernzerhof (PBE) exchange correlation function were selected for analysis [44]. The energy convergence accuracy, maximum force and maximum displacement were set as 1 × 10 −5 Ha, 2 × 10 −3 Ha/Å and 5 × 10 −3 Ha/Å, respectively. Considering the weak interaction between atoms in calculations, such as the van der Waals force, Grimme’s DFT-D was applied [45,46,47,48]. Double numerical polarization (DNP) was also selected at the basic setting of 3.5.
A 4 × 4 × 1 GeSe monolayer supercell was constructed and optimized using the highest precision setting in the material studio software. The k-point samples of the Monkhorst–Pack grid were sampled at 6 × 6 × 1 for geometry optimization and 10 × 10 × 1 for electronic optimization. The size of the vacuum space between two adjacent sheets was set at the value of 15 Å in order to avoid interactions between periodic images. A self-consistent loop energy of 10−6 Ha, a global orbital cut-off radius of 5.0 Å and a smearing of 0.005 Ha were applied. In order to study the role of MoS2 doping in the adsorption of CO2 to the GeSe monolayer, we compared the most stable configurations of CO2 that were adsorbed to the MoS2–GeSe monolayer. Our strategy was to calculate all typical adsorption structures and then the maximum adsorption energy would correspond to the most stable adsorption configuration.
Eb was determined using Equation (1), as follows:
Eb = EMoS2–GeSeEGeSeEMoS2
where EMoS2–GeSe, EGeSe and EMoS2 represent the total energies of MoS2–GeSe, pure GeSe and MoS2 units, respectively.
Eads was calculated using Equation (2), as follows:
Eads = ECO2/MoS2–GeSeECO2EMoS2–GeSe
where ECO2/MoS2–GeSe, ECO2 and EMoS2–GeSe represent the energy of the total systems, CO2 and the separated systems, respectively.
The charge amount was obtained using the Milliken charge analysis and ΔQ was determined using Equation (3), as follows:
ΔQ = QaQb
where Qa and Qb represent the total charge of the gas molecules after and before adsorption, respectively.

3. Results

As shown in Figure 1d, CO2 is a common compound in air that is composed of two oxygen atoms and a carbon atom, which are attached by a polar covalent bond. In Figure 1a,e, the ball–stick model structure of the pure GeSe monolayer is shown. GeSe monolayer are serrated black crystals that are similar to black phosphorus and belong to the orthogonal crystal system. After structural optimization, the Ge–Se bond lengths of the most stable GeSe monolayers were maintained at 2.641 and 2.544 Å. As shown in Figure 1b,f, the MoS2 units were stably adsorbed to the GeSe monolayer surface and the Eb value reached up to −5.994 eV, which indicated that a stable doping structure was formed with great thermal stability. In addition, the introduction of MoS2 units caused the Ge atoms in the GeSe monolayer to shift upward, which was produced by the strong interaction between the MoS2 units and the GeSe monolayer. As shown in Figure 1c,g, a stable structure with the maximum Eads value was finally selected after the calculation of eight different sites. For the metal oxide materials, we found the four intrinsic structures of ZnO, SnO2, TiO2 and WO3 through a literature review and compared and explored the gas adsorption characteristics of CO2 to these common metal oxides. The specific parameters are shown in Table 1. It was obvious that the Eads values of several metal oxides were much lower than those of the MoS2–GeSe monolayer (−0.955 eV) using DFT calculations, which indicated that the MoS2–GeSe monolayer had excellent response characteristics for CO2 detection. For pure ZnO, WO3 had some positive Eads values, which indicated that the reaction could not be carried out spontaneously nor could that material be used as a sensor.
