Next Article in Journal
Fe3O4-PVDF Composite Network for Dendrite-Free Lithium Metal Batteries
Next Article in Special Issue
First-Principles Dynamics Investigation of Germanium as an Anode Material in Multivalent-Ion Batteries
Previous Article in Journal
Convergence Gas Sensors with One-Dimensional Nanotubes and Pt Nanoparticles Based on Ultraviolet Photonic Energy for Room-Temperature NO2 Gas Sensing
Previous Article in Special Issue
First-Principles Investigation of Size Effects on Cohesive Energies of Transition-Metal Nanoclusters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

DFT Study of Zn-Modified SnP3: A H2S Gas Sensor with Superior Sensitivity, Selectivity, and Fast Recovery Time

1
Key Laboratory of Optoelectronic Technology & Systems (Chongqing University), Ministry of Education, Chongqing 400044, China
2
Centre for Intelligent Sensing Technology, College of Optoelectronic Engineering, Chongqing University, Chongqing 400044, China
3
School of Microelectronics, Southern University of Science and Technology, Shenzhen 518055, China
4
Faculty of Science and Engineering, Swansea University, Swansea SA1 8EN, UK
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(20), 2781; https://doi.org/10.3390/nano13202781
Submission received: 24 September 2023 / Revised: 11 October 2023 / Accepted: 15 October 2023 / Published: 17 October 2023
(This article belongs to the Special Issue First-Principle Calculation Study of Nanomaterials)

Abstract

:
The adsorption properties of Cu, Ag, Zn, and Cd-modified SnP3 monolayers for H2S have been studied using density functional theory (DFT). Based on phonon spectrum calculations, a structurally stable intrinsic SnP3 monolayer was obtained, based on which four metal-modified SnP3 monolayers were constructed, and the band gaps of the modified SnP3 monolayers were significantly reduced. The adsorption capacity of Cu, Zn-modified SnP3 was better than that of Ag, Cd-modified SnP3. The adsorption energies of Cu-modified SnP3 and Zn-modified SnP3 for H2S were −0.749 eV and −0.639 eV, respectively. In addition, Cu-modified SnP3 exhibited chemisorption for H2S, while Zn-modified SnP3 exhibited strong physisorption, indicating that it can be used as a sensor substrate. Co-adsorption studies showed that ambient gases such as N2, O2, and H2O had little effect on H2S. The band gap change rate of Zn-modified SnP3 after adsorption of H2S was as high as −28.52%. Recovery time studies based on Zn-modified SnP3 showed that the desorption time of H2S was 0.064 s at 298 K. Therefore, Zn-modified SnP3 can be used as a promising sensor substrate for H2S due to its good selectivity, sensitivity, and fast recovery time.

1. Introduction

With the increasing advancement of urbanization and industrialization, the domestic sewage, industrial sewage and runoff sewage gathered by pipelines produce irritant gases such as hydrogen sulfide (H2S) [1]. In addition, H2S as an industrial waste gas, its emission can pollute the atmosphere [2]. H2S is a highly toxic and corrosive gas, and as a kind of flammable hazardous chemical, the leakage of H2S may result in significant economic and property losses or even casualties. Therefore, it is critical to investigate novel solutions for detecting H2S gas.
With its successful discovery in 2004, graphene has demonstrated considerable promise in a variety of technical areas. Due to their substantial theoretical specific surface area, two-dimensional (2D) materials such as graphene offer excellent options for extremely sensitive sensing and detection devices [3]. However, after years of research, the sensing adsorption behavior of substrates such as transition metal dichalcogenides (TMDCs) and graphene has been systematically studied [4,5,6,7].
SnP3, a layered material made of Sn and P, has been studied and it was pointed out in the 1970s that it can be obtained by slow cooling to room temperature after heat treatment for 2 days at 575 °C [8]. Based on theoretical calculations, it has been found that double and single-layer SnP3 have relatively low cleavage energies of 0.45 J/m2 and 0.71 J/m2, separately, approached by phosphorene (0.36 J/m2) and graphene (0.32 J/m2) [9]. Therefore, SnP3 is a novel 2D substance that is easy to peel off. A 2D SnP3 not only has high thermodynamic stability [10], but also has excellent carrier mobility [11,12]. For potassium-ion, lithium-ion, and sodium-ion batteries, SnP3 has been used as an anode material [13,14,15]. Application studies in gas sensing have shown that intrinsic SnP3 has a strong adsorption effect for NO2 and NO [16,17,18].
In the study of gas sensing in 2D materials, computational works are frequently conducted by using density functional theory (DFT) [19,20,21]. Moreover, doped 2D materials can enhance the adsorption capacity of gases, according to DFT computational studies. Besides the impurity doping, the doping could be realized by metallic gating, and high doping carrier densities of the order of 1014 cm−2 have been achieved in 2D materials by electrolyte gating [22]. Most of the time, impurity doping could offer better selectivity, in comparison with metallic gating. For example, the interaction of O2 and CO is significantly enhanced by Au-doped graphene [23], As-doped WSe2 shows a significant increase in NO2 adsorption [24], and the adsorption ability of N2 by InN decorated with Ni atom is significantly improved [25]. However, only a few works on the doping adsorption of SnP3 monolayers, for example, indium-doped SnP3 monolayer is used for CO2 adsorption [26], chromium-doped SnP3 monolayer is used for SO2 adsorption [27], and palladium-doped SnP3 monolayer is used for H2 adsorption research [28].
In this work, four transition metals were selected as doping elements. Cu and Ag have similar properties because they belong to group IB. Zn and Cd have similar properties as they belong to group IIB. In addition, Cu (1.35 Å) and Zn (1.35 Å) have the same atomic radius, and Ag (1.60 Å) and Cd (1.55 Å) have similar atomic radius. By studying Cu, Ag, Zn, and Cd, the doping effects of elements in the same and adjacent groups can be compared. Therefore, the adsorption properties of H2S on SnP3 monolayers modified with Cu, Ag, Zn, and Cd atoms were studied for the first time in this paper using DFT calculations, providing theoretical support for the application of SnP3 monolayer material in gas sensing.

