Next Article in Journal
A Scale Factor Calibration Method for MEMS Resonant Accelerometers Based on Virtual Accelerations
Next Article in Special Issue
Protein-Based Anchoring Methods for Nucleic Acid Detection in Lateral Flow Format Assays
Previous Article in Journal
Investigation of Parallel and Orthogonal MIMO Antennas with Two-Notched Structures for Ultra-Wideband Application
Previous Article in Special Issue
Research on Integrated 3D Printing of Microfluidic Chips
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two−Dimensional Planar Penta−NiPN with Ultrahigh Carrier Mobility and Its Potential Application in NO and NO2 Gas Sensing

1
Wuhan Second Ship Design and Research Institute, Wuhan 430205, China
2
College of Railway Rolling Stock, Wuhan Railway Vocational College of Technology, Wuhan 430205, China
3
School of Science, Wuhan University of Technology, Wuhan 430070, China
4
School of Integrated Circuits, Peking University, Beijing 100871, China
*
Authors to whom correspondence should be addressed.
Micromachines 2023, 14(7), 1407; https://doi.org/10.3390/mi14071407
Submission received: 19 June 2023 / Revised: 6 July 2023 / Accepted: 7 July 2023 / Published: 12 July 2023
(This article belongs to the Special Issue Recent Progress of Lab-on-a-Chip Assays)

Abstract

:
Two−dimensional (2D) materials with novel structures and electronic properties are promising candidates for the next generation of micro− and nano−electronic devices. Herein, inspired by the recent experimental synthesis of penta−NiN2 (ACS Nano, 2021, 15, 13539–13546), we propose for the first time a novel ternary penta−NiPN monolayer with high stability by partial element substitution. Our predicted penta−NiPN monolayer is a quasi−direct bandgap (1.237 eV) semiconductor with ultrahigh carrier mobilities (103–105 cm2V−1s−1). Furthermore, we systematically studied the adsorption properties of common gas molecules (CO, CO2, CH4, H2, H2O, H2S, N2, NO, NO2, NH3, and SO2) on the penta−NiPN monolayer and its effects on electronic properties. According to the energetic, geometric, and electronic analyses, the penta−NiPN monolayer is predicted to be a promising candidate for NO and NO2 molecules. The excellent electronic properties of and the unique selectivity of the penta−NiPN monolayer for NO and NO2 adsorption suggest that it has high potential in advanced electronics and gas sensing applications.