As shown in Figure 2a, the overall TDOS of the MoS2–GeSe monolayer shifted to the right after doping with MoS2 units, especially near the Fermi level, which indicated that the conductivity of the GeSe monolayer decreased. In addition, a new peak appeared at −6 eV, which lays the foundation for the gas adsorption reaction. After the CO2 adsorption by the MoS2–GeSe monolayer, the peak amplitudes at −8.5 eV, −4 eV, −3 eV and 1.7 eV in the TDOS of the system increased slightly, which depended on the small regional state coating. This reflected the direct weak interactions between the two-dimensional materials and gases. We performed a PDOS analysis of the MoS2–GeSe monolayer decomposition (Figure 2b). Obviously, the adsorbed CO2, doping MoS2 units and modified GeSe had very obvious overlapping peaks at −8 eV and−4 eV, which indicated the existence of multi-electron orbital hybridization. After adsorption, we also found that the gas molecules did not produce new hybrid orbitals but rather shifted left overall. Therefore, we could conclude that the doped metal oxide had a tiny effect on the charge redistribution of the gas. In other words, the adsorbed CO2 could not redistribute the surface charge of the MoS2–GeSe monolayer or the electrical conductivity of the system, which revealed the corresponding response mechanisms of the gas sensor. As shown in Figure 2d, all elements showed high levels of hybridization and the orbitals were distributed between −10 and 2 eV, where Ge–p, Se–p, O–p and C–p orbitals contributed greatly to the charge recombination. The insertion of the p orbitals of O atoms into the MoS2 units caused electronic transitions in the gas molecules near the Fermi level, which formed new coupled states. In short, this indicated that there was an interaction between CO2, the MoS2 units and the GeSe monolayer; however, it was generally a weak interaction, which conformed to the application prospects of the MoS2–GeSe monolayer as a gas sensor.
In Figure 3a,d, the gain and loss of electron density are indicated by the blue and red regions, respectively. From the DCD distribution and ΔQ values, the gas molecules contributed 0.266 e to the material surface as electron donors and the electron density changed significantly between the different distribution regions. The electron density of the O atoms in CO2 increased greatly, while the region around the C atom lost electrons. The analysis indicated that the C atom of the CO2 molecule provided electrons not only to the two O atoms, but also to the O atom in the doped MoS2 units. It was evident from this transfer that the adsorption reaction was both exothermic and self-generated, which guaranteed stable and efficient adsorption effects. Figure 3b,e represents the redistribution of the charge density of CDD after adsorption. The adsorbed CO2 was charged positively overall and had purple areas around it, which indicated strong electron drainage. The separated charge build-up and exhaustion also confirmed the weak interaction between the gas and the adsorbent, with no new chemical bonds taking shape. As shown in Figure 3c,f, the ELF values were between −1 and 1, where −1 and 1 stand for very low electron density and completely limited electron density, respectively. It was found that stable O–Ge and Mo–Se bonds were formed between the MoS2 units and GeSe monolayer, which further confirmed the stability of the doping structure. In addition, there was no significant regional fusion between the CO2 and the MoS2–GeSe monolayer, which suggested that the adsorption reaction was physical adsorption and showed the uniqueness of the MoS2–GeSe monolayer as a gas sensor that allows for a faster response and recovery.
The energy of electrons in solids cannot be continuously valued, but some discontinuous energy bands can be valued. To conduct electricity, there must be free electrons or holes. The energy of free electrons is called the conduction band and the energy of free holes is called the valence band. When the bound electrons become free electrons or holes, they must obtain enough energy from the valence band for the conduction band. The minimum energy is Eg. In short, Eg refers to the energy transfer between the lowest level of the conduction band and the highest level of the valence band [54,55,56]:
σ ∝ A × exp (−ΔEg/2kT)
where σ is the conductivity, k is the Boltzmann constant (1.38 × 10−23 J/K), A is a fixed parameter and T is the temperature. The change in Eg after an adsorption reaction can reflect the change in system conductivity and can further reflect the change in resistance value in practical applications. The Eg value of the MoS2–GeSe monolayer decreased after CO2 adsorption, which indicated that the conductivity of the GeSe-based monolayer increased and the resistance value decreased. As shown in Figure 4a,c, the doping with MoS2 units reduced the conductivity of the GeSe monolayer and increased the Eg value from 0.725 eV to 0.786 eV, which indicated that the MoS2–GeSe monolayer maintained the excellent semiconductor characteristics and gas response characteristics of pure GeSe. However, the Eg value of the pure GeSe monolayer after CO2 adsorption changed very little (from 0.725 eV to 0.753 eV), which indicated that CO2 had little effect on the electrical conductivity of the system (Figure 4b). In actual detection, it could be impossible to detect the resistance change of GeSe-based monolayer gas sensors using high precision instruments. The adsorption of CO2 reduced the Eg value of the MoS2–GeSe monolayer to 0.667 eV, which indicated that the difficulty of electron transition from the top of the valence band to the bottom of the conduction band was reduced (Figure 4d). This helped us to observe the resistance change that was caused by the gas concentration more clearly. In contrast to the pure GeSe monolayer, the MoS2–GeSe monolayer could be used as a gas sensor for CO2 detection.
WFs are a property of particular material surfaces, which indicate the minimum energy that is required for the surface to immediately move electrons to a point within a vacuum [57,58]. In gas interaction, this determines the aligned contact barrier between the gas molecules and specific surfaces. As shown in Figure 5a, the doping with MoS2 units increased the WFs of the GeSe monolayer to 4.653 eV, which hindered the electronic affinity of the system. This phenomenon further proved that the electronic transition of the GeSe monolayer was more difficult and that the chemical properties of the system were more stable after doping with MoS2 units. However, the introduction of CO2 reduced the WFs of the system to 4.544 eV, which indicated that after the CO2 was adsorbed, the difficulty of electron spillover from the vacuum reduced. This further guided the adsorption of the MoS2–GeSe monolayer to CO2. It also reflected the electrostatic attraction between the gas molecules and the adsorption system. It was this simultaneous attraction that led to the interaction between the CO2 and the MoS2–GeSe monolayer. In conclusion, gas adsorption could cause changes in WFs, which also proved the possibility of using a Kelvin probe microscope to detect WFs in the field of gas sensing. Zhang et al. also studied the process of gas adsorption using WFs [33,53,55,56,57,58,59,60,61]. Figure 5b presents the theoretical recovery time of CO2 at different temperatures that were calculated. The recovery times of sensing materials are closely related to the exposed ambient temperature. The calculation formula that we used was as follows [62,63]:
τ = v0−1 × exp (−Eads/kT)
where k is the Boltzmann constant, τ is the recovery time and v0 is the attempt frequency (1012 s−1). The thermal environment offers more energy for the desorption process and fosters the further diffusion of molecules. At 298 K, the recovery time of CO2 was 14400 s, which caused the sensor to become unrecyclable. At 398 K, the desorption time of CO2 decreased to 1.25 s, which further illustrated the excellent performance of the MoS2–GeSe monolayer as a gas sensor. We inferred that 398 K could be used as the optimum temperature for the sensor, which would make it possible to apply the MoS2–GeSe monolayer in agriculture.

4. Conclusions

In this work, the sensing properties of an MoS2–GeSe monolayer for CO2 detection were systematically studied based on DFT. Compared to the pure GeSe monolayer, the MoS2–GeSe monolayer had a higher sensitivity to CO2 and the variation in Eg increased by 661% (0.119/0.018 × 100%). In practical applications, the MoS2–GeSe monolayer would have a greater change in resistance values as a chemical resistance sensor than during testing. The main parameters are shown in Table 2. The large Eb value (−5.994 eV) indicated that the MoS2 units and GeSe monolayer had excellent binding abilities. The thermal stability of the doped system was robust enough and the sensor could deal with changes in the ambient operating temperature. As can be seen from the DOS, the enriched orbital hybridization demonstrated the interaction between the CO2 and the active atomic groups. The ideal Eads (−0.955 eV), ΔQ (0.266 e) and recovery time (1.25 s in 398 K) not only ensured the detection ability of the MoS2–GeSe monolayer as a sensor, but also ensured its excellent recyclability. This is an advantage that most two-dimensional material gas sensors do not have. Based on this theoretical investigation, we will proceed to test flexible patch gas sensors for the monitoring of crop growth.