2. Computational Details

Herein, adsorption characteristics of the SnP3 monolayer were investigated using the DMol3 module based on DFT calculations [29,30]. The exchange correlation in the module configuration was described by the PBE (Perdew–Burke–Ernzerhof) functional in the GGA (generalized gradient approximation) [31,32]. We adjusted the dispersion interactions using the TS (Tkatchenko–Scheffler) technique (a method for DFT-D) [33,34], and checked the spin unrestricted (spin-polarization) [35]. The energy had a convergence tolerance of 10−5 Ha (1 Ha is approximately equal to 27.2114 eV), the maximum displacement convergence tolerance was set to 5 × 10−3 Å, and the convergence tolerance for maximum force was set to 2 × 10−3 Ha/Å [36]. DSPP (DFT Semi-core Pseudopots) was used as the core treatment technique for introducing relativistic corrections [37]. The atomic orbital used a dual numerical basis set having the orbital polarization function [38]. Smearing was set to 0.005 Ha to accelerate convergence [17]. After convergence tests, the geometric structure was optimized with a k-point set to 6 × 6 × 1, and the properties were calculated using a more intensive k-point grid of 12 × 12 × 1 [27,28]. Intrinsic and doped SnP3 supercell structures of 2 × 2 × 1 size with a vacuum area of 20 Å were created for adsorption studies.
After geometry optimization, the energy of the structure was obtained. The following equation was utilized to compute the adsorption energy (Ea) of gas molecules on SnP3 substrates:
E a = E g a s / s u b E g a s E s u b
where Egas/sub is the system’s total energy, Egas and Esub are the energies of gas molecules and substrates, respectively. The substrates are more strongly adsorbed to the gas molecules, the greater the absolute values of Ea. Adsorption can typically happen on its own if the adsorption energy Ea is negative.
The relevant properties were analyzed by DFT module calculations. To calculate the transfer charge (Q) of a gas molecule after adsorption by a substrate, Mulliken population analysis was utilized to determine the charge gain or loss of individual atoms in the gas molecule [30,39]. When Q is positive, the whole gas molecule consumes electrons; when Q is negative, the whole gas molecule accumulates electrons [40]. The shortest distance between the surface and the adsorbed gas molecule in the adsorption system was indicated by the adsorption distance (d). A closer adsorption distance indicates a stronger interaction, and conversely, a farther adsorption distance indicates a weaker interaction. The interactions between the substrates and the doped atoms, as well as between the substrates and the adsorbed gas molecules, were investigated through analyzing the density of states (DOS). The transfer charge between substrates and gas molecules was studied using the charge density difference (CDD). A description of whether a bond forms between the gas molecule and the substrate could be given by the electron localization function (ELF) [41,42].

3. Results and Discussion

3.1. Establishment and Analysis of SnP3 Monolayer

First, based on previous research reports [27], a supercell of intrinsic SnP3 monolayer with a size of 2 × 2 × 1 was constructed in this paper for adsorption studies, as shown in Figure 1. The lattice parameter a = b = 7.359 Å of its unit cell is comparable to a = b = 7.37 Å in the literature [9]. The top view shows that the structure contains 8 Sn atoms and 24 P atoms. The side view shows that the structure is not flat, but a sandwich structure consisting of P atoms in the center and Sn atoms on both sides. Each P atom creates a P–Sn bond with an adjacent Sn atom and two P–P bonds with two adjacent P atoms, while three Sn–P bonds are formed between each Sn atom and the three nearby P atoms. With a thickness of 2.888 Å, the SnP3 monolayer is similar to that of 2.76 Å in the literature [28].
The intrinsic SnP3 band structure is shown in Figure 2a, where the horizontal coordinates are the K points taken based on the symmetry of the structural model, and the vertical coordinates indicate the energy, with the position of the zero-point corresponding to the Fermi energy level. There is a band gap of 0.498 eV for the SnP3 monolayer structure, which is consistent with the findings of earlier research [31,43]. To verify the stability of the constructed monolayer structure, we calculated the phonons using the finite displacement method in the CASTEP module. The same functional and DFT-D corrections as in the DMol3 were used in the module setup, and additionally, an energy cutoff of 500 eV and ultrasoft pseudopotentials were set. The phonon spectrum can reflect whether the material exists stably or not, if the phonon spectrum of the material has no imaginary frequency, it can be proved that the material can exist stably. As shown in Figure 2b, there is no imaginary frequency in the phonon spectrum of the intrinsic SnP3 structure, which can be a good indication that the structure is stable. On this basis, we will construct several metal doped structures.
Next, the modified doping model was constructed based on intrinsic SnP3. Four transition metal atoms, Cu, Ag, Zn, and Cd were selected as doping elements. As depicted in Figure 1, four distinct modification locations were taken into account in the construction of the metal atom doping structure, namely, the TP location above the P atom, the TH1 location above the hexagonal ring composed of P atoms, the TSn location above the Sn atom, and the TH2 location above the hexagonal ring made up of P and Sn atoms. Usually, the stability of the doped structure can be evaluated by calculating the formation energy [44], and the formation energy (Ef) can be calculated according to the following equation:
E f = E X S n P 3 E S n P 3 E X
where EXSnP3, ESnP3 and EX are the energies of the X-modified SnP3 (X = Cu, Ag, Zn, Cd), SnP3, and a metal atom, respectively. The more negative the formation energy, the more stable the doped structure. It can be seen that the energy of the doped structure determines the formation energy, since the energy of SnP3 and metal atoms are certain. After comparison, the doped structure with the metal atom modified above the Sn atom has the maximum absolute value of energy, that is, the most stable structure, so the TSn site is the best modification site, which is consistent with the doping site in the literature [28].
The stable structures of the SnP3 monolayers modified by four metal atoms (Cu–SnP3, Ag–SnP3, Zn–SnP3, Cd–SnP3) are shown in Figure 3. The average lengths of the Cu–P, Ag–P, Zn–P, and Cd–P bonds formed by the modified metal atoms with three P atoms are 2.275 Å, 2.622 Å, 2.480 Å, and 2.812 Å, respectively. In addition, the distances of the modified metal atoms from the Sn atoms directly below them are 2.943 Å, 3.576 Å, 3.324 Å, and 3.761 Å, respectively. Compared with the intrinsic SnP3 monolayer, the thickness of doped SnP3 monolayers increased to 3.278 Å, 3.303 Å, 3.353 Å, and 3.278 Å, respectively.
The modified SnP3 monolayer energy band structures are shown in Figure 4. The band gap of intrinsic SnP3 is 0.498 eV. The band gap of the atom-doped SnP3 monolayer is reduced, and the energy bands of Cu–SnP3 and Ag–SnP3 cross the Fermi energy level and exhibit metallic properties, and the band gaps of Zn–SnP3 and Cd–SnP3 are 0.291 eV and 0.327 eV, respectively.
The DOS of intrinsic and four doped SnP3 monolayers are depicted in Figure 5 and Figure 6. The zero point of the energy coordinate indicates the Fermi energy level. Below zero is the valence band, and all positions are filled with electrons. The material properties have a strong correlation with the energy band structure close to the Fermi energy level. By studying the TDOS (total DOS) and the PDOS (projected DOS), we analyzed the effect of Cu, Ag, Zn, and Cd doping on the SnP3 monolayer, and the interactions between the SnP3 monolayer and the doped transition metal atoms.
Figure 5a,b and Figure 6a,b illustrate the TDOS of four doped SnP3 monolayers compared with the intrinsic SnP3, respectively. It is obvious that the TDOS of the doped SnP3 monolayers is shifted to the left with respect to the intrinsic monolayer, and combined with the energy band structure in Figure 4, indicates that doping reduces the band gap, which results in enhanced metal characteristics and increased conductivity. Moreover, as the band gap narrows, the valence band electrons can move more readily into the conduction band [45].
The PDOS of Cu–SnP3 is seen in Figure 5c. The P-3p and Cu-3d orbits significantly superimpose between −4.682 eV and −0.361 eV. It shows that the P-3p and Cu-3d orbitals have strong interactions and orbital hybridization mostly happens between them, which will cause the electron distribution to alter, and denotes the formation of the Cu–P bond. It demonstrates that the Cu atom modified on the SnP3 monolayer surface has a stable structure. The PDOS of Ag–SnP3 is shown in Figure 5d. It is clear that the Ag-4d and P-3p orbits significantly overlap between −4.831 eV and −1.913 eV, and the Ag-5s and P-3p orbits overlap slightly at the peak of 2.082 eV. It shows that Ag-4d, Ag-5s, and P-3p orbitals have strong interactions. Orbital hybridization mostly happens between Ag-4d and P-3p, which will cause the electron distribution to alter, and indicates the formation of the Ag–P bond. It shows that the Ag atom modified on the SnP3 monolayer surface has a stable structure.
The PDOS of Zn–SnP3 is depicted in Figure 6c. The Zn-3d and P-3s, P-3p orbits overlap between −8.374 eV and −4.913 eV, and the Zn-4s and P-3p orbital peaks overlap at −4.580 eV, −1.212 eV, and 1.151 eV. It indicates that all orbitals interact with each other. Orbital hybridization mostly happens between Zn-4s and P-3p, which will cause the electron distribution to alter, and indicates the formation of the Zn–P bond. It shows that the Zn atom modified on the SnP3 monolayer surface has a stable structure. The PDOS of Cd–SnP3 is seen in Figure 6d. The Cd-4d and P-3s, P-3p orbits overlap between −9.852 eV and −7.281 eV, and the peaks of the Cd-5s and P-3p orbits coincide at −4.276 eV, −1.456 eV, and 0.973 eV, which means there are interactions between the orbitals. Orbital hybridization mostly happens between P-3p and Cd-5s, which will cause the electron distribution to alter, and indicates the formation of the Cd–P bond. It shows that the Cd atom modified on the SnP3 monolayer surface has a stable structure.

