Next Article in Journal
Multimodal Summarization of User-Generated Videos
Next Article in Special Issue
Strain Induced Topological Insulator Phase in CsPbBrxI3−x (x = 0, 1, 2, and 3) Perovskite: A Theoretical Study
Previous Article in Journal
A Wearable System for the Estimation of Performance-Related Metrics during Running and Jumping Tasks
Previous Article in Special Issue
Mini-Review: Recent Technologies of Electrode and System in the Enzymatic Biofuel Cell (EBFC)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influences of Work Function Changes in NO2 and H2S Adsorption on Pd-Doped ZnGa2O4(111) Thin Films: First-Principles Studies

1
Center for General Education, China Medical University, Taichung 40402, Taiwan
2
Master Program for Digital Health Innovation, China Medical University, Taichung 40402, Taiwan
3
Graduate Institute of Precision Engineering, National Chung Hsing University, Taichung 402, Taiwan
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(11), 5259; https://doi.org/10.3390/app11115259
Submission received: 30 April 2021 / Revised: 27 May 2021 / Accepted: 1 June 2021 / Published: 5 June 2021
(This article belongs to the Special Issue Selected Papers from ISET 2020 and ISPE 2020)

Abstract

:

Featured Application

This work provides a detailed analysis of the gas-sensing performance of Pd-doped ZnGa2O4-based gas sensors.

Abstract

The work function variations of NO2 and H2S molecules on Pd-adsorbed ZnGa2O4(111) were calculated using first-principle calculations. For the bonding of a nitrogen atom from a single NO2 molecule to a Pd atom, the maximum work function change was +1.37 eV, and for the bonding of two NO2 molecules to a Pd atom, the maximum work function change was +2.37 eV. For H2S adsorption, the maximum work function change was reduced from −0.90 eV to −1.82 eV for bonding sulfur atoms from a single and two H2S molecules to a Pd atom, respectively. Thus, for both NO2 and H2S, the work function change increased with an increase in gas concentration, showing that Pd-decorated ZnGa2O4(111) is a suitable material in NO2/H2S gas detectors.