1. Introduction

Sensors are electronic devices designed to detect and perceive their surroundings or specific substances, such as gas, light, temperature, humidity, the human body, etc., and they received significant concern because of their importance to the information society [1,2,3]. They can convert analog signals into digital signals, then send them to the central processor for processing, and finally realize the detection of the calibration object. Due to their importance and irreplaceability in today’s society, the development of new sensors is a popular research topic.
Currently, gas sensors are among the most widely studied type of sensors. High−performance gas sensors are in great demand for industrial production, environmental monitoring, and clinical medicine. Two−dimensional (2D) materials have emerged as promising gas−sensitive materials due to their large surface area, high surface activity, and abundant surface active sites [4,5,6,7]. For example, Ou et al. proved that 2D SnS2 is a potential material for selective and reversible NO2 detection at low operating temperatures [8], while Qin et al. developed a flexible paper substrate sensor based on 2D WS2−x for NH3 detection at room temperature [9]. So far, many types of gas detection have been realized experimentally or theoretically based on various 2D materials, such as NO, CO, SO2, NH3, H2O, and others [10,11,12,13]. However, developing a high−sensitivity, fast−response, and completely desorbed gas sensor based on 2D materials under harsh working conditions remains a challenge. Nonetheless, the search for new 2D materials with excellent electronic properties that can be used in gas sensors remains a major research focus.
Recently, Bykov et al. [14] successfully experimentally realized penta−NiN2 with an ideal Cairo tessellation via a high−pressure route. Notably, this penta−NiN2 has a typical layered structure and is stable at room temperature, suggesting that a penta−NiN2 monolayer can be obtained by mechanical exfoliation, similar to that with graphene or black phosphorene [15,16]. Theoretical studies have shown that the penta−NiN2 monolayer is a direct bandgap semiconductor with moderate thermal conductivity and outstanding mechanical properties, and it can be applied to bifunctional oxygen electrocatalysts and gas sensors [17,18,19,20,21]. In fact, before the experimental synthesis of penta−NiN2, similar structures had been studied extensively, such as penta−MX2 (M = Ni, Pd, Pt; X = N, P, As, Sb) [17,22,23,24,25,26,27]. This family of 2D materials demonstrated excellent properties, such as a suitable bandgap, a high optical absorption coefficient, ultrahigh carrier mobility, and so on [22,23,24,25]. However, most research on this family of 2D materials focused on binary systems, and previous studies showed that the properties of systems can be effectively improved by introducing homologous elements, such as α−P1−xAsx [28]. Therefore, the successful experimental realization of penta−NiN2 provides a material basis for further study of relevant ternary systems.
In this work, we predicted a new planar ternary penta−NiPN monolayer by atomic substitution of the penta−NiN2 monolayer using first−principles calculations. After careful stability analysis, we focused on electronic structures and adsorption properties for 12 kinds of common gases. Electronic property analysis showed that the penta−NiPN monolayer is a quasi−direct bandgap semiconductor (1.237 eV) with ultrahigh carrier mobility (up to 105 cm2V−1s−1). The results of gas adsorption revealed that the penta−NiPN monolayer exhibits good selectivity for NO and NO2, indicating its potential as a gas−sensitive material for NO and NO2.

2. Methods

All structural optimization and electronic properties in this work were obtained via the Vienna Ab initio Simulation Package (VASP) [29,30], based on density functional theory (DFT) [31,32]. The cutoff energy was set to 500 eV, and the generalized gradient approximation (GGA) [33] with Perdew–Burke–Ernzerhof (PBE) functional form was used to describe the exchange–correlation energy. However, to overcome the bandgap underestimation problem of GGA–PBE for semiconductors or insulators, the hybrid functional HSE06 [34] was also employed for electronic structure calculations. Moreover, the valence electrons of different elements in the calculation were set as follows: 1s for H, 2s and 2p for C/N/O, 3s and 3p for P/S, and 4d and 5s for Ni. The convergence accuracy of total energy and force in the calculation was set a 10−7 eV and 0.005 eV/Å per atom, respectively. For geometry optimization, a 12 × 12 × 1 and a 5 × 5 × 1 k−point grid, following the scheme of Monkhorst–Pack [35], was used to sample the Brillouin zones for unit cells of the penta−NiPN monolayer and the gas adsorption supercell model, respectively. Denser k−point grids were used for self−consistent and electronic property calculations. Grimme’s DFT–D2 [36] was adopted to describe the van der Waals (vdW) corrections in gas adsorption. A vacuum layer of 20 Å length was introduced into all models to avoid interactions between adjacent layers. Phonon dispersion for the new predicted penta−NiPN monolayer was performed using VASP and Phonopy code [37], based on density functional perturbation theory (DFPT). Additionally, ab initio molecular dynamics (AIMD) simulations lasting for 5 ps at room temperature were employed to evaluate the thermal stability of the penta−NiPN monolayer using a 4 × 4 × 1 supercell.