3.2. Study on Adsorption of H2S by SnP3 Monolayer

In constructing the intrinsic SnP3 monolayer adsorption system for H2S, the initial separation between the H2S and the SnP3 monolayer was adjusted to 3 Å. As shown in Figure S1, various adsorption positions of the H2S molecule are considered, including H2S molecules placed parallel or vertically above the Sn atom, P atom, and the hexagonal ring. According to modular calculations, the intrinsic SnP3 monolayer adsorbed H2S has the most stable structure, as illustrated in Figure S1b. Similarly, the system of X–SnP3 (X = Cu, Ag, Zn, Cd) for H2S was constructed using the above method. Since the doped metal atoms act as interacting active sites between the substrate and the adsorbed gases [46], the H2S molecule was placed parallel or vertically on the doped metal atoms in the construction of the adsorption systems of the four doped SnP3 monolayers to H2S.
As shown in Figure 7, the most stable structures of the intrinsic as well as the four doped SnP3 monolayers for H2S adsorption are displayed. Table 1 summarizes the results of H2S adsorbed on the intrinsic SnP3 and the SnP3 doped with four metal atoms. It includes the adsorption energy (Ea) of the SnP3 monolayer to the H2S molecule, the transfer charge (Q) of the H2S molecule, and the shortest adsorption distance (d) between the SnP3 monolayer and the H2S. According to Equation (1), the adsorption energy of intrinsic SnP3 for H2S is −0.392 eV. The transfer charge and adsorption distance are 0.024 e and 2.574 Å, respectively. The X–SnP3 (X = Cu, Ag, Zn, Cd) monolayer adsorption energies for H2S are −0.749 eV, −0.595 eV, −0.639 eV, and −0.402 eV, correspondingly. The absolute values of adsorption energies are Cu–SnP3 > Zn–SnP3 > Ag–SnP3 > Cd–SnP3. The transfer charges of H2S molecules amounted to 0.272 e, 0.211 e, 0.234 e, and 0.187 e, correspondingly. The transfer charges show numerically Cu–SnP3 > Zn–SnP3 > Ag–SnP3 > Cd–SnP3. The adsorption distances, which are 2.336 Å, 2.601 Å, 2.520 Å, and 2.877 Å, showed numerically that Cu–SnP3 < Zn–SnP3 < Ag–SnP3 < Cd–SnP3.
Compared with the intrinsic SnP3 monolayer, the transfer charge and adsorption energy values of the four doped SnP3 monolayers are increased, and the adsorption distances between Cu–SnP3, Zn–SnP3, and H2S molecule are smaller than those of the intrinsic SnP3 monolayer. The larger the value of the transfer charge and adsorption energy, and the closer the adsorption distance, the better the adsorption effect is considered. Therefore, based on the information in Table 1, the SnP3 monolayer doped with four kinds of metals can improve the adsorption of H2S, only Cu–SnP3 and Zn–SnP3 can significantly improve the adsorption of H2S, while Ag–SnP3 and Cd–SnP3 have a smaller enhancing effect on the adsorption of H2S.
Next, to study the effects of adsorbed H2S on the electronic structure of the SnP3 monolayers, DOS, CDD, and ELF analyses were performed for the adsorption system. Since Cu–SnP3 and Zn–SnP3 have better adsorption capacities for H2S than Ag–SnP3 and Cd–SnP3, here we only took Cu–SnP3 and Zn–SnP3 as examples to study the adsorption of H2S.
As illustrated in Figure 8a,b, the TDOS of Cu–SnP3 and Zn–SnP3 monolayers is not significantly shifted after H2S adsorption, but there is an increase at some positions (marked by circles), and the increased positions correspond exactly to the S-3s, 3p orbitals of the PDOS in Figure 8c,d. From Figure 8c, the Cu-3d, 4s, and S-3p orbital peaks overlap at −6.362 eV, −4.708 eV, −4.197 eV, −2.762 eV, and 2.663 eV. The overlap of the same waveform of the density of states peak represents the hybridization between the adsorbed molecule and the substrate, thus indicating a strong interaction of the S atom in H2S with the substrate-modified Cu atom. In addition, as seen in Figure 8d, the Zn-3d, 4s, and S-3p orbital peaks overlap at −7.465 eV, −6.069 eV, −4.745 eV, −4.065 eV, and 2.528 eV, suggesting that the doped Zn atom interacts strongly with the S atom in the H2S molecule.
Figure 9a–d shows the top and side perspectives of Cu–SnP3 and Zn–SnP3 structures that are most stable to H2S adsorption, respectively. The CDD of H2S adsorbed by Cu–SnP3 and Zn–SnP3 is seen in Figure 9e,f. In the study of adsorption, CDD can be used to observe the direction of charge transfer in space after molecular adsorption. The isosurface can be obtained by connecting the points with the same charge gain and loss probability. Here, we set the isosurface value to 0.015 e/Å3, which can obtain a better graphical effect. The yellow and blue parts of the CDD illustration indicate charge consumption and accumulation, respectively. The charge of H atoms inside the H2S molecule is consumed, while the charge of the S atom is accumulated. However, the charge density of the H2S molecule is reduced from the overall view of the adsorption system, indicating that the substrate receives the charge from H2S. As a comparison, the CDD of the intrinsic SnP3 adsorbed H2S is displayed in Figure S2c, showing little transfer charge between the substrate and H2S molecule. From Table 1, the charge transfer between H2S and Cu–SnP3, and Zn–SnP3 was calculated by Mulliken population analysis to be 0.272 e, and 0.234 e, separately. The positive transfer charge value shows that the H2S molecule is positively charged, which implies that the H2S molecule loses electrons and transfers to the substrate, so the population analysis value is consistent with the CDD result.
Electron localization function (ELF) is one of the means to study electronic structure, which can characterize the localization degree of electrons. ELF is used to show the distribution of electrons outside the nucleus and analyze the properties of electrons near the nucleus and bonding regions. The range of the ELF value is 0 to 1: when ELF = 1, the electron is completely localized; when ELF = 0, it corresponds to complete delocalization of the electron, which also means that there is no electron at that place [47]. Figure 9g,h formed along the lines in Figure 9a,b show the ELF section plots of Cu–SnP3 and Zn–SnP3 adsorbing H2S. In the ELF diagram, the closer the blue area is, the less and more dispersed the electrons are. The closer the red area is, the more concentrated the electrons are. From the dashed box in Figure 9g, the overlap of electron localization between the doped Cu atom and the S atom in H2S, indicates the existence of shared electrons and the formation of a chemical bond between them, suggesting that Cu–SnP3 is chemically adsorbed to H2S. As shown in Figure 9h, it is clear from the dotted box that the Zn atom and the S atom electron localization edges are very close to each other with strong interactions, indicating that Zn–SnP3 is a strong physisorption for H2S. As a comparison, the ELF of H2S adsorption by intrinsic SnP3 is shown in Figure S2d, which shows that the electron localization is not overlapped between the H2S molecule and the SnP3 and is far away, indicating that intrinsic SnP3 has only weak physical adsorption of H2S.

3.3. Study on Co-Adsorption of H2S with Ambient Gases

To study the interference of ambient gases with H2S, we calculated isotherm adsorption curves of the SnP3 monolayer for H2S as well as for the three interfering gases (N2, O2, and H2O) using the Metropolis method in the Sorption module. Cu–SnP3 and Zn–SnP3 were chosen as the adsorption substrates because of their better adsorption capacity for H2S than Ag–SnP3 and Cd–SnP3. As shown in Figure 10, the horizontal axis represents the pressure, and the vertical axis represents the adsorption capacity. When the pressure is between 0 and 200 kPa and the temperature is 298 K, the adsorption capacity of the substrate for several gases improves with increasing pressure. Notably, the adsorption of H2S by the substrate was significantly greater than that of N2, O2, and H2O throughout the co-adsorption process. At pressures ranging from 20 to 200 kPa, the adsorption capacity of Cu–SnP3 and Zn–SnP3 for H2S was approximately six to seven times that of N2, O2, and H2O. The results showed that Cu–SnP3 and Zn–SnP3 had good selectivity for H2S, which provided a strong theoretical support for the adsorption of H2S in the ambient environment.

3.4. Study on Sensing of H2S by Zn–SnP3

The above studies have shown that Zn–SnP3 has strong physical adsorption on H2S, therefore Zn–SnP3 is a potentially sensitive material for the detection of H2S. Sensor devices can be fabricated from Zn–SnP3 materials and applied to the detection of the target gas H2S. Typically, when a sensitive material adsorbs a target substance, the electrical conductivity of that material changes at the macro level. The conductivity (σ) can be described by the following relation [48]:
σ exp ( E g 2 T K B )
where Eg in the relation represents the band gap, and T and KB are the temperature and Boltzmann constant, respectively. It is easy to notice from this relation that only the parameter Eg affects the conductivity of the adsorption system at a certain temperature. The change in band gap after gas adsorption directly affects the change in conductivity, and the more obvious the change in conductivity, the easier the signal of current or voltage in the device can be detected. The band gap of Zn–SnP3 is 0.291 eV, as shown in Figure 4. The energy band structure of Zn–SnP3 after adsorption of H2S is shown in Figure 11, with the band gap of 0.208 eV.
The adsorption energy, transfer charge, band gap, and band gap change rate of H2S adsorbed by different substrate materials were summarized, as shown in Table 2. It can be seen from the table that compared with the intrinsic substrate, the absolute values of adsorption energy and transfer charge of H2S on doped substrate have increased, indicating the enhancement of adsorption performance. However, it does not mean that these doped substrate materials can be used as sensitive materials for sensors. For example, after adsorption of H2S by Zn–MoSe2, the band gap change rate was −7.06%, indicating its low sensitivity to H2S, but due to its larger adsorption energy (−1.361 eV), it can be applied to gas elimination [49]. In this work, the band gap change rate of Zn–SnP3 after adsorption of H2S is as high as −28.52%, significantly higher than that of other references, which strongly suggests that Zn–SnP3 can be used as a sensitive material for H2S.
In addition, the recovery time is one of the important indicators of the gas sensor. Generally, a good gas sensor should have a short recovery time, and the recovery time (τ) can be estimated from Van’t Hoff–Arrhenius expression for the rate constant, which can be calculated by the following definition [51]:
τ = γ 1 exp ( E a / T K B )
where according to transition state theory the attempt frequency γ is commonly assumed to be ‘a typical value’ in the range of 1012–1013 s−1 [52], which is taken here to be 1012 s−1. Ea represents the adsorption energy with the values shown in Table 1. T is the temperature, in K. KB denotes the Boltzmann constant with a value of 8.62 × 10−5 eV K−1. The recovery time increases with the absolute value of the adsorption energy and decreases with the temperature. The recovery time of the H2S gas molecule at different temperatures is shown in Figure 12. The desorption time at 298 K is only 0.064 s, indicating that Zn–SnP3 has a fast recovery rate for H2S at room temperature. As a comparison, the recovery time after H2S adsorption on Cu–SnP3 was calculated, as shown in Figure S3, which shows that the desorption time of H2S adsorbed on Cu–SnP3 is significantly longer than that of Zn–SnP3. In addition, the attempt frequency γ can be associated with the vibrational frequency [51]. Some studies have shown that light exposure is related to gas desorption [53,54], which may be due to the fact that different wavelengths of light change the vibrational frequency, which in turn has an effect on the attempt frequency and alters the recovery time.

4. Conclusions

Overall, this work has used the density functional theory (DFT) to study in detail the adsorption of intrinsic SnP3 and metal atom modified SnP3 to irritant H2S gases. The monolayers of X–SnP3 (X = Cu, Ag, Zn, Cd) with stable structures were constructed, and all four atoms were modified at the same position, showing good consistency. The intrinsic SnP3 monolayer showed only physical adsorption and weak interaction for H2S, while the four doped monolayers showed enhanced adsorption capacity for H2S. Because the adsorption capacity of Cu–SnP3 and Zn–SnP3 is better than that of Ag–SnP3 and Cd–SnP3, the adsorption mechanism of Cu–SnP3 and Zn–SnP3 monolayer to H2S was further investigated. The findings indicate that the strong interactions between gas molecules and Cu–SnP3 and Zn–SnP3 monolayers result from the hybridization of orbitals and transfer of charges between gas molecules and modified Cu and Zn atoms. The Cu–SnP3 monolayer shows chemical adsorption for H2S, while the Zn–SnP3 monolayer shows physical adsorption with a strong interaction for H2S. Co-adsorption studies of ambient gases (N2, O2, and H2O) with H2S showed that SnP3 has good selectivity for H2S. The study of recovery time showed that at 298 K, 398 K, and 498 K, the desorption time of H2S was 0.064 s, 1.227 × 10−4 s, and 2.915 × 10−6 s. It is indicated that H2S has a rapid desorption ability at room temperature. In this work, the Zn–SnP3 monolayer was developed based on metal doping with good selectivity and sensitivity to H2S, which can provide theoretical support for the application of 2D materials in gas sensing.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13202781/s1.

Author Contributions

Conceptualization, H.C.; methodology, C.G. and Y.L.; software, H.C., C.G. and L.L.; validation, H.C.; investigation, P.W. and Z.W.; data curation, H.C. and P.W.; writing—original draft preparation, H.C.; writing—review and editing, L.L. and Y.L.; supervision, Z.W. and Y.L.; project administration, H.Y. and Y.L.; funding acquisition, Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Natural Science Foundation of Chongqing for Distinguished Young Scholars (cstc2021jcyj-jqX0014), and Chongqing Entrepreneurship and Innovation Support Program (CX201803), and the National Key Research and Development Program of China (Grant No.2021YFB2800203).

Data Availability Statement

The data will be furnished upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tsang, Y.F.; Wang, L.; Chua, H. Simultaneous hydrogen sulphide and ammonia removal in a biotrickling filter: Crossed inhibitory effects among selected pollutants and microbial community change. Chem. Eng. J. 2015, 281, 389–396. [Google Scholar] [CrossRef]
  2. Li, D.; Zu, X.; Ao, D.; Tang, Q.; Fu, Y.; Guo, Y.; Bilawal, K.; Faheem, M.B.; Li, L.; Li, S.; et al. High humidity enhanced surface acoustic wave (SAW) H2S sensors based on sol–gel CuO films. Sens. Actuators B Chem. 2019, 294, 55–61. [Google Scholar] [CrossRef]
  3. Yu, J.-G.; Yu, L.-Y.; Yang, H.; Liu, Q.; Chen, X.-H.; Jiang, X.-Y.; Chen, X.-Q.; Jiao, F.-P. Graphene nanosheets as novel adsorbents in adsorption, preconcentration and removal of gases, organic compounds and metal ions. Sci. Total Environ. 2015, 502, 70–79. [Google Scholar] [CrossRef] [PubMed]
  4. Choi, W.; Choudhary, N.; Han, G.H.; Park, J.; Akinwande, D.; Lee, Y.H. Recent development of two-dimensional transition metal dichalcogenides and their applications. Mater. Today 2017, 20, 116–130. [Google Scholar] [CrossRef]
  5. Shinde, P.V.; Saxena, M.; Singh, M.K. Chapter 11—Recent Developments in Graphene-Based Two-Dimensional Heterostructures for Sensing Applications. In Fundamentals and Sensing Applications of 2D Materials; Hywel, M., Rout, C.S., Late, D.J., Eds.; Woodhead Publishing: Sawston, UK, 2019; pp. 407–436. [Google Scholar]
  6. Kumar, R.; Goel, N.; Hojamberdiev, M.; Kumar, M. Transition metal dichalcogenides-based flexible gas sensors. Sens. Actuators A Phys. 2020, 303, 111875. [Google Scholar] [CrossRef]
  7. Ju, L.; Tang, X.; Li, X.; Liu, B.; Qiao, X.; Wang, Z.; Yin, H. NO2 Physical-to-Chemical Adsorption Transition on Janus WSSe Monolayers Realized by Defect Introduction. Molecules 2023, 28, 1644. [Google Scholar] [CrossRef]
  8. Gullman, J.; Olofsson, O. The crystal structure of SnP3 and a note on the crystal structure of GeP3. J. Solid State Chem. 1972, 5, 441–445. [Google Scholar] [CrossRef]
  9. Ghosh, B.; Puri, S.; Agarwal, A.; Bhowmick, S. SnP3: A Previously Unexplored Two-Dimensional Material. J. Phys. Chem. C 2018, 122, 18185–18191. [Google Scholar] [CrossRef]
  10. Slassi, A.; Sorokin, P.B.; Pershin, A. Ohmic/Schottky barrier engineering in metal/SnP3 heterostructures. J. Alloys Compd. 2020, 831, 154800. [Google Scholar] [CrossRef]
  11. Sun, S.; Meng, F.; Wang, H.; Wang, H.; Ni, Y. Novel two-dimensional semiconductor SnP3: High stability, tunable bandgaps and high carrier mobility explored using first-principles calculations. J. Mater. Chem. A 2018, 6, 11890–11897. [Google Scholar] [CrossRef]
  12. Feng, L.-P.; Li, A.; Wang, P.-C.; Liu, Z.-T. Novel Two-Dimensional Semiconductor SnP3 with High Carrier Mobility, Good Light Absorption, and Strong Interlayer Quantum Confinement. J. Phys. Chem. C 2018, 122, 24359–24367. [Google Scholar] [CrossRef]
  13. Fan, X.; Mao, J.; Zhu, Y.; Luo, C.; Suo, L.; Gao, T.; Han, F.; Liou, S.-C.; Wang, C. Superior Stable Self-Healing SnP3 Anode for Sodium-Ion Batteries. Adv. Energy Mater. 2015, 5, 1500174. [Google Scholar] [CrossRef]
  14. Park, J.W.; Park, C.M. Electrochemical Li Topotactic Reaction in Layered SnP3 for Superior Li-Ion Batteries. Sci. Rep. 2016, 6, 35980. [Google Scholar] [CrossRef]
  15. Verma, R.; Didwal, P.N.; Ki, H.S.; Cao, G.; Park, C.J. SnP3/Carbon Nanocomposite as an Anode Material for Potassium-Ion Batteries. ACS Appl. Mater. Interfaces 2019, 11, 26976–26984. [Google Scholar] [CrossRef] [PubMed]
  16. Niu, F.; Cai, M.; Pang, J.; Li, X.; Yang, D.; Zhang, G. Gas molecular adsorption effects on the electronic and optical properties of monolayer SnP3. Vacuum 2019, 168, 108823. [Google Scholar] [CrossRef]
  17. Du, J.; Jiang, G. First-principle study on monolayer and bilayer SnP3 sheets as the potential sensors for NO2, NO, and NH3 detection. Nanotechnology 2020, 31, 325504. [Google Scholar] [CrossRef]
  18. Ramzan, M.S.; Kuc, A.B.; Kim, H.S. Electronic fingerprint mechanism of NOx sensor based on single-material SnP3 logical junction. NPJ Comput. Mater. 2022, 8, 220. [Google Scholar] [CrossRef]
  19. Gao, H.; Liu, Z. DFT study of NO adsorption on pristine graphene. RSC Adv. 2017, 7, 13082–13091. [Google Scholar] [CrossRef]
  20. Zeng, Y.; Lin, S.; Gu, D.; Li, X. Two-Dimensional Nanomaterials for Gas Sensing Applications: The Role of Theoretical Calculations. Nanomaterials 2018, 8, 851. [Google Scholar] [CrossRef]
  21. Lakshmy, S.; Sanyal, G.; Kalarikkal, N.; Chakraborty, B. Pristine and metal decorated biphenylene monolayer for enhanced adsorption of nitrobenzene: A DFT approach. Appl. Surf. Sci. 2023, 613, 155995. [Google Scholar] [CrossRef]
  22. Meng, R.; da Costa Pereira, L.; Locquet, J.-P.; Afanas’ev, V.; Pourtois, G.; Houssa, M. Hole-doping induced ferromagnetism in 2D materials. NPJ Comput. Mater. 2022, 8, 230. [Google Scholar] [CrossRef]
  23. Zhou, M.; Lu, Y.H.; Cai, Y.Q.; Zhang, C.; Feng, Y.P. Adsorption of gas molecules on transition metal embedded graphene: A search for high-performance graphene-based catalysts and gas sensors. Nanotechnology 2011, 22, 385502. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, Z.; Gao, C.; Hou, S.; Yang, H.; Shao, Z.; Xu, S.; Ye, H. A DFT study of As doped WSe2: A NO2 sensing material with ultra-high selectivity in the atmospheric environment. Mater. Today Commun. 2021, 28, 102654. [Google Scholar] [CrossRef]
  25. Tao, L.; Dastan, D.; Wang, W.; Poldorn, P.; Meng, X.; Wu, M.; Zhao, H.; Zhang, H.; Li, L.; An, B. Metal-Decorated InN Monolayer Senses N2 against CO2. ACS Appl. Mater. Interfaces 2023, 15, 12534–12544. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, Y.; Zhang, W.; Feng, Y.; Ma, J. Promoted CO2 electroreduction over indium-doped SnP3: A computational study. J. Energy Chem. 2020, 48, 1–6. [Google Scholar] [CrossRef]
  27. Liu, Y.; Zhou, Q.; Mi, H.; Wang, J.; Zeng, W. Gas-sensing mechanism of Cr doped SnP3 monolayer to SF6 partial discharge decomposition components. Appl. Surf. Sci. 2021, 546, 149084. [Google Scholar] [CrossRef]
  28. Wang, H.; Hu, X.; Liu, B.; Duan, D. Pd-doped SnP3 monolayer: A new 2D buddy for sensing typical dissolved gases in transformer oil. Appl. Surf. Sci. 2021, 568, 150893. [Google Scholar] [CrossRef]
  29. Delley, B. DFT studies: From molecules and molecular environments to surfaces and solids. Comp. Mater. Sci. 2000, 17, 122–126. [Google Scholar] [CrossRef]
  30. Yang, H.; Li, J.; Shao, Z.; Tan, C.; Gao, C.; Cui, H.; Tang, X.; Liu, Y.; Ye, H.; Zhang, G. Strain-engineered S-HfSe2 monolayer as a promising gas sensor for detecting NH3: A first-principles study. Surf. Interfaces 2022, 34, 102317. [Google Scholar] [CrossRef]
  31. Zhu, X.L.; Liu, P.F.; Zhang, J.; Zhang, P.; Zhou, W.X.; Xie, G.; Wang, B.T. Monolayer SnP3: An excellent p-type thermoelectric material. Nanoscale 2019, 11, 19923–19932. [Google Scholar] [CrossRef]
  32. Song, H.; Zhang, X.; Yuan, P.; Hu, W.; Gao, Z. First-principles study on bilayer SnP3 as a promising thermoelectric material. Phys. Chem. Chem. Phys. 2022, 24, 29693–29699. [Google Scholar] [CrossRef]
  33. Tkatchenko, A.; Scheffler, M. Accurate molecular van der Waals interactions from ground-state electron density and free-atom reference data. Phys. Rev. Lett. 2009, 102, 073005. [Google Scholar] [CrossRef]
  34. Tkatchenko, A.; DiStasio, R.A., Jr.; Head-Gordon, M.; Scheffler, M. Dispersion-corrected Moller-Plesset second-order perturbation theory. J. Chem. Phys. 2009, 131, 094106. [Google Scholar] [CrossRef] [PubMed]
  35. Liu, Y.; Ren, M.; Song, B.; Dong, M. A DFT study of toxic gases (NH3, C2H2, NO) adsorption and detection on metal oxides (CuO, Ag2O, In2O3) modified MoTe2 monolayer. Appl. Surf. Sci. 2023, 622, 156858. [Google Scholar] [CrossRef]
  36. Gao, C.; Zhang, Y.; Yang, H.; Liu, Y.; Liu, Y.; Du, J.; Ye, H.; Zhang, G. A DFT study of In doped Tl2O: A superior NO2 gas sensor with selective adsorption and distinct optical response. Appl. Surf. Sci. 2019, 494, 162–169. [Google Scholar] [CrossRef]
  37. Cui, H.; Jiang, J.; Gao, C.; Dai, F.; An, J.; Wen, Z.; Liu, Y. DFT study of Cu-modified and Cu-embedded WSe2 monolayers for cohesive adsorption of NO2, SO2, CO2, and H2S. Appl. Surf. Sci. 2022, 583, 152522. [Google Scholar] [CrossRef]
  38. Delley, B. Hardness conserving semilocal pseudopotentials. Phys. Rev. B 2002, 66, 155125. [Google Scholar] [CrossRef]
  39. Mulliken, R.S. Electronic Population Analysis on LCAO–MO Molecular Wave Functions. I. J. Chem. Phys. 1955, 23, 1833–1840. [Google Scholar] [CrossRef]
  40. Liu, L.; Zhang, G.; Wang, Z.; Yuan, J.; Tan, S.; Li, Y. Mo2C-Based Microfluidic Gas Sensor Detects SF6 Decomposition Components: A First-Principles Study. Chemosensors 2022, 10, 368. [Google Scholar] [CrossRef]
  41. Becke, A.D.; Edgecombe, K.E. A simple measure of electron localization in atomic and molecular systems. J. Chem. Phys. 1990, 92, 5397–5403. [Google Scholar] [CrossRef]
  42. Savin, A.; Nesper, R.; Wengert, S.; Fässler, T.F. ELF: The Electron Localization Function. Angew. Chem. Int. Ed. 1997, 36, 1808–1832. [Google Scholar] [CrossRef]
  43. Zhang, N.; Li, X.; Ruan, S.; Chen, X.; Li, S.; Hu, T. First-Principles Study of Electronic Properties of Substitutionally Doped Monolayer SnP3. Materials 2022, 15, 2462. [Google Scholar] [CrossRef] [PubMed]
  44. Meng, R.; Houssa, M. Ferromagnetism in two-dimensional metal dibromides induced by hole-doping. Sci. Rep. 2023, 13, 11521. [Google Scholar] [CrossRef]
  45. Xu, L.; Gui, Y.; Li, W.; Li, Q.; Chen, X. Gas-sensing properties of Ptn-doped WSe2 to SF6 decomposition products. J. Ind. Eng. Chem. 2021, 97, 452–459. [Google Scholar] [CrossRef]
  46. Wei, H.; Gui, Y.; Kang, J.; Wang, W.; Tang, C. A DFT Study on the Adsorption of H2S and SO2 on Ni Doped MoS2 Monolayer. Nanomaterials 2018, 8, 646. [Google Scholar] [CrossRef]
  47. Zhang, X.; Hou, M.; Cen, W.; Jiao, W. A DFT study of H2S, H2O, SO2 and CH4 Adsorption Behavior on Graphene Surface Decorated with Alkaline Earth Metals. Surf. Sci. 2022, 726, 122178. [Google Scholar] [CrossRef]
  48. Gao, C.; Liu, X.; Yang, H.; Zhou, Q.; Zhang, Y.; Yang, A.; Ye, H.; Liu, Y.; Zhang, L.; Zhang, G. A theoretical study of Ti-MoSe2 as a noninvasive type-1 diabetes diagnosis material for detecting acetone from exhaled breath. Vacuum 2020, 182, 109729. [Google Scholar] [CrossRef]
  49. Ayesh, A.I. DFT investigation of H2S and SO2 adsorption on Zn modified MoSe2. Superlattices Microstruct. 2022, 162, 107098. [Google Scholar] [CrossRef]
  50. Li, B.; Zhou, Q.; Peng, R.; Liao, Y.; Zeng, W. Adsorption of SF6 decomposition gases (H2S, SO2, SOF2 and SO2F2) on Sc-doped MoS2 surface: A DFT study. Appl. Surf. Sci. 2021, 549, 149271. [Google Scholar] [CrossRef]
  51. Kokalj, A. Formation and structure of inhibitive molecular film of imidazole on iron surface. Corros. Sci. 2013, 68, 195–203. [Google Scholar] [CrossRef]
  52. Jacobs, T.D.B.; Gotsmann, B.; Lantz, M.A.; Carpick, R.W. On the Application of Transition State Theory to Atomic-Scale Wear. Tribol. Lett. 2010, 39, 257–271. [Google Scholar] [CrossRef]
  53. Chen, R.J.; Franklin, N.R.; Kong, J.; Cao, J.; Tombler, T.W.; Zhang, Y.; Dai, H. Molecular photodesorption from single-walled carbon nanotubes. Appl. Phys. Lett. 2001, 79, 2258–2260. [Google Scholar] [CrossRef]
  54. Aasi, A.; Ebrahimi Bajgani, S.; Panchapakesan, B. A first-principles investigation on the adsorption of octanal and nonanal molecules with decorated monolayer WS2 as promising gas sensing platform. AIP Adv. 2023, 13, 025157. [Google Scholar] [CrossRef]
Figure 1. SnP3 monolayer structure, top and side views, with solid circles representing the modification sites.
Figure 1. SnP3 monolayer structure, top and side views, with solid circles representing the modification sites.
Nanomaterials 13 02781 g001
Figure 2. The (a) band structure and (b) phonon spectrum of the intrinsic SnP3 monolayer.
Figure 2. The (a) band structure and (b) phonon spectrum of the intrinsic SnP3 monolayer.
Nanomaterials 13 02781 g002
Figure 3. Optimized structures of (a) Cu–SnP3, (b) Ag–SnP3, (c) Zn–SnP3, and (d) Cd–SnP3.
Figure 3. Optimized structures of (a) Cu–SnP3, (b) Ag–SnP3, (c) Zn–SnP3, and (d) Cd–SnP3.
Nanomaterials 13 02781 g003
Figure 4. Energy band structures of (a) Cu–SnP3, (b) Ag–SnP3, (c) Zn–SnP3, and (d) Cd–SnP3.
Figure 4. Energy band structures of (a) Cu–SnP3, (b) Ag–SnP3, (c) Zn–SnP3, and (d) Cd–SnP3.
Nanomaterials 13 02781 g004
Figure 5. TDOS of (a) Cu–SnP3 and (b) Ag–SnP3. PDOS of (c) Cu–SnP3 and (d) Ag–SnP3.
Figure 5. TDOS of (a) Cu–SnP3 and (b) Ag–SnP3. PDOS of (c) Cu–SnP3 and (d) Ag–SnP3.
Nanomaterials 13 02781 g005
Figure 6. TDOS of (a) Zn–SnP3 and (b) Cd–SnP3. PDOS of (c) Zn–SnP3 and (d) Cd–SnP3.
Figure 6. TDOS of (a) Zn–SnP3 and (b) Cd–SnP3. PDOS of (c) Zn–SnP3 and (d) Cd–SnP3.
Nanomaterials 13 02781 g006
Figure 7. Optimized structures of (a) intrinsic SnP3, (b) Cu–SnP3, (c) Ag–SnP3, (d) Zn–SnP3, and (e) Cd–SnP3 for H2S adsorption.
Figure 7. Optimized structures of (a) intrinsic SnP3, (b) Cu–SnP3, (c) Ag–SnP3, (d) Zn–SnP3, and (e) Cd–SnP3 for H2S adsorption.
Nanomaterials 13 02781 g007
Figure 8. TDOS after adsorption of H2S by (a) Cu–SnP3 and (b) Zn–SnP3. PDOS after adsorption of H2S by (c) Cu–SnP3 and (d) Zn–SnP3.
Figure 8. TDOS after adsorption of H2S by (a) Cu–SnP3 and (b) Zn–SnP3. PDOS after adsorption of H2S by (c) Cu–SnP3 and (d) Zn–SnP3.
Nanomaterials 13 02781 g008
Figure 9. Top view of (a) H2S/Cu–SnP3 and (b) H2S/Zn–SnP3. Side view of (c) H2S/Cu–SnP3 and (d) H2S/Zn–SnP3. CDD of (e) H2S/Cu–SnP3 and (f) H2S/Zn–SnP3. ELF of (g) H2S/Cu–SnP3 and (h) H2S/Zn–SnP3.
Figure 9. Top view of (a) H2S/Cu–SnP3 and (b) H2S/Zn–SnP3. Side view of (c) H2S/Cu–SnP3 and (d) H2S/Zn–SnP3. CDD of (e) H2S/Cu–SnP3 and (f) H2S/Zn–SnP3. ELF of (g) H2S/Cu–SnP3 and (h) H2S/Zn–SnP3.
Nanomaterials 13 02781 g009
Figure 10. Adsorption isotherms of H2S, N2, O2, and H2O on (a) Cu–SnP3 and (b) Zn–SnP3 at pressures ranging from 0 to 200 kPa and at temperatures of 298 K.
Figure 10. Adsorption isotherms of H2S, N2, O2, and H2O on (a) Cu–SnP3 and (b) Zn–SnP3 at pressures ranging from 0 to 200 kPa and at temperatures of 298 K.
Nanomaterials 13 02781 g010
Figure 11. Band structure of Zn–SnP3 after adsorption of H2S.
Figure 11. Band structure of Zn–SnP3 after adsorption of H2S.
Nanomaterials 13 02781 g011
Figure 12. The recovery time of H2S at 298 K, 398 K, and 498 K, respectively.
Figure 12. The recovery time of H2S at 298 K, 398 K, and 498 K, respectively.
Nanomaterials 13 02781 g012
Table 1. Parameters of H2S molecule adsorption by five SnP3 substrates. Ea, Q, and d stand for the adsorption energy, transfer charge, and shortest adsorption distance, respectively.
Table 1. Parameters of H2S molecule adsorption by five SnP3 substrates. Ea, Q, and d stand for the adsorption energy, transfer charge, and shortest adsorption distance, respectively.
SystemEa (eV)Q (e)d (Å)
H2S/SnP3−0.3920.0242.574 (P–H)
H2S/Cu–SnP3−0.7490.2722.336 (Cu–S)
H2S/Ag–SnP3−0.5950.2112.601 (Ag–S)
H2S/Zn–SnP3−0.6390.2342.520 (Zn–S)
H2S/Cd–SnP3−0.4020.1872.877 (Cd–S)
Table 2. Summary of adsorption energy (Ea), transfer charge (Q), band gap of the substrate before and after adsorption of the gases (Eg1 and Eg2), change in band gap (ΔEg = Eg2Eg1), and rate of change in band gap (ΔEg/Eg1) for H2S adsorption on different substrates.
Table 2. Summary of adsorption energy (Ea), transfer charge (Q), band gap of the substrate before and after adsorption of the gases (Eg1 and Eg2), change in band gap (ΔEg = Eg2Eg1), and rate of change in band gap (ΔEg/Eg1) for H2S adsorption on different substrates.
SubstrateEa (eV)Q (e)Eg1 (eV)Eg2 (eV)ΔEg (eV)ΔEg/Eg1Reference
MoSe2−0.250−0.0631.6091.605−0.004−0.25%[49]
SnP3−0.3630.0500.5460.543−0.003−0.5%[17]
MoS2/02.062.03−0.03−1.5%[50]
Zn–MoSe2−1.361−0.3230.3260.303−0.023−7.06%[49]
Zn–SnP3−0.6390.2340.2910.208−0.083−28.52%This work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cui, H.; Gao, C.; Wang, P.; Li, L.; Ye, H.; Wen, Z.; Liu, Y. DFT Study of Zn-Modified SnP3: A H2S Gas Sensor with Superior Sensitivity, Selectivity, and Fast Recovery Time. Nanomaterials 2023, 13, 2781. https://doi.org/10.3390/nano13202781

AMA Style

Cui H, Gao C, Wang P, Li L, Ye H, Wen Z, Liu Y. DFT Study of Zn-Modified SnP3: A H2S Gas Sensor with Superior Sensitivity, Selectivity, and Fast Recovery Time. Nanomaterials. 2023; 13(20):2781. https://doi.org/10.3390/nano13202781

Chicago/Turabian Style

Cui, Hongyuan, Chenshan Gao, Pengwei Wang, Lijie Li, Huaiyu Ye, Zhongquan Wen, and Yufei Liu. 2023. "DFT Study of Zn-Modified SnP3: A H2S Gas Sensor with Superior Sensitivity, Selectivity, and Fast Recovery Time" Nanomaterials 13, no. 20: 2781. https://doi.org/10.3390/nano13202781

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