Next Article in Journal
Modeling and Speed Tuning of a PMSM with Presence of Fissure Using Dragonfly Algorithm
Next Article in Special Issue
Study on Electromigration Effects and IMC Formation on Cu–Sn Films Due to Current Stress and Temperature
Previous Article in Journal
Nanoporous Activated Carbon Derived via Pyrolysis Process of Spent Coffee: Structural Characterization. Investigation of Its Use for Hexavalent Chromium Removal
Previous Article in Special Issue
New Ternary Compounds of the Composition Cu2SnTi3 and Their Crystal Structures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption of NO2 and H2S on ZnGa2O4(111) Thin Films: A First-Principles Density Functional Theory Study

1
Center for General Education, China Medical University, Taichung 404, Taiwan
2
Graduate Institute of Precision Engineering, National Chung Hsing University, Taichung 402, Taiwan
*
Author to whom correspondence should be addressed.
Appl. Sci. 2020, 10(24), 8822; https://doi.org/10.3390/app10248822
Submission received: 4 November 2020 / Revised: 29 November 2020 / Accepted: 5 December 2020 / Published: 9 December 2020
(This article belongs to the Special Issue Selected Papers from ISET 2020 and ISPE 2020)

Abstract

:

Featured Application

This work offers a detailed description of gas-sensing performance of ZnGa2O4-based gas sensors.

Abstract

We performed first-principles total-energy density functional calculations to study the reactions of NO2 and H2S molecules on Ga–Zn–O-terminated ZnGa2O4(111) surfaces. The adsorption reaction and work functions of eight NO2 and H2S adsorption models were examined. The bonding of the nitrogen atom from a single NO2 molecule to the Ga atom of the Ga–Zn–O-terminated ZnGa2O4(111) surfaces exhibited a maximum work function change of +0.97 eV. The bond joining the sulfur atom from a single H2S molecule and the Ga atom of Ga–Zn–O-terminated ZnGa2O4(111) surfaces exhibited a maximum work function change of −1.66 eV. Both results concur with previously reported experimental observations for ZnGa2O4-based gas sensors.

1. Introduction

Household and industrial gas sensors are of great significance in artificial intelligence systems [1]. However, several key problems and challenges persist in the development of sensing components. The sensor has an operating temperature that is too high for it to be used as a wearable device; moreover, wearable devices are subject to moisture, which weakens their sensing efficiency. Other problems are the difficulty of distinguishing the composition and concentration of the mixing gases and the inaccuracy of sensing results. Metal oxide compounds, such as tin oxide (SnO2), gallium oxide (Ga2O3), and titanium dioxide (TiO2) semiconductor materials, have excellent potential for sensing applications and can detect harmful and toxic gases in the temperature range of 300–550 °C [2]. Kolmakov et al., reported that tin oxide nanowire sensors can sense the oxidizing gas of oxygen (O2) and the reductive gas of carbon monoxide (CO) at temperatures between 200 and 280 °C [3]. Schwebel et al., reported that β-Ga2O3 films can be used to sense reductive gases such as CO, hydrogen (H2), methane (CH4), nitric oxide (NO), and ammonia (NH3) [4]. Liu et al., reported that Ga2O3 nanowire gas sensors exhibit a reversible response to the oxidizing gases of O2 and reductive gas of CO in a working temperature range of 100–500 °C [5]. The maximum response of the Ga2O3 nanowire gas sensors for CO gas is four times larger than that for hydrogen, ammonia, or hydrogen sulfide (H2S) gas.
N-type semiconductors of ZnGa2O4 thin films have been developed, and ZnGa2O4 materials have been successfully used as gas sensors. Satyanarayana et al., reported that spinel ZnGa2O4 films can be used to sense liquid petroleum gas (LPG) at temperatures ranging from 200 to 400 °C. ZnGa2O4 sensors doped with palladium at approximately 320 °C have high selectivity to LPG, but poor sensitivity to CH4 and CO [6]. Furthermore, Jiao et al., reported the use of ZnGa2O4 prepared by spraying coprecipitation for sensing LPG, CO, ethanol (C2H5OH), and methane [7]. A novel approach described in Chen et al., (2010) [8] is based on the use of ZnGa2O4/ZnO core–shell nanowires. Sensitivity to nitrogen dioxide (NO2) gas can be considerably improved by using ZnGa2O4/ZnO core–shell nanowires rather than pure ZnO nanowires. Despite the introduction of new approaches to developing gas sensors, the fundamental interaction between gas molecules and metal oxide compounds remains unclear.
Density functional theory (DFT) calculations have been widely used in the fields of atomic and molecular adsorption on surfaces and the gas-sensing mechanism [9,10]. For example, water adsorption on the carbon nanotube field-effect transistor (CNTFET) explains the humidity-induced hysteresis, and it can be reasoned that water adsorption on the CNTFET leads to an increased chemical activity, the modification of the Schottky barrier, and the adsorption of some ionic substances in the vicinity of the carbon nanotube [9]. The gas-sensing mechanism of ZnO applied to H2, NH3, CO, and C2H5OH was established by conducting first-principles DFT calculations [10]. When the gas molecules, i.e., H2, NH3, CO, and C2H5OH molecules, become incident upon the ZnO(10-10) surface, the adsorption-induced reconstruction of the ZnO surface and charge transfer from the gas molecules to the ZnO surface control the sensing process evaluated by the change of electronic conductance of ZnO. Vorobyeva et al., reported that the resistance response of ZnO at operating temperature 450 °C increased (decreased) in the presence of NO2 (H2S) [11]. Moreover, the sensor responses of ZnO and ZnO with gallium contents of 0.5% and 4.0 at% show that an increase of gallium caused a monotonous decrease of the sensor response to H2S due to the enhanced electron-donor ability of surface oxygen anions, the H–S bond-breaking in the H2S molecule, as well as the decrease of the H2S adsorption. Recently, we demonstrated epitaxial growth of ZnGa2O4 thin film grown on the sapphire substrate using a metalorganic chemical vapor deposition (MOCVD) technique that yields a high-selectivity gas sensor [12]. The ZnGa2O4 gas sensor has superior selectivity to NO at the operating temperature of 300 °C. Reactions of NO molecules on Ga–Zn–O-terminated ZnGa2O4(111) surfaces were modeled and carried out using a first-principles density functional theory method. The NO molecules combine with the gallium atoms on the ZnGa2O4(111) surface to produce N–Ga bonds, leading to the work function changes. The sensor response can be gained from the changes in the work functions, indicating that the N–Ga bonding exhibits a very sensitive adsorption response. In this study, we developed adequate models for accurately predicting toxic NO2 oxidizing gases and H2S reducing gases adsorbed on the ZnGa2O4(111) surface, which offer a detailed description of gas-sensing performance of the use of ZnGa2O4-based thin-film sensors, which may be probed experimentally using phenomenological techniques of sensor characterization such as chemical components, sensing layers, and surface modification by metal doping.

2. Computational Details

A series of ab initio calculations were performed to evaluate the adsorption reactions and work functions of NO2 and H2S on ZnGa2O4(111) surfaces. The ab initio theoretical results were implemented in the Vienna ab initio simulation package [13,14], and the exchange correlation function was predicted using the generalized gradient approximation (GGA) with the Perdew–Wang (PW91) correction [15,16]. The crystal structure of ZnGa2O4 is displayed in Figure 1. The space group for ZnGa2O4 is Fd-3m, which contains 56 atoms, comprising 8 Zn, 16 Ga, and 32 O atoms. The cutoff energy was set as 400 eV. To simulate the change of the work function with and without an NO2 or H2S molecule, we developed a supercell of ZnGa2O4 along the (111) direction. This supercell contained 112 atoms, and the vacuum was set to 20 Å. The stoichiometry of all supercells was fixed at Zn16Ga32O64. We used the preferred Ga–Zn–O-terminated surface of ZnGa2O4(111) with a surface energy of 0.10 eV/Å2 [17]. A gamma-centered 3 × 3 × 1 Monkhorst–Pack grid was used. This supercell was fully relaxed until the force acting on each atom was less than 0.001 eV/Å.
To determine the most favorable adsorption site of NO2 (H2S) molecules on the ZnGa2O4(111) surface, we labeled the surface atoms as Ga3c, Zn3c, O3c, and O4c in the top and side views of Figure 2.
We first calculated the work function using the formula [18]
Φ Z G O = E V A C E F
where Φ Z G O is the work function for the clean ZnGa2O4(111) surface and E V A C and E F are the energy of the vacuum and the Fermi energy, respectively. Similarly, when the NO2 (H2S) molecule was adsorbed on the ZnGa2O4(111) surface, we also calculated the work function Φ N O 2 and Φ H 2 S , which in turn determined the work function differences Δ Φ (eV) between Φ Z G O and Φ N O 2 ( Φ H 2 S ). The binding energy E B (eV/u. c.) is also an important factor determining gas-sensing performance and can be defined as
E B = E t o t a l E Z G O E N O 2 / H 2 S
where E t o t a l is the total energy of the NO2 (H2S) molecule adsorbed on the ZnGa2O4(111) surface, and E Z G O and E N O 2 / H 2 S are the total energies of the slab of ZnGa2O4(111) surface and the free or isolated NO2 (H2S) molecule, respectively.
First, we constructed eight adsorption models, denoted N1, N2, N3, N4, O1, O2, O3, and O4. Here, N (O) represents a nitrogen (oxygen) atom in NO2, and numbers “1”, “2”, “3”, and “4” respectively indicate the initial adsorbed Ga3c, Zn3c, O3c, and O4c atom on the ZnGa2O4(111) surface. The initial distance of the NO2 molecule to the ZnGa2O4(111) surface was set as the sum of the van der Waal radii of N (O) for the NO2 molecule and Ga (Zn or O) for surface atoms. Similarly, for the H2S molecule adsorbed on the ZnGa2O4(111) surface, we also developed eight adsorption models, denoted S1, S2, S3, S4, H1, H2, H3, and H4. Here, S (H) represents a sulfur (hydrogen) atom in H2S. The initial distance of the H2S molecule to the ZnGa2O4(111) surface was also set as the sum of the van der Waal radii of H (S) and Ga (Zn or O) for surface atoms. We performed structure optimization in each model until the force acting on each atom was less than 0.001 eV/Å, yielding optimized atomic structures.
The gas sensitivity could be determined from the ratio of the resistance in the presence of the investigated gas (Rg) to the resistance in the reference gas (Ra), which is usually air [19]. The relation of gas sensitivity to the work function difference Δ Φ is represented by the following equation:
Δ Φ = Δ X + k T ln ( R g / R a )
where Δ X denotes the change in electron affinity and kT denotes the product of the Boltzmann constant k and the temperature T. The upward bending of the surface band due to the adsorption of oxidizing gas on ZnGa2O4(111) results in the depletion of free charge carriers on the surface, thus leading to the formation of a region with high ohmic resistance or a positive work function difference. However, the resistance of ZnGa2O4(111) gas sensors decreases under exposure to reducing gases because of the downward bending of the surface band, resulting in a negative work function difference.

3. Results and Discussion

The most favorable configurations for NO2 on ZnGa2O4 (111) surfaces are shown in Figure 3. In Model N1, the nitrogen atom of the NO2 molecule was adsorbed on the Ga3c atom of the ZnGa2O4(111) surface to yield a N–Ga bond with a bond length of 2.01 Å. In Model N2, the NO2 molecule moved away from the originally adsorbed Zn3c atom to approach the neighboring Ga atom of the ZnGa2O4(111) surface to form an O–Ga bond with a bond length of 1.90 Å. In Model N3, the NO2 molecule moved away from the originally adsorbed O3c atom toward the neighboring Ga and Zn atoms of the ZnGa2O4(111) surface to form N–Ga and O–Zn bonds with bond lengths of 2.05 Å and 2.30 Å, respectively. In Model N4, the nitrogen (oxygen) atom of the NO2 molecule adsorbed on the Ga3c (Zn3c) atom of the ZnGa2O4(111) surface to form an N–Ga (O–Zn) bond with a bond length of 2.06 Å (2.21 Å). In Model O1, the two O atoms of the NO2 molecule adsorbed on the Ga3c atom of the ZnGa2O4(111) surface to yield two O–Ga bonds with bond lengths of 2.07 Å and 2.13 Å. Similarly, in Model O2 (O3), the NO2 molecule moved away from the originally adsorbed Zn3c (O3c) atom toward the neighboring Ga3c atom of the ZnGa2O4(111) surface to form two O–Ga bonds with bond lengths of 2.10 (2.13) Å and 2.11 (2.08) Å. In Model O4, an O atom of the NO2 molecule was adsorbed on the Ga3c atom of the ZnGa2O4(111) surface to form an O–Ga bond with a bond length of 1.92 Å. The vacuum energy, Fermi energy, work functions of ZnGa2O4(111) with and without NO2, and work function difference among NO2 adsorption models are presented in Table 1. The work function increased when the NO2 molecule was adsorbed on the ZnGa2O4(111) surface. The increase in work function ranged from +0.29 eV to +0.97 eV, which concurs with previously reported experimental results of similar resistance responses of ZnO(Ga) samples to NO2 [8,11]. The NO2 molecule preferred bonding with the Ga3c or Zn3c atom. Model N1 had the maximum work function difference of +0.97 eV, exhibiting increased sensitivity for detecting NO2 molecules.
Figure 4 displays the lowest energy configurations for H2S on ZnGa2O4(111) surfaces. In Models S1 and S2, the sulfur atom of the H2S molecule adsorbed on the Ga3c and Zn3c atoms of the ZnGa2O4(111) surface to yield S–Ga and S–Zn bonds with bond lengths of 2.46 Å and 2.50 Å, respectively. In Model S3, the H2S molecule adsorbed on the O3c atom of the ZnGa2O4(111) surface and dissociated to HS and H+ ions to form H–O and S–Ga bonds with bond lengths of 0.97 Å and 2.22 Å, respectively. In Model S4, the H2S molecule moved away from the originally adsorbed O4c atom of the ZnGa2O4(111) surface to form an S–Ga bond with a bond length of 2.46 Å. In Models H1 and H2, the H2S molecules adsorbed on Ga3c and Zn3c atoms of the ZnGa2O4(111) surface and dissociated to H+ and HS ions to form an H–O bond and an S–Ga bond with bond lengths of 0.97 Å and 2.22 Å, respectively. In Model H3, the hydrogen atom of the H2S molecule was adsorbed on the O3c atom of the ZnGa2O4(111) surface to yield an H–O bond with a bond length of 1.25 Å. In Model H4, the H2S molecule adsorbed on the O4c atom of the ZnGa2O4(111) surface. However, its adsorption was unstable and therefore not ideal, with atomic distances of 3.12 Å and 2.66 Å for H–Ga and H–Zn, respectively.
The vacuum energy, Fermi energy, work functions of ZnGa2O4(111) with and without H2S, and work function difference among H2S adsorption models are listed in Table 2. The work function difference among Models S1, S2, and S4 decreased when the H2S molecules adsorbed on the Ga3c and Zn3c atoms of the ZnGa2O4(111) surface. The decrease in work function ranged from −0.31 to −1.66 eV, which concurs with the experimental results of other studies showing similar resistance responses of ZnO(Ga) samples to H2S [11]. Similarly, the H2S molecule preferred bonding with the Ga3c or Zn3c atom. Model S1 had the maximum work function difference of −1.66 eV, showing increased sensitivity for detecting H2S molecules. Models S3, H1, H2, H3, and H4 had positive work function differences, possibly because of the heterolytic break of the H–S bond in the H2S, resulting in the tendency of forming H–O bonds on the surface of ZnGa2O4. The H–S bond-breaking leading to the decrease of the H2S adsorption might explain the observed change in the decreased sensitivity for detecting H2S molecules, in agreement with the experimental observation [11].

4. Conclusions

The adsorption reactions and work functions of NO2 and H2S on ZnGa2O4(111) surfaces were studied using first-principles DFT–GGA calculations. Our results showed that the bonding of the nitrogen (sulfur) atom from a single NO2 (H2S) molecule to the Ga atom of Ga–Zn–O-terminated ZnGa2O4(111) surfaces exhibited the highest work function change of +0.97 eV (−1.66 eV). Experiments on the resistance responses of ZnO(Ga) samples to NO2 (H2S) molecules revealed that sensitivity responses to NO2 (H2S) molecules of ZnGa2O4-based thin-film sensors exhibit the same trend as that of positive (negative) work function differences [11]. In our favorable configuration, both NO2 and H2S molecules preferred bonding with the Ga3c atom of Ga–Zn–O-terminated ZnGa2O4(111) surfaces. The results demonstrate the sensitivity responses to NO2 and H2S molecules of ZnGa2O4-based thin-film sensors, which concur with experimental observations of ZnGa2O4-based gas sensors. The thin-film pretreatment technology, such as the thin-film coatings resulting in a Ga-terminated surface, could enhance the sensitivity responses of future gas-sensing devices.

Author Contributions

Conceptualization, J.-C.T. and P.-L.L.; methodology, Y.-H.C. and D.-Y.W.; validation, Y.-H.C. and D.-Y.W.; formal analysis, J.-C.T., Y.-H.C., D.-Y.W., and P.-L.L.; investigation, Y.-H.C. and D.-Y.W.; writing—original draft preparation, J.-C.T. and P.-L.L.; writing—review and editing, J.-C.T. and P.-L.L.; supervision, P.-L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology (MOST), Taiwan, grant numbers 109-2221-E-005-042 and 108-2221-E-005-001.

Acknowledgments

Computational studies were performed using the resources of the National Center for High Performance Computing, Taiwan. This manuscript was edited by Wallace Academic Editing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yamazoe, N. Toward Innovations of Gas Sensor Technology. Sens. Actuators B Chem. 2005, 108, 2–14. [Google Scholar] [CrossRef]
  2. Meixner, H.; Lampe, U. Metal Oxide Sensors. Sens. Actuators B Chem. 1996, 33, 198–202. [Google Scholar] [CrossRef]
  3. Kolmakov, A.; Zhang, Y.; Cheng, G.; Moskovits, M. Detection of CO and O2 using Tin Oxide Nanowire Sensor. Adv. Mater. 2003, 15, 997–1000. [Google Scholar] [CrossRef]
  4. Schwebel, T.; Fleischer, M.; Meixner, H. A Selective Temperature Compen-Sated O2 Sensor based on Ga2O3 Thin Films. Sens. Actuators B Chem. 2002, 65, 176–180. [Google Scholar] [CrossRef]
  5. Liu, Z.; Yamazaki, T.; Shen, Y.; Kikuta, T.; Nakatani, N.; Li, Y. O2 and CO Sensing of Ga2O3 Multiple Nanowire Gas Sensors. Sens. Actuators B Chem. 2008, 129, 666–670. [Google Scholar] [CrossRef]
  6. Satyanarayana, L.; Reddy, C.V.G.; Manorama, S.V.; Rao, V.J. Liquid-Petroleum-Gas Sensor Based on a Spinel Semiconductor ZnGa2O4. Sens. Actuators B Chem. 1998, 46, 1–7. [Google Scholar] [CrossRef]
  7. Jiao, Z.; Ye, G.; Chew, F.; Li, M.; Liu, J. The Preparation of ZnGa2O4 Nano Crystals by Spray Coprecipitation and its Gas Sensitive Characteristics. Sensors 2002, 2, 71–78. [Google Scholar] [CrossRef] [Green Version]
  8. Chen, I.-C.; Lin, S.-S.; Lin, T.-J.; Hsu, C.-L.; Hsueh, T.J.; Shieh, T.-Y. The Assessment for Sensitivity of a NO2 Gas Sensor with ZnGa2O4/ZnO Core-Shell Nanowires—a Novel Approach. Sensors 2010, 10, 3057–3072. [Google Scholar] [CrossRef] [PubMed]
  9. Sung, D.; Hong, S.; Kim, Y.-H.; Park, N.; Kim, S.; Maeng, S.L.; Kim, K.-C. Ab initio study of the Effect of Water Adsorption on the Carbon Nanotube Field-Effect Transistor. Appl. Phys. Lett. 2005, 87, 243110. [Google Scholar] [CrossRef] [Green Version]
  10. Yuan, Q.; Zhao, Y.-P.; Li, L.; Wang, T. Ab Initio Study of ZnO-Based Gas-Sensing Mechanisms: Surface Reconstruction and Charge Transfer. J. Phys. Chem. 2009, C113, 6107–6113. [Google Scholar] [CrossRef] [Green Version]
  11. Vorobyeva, N.; Rumyantseva, M.; Filatova, D.; Konstantinova, E.; Grishina, D.; Abakumov, A.; Turner, S.; Gaskov, A. Nanocrystalline ZnO(Ga): Paramagnetic Centers, Surface Acidity and Gas Sensor Properties. Sens. Actuators B Chem. 2013, 182, 555–564. [Google Scholar] [CrossRef]
  12. Wu, M.-R.; Li, W.-Z.; Tung, C.-Y.; Huang, C.-Y.; Chiang, Y.-H.; Liu, P.-L.; Horng, R.-H. NO gas sensor based on ZnGa2O4 epilayer grown by metalorganic chemical vapor deposition. Sci. Rep. 2019, 9, 7459. [Google Scholar] [CrossRef] [PubMed]
  13. Kresse, G.; Furthmüller, J. Efficiency of Ab-initio Total Energy Calculations for Metals and Semiconductors using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  14. Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab initio Total-Energy Calculations using a Plane-Wave basis Set. Phys. Rev. 1996, B54, 11169–11186. [Google Scholar] [CrossRef] [PubMed]
  15. Kresse, G.; Hafner, J. Norm-Conserving and Ultrasoft Pseudopotentials for First-Row and Transition Elements. J. Phys. Condens. Matter 1997, 6, 8245–8257. [Google Scholar] [CrossRef]
  16. Perdew, J.P.; Wang, Y. Accurate and Simple Density Functional for the Electronic Exchange Energy: Generalized Gradient Approximation. Phys. Rev. 1986, B3, 8800–8802. [Google Scholar] [CrossRef] [PubMed]
  17. Jia, C.; Fan, W.; Yang, F.; Zhao, X.; Sun, H.; Li, P.; Liu, L. A Theoretical Study of Water Adsorption and Decomposition on Low-Index Spinel ZnGa2O4 Surfaces: Correlation between Surface Structure and Photocatalytic Properties. Langmuir 2013, 29, 7025–7037. [Google Scholar] [CrossRef] [PubMed]
  18. Kahn, A. Fermi Level Work Function and Vacuum Level. Mater. Horiz. 2016, 3, 7–10. [Google Scholar] [CrossRef]
  19. Sahm, T.; Gurlo, A.; Barsan, N.; Weimar, U. Basics of oxygen and SnO2 Interaction Work Function Change and Conductivity Measurements. Sens. Actuators B Chem. 2006, 118, 78–83. [Google Scholar] [CrossRef]
Figure 1. Atomistic representation of bulk ZnGa2O4. Atoms are represented by spheres: Zn (purple, large), Ga (brown, medium-sized), and O (red, small).
Figure 1. Atomistic representation of bulk ZnGa2O4. Atoms are represented by spheres: Zn (purple, large), Ga (brown, medium-sized), and O (red, small).
Applsci 10 08822 g001
Figure 2. Top and side views of the ZnGa2O4(111) surface. Surface atoms are labeled Ga3c, Zn3c, O3c, and O4c. Atoms are represented by spheres: Zn (purple, large), Ga (brown, medium-sized), and O (red, small).
Figure 2. Top and side views of the ZnGa2O4(111) surface. Surface atoms are labeled Ga3c, Zn3c, O3c, and O4c. Atoms are represented by spheres: Zn (purple, large), Ga (brown, medium-sized), and O (red, small).
Applsci 10 08822 g002
Figure 3. Top and side views of the favorable configurations for NO2 on the ZnGa2O4(111) surface. Atoms are represented by spheres: Zn (purple, large), Ga (brown, large), S (yellow, medium-sized), O (red, medium-sized), and H (white, small). Equilibrium bond lengths between molecular and surface atoms are given in Angstrom.
Figure 3. Top and side views of the favorable configurations for NO2 on the ZnGa2O4(111) surface. Atoms are represented by spheres: Zn (purple, large), Ga (brown, large), S (yellow, medium-sized), O (red, medium-sized), and H (white, small). Equilibrium bond lengths between molecular and surface atoms are given in Angstrom.
Applsci 10 08822 g003
Figure 4. Top and side views of the most favorable configurations for H2S on the ZnGa2O4(111) surface. Atoms are represented by spheres: Zn (purple, large), Ga (brown, large), S (yellow, medium-sized), O (red, medium-sized), and H (white, small). Equilibrium bond lengths between molecular and surface atoms are given in Angstrom.
Figure 4. Top and side views of the most favorable configurations for H2S on the ZnGa2O4(111) surface. Atoms are represented by spheres: Zn (purple, large), Ga (brown, large), S (yellow, medium-sized), O (red, medium-sized), and H (white, small). Equilibrium bond lengths between molecular and surface atoms are given in Angstrom.
Applsci 10 08822 g004
Table 1. Calculated vacuum energy E V A C , Fermi energy E F , work function of NO2 Φ N O 2 , work function of ZnGa2O4(111) Φ Z G O , and work function difference Δ Φ ( Δ Φ = Φ N O 2 Φ Z G O ) among the eight NO2 adsorption models. All energies are presented in eV.
Table 1. Calculated vacuum energy E V A C , Fermi energy E F , work function of NO2 Φ N O 2 , work function of ZnGa2O4(111) Φ Z G O , and work function difference Δ Φ ( Δ Φ = Φ N O 2 Φ Z G O ) among the eight NO2 adsorption models. All energies are presented in eV.
ModelsAdsorption Sites E V A C E F Φ N O 2 Φ Z G O ΔΦ
ZnGa2O4(111)-0.66−3.38-4.04-
N1Ga3c1.45−3.565.01-+0.97
N2Ga3c1.04−3.294.33-+0.29
N3Ga3c, Zn3c1.12−3.314.43-+0.39
N4Ga3c, Zn3c1.09−3.314.40-+0.36
O1Ga3c0.98−3.414.39-+0.35
O2Ga3c1.01−3.364.37-+0.33
O3Ga3c1.07−3.314.38-+0.34
O4Ga3c1.04−3.294.33-+0.29
Table 2. Calculated vacuum energy EVAC, Fermi energy EF, work function of H2S Φ H 2 S , work function of ZnGa2O4(111) Φ Z G O , and work function difference ΔΦΦ = Φ H 2 S Φ Z G O ) among the eight H2S adsorption models. All energies are presented in eV.
Table 2. Calculated vacuum energy EVAC, Fermi energy EF, work function of H2S Φ H 2 S , work function of ZnGa2O4(111) Φ Z G O , and work function difference ΔΦΦ = Φ H 2 S Φ Z G O ) among the eight H2S adsorption models. All energies are presented in eV.
ModelsAdsorption Sites E V A C E F Φ H 2 S Φ Z G O ΔΦ
ZnGa2O4(111)-0.66−3.38-4.04-
S1Ga3c−0.37−2.752.38-−1.66
S2Zn3c0.54−3.193.73-−0.31
S3Ga3c, O3c0.91−3.324.23-+0.19
S4Ga3c−0.51−2.912.40-−1.64
H1Ga3c, O3c0.90−3.324.22-+0.18
H2Ga3c, O3c0.94−3.324.26-+0.22
H3O3c0.79−3.374.16-+0.12
H4-0.85−3.474.32-+0.28
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tung, J.-C.; Chiang, Y.-H.; Wang, D.-Y.; Liu, P.-L. Adsorption of NO2 and H2S on ZnGa2O4(111) Thin Films: A First-Principles Density Functional Theory Study. Appl. Sci. 2020, 10, 8822. https://doi.org/10.3390/app10248822

AMA Style

Tung J-C, Chiang Y-H, Wang D-Y, Liu P-L. Adsorption of NO2 and H2S on ZnGa2O4(111) Thin Films: A First-Principles Density Functional Theory Study. Applied Sciences. 2020; 10(24):8822. https://doi.org/10.3390/app10248822

Chicago/Turabian Style

Tung, Jen-Chuan, Yi-Hung Chiang, Ding-Yuan Wang, and Po-Liang Liu. 2020. "Adsorption of NO2 and H2S on ZnGa2O4(111) Thin Films: A First-Principles Density Functional Theory Study" Applied Sciences 10, no. 24: 8822. https://doi.org/10.3390/app10248822

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