3. Results and Discussion

3.1. Structure and Stability

The penta−NiPN monolayer, shown in Figure 1a, is a perfect single atomic thickness planar 2D material, similar to graphene. The penta−NiPN monolayer consists of Ni2P2N pentagons and Ni2PN2 pentagons with a calculated lattice constant a/b of 4.995/5.011 Å (see Table 1). This value is larger than that of the penta−NiN2 monolayer (4.53 Å) [17] and smaller than that of the penta−NiP2 monolayer (5.55 Å) [22]. The penta−NiPN monolayer has lower symmetry (space group: Pb21m, No.26) than a single element system (space group: P4/mbm, No.127) due to the existence of two non−metallic elements (N and P). As a result, two kinds of unequal Ni−N bonds and Ni−P bonds with bond lengths of 1.929/1.910 Å and 2.125/2.107 Å, respectively, exist in the penta−NiPN monolayer. The bond lengths in the penta−NiPN monolayer are between the lengths of the Ni−N and Ni−P bonds in the penta−NiN2 monolayer (1.88 Å) and the penta−NiP2 monolayer (2.16 Å). The N−P bond length in the penta−NiPN monolayer is 1.605 Å, while the N−N bond length and P−P bond length in penta−NiN2 monolayer and penta−NiP2 monolayer are 1.24 Å and 2.11 Å, respectively.
The ideal pentagonal Cairo tiling maintains bond angles of 90° and 120°, as depicted in Figure 1b. It is clear that pentagonal monolayers consisting of two (e.g., penta−NiN2 and −NiP2) or three elements cannot form a perfect pentagonal Cairo tile. Figure 1b shows the corresponding bond lengths and bond angles of Ni2N3, Ni2P2N/Ni2PN2, and Ni2P3 pentagons in the penta−NiN2, NiPN, and NiP2 monolayers. For the Ni2N3 and Ni2P3 pentagons, the corresponding bond angles of N−Ni−N (90°) and P−Ni−P (90.01°) coincide with the ideal pentagonal Cairo tile. However, among other bond angles, such as Ni−N(P)−N(P) and Ni−N(P)−Ni, the results largely deviate from the ideal pentagonal Cairo tile. Among pentagonal monolayers, the penta−NiPN monolayer is notably more complex than the penta−NiN2 and −NiP2 monolayers. Due to the introduction of two non−equivalent non−metallic elements, there exist non−equivalent pentagons (Ni2P2N and Ni2PN2) in the lattice (as seen in Figure 1b). Generally, the Ni2P2N and Ni2PN2 pentagons undergo varying degrees of distortion compared to the ideal pentagonal Cairo tiling. Naturally, this atomic divergence leads to an array of unique characteristics in this predicted monolayer.
Structural stability forms the foundation for all subsequent studies. We then focus on the stability of the predicted penta−NiPN monolayer, covering its kinetic, thermodynamic, and mechanical stability. Based on DFPT, we initially calculated the phonon dispersion of the penta−NiPN monolayer, as displayed in Figure 1d. The negligible imaginary frequency of the phonon spectrum (occurring near the Γ−point) is primarily due to calculation errors. Despite this fact, the kinetic stability of the penta−NiPN monolayer is unmistakable. Additionally, the phonon density of states (DOS) analysis shows that the low−frequency portion predominantly arises from Ni−P bonds, whereas the high−frequency portion is dominated by Ni−N and N−P bonds. This result indicates the robust bonding traits of the penta−NiPN monolayer. The thermodynamic stability analysis of the penta−NiPN monolayer at room temperature (300 K) was performed via ab initio molecular dynamics simulations. As shown in Figure 1e, the inherent energy fluctuation of the system, covering a simulation time of 5 ps, was within 0.025 eV/atom. No observable structural collapse occurred in the atomic structure of the final state, indicating the high thermodynamic stability of the predicted penta−NiPN monolayer at room temperature. Cohesive energy (Ecoh), as another thermodynamic indicator, can evaluate the predicted system’s relative realizability under experimental conditions. The Ecoh of the penta−NiPN monolayer is defined as E coh = 2 E Ni + 2 E P + 2 E N E NiPN / 6 , where E Ni / E P / E N and E NiPN are the energy