1. Introduction

Gas sensors for household and industrial usages have attracted much attention since 1970 [1]. Toxic gases, such as hydrogen sulfide (H2S), sulfur dioxide (SO2), nitrogen dioxide (NO2), carbon monoxide (CO), and nitric oxide (NO), are harmful to plants, organisms, and people’s health. Therefore, finding a suitable sensor for toxic gases is an emerging and interesting research field. Experimentally, several key problems and challenges persist in the development of sensing components. For example, the sensor’s working or operating temperature may be too high or too low, which restricts its use. The size of the gas-sensing application is also a key factor for a wearable device. Moreover, distinguishing the component and concentration of a specified gas from mixing gases is difficult.
The realization of the sensor response to functional or target gases can significantly enhance applications of the biomimetic response and controlling, as well as autonomously release functional gases [2,3]. The conjugated bimetallopolymer with Pt(II)-acetylide and Ru(II)-porphyrin-pyridyl is polymerized with oligo (phenylene ethynylene) (OPE) backbones and insulated by permethylated α-cyclodextrins (PM α-CDs) [2]. The conjugated bimetallopolymer is depolymerized by Ru complexes, which behave as acceptors with carbon monoxide (CO) reactivity. The depolymerization of the non-radiative conjugated polymer to luminescent monomers of Pt-acetylide complexes is the first self-activation threshold for the activation of the system. The luminescent intensity increase with the increase in CO concentration is an autonomous response above the threshold. Approaching constant amounts of luminescent Pt-monomers is the second self-regulation threshold for the rate-determining step of the Ru complex. The dual self-controlling system of artificial biomimetic materials can be used as sensors, catalytic systems, and machines transporting materials.
The charge-compensating anions HS and NO2− and hydrated water (mH2O) in MgyAl(OH)2(y+1) (y is in the range of 2~4) layered double hydroxides (Mg/Al(y/1) LDHs) are used as the gas release system for medical applications [3]. Interlayer HS and NO2− in LDHs are protonated with aerial CO2 and H2O, releasing H2S or HNO2 under air. NO and NO2 can be further generated through an automatic disproportionation reaction of HNO2. LDHs (Mg/Al = 2/1 or 3/1) mixed with NaHS, Na2S, and NaNO2, i.e., NaHS-Mg/Al(2/1), NaHS-Mg/Al(3/1), Na2S-Mg/Al(2/1), Na2S-Mg/Al(3/1), NaNO2-Mg/Al(2/1), and NaNO2-Mg/Al(3/1) LDHs, have been successfully synthesized to release p.p.m.-level H2S and NO derived from aerial CO2-stimulated anion exchange, respectively. NaHS-Mg/Al(2/1) and NaNO2-Mg/Al(3/1) LDHs showed the best release profile for H2S and NO in a rather flat and elongated gas release, respectively. NaNO2-Mg/Al(3/1) LDHs have been demonstrated to produce a battery-free respirator for inhaled NO, which is a selective and quick-acting pulmonary vasodilator after inhalation.
Metal oxide semiconductors [4], such as SnO2, Ga2O3, and WO3, have been widely studied and used as gas sensors [5,6,7,8,9,10,11]. The important features of metal oxide semiconductor sensors are reversible interactions between gas and the surface of metal oxide semiconductors [10]. The SnO2, Ga2O3, WO3 and other metal oxide semiconductor sensors can be worked stably at a high temperature from 200 °C to 900 °C and can be used for a long time. In recent years, SnO2, Ga2O3, WO3, and other sensors have been determined as sensors for detecting CO, H2, CH4, O3, NO2, H2S, SO2, and other gases [11]. The metal oxide semiconductor sensors with good thermal stability and long-term working stability have greatly aroused the interest of researchers and have been extensively studied [7,8,9,10] in the past decade. The gas adsorption reaction is converted into the sensitivity response of the sensor to detect gas by measuring the work function, capacitance, conductivity [10,12,13,14,15], optical characteristics, and other parameters of the sensing material surface. Kolmakov et al. reported that nanowire sensors made by SnO2 can detect CO at temperatures between 200 °C and 280 °C [16]. The β-Ga2O3 films can be used to detect several gases, such as CO, hydrogen, methane (CH4), NO, and ammonia (NH3) [17].
The structure of ZnGa2O4 is spinel (space group: F 4 ¯ 3 m ). Both ZnGa2O4 and n-type semiconductor ZnGa2O4 thin films have been successfully developed as gas sensors. Satyanarayana et al. [13] reported that ZnGa2O4 films can be used to sense liquid petroleum gas (LPG). When the ZnGa2O4 sensor is further doped with palladium at approximately 320 °C, the sensitivity of detecting LPG increases, but the sensitivity to CH4 and CO [13] gas becomes poor. Despite the introduction of new approaches for developing gas sensors, the fundamental interaction between gas molecules and metal oxide compounds remains unclear. In recent decades, the density functional theory (DFT) has been widely used in the fields of the gas-sensing mechanism, adsorption of atoms and molecules on surfaces, and surface reconstruction [6,7]. For example, the gas-sensing mechanism of ZnO applied to H2, NH3, CO, and ethanol (C2H5OH) was studied by first-principles DFT calculations [7]. Due to the adsorption of surface molecules, the ZnO surface is reconstructed. The charge is transferred from the gas to the ZnO surface and vice versa, which affects the conductivity of the ZnO in gas detection. Vorobyeva et al. reported that the resistance of ZnO increases (decreases) in the presence of NO2 (H2S) [8] adsorption. Recently, we demonstrated that the use of the metal organic chemical vapor deposition (MOCVD) technique to grow epitaxial ZnGa2O4 films on sapphire substrates can produce highly selective gas sensors [9].
Reactions of NO2 and H2S molecules on Ga-Zn-O terminated ZnGa2O4(111) surfaces were modeled and carried out using a first-principles density functional theory method [18]. Based on our previous theoretical study, we report on the palladium adsorbed on the ZnGa2O4(111) surfaces and the work function changes with and without NO2 (H2S) molecules. In principle, the sensitivity response of the sensor to detect the target gas can be determined by the following work function changes Δ Φ according to
Δ Φ = Δ X + k T ln ( R g / R a )
where ΔX is the change in electron affinity, and kT is the product of the Boltzmann’s constant k and the temperature T. Rg is the resistance in the presence of the target gas. Ra is the resistance in the reference gas, which is usually air. We hope that this work can help experimental researchers to effectively develop gas-sensing devices.

2. Computational Details

The work functions of NO2 and H2S adsorbed on Pd-decorated ZnGa2O4(111) (Pd-ZGO) surfaces were systematically studied. All calculations were performed by using the density functional theory as implemented in the Vienna ab initio simulation package (VASP) [19,20] with the exchange correlation function described by generalized gradient approximation (GGA) and the Perdew–Wang (PW91) correction [21,22]. The crystal structure of ZnGa2O4 is displayed in Figure 1a. We used a conventional crystal structure containing 56 atoms, i.e., 8 Zn, 16 Ga, and 32 O atoms. The cutoff energy was set to 500 eV, and the self-consistent total energy criterion was 10 5 eV/unit cell.
To understand the work function changes when NO2 (H2S) is adsorbed on Pd-ZGO surfaces, we built a supercell containing 112 atoms to simulate ZnGa2O4(111) surfaces with a vacuum of 20 Å, which should be wide enough to decouple the nearest surface interaction. This preferred Ga-Zn-O-terminated surface was selected from the previous study on ZnGa2O4, and its surface energy was 0.01 eV/Å2 [23]. In order to describe the effect of changes in the work function of the NO2 and H2S on the surface of Pd-ZGO, we firstly calculated the work function Φ S using the following formula [24]:
Φ S = E V A C E F
where Φ S is the work function of ZnGa2O4(111) surface. E V A C and E F are the energy of the vacuum and the Fermi energy, respectively.
Second, we constructed four adsorption models of Pd atoms, denoted as Pd-ZGOi (i = 1,2,3,4). Here, the superscripts 1, 2, 3, and 4 in Pd-ZGO, respectively, indicate the positions of the initial adsorbed sites Ga3C, Zn3C, O3C, and O4C on the ZnGa2O4(111) surface, as shown in Figure 1b. The initial distance from the Pd atom to the ZnGa2O4(111) surface was set as the sum of the van der Waals radii of the Pd atom and Ga (Zn or O) surface atoms, as shown in Figure 2. Third, we considered one NO2 (H2S) molecule and two NO2 (H2S) molecules adsorbed on the surface Pd atoms and compared the corresponding results of the calculated work functions in the second step, denoted as nNO2-Pd-ZGOi and nH2S-Pd-ZGOi (n = 1,2; i = 1,2,3,4). Here, n represents the number of molecules. All supercells were fully relaxed until the force attraction on each atom was less than 0.001 eV/Å, and a 3 × 3 × 1 Monkhorst–Pack grid was used. The initial distance from NO2 (H2S) molecules to Pd surface atoms was also set to the sum of the van der Waals radii of N (S) from the NO2 (H2S) molecules and the Pd surface atoms. The van der Waals distances of H, N, O, S, Zn, Ga, and Pd atom were 1.20 Å, 1.55 Å, 1.52 Å, 1.80 Å 1.39 Å, 1.87 Å, and 1.63 Å, respectively.
When the NO2 (H2S) molecules were adsorbed on the ZnGa2O4(111) and Pd-ZGO(111) surface, we calculated the work function Φ S ,   gas , which in turn determined the work function differences Δ Φ (eV) between Φ S ,   gas and Φ S . The difference in these obtained work functions is the key factor of gas sensor performance [25]. Finally, the adsorption response of the sensor is closely related to the energy lost or gained by the corresponding gas adsorption. The lost or gained energy of the adsorbed gas can be determined by the adsorption energy. The adsorption energy is the energy released when NO2 (H2S) molecules are adsorbed on the surface of Pd-ZGOi. The adsorption energy ΔE was calculated as follows:
ΔE = Egas+Pd-ZGO − (Egas + EPd-ZGO)
where Egas+Pd-ZGO represents the energy of the Pd-ZGOi surface with the adsorbed NO2 (H2S) molecules. Egas represents the energy of free NO2 (H2S) molecules, and EPd-ZGO represents the energy of the Pd-ZGOi surface. Whether the adsorption energy is related to the work function is also a question to be discussed in this study.

3. Results and Discussion

3.1. Structures

The calculated equilibrium bond lengths of Pd-ZGOi, 1NO2-Pd-ZGOi, 2NO2-Pd-ZGOi, 1H2S-Pd-ZGOi, and 2H2S-Pd-ZGOi (i = 1, 2, 3, 4) are listed in Table 1. The calculated Pd-Ga equilibrium bond lengths are 2.32 Å and 2.37 Å, respectively, for Pd-ZGO1 and Pd-ZGO4 structures; the calculated Pd-Zn equilibrium bond lengths are 2.57 Å and 2.66 Å, respectively, for Pd-ZGO2 and Pd-ZGO4 structures; the calculated Pd-O equilibrium bond length is 2.04 Å for the Pd-ZGO3 structure. The equilibrium bond length of Pd-Ga or Pd-Zn is significantly larger than that of Pd-O (see Figure 1 and Table 1). This is because the valence electron for the O atom is negative, while it is positive for Pd, Ga, and Zn atoms. The Wigner–Seitz radii of Pd, Zn, Ga, and O atoms are 1.434 Å, 1.217 Å, 1.402 Å, and 0.820 Å, respectively.
The bond length of Pd-Zn is larger than that of Pd-Ga because it is dominated by the Coulomb interaction. In other words, the electronic configurations of Zn and Ga are [Ar] 3d104p2 and [Ar] 3d104s24p1, respectively. When the NO2 molecule is adsorbed on the Pd-ZGO structure, the Pd atom is a bridge connecting the NO2 molecule and the ZGO(111) surface. The calculated equilibrium bond length is completely different from Pd-ZGOi. In the 1NO2-Pd-ZGO1 structure, the bond length of Pd-Ga is 2.53 Å, where it is 2.32 Å in the absence of the NO2 molecule without Pd atoms. Note that in the initial structure of 1NO2-Pd-ZGO1, the distance between the NO2 molecule and the surface Pd atom is 3.18 Å. The electronic configuration of the Pd atom is [Kr]4d95s1. This implies the attraction between Pd and N atoms. As a result, the bond length of N-Pd is reduced to 2.01 Å, while the bond length for Pd-Ga is increased. Generally, the N-Pd bond lengths in nNO2-Pd-ZGOi are about 2 Å.
In the 1NO2-Pd-ZGO2 and 1NO2-Pd-ZGO4 structures, Pd atoms do not remain on the top of the Zn3C and O4C sites. Therefore, the calculated bond lengths change significantly, especially for the O4C site (see Figure 3). It seems that Pd atoms are attracted by Ga atoms on the ZGO(111) surface. In one and two NO2 molecules attracted to Pd atoms, the results of 1NO2-Pd-ZGO1 (1NO2-Pd-ZGO3) and 2NO2-Pd-ZGO1 (2NO2-Pd-ZGO3) are similar, while Pd atoms are retained on the top of the Ga3C and O3C sites. When two NO2 molecules are adsorbed into the Pd atom, the calculated bond length of N-Pd is larger. In the 2NO2-Pd-ZGO4 structure, the interaction between the Pd and surface Zn (O) atom disappears. Generally, when NO2 molecules are adsorbed on Pd atoms, the bond length of Pd-Ga, Pd-Zn, and Pd-O will increase.
When H2S molecules were adsorbed on the surface of Pd-ZGO, as shown in Figure 4, we first noticed an increase in the bond lengths of Pd-Ga, Pd-Zn, and Pd-O. We also found that in the 1H2S-Pd-ZGO1 and 1H2S-Pd-ZGO4 structures, the Pd atom is approximately at the top of Ga3C site, and the distances for the S-Pd and Pd-Ga are equal. The bond length of S-Pd in nH2S-Pd-ZGOi (i = 1,2,3,4) is larger than that of N-Pd in nNO2-Pd-ZGOi. This difference is mainly due to the van der Waals radii.

3.2. Work Functions

Table 2 shows that the work function decreases when Pd atoms are adsorbed on the ZGO surface. The reduction in work function ranges from −0.44 eV to −0.15 eV. The larger magnitude of Δ Φ is better. Therefore, the Pd-ZGO1 structure, i.e., Pd on the top of the Ga3C site, is the most sensitive.
When a NO2 molecule was adsorbed on the surface of Pd-ZGOi, we found that 1NO2-Pd-ZGO1 had the largest Δ Φ (+1.37 eV). In the 1NO2-Pd-ZGO1 structure, N-Pd and Pd-Ga have the largest bond lengths. The magnitude of Δ Φ ranges from +1.22 eV to +1.37 eV in 1NO2-Pd-ZGOi. This increase in Δ Φ is further increased as the number of NO2 molecules becomes two. Note that when H2S is adsorbed on the ZGO surface, the work function difference Δ Φ changes significantly to −1.69 eV. When H2S molecules are adsorbed on the surface of Pd-ZGOi, the magnitude of Δ Φ is about half that of H2S-ZGO, ranging from −0.49 eV to −0.90 eV in 1H2S-Pd-ZGOi. The 1H2S-Pd-ZGO3 structure has the largest Δ Φ (−0.90 eV). When two H2S molecules are adsorbed on the surface of Pd-ZGOi, the magnitudes of Δ Φ of 2H2S-Pd-ZGO1 (−1.77 eV) and 2H2S-Pd-ZGO4 (−1.82 eV) are larger than that of H2S-ZGO without the Pd atom (−1.69 eV), showing that the sensitivity of the ZGO surface doped with palladium becomes higher.

3.3. Adsorption Energies

Table 2 also shows adsorption energies of the NO2 and H2S molecules on the surface of ZGO with and without Pd atoms. It can be clearly seen that a NO2 (H2S) molecule is adsorbed to the ZGO surface, which is endothermic by 0.64 eV (2.32 eV). In contrast with Pd atoms adsorbed on the surface of ZGO, i.e., Pd-ZGOi, the NO2 and H2S molecules on the surface of Pd-ZGOi have exothermic or negative adsorption energy. It should be noted that the spontaneous reaction occurs in the direction where the Gibbs free energy change, ΔG, decreases. The Gibbs free energy change consists of enthalpy change ΔH and entropy change ΔS by the following equation:
ΔG = ΔHTΔS
The enthalpy change is defined as the sum of the internal energy change ΔE and the product of the pressure P and volume change ΔV, as follows:
ΔH = ΔE + PΔV
Therefore, the Gibbs free energy change can be expressed by the following equation:
ΔG = ΔE + PΔVTΔS
In our study, we ignored the contribution of entropy change, and the volume change in the Pd-ZGOi structures was negligible. The Gibbs free energy change can be further simplified as the internal energy change caused by the kinetic energy, potential energy, and chemical energy of the target system. The negative Gibbs free energy change, its negative internal energy change, or its negative adsorption energies, ΔE, provides a means of spontaneous physical and chemical changes to take place without any external help. On the contrary, the positive adsorption energy indicates that the corresponding endothermic process can be supported by the increase in temperature. A higher positive adsorption energy means that more energy is required from the outside world, which also indicates that the endothermic reaction does not occur easily. Our results on the surface of ZGO show that the adsorption of NO2 molecules on the surface is energetically more favorable than the adsorption of H2S molecules on the surface by 1.68 eV.
Our results on NO2 (H2S) molecules adsorbed on the surface of Pd-ZGOi show that the adsorption reaction is spontaneous and exothermic. Most importantly, the structures of Pd atoms remaining at the top of the Zn3C site, namely, 1NO2-Pd-ZGO2, 2NO2-Pd-ZGO2, 1H2S-Pd-ZGO2, and 2H2S-Pd-ZGO2, have the overall lowest adsorption energy for NO2 and H2S molecules, regardless of the number of NO2 and H2S molecules. It is interesting to note that the magnitudes of adsorption energy of 1NO2-Pd-ZGO2 (−2.94 eV) and 2NO2-Pd-ZGO2 (−3.72 eV) are larger than that of 1H2S-Pd-ZGO2 (−2.89 eV) and 2H2S-Pd-ZGO4 (−2.44 eV), showing that the surface of ZGO doped with palladium benefits the adsorption reaction to increase the amount of NO2. However, the adsorption reaction on the surface of Pd-ZGOi is unfavorable for the increase in the amount of H2S. The largest adsorption energy occurs in the structures of the Pd atoms remaining on the top of the Zn3C site, but the largest work function difference occurs in the structures of Pd atoms remaining on the top of the Ga3C site. Overall, the difference in work functions is not closely related to their respective adsorption energies.

4. Conclusions

Using the framework of density functional theory, we studied the adsorption reactions and work functions of NO2 and H2S on Pd-ZnGa2O4(111) surfaces. Our previous studies 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 exhibits the highest work function change of +0.92 eV (−1.69 eV). In this study, we found that the work function change of one NO2 (H2S) molecule on the Pd-ZGO(111) surface is enhanced (reduced) to +1.37 eV (−0.90 eV). This difference is mainly due to the decrease in work function when the palladium is adsorbed on the ZGO(111) surface. The work function difference for two NO2 (H2S) molecules adsorbed on the surface of Pd-ZGOi is about twice as large as that for one NO2 (H2S) molecule, showing a higher work function difference, i.e., a higher sensitivity. The resistance response experiment of ZnO(Ga) samples to NO2 (H2S) molecules shows that the sensitivity response of ZnGa2O4-based thin-film sensors to NO2 (H2S) molecules presents the same trend as the positive (negative) work function differences [8]. Adsorption energies of NO2 (H2S) molecules on the surface of ZGO with and without Pd atoms were analyzed to explain the exothermic and endothermic adsorption reactions, but no fully satisfactory explanation related to their respective work function differences can be provided at this time.

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., D.-Y.W. and P.-L.L.; investigation, Y.-H.C. and D.-Y.W.; writing—original draft preparation, J.-C.T., Y.-H.C. and P.-L.L.; writing—review and editing, J.-C.T., Y.-H.C. 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. It was also funded by the China Medical University, grant number CMU-109-N24.

Acknowledgments

Computational studies were performed using the resources of the National Center for High Performance Computing, Taiwan.

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. Hiroshi, M.; Takuya, Y.; Hiromichi, V.M.; Maning, L.; Yasuhiro, T.; Tetsuaki, F.; Yasushi, T.; Jun, T. Insulated conjugated bimetallopolymer with sigmoidal response by dual self-controlling system as a biomimetic material. Nat. Commun. 2020, 11, 408. [Google Scholar]
  3. Shinsuke, I.; Nobuo, I. Controlled release of H2S and NO gases through CO2-stimulated anion exchange. Nat. Commun. 2020, 11, 453. [Google Scholar]
  4. Meixner, H.; Lampe, U. Metal Oxide Sensors. Sens. Actuators B Chem. 1996, 33, 198–202. [Google Scholar] [CrossRef]
  5. 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]
  6. 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]
  7. 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]
  8. 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]
  9. 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]
  10. 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]
  11. Korotcenkov, G. Metal oxides for solid-state gas sensors: What determines our choice. Mater. Sci. Eng. B. 2007, 139, 1–23. [Google Scholar] [CrossRef]
  12. Wang, C.X.; Yin, L.W.; Zhang, L.Y.; Xiang, D.; Gao, R. Metal Oxide Gas Sensors: Sensitivity and Influencing Factors. Sensors 2010, 10, 2088–2106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. 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]
  14. 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]
  15. Schierbaum, K.D.; Kirner, U.K.; Geiger, J.F.; Gopel, W. Schottky-barrier and conductivity gas sensors based upon Pd/SnO2 and Pt/TiO2. Sens. Actuators B Chem. 1991, 4, 87–94. [Google Scholar] [CrossRef]
  16. 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]
  17. Schwebel, T.; Fleischer, M.; Meixner, H. A Selective Temperature Compensated O2 Sensor based on Ga2O3 Thin Films. Sens. Actuators B Chem. 2002, 65, 176–180. [Google Scholar] [CrossRef]
  18. Tung, J.-C.; Chiang, Y.-H.; Wang, D.-Y.; Liu, P.-L. Adsorption of NO2 and H2S on ZnGa2O4(111) Thin Films: A First-Principle Density Functional Theory Study. Appl. Sci. 2020, 10, 8822. [Google Scholar] [CrossRef]
  19. 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]
  20. 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]
  21. 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]
  22. 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]
  23. 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]
  24. Kahn, A. Fermi Level Work Function and Vacuum Level. Mater. Horiz. 2016, 3, 7–10. [Google Scholar] [CrossRef]
  25. 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. Left and right panels illustrate bulk ZnGa2O4 crystal structure and top view of the adsorption sites O3C, Zn3C, O4C, and Ga3C on ZnGa2O4(111) surface, respectively. These sites are labeled for the adsorption sites of the Pd atom.
Figure 1. Left and right panels illustrate bulk ZnGa2O4 crystal structure and top view of the adsorption sites O3C, Zn3C, O4C, and Ga3C on ZnGa2O4(111) surface, respectively. These sites are labeled for the adsorption sites of the Pd atom.
Applsci 11 05259 g001
Figure 2. Side view of the Pd atom on ZnGa2O4(111) surface. Pd-ZGOi, showing Pd atoms located on the top of Ga3C, Zn3C, O3C, and O4C sites for i = 1, 2, 3, and 4, respectively. The bond lengths of Pd-ZGOi are listed in Table 1.
Figure 2. Side view of the Pd atom on ZnGa2O4(111) surface. Pd-ZGOi, showing Pd atoms located on the top of Ga3C, Zn3C, O3C, and O4C sites for i = 1, 2, 3, and 4, respectively. The bond lengths of Pd-ZGOi are listed in Table 1.
Applsci 11 05259 g002
Figure 3. Side view of the nNO2 molecules on Pd-ZnGaOi surface. The equilibrium bond lengths of nNO2-Pd-ZGOi are listed in Table 1.
Figure 3. Side view of the nNO2 molecules on Pd-ZnGaOi surface. The equilibrium bond lengths of nNO2-Pd-ZGOi are listed in Table 1.
Applsci 11 05259 g003
Figure 4. Side view of the nH2S molecules on Pd-ZnGaOi surface. The equilibrium bond lengths of nH2S-Pd-ZGOi are listed in Table 1.
Figure 4. Side view of the nH2S molecules on Pd-ZnGaOi surface. The equilibrium bond lengths of nH2S-Pd-ZGOi are listed in Table 1.
Applsci 11 05259 g004
Table 1. Calculated equilibrium bond length (Å) for the Pd-ZGOi, nNO2-Pd-ZGOi, and nH2S-Pd-ZGOi structures.
Table 1. Calculated equilibrium bond length (Å) for the Pd-ZGOi, nNO2-Pd-ZGOi, and nH2S-Pd-ZGOi structures.
Pd-ZGOiBond Length (Å)
Pd-GaPd-ZnPd-O
Pd-ZGO12.32--
Pd-ZGO2-2.57-
Pd-ZGO3--2.04
Pd-ZGO42.372.66-
nNO2-Pd-ZGOiBond Length (Å)
N-PdPd-GaPd-ZnPd-O
1NO2-Pd-ZGO12.012.53--
1NO2-Pd-ZGO21.942.34-2.17
1NO2-Pd-ZGO31.96--2.10
1NO2-Pd-ZGO41.962.312.862.48
2NO2-Pd-ZGO12.042.042.51--
2NO2-Pd-ZGO22.032.212.442.822.78
2NO2-Pd-ZGO32.112.05--2.15
2NO2-Pd-ZGO42.042.122.46--
nH2S-Pd-ZGOiBond Length (Å)
S-PdPd-GaPd-ZnPd-O
1H2S-Pd-ZGO12.372.38--
1H2S-Pd-ZGO22.262.822.892.18
1H2S-Pd-ZGO32.252.62-2.11
1H2S-Pd-ZGO42.382.38--
2H2S-Pd-ZGO12.522.542.34--
2H2S-Pd-ZGO22.412.602.812.882.37
2H2S-Pd-ZGO32.312.99--2.13
2H2S-Pd-ZGO42.542.562.46--
Table 2. Calculated work functions Φ S and Φ S , gas , work function differences Δ Φ , and adsorption energies ΔE for the ZGO surfaces, NO2 and H2S molecules on ZGO surfaces, Pd-adsorbed ZGO surfaces, and NO2/H2S on Pd-adsorbed ZGO surfaces.
Table 2. Calculated work functions Φ S and Φ S , gas , work function differences Δ Φ , and adsorption energies ΔE for the ZGO surfaces, NO2 and H2S molecules on ZGO surfaces, Pd-adsorbed ZGO surfaces, and NO2/H2S on Pd-adsorbed ZGO surfaces.
Model Φ S   ( eV ) Φ S , gas ( eV ) Δ Φ   ( eV ) Δ E   ( eV )
ZGO4.09---
NO2-ZGO-5.01+0.92+0.64
H2S-ZGO-2.40−1.69+2.32
Pd-ZGO13.65-−0.44-
Pd-ZGO23.84-−0.25-
Pd-ZGO33.94-−0.15-
Pd-ZGO43.75-−0.34-
1NO2-Pd-ZGO1-5.02+1.37−1.30
1NO2-Pd-ZGO2-5.06+1.22−2.94
1NO2-Pd-ZGO3-5.23+1.29−2.25
1NO2-Pd-ZGO4-5.08+1.33−2.14
2NO2-Pd-ZGO1-6.02+2.37−2.64
2NO2-Pd-ZGO2-5.61+1.77−3.72
2NO2-Pd-ZGO3-5.38+1.44−3.06
2NO2-Pd-ZGO4-5.62+1.87−2.53
1H2S-Pd-ZGO1-2.93−0.72−0.73
1H2S-Pd-ZGO2-3.16−0.68−2.89
1H2S-Pd-ZGO3-3.04−0.90−1.85
1H2S-Pd-ZGO4-3.26−0.49−0.95
2H2S-Pd-ZGO1-1.88−1.77−0.45
2H2S-Pd-ZGO2-2.58−1.26−2.44
2H2S-Pd-ZGO3-2.62−1.32−1.80
2H2S-Pd-ZGO4-1.93−1.82−1.03
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.; Wang, D.-Y.; Chen, Y.-H.; Liu, P.-L. Influences of Work Function Changes in NO2 and H2S Adsorption on Pd-Doped ZnGa2O4(111) Thin Films: First-Principles Studies. Appl. Sci. 2021, 11, 5259. https://doi.org/10.3390/app11115259

AMA Style

Tung J-C, Wang D-Y, Chen Y-H, Liu P-L. Influences of Work Function Changes in NO2 and H2S Adsorption on Pd-Doped ZnGa2O4(111) Thin Films: First-Principles Studies. Applied Sciences. 2021; 11(11):5259. https://doi.org/10.3390/app11115259

Chicago/Turabian Style

Tung, Jen-Chuan, Ding-Yuan Wang, Yu-Hsuan Chen, and Po-Liang Liu. 2021. "Influences of Work Function Changes in NO2 and H2S Adsorption on Pd-Doped ZnGa2O4(111) Thin Films: First-Principles Studies" Applied Sciences 11, no. 11: 5259. https://doi.org/10.3390/app11115259

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