Author Contributions

Conceptualization, X.G.; software, Y.L.; data curation, X.G.; writing—original draft preparation, X.G.; writing—review and editing, Y.L.; funding acquisition, Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Science and Technology Support Plan Project of Guizhou Province: [2019] No. 2384 and Science and Technology Support Plan Project of Guizhou Province: [2021] General 171.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Alvarado, K.A.; Mill, A.; Pearce, J.M.; Vocaet, A.; Denkenberger, D. Scaling of greenhouse crop production in low sunlight scenarios. Sci. Total Environ. 2020, 707, 136012. [Google Scholar] [CrossRef] [PubMed]
  2. Rodríguez, F.; Berenguel, M.; Guzmán, J.L.; Ramírez-Arias, A. Modeling and Control of Greenhouse Crop Growth; Springer: Berlin/Heidelberg, Germany, 2015. [Google Scholar]
  3. Timmermans, G.H.; Hemming, S.; Baeza, E.; Van Thoor, E.A.; Schenning, A.P.; Debije, M.G. Advanced optical materials for sunlight control in greenhouses. Adv. Opt. Mater. 2020, 8, 2000738. [Google Scholar] [CrossRef]
  4. Körner, C. Plant CO2 responses: An issue of definition, time and resource supply. New Phytol. 2006, 172, 393–411. [Google Scholar] [CrossRef] [PubMed]
  5. Pritchard, S.G.; Rogers, H.H.; Prior, S.A.; Peterson, C.M. Elevated CO2 and plant structure: A review. Glob. Change Biol. 1999, 5, 807–837. [Google Scholar] [CrossRef] [Green Version]
  6. Gerhart, L.M.; Ward, J.K. Plant responses to low [CO2] of the past. New Phytol. 2010, 188, 674–695. [Google Scholar] [CrossRef]
  7. Leakey, A.D.; Ainsworth, E.A.; Bernacchi, C.J.; Rogers, A.; Long, S.P.; Ort, D.R. Elevated CO2 effects on plant carbon, nitrogen, and water relations: Six important lessons from FACE. J. Exp. Bot. 2009, 60, 2859–2876. [Google Scholar] [CrossRef]
  8. Morison, J.; Lawlor, D. Interactions between increasing CO2 concentration and temperature on plant growth. Plant Cell Environ. 1999, 22, 659–682. [Google Scholar] [CrossRef] [Green Version]
  9. Bazzaz, F.; McConnaughay, K. Plant plant interactions in elevated CO2 environments. Aust. J. Bot. 1992, 40, 547–563. [Google Scholar] [CrossRef]
  10. Makino, A.; Mae, T. Photosynthesis and plant growth at elevated levels of CO2. Plant Cell Physiol. 1999, 40, 999–1006. [Google Scholar] [CrossRef]
  11. Cui, H.; Zhang, X.; Li, Y.; Chen, D.; Zhang, Y. First-principles insight into Ni-doped InN monolayer as a noxious gases scavenger. Appl. Surf. Sci. 2019, 494, 1267. [Google Scholar] [CrossRef]
  12. Cui, H.; Zhang, X.; Zhang, G.; Tang, J. Pd-doped MoS_2 monolayer: A promising candidate for DGA in transformer oil based on DFT method. Appl. Surf. Sci. 2019, 470, 1035–1042. [Google Scholar] [CrossRef]
  13. Cui, H.; Zheng, K.; Xie, Z.; Yu, J.; Chen, X. Tellurene Nanoflake-Based NO2 Sensors with Superior Sensitivity and a Sub-Parts-per-Billion Detection Limit. ACS Appl. Mater. Interfaces 2020, 12, 47704–47713. [Google Scholar] [CrossRef] [PubMed]
  14. Hao, C.; Guozhi, Z.; Xiaoxing, Z.; Ju, T. Rh-doped MoSe2 as toxic gas scavenger: A first-principles study. Nanoscale Adv. 2018, 1, 772–780. [Google Scholar]
  15. Heping, C.; Kai, Z.; Yingying, Z.; Huaiyu, Y.; Xianping, C. Superior Selectivity and Sensitivity of C3N Sensor in Probing Toxic Gases NO2 and SO2. IEEE Electron Device Lett. 2017, 39, 284–287. [Google Scholar]
  16. Zhang, X.; Yu, L.; Wu, X.; Hu, W. Experimental Sensing and Density Functional Theory Study of H2S and SOF2 Adsorption on Au-Modified Graphene. Adv. Sci. 2015, 2, 1500101. [Google Scholar] [CrossRef]
  17. Gui, Y.; Sun, H.; Wei, H.; Duan, S.; Tang, C.; Zhang, X. Effect of nickel doping on adsorption of SF6 decomposition products over MoS2 surface. JOM 2019, 71, 3971–3979. [Google Scholar] [CrossRef]
  18. Jiang, T.; He, Q.; Bi, M.; Chen, X.; Sun, H.; Tao, L. First-principles calculations of adsorption sensitivity of Au-doped MoS2 gas sensor to main characteristic gases in oil. J. Mater. Sci. 2021, 56, 13673–13683. [Google Scholar] [CrossRef]
  19. Peng, Z.; Tao, L.-Q.; Zheng, K.; Yu, J.; Wang, G.; Sun, H.; Zhu, C.; Zou, S.; Chen, X. Gas Sensor Based on Semihydrogenated and Semifluorinated h-BN for SF₆ Decomposition Components Detection. IEEE Trans. Electron Devices 2021, 68, 1878–1885. [Google Scholar] [CrossRef]
  20. Sun, H.; Gui, Y.; Wei, H.; Long, Y.; Wang, Q.; Tang, C. DFT study of SF6 decomposed products on Pd–TiO2: Gas sensing mechanism study. Adsorption 2019, 25, 1643–1653. [Google Scholar] [CrossRef]
  21. Xia, S.-Y.; Tao, L.-Q.; Jiang, T.; Sun, H.; Li, J. Rh-doped h-BN monolayer as a high sensitivity SF6 decomposed gases sensor: A DFT study. Appl. Surf. Sci. 2021, 536, 147965. [Google Scholar] [CrossRef]
  22. Liu, Z.; Gui, Y.; Xu, L.; Chen, X. Adsorption and sensing performances of transition metal (Ag, Pd, Pt, Rh, and Ru) modified WSe2 monolayer upon SF6 decomposition gases (SOF2 and SO2F2). Appl. Surf. Sci. 2022, 581, 152365. [Google Scholar] [CrossRef]
  23. Liu, Z.; Gui, Y.; Xu, L.; Chen, X. Adsorption and gas-sensing properties of Aun (n = 1–3) cluster doped MoTe2 for NH3, NO2, and SO2 gas molecules. Surf. Interfaces 2022, 30, 101883. [Google Scholar] [CrossRef]
  24. Wang, X.; Gui, Y.; Ding, Z.; Xu, H.; Zeng, H.; Chen, X. Density functional theory study of Pd, Pt, and Au modified GeSe for adsorption and sensing of dissolved gases in transformer oil. Surf. Interfaces 2022, 31, 101994. [Google Scholar] [CrossRef]
  25. Wang, X.; Gui, Y.; Sun, N.; Ding, Z.; Chen, X. A DFT calculation: Gas sensitivity of defect GeSe to air decomposition products (CO, NO and NO2). IEEE Sens. J. 2022, 19, 905. [Google Scholar]
  26. Hu, Y.; Zhang, S.; Sun, S.; Xie, M.; Cai, B.; Zeng, H. GeSe monolayer semiconductor with tunable direct band gap and small carrier effective mass. Appl. Phys. Lett. 2015, 107, 122107. [Google Scholar] [CrossRef]
  27. Li, Z.; Liu, X.; Wang, X.; Yang, Y.; Liu, S.-C.; Shi, W.; Li, Y.; Xing, X.; Xue, D.-J.; Hu, J.-S. Strain-engineering the in-plane electrical anisotropy of GeSe monolayers. Phys. Chem. Chem. Phys. 2020, 22, 914–918. [Google Scholar] [CrossRef]
  28. Zhao, H.; Mao, Y.; Mao, X.; Shi, X.; Xu, C.; Wang, C.; Zhang, S.; Zhou, D. Band structure and photoelectric characterization of GeSe monolayers. Adv. Funct. Mater. 2018, 28, 1704855. [Google Scholar] [CrossRef]
  29. Lv, X.; Wei, W.; Mu, C.; Huang, B.; Dai, Y. Two-dimensional GeSe for high performance thin-film solar cells. J. Mater. Chem. A 2018, 6, 5032–5039. [Google Scholar] [CrossRef]
  30. Mao, Y.; Long, L.; Yuan, J.; Zhong, J.; Zhao, H. Toxic gases molecules (NH3, SO2 and NO2) adsorption on GeSe monolayer with point defects engineering. Chem. Phys. Lett. 2018, 706, 501–508. [Google Scholar] [CrossRef]
  31. Sun, H.; Tao, L.-Q.; Li, T.; Gao, X.; Sang, T.; Li, Y.; Wei, Y.; Wang, G.; Peng, Z.; Gui, Y. TiO2–Doped GeSe Monolayer: A highly selective gas sensor for SF6 decomposed species detection based on DFT method. Appl. Surf. Sci. 2022, 572, 151212. [Google Scholar] [CrossRef]
  32. Sun, H.; Tao, L.-Q.; Zhang, F.; Li, T.; He, R.; Gao, X.; Sang, T.; Wang, G.; Peng, Z.; Zhu, C. Metal Oxide Nanoparticles (XO, X= Cu, Zn, Ni) Doped GeSe Monolayer: Theoretical Exploration of a Novel H2S Gas Sensor for Health and Industrial Monitoring. IEEE Sens. J. 2021, 21, 26542–26548. [Google Scholar] [CrossRef]
  33. Wang, Y.; Li, T.; Peng, Y.; Gui, Y.; Sun, H. Pd and Pt decorated GeSe monolayers as promising materials for SOF2 and SO2F2 sensing. Appl. Surf. Sci. 2021, 560, 150028. [Google Scholar] [CrossRef]
  34. Kwon, S.; Choi, J.I.; Lee, S.G.; Jang, S.S. A density functional theory (DFT) study of CO2 adsorption on Mg-rich minerals by enhanced charge distribution. Comput. Mater. Sci. 2014, 95, 181–186. [Google Scholar] [CrossRef]
  35. Wu, C.; Yang, W.; Wang, J.; Kannaiyan, R.; Gates, I.D. CO2 adsorption and dissociation on single and double iron atomic molybdenum disulfide catalysts: A DFT study. Fuel 2021, 305, 121547. [Google Scholar] [CrossRef]
  36. Yan, X.; Li, Y.; Zhao, J.; Wang, Z. Density Functional Theory Study on CO2 Adsorption by Ce-Promoted CaO in the Presence of Steam. Energy Fuels 2020, 34, 6197–6208. [Google Scholar] [CrossRef]
  37. Ye, J.; Liu, C.; Ge, Q. DFT Study of CO2 Adsorption and Hydrogenation on the In2O3 Surface. J. Phys. Chem. C 2011, 116, 7817–7825. [Google Scholar] [CrossRef]
  38. Wang, C.; Fang, Y.; Duan, H.; Liang, G.; Long, M. DFT study of CO2 adsorption properties on pristine, vacancy and doped graphenes. Solid State Commun. 2021, 5948, 114436. [Google Scholar] [CrossRef]
  39. Wang, G.; Zheng, K.; Huang, Y.; Yu, J.; Tao, L.Q. Investigation of the positive effect of doping Al atom to the adsorption of CO2 on BN nanosheets: A DFT study. Phys. Chem. Chem. Phys. 2020, 22, 17. [Google Scholar] [CrossRef]
  40. He, X.; Gui, Y.; Liu, K.; Xu, L. Comparison of sensing and electronic properties of C2H2 on different transition metal oxide nanoparticles (Fe2O3, NiO, TiO2) modified BNNT (10, 0). Appl. Surf. Sci. 2020, 521, 146463. [Google Scholar] [CrossRef]
  41. He, X.; Gui, Y.; Xie, J.; Liu, X.; Wang, Q.; Tang, C. A DFT study of dissolved gas (C2H2, H2, CH4) detection in oil on CuO-modified BNNT. Appl. Surf. Sci. 2020, 500, 144030. [Google Scholar] [CrossRef]
  42. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.; Refson, K.; Payne, M.C. First principles methods using CASTEP. Z. Für Krist. Cryst. Mater. 2005, 220, 567–570. [Google Scholar] [CrossRef] [Green Version]
  43. Delley, B. Time dependent density functional theory with DMol3. J. Phys. Condens. Matter 2010, 22, 384208. [Google Scholar] [CrossRef] [PubMed]
  44. Grimme, S. Semiempirical GGA—Type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef] [PubMed]
  45. Antony, J.; Grimme, S. Density functional theory including dispersion corrections for intermolecular interactions in a large benchmark set of biologically relevant molecules. Phys. Chem. Chem. Phys. 2006, 8, 5287–5293. [Google Scholar] [CrossRef] [PubMed]
  46. Grimme, S. Density functional theory with London dispersion corrections. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2011, 1, 211–228. [Google Scholar] [CrossRef]
  47. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [Green Version]
  48. Moellmann, J.; Ehrlich, S.; Tonner, R.; Grimme, S. A DFT-D study of structural and energetic properties of TiO2 modifications. J. Phys. Condens. Matter 2012, 24, 424206. [Google Scholar] [CrossRef]
  49. Xia, Y.; Pan, A.; Su, Y.-Q.; Zhao, S.; Li, Z.; Davey, A.K.; Zhao, L.; Maboudian, R.; Carraro, C. In-situ synthesized N-doped ZnO for enhanced CO2 sensing: Experiments and DFT calculations. Sens. Actuators B Chem. 2022, 357, 131359. [Google Scholar] [CrossRef]
  50. Baltanás, M.A.; Reimers, W.G.; Branda, M.M. CO, CO2 and H2 adsorption on ZnO, CeO2, and ZnO/CeO2 surfaces: DFT simulations. J. Mol. Modeling 2014, 20, 2270-1–2270-10. [Google Scholar]
  51. Wang, D.; Chen, Y.; Liu, Z.; Ling, L.; Shi, C.; Qin, H.; Hu, J. CO2-sensing properties and mechanism of nano-SnO2 thick-film sensor. Sens. Actuators B Chem. 2016, 227, 73–84. [Google Scholar] [CrossRef]
  52. Chen, H.-Y.T.; Tosoni, S.; Pacchioni, G. A DFT study of the acid–base properties of anatase TiO2 and tetragonal ZrO2 by adsorption of CO and CO2 probe molecules. Surf. Sci. 2016, 652, 163–171. [Google Scholar] [CrossRef]
  53. Li, J.-H.; Wu, J.; Yu, Y.-X. DFT exploration of sensor performances of two-dimensional WO3 to ten small gases in terms of work function and band gap changes and I-V responses. Appl. Surf. Sci. 2021, 546, 149104. [Google Scholar] [CrossRef]
  54. Bundgaard, E.; Krebs, F.C. Low band gap polymers for organic photovoltaics. Sol. Energy Mater. Sol. Cells 2007, 91, 954–985. [Google Scholar] [CrossRef] [Green Version]
  55. Soukoulis, C.M. Photonic Band Gap Materials; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2012; Volume 315. [Google Scholar]
  56. Yablonovitch, E. Photonic band-gap structures. JOSA B 1993, 10, 283–295. [Google Scholar] [CrossRef]
  57. Lang, N.; Kohn, W. Theory of metal surfaces: Work function. Phys. Rev. B 1971, 3, 1215. [Google Scholar] [CrossRef]
  58. Michaelson, H.B. The work function of the elements and its periodicity. J. Appl. Phys. 1977, 48, 4729–4733. [Google Scholar] [CrossRef] [Green Version]
  59. Sun, H.; Tao, L.Q.; Li, T.; Gao, X.; Wang, G.; Peng, Z.; Zhu, C.; Zou, S.; Gui, Y.; Xia, S.Y.; et al. Sensing Characteristics of Toxic C₄F₇N Decomposition Products on Metallic- Nanoparticle Co-Doped BN Monolayer: A First Principles Study. IEEE Sens. J. 2021, 21, 13082–13089. [Google Scholar] [CrossRef]
  60. Zhou, Q.; Zhang, G.; Tian, S.; Zhang, X. First-Principles Insight into Pd-Doped ZnO Monolayers as a Promising Scavenger for Dissolved Gas Analysis in Transformer Oil. ACS Omega 2020, 5, 17801–17807. [Google Scholar] [CrossRef]
  61. Cui, H.; Yan, C.; Jia, P.; Cao, W. Adsorption and sensing behaviors of SF6 decomposed species on Ni-doped C3N monolayer: A first-principles study. Appl. Surf. Sci. 2020, 512, 145759. [Google Scholar] [CrossRef]
  62. Cui, H.; Jia, P.; Peng, X.; Li, P. Adsorption and sensing of CO and C2H2 by S-defected SnS2 monolayer for DGA in transformer oil: A DFT study. Mater. Chem. Phys. 2020, 249, 123006. [Google Scholar] [CrossRef]
  63. Chen, D.; Zhang, X.; Tang, J.; Cui, Z.; Cui, H. Pristine and Cu decorated hexagonal InN monolayer, a promising candidate to detect and scavenge SF6 decompositions based on first-principle study. J. Hazard. Mater. 2019, 363, 346–357. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Geometric structures of the GeSe monolayer (a,e), the MoS2–GeSe monolayer (b,f), the CO2/MoS2–GeSe monolayer (c,g) and CO2 (d). The red circle is MoS2 unit.
Figure 1. Geometric structures of the GeSe monolayer (a,e), the MoS2–GeSe monolayer (b,f), the CO2/MoS2–GeSe monolayer (c,g) and CO2 (d). The red circle is MoS2 unit.
Sensors 22 03860 g001
Figure 2. The TDOS and PDOS analyses of the CO2/MoS2–GeSe monolayer. The Fermi level is set at zero. (a) TDOS of CO2/MoS2–GeSe monolayer. (b) PDOS of CO2/MoS2–GeSe monolayer. (c) PDOS of two CO2. (d) PDOS of CO2/MoS2–GeSe monolayer with atoms.
Figure 2. The TDOS and PDOS analyses of the CO2/MoS2–GeSe monolayer. The Fermi level is set at zero. (a) TDOS of CO2/MoS2–GeSe monolayer. (b) PDOS of CO2/MoS2–GeSe monolayer. (c) PDOS of two CO2. (d) PDOS of CO2/MoS2–GeSe monolayer with atoms.
Sensors 22 03860 g002
Figure 3. DCD (a), CDD (b) and ELF (c) on the MoS2–GeSe monolayer and DCD (d), CDD (e) and ELF (f) on the CO2/MoS2–GeSe monolayer.
Figure 3. DCD (a), CDD (b) and ELF (c) on the MoS2–GeSe monolayer and DCD (d), CDD (e) and ELF (f) on the CO2/MoS2–GeSe monolayer.
Sensors 22 03860 g003
Figure 4. The Eg values of the pure GeSe monolayer (a), the CO2/GeSe monolayer (b), the MoS2–GeSe monolayer (c) and the CO2/MoS2–GeSe monolayer (d); the change trend of Eg (e).
Figure 4. The Eg values of the pure GeSe monolayer (a), the CO2/GeSe monolayer (b), the MoS2–GeSe monolayer (c) and the CO2/MoS2–GeSe monolayer (d); the change trend of Eg (e).
Sensors 22 03860 g004
Figure 5. Change trend of WFs (a); recovery time at different temperatures (b).
Figure 5. Change trend of WFs (a); recovery time at different temperatures (b).
Sensors 22 03860 g005
Table 1. The Eads values of different materials.
Table 1. The Eads values of different materials.
2D MaterialsGasEads (eV)Ref.
ZnOCO2+0.01[49]
ZnOCO2+0.44[50]
SnO2CO2−0.663[51]
TiO2CO2−0.17[52]
ZrO2CO2−0.05[52]
WO₃CO20.22[53]
Table 2. Main parameters.
Table 2. Main parameters.
EbEadsΔQWFsRecovery Time
MoS2–GeSe Monolayer−5.994 eV//4.653 eV/
CO2/MoS2–GeSe Monolayer/−0.955 eV0.2664.544 eV14,400 s (298 K)
28.2 s (358 K)
1.25 s (398 K)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gao, X.; Li, Y. Monitoring Gases Content in Modern Agriculture: A Density Functional Theory Study of the Adsorption Behavior and Sensing Properties of CO2 on MoS2 Doped GeSe Monolayer. Sensors 2022, 22, 3860. https://doi.org/10.3390/s22103860

AMA Style

Gao X, Li Y. Monitoring Gases Content in Modern Agriculture: A Density Functional Theory Study of the Adsorption Behavior and Sensing Properties of CO2 on MoS2 Doped GeSe Monolayer. Sensors. 2022; 22(10):3860. https://doi.org/10.3390/s22103860

Chicago/Turabian Style

Gao, Xin, and Yunwu Li. 2022. "Monitoring Gases Content in Modern Agriculture: A Density Functional Theory Study of the Adsorption Behavior and Sensing Properties of CO2 on MoS2 Doped GeSe Monolayer" Sensors 22, no. 10: 3860. https://doi.org/10.3390/s22103860

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