of a single Ni/P/N atom and the total energy of penta−NiPN monolayer, respectively. According to the definition, a higher Ecoh value signifies greater stability. The calculated Ecoh of the penta−NiPN monolayer is 4.55 eV, which is higher than that of the NiP2 monolayer (4.09 eV, 3.944 eV) [22,23], silicene (3.94 eV) [24], and phosphorene (3.477 eV) [38] but slightly lower than that of penta−NiN2 (4.98 eV) [17]. As silicene and phosphorene have been obtained successfully in experiments, the high−realization potential of the penta−NiPN monolayer is likewise promising, especially considering that Bykov et al. [14] recently achieved a room−temperature stable penta−NiN2 layer experimentally. Therefore, it is reasonable to anticipate the experimental realization of a penta−NiPN monolayer (such as via element substitution doping) in the near future.
In addition to kinetics and thermodynamic stability, mechanical stability is another critical factor to consider for newly predicted monolayers. According to the Born–Huang criterion [39], stable 2D materials should satisfy: C 11 C 22 C 12 2 > 0 and C 66 > 0 , where C11, C22, C12, and C66 are the independent elastic constants of the predicted monolayer. For the penta−NiPN monolayer, the calculated C11, C22, C12, and C66 are 158.25 N m−1, 154.64 N m−1, 31.85 N m−1, and 41.22 N m−1, respectively. These values confirm that the predicted penta−NiPN monolayer in this work possesses good mechanical stability.
Furthermore, Young’s modulus (Y) and Poisson’s ratio (υ) are important indices for measuring the mechanical properties of materials. We further evaluate these properties for the predicted penta−NiPN monolayer using the calculated independent elastic constants (C11, C22, C12, and C66) above. The angle dependent in−plane Young’s modulus Y(θ) and Poisson’s ratio υ(θ) can be expressed as follows [40]:
Y ( θ ) = C 11 C 12 C 12 2 C 11 sin 4 θ + A sin 2 θ cos 2 θ + C 22 cos 4 θ
υ ( θ ) = C 12 sin 4 θ B sin 2 θ cos 2 θ + C 12 cos 4 θ C 11 sin 4 θ + A sin 2 θ cos 2 θ + C 22 cos 4 θ
where A = ( C 11 C 22 C 12 2 ) / C 66 - 2 C 12 and B = C 11 + C 22 ( C 11 C 22 C 12 2 ) / C 66 . As shown in Figure 2, the Young’s modulus and Poisson’s ratio of the penta−NiPN monolayer are anisotropic and exhibit angle dependence, different from those of the penta−NiN2 and −NiP2 monolayer. The maximum Young’s modulus is 151.69 N m−1 along the x direction (Y11, θ = 0°/180°), while the Y22 (148.23 N m−1, θ = 90°/270°) is slightly smaller than that of Y11. The minimum Young’s modulus is 114.66 N m−1 along the diagonal direction (θ = 45°/135°/225°/315°). The Young’s modulus of the penta−NiPN monolayer is lower than that of the penta−NiN2 monolayer (168.8 N m−1) [17], but higher or comparable to that of the penta−NiP2 monolayer (122.19 N m−1) [22]. The corresponding Poisson’s ratio of the penta−NiPN monolayer is shown in Figure 2b. In contrast to the Young’s modulus, the minimum value of Poisson’s ratio is obtained in the axial direction υ22 (θ = 90°/270°), where the value is 0.201, while υ11 (0.206, θ = 0°/180°) is slightly higher than υ22. In addition, the penta−NiPN monolayer has a maximum Poisson’s ratio value of 0.391 in the diagonal direction (θ = 45°/135°/225°/315°). The minimum Poisson’s ratio of the penta−NiPN monolayer is much higher than that of the penta−NiN2 monolayer (0.130) [17] but comparable to that of the penta−NiP2 monolayer (0.22). Therefore, in general, the penta−NiPN monolayer is less stiff and more flexible than the penta−NiN2 and −NiP2 monolayers. With the intrinsic anisotropy of the penta−NiPN monolayer considered, it can be anticipated that penta−NiPN has a more diverse and adjustable set of mechanical properties than the penta−NiN2 and −NiP2 monolayers.
In short, the penta−NiPN monolayer is predicted to be robust and stable, providing a strong theoretical basis for its design, and it has also been shown to be feasible for experimental investigation. Its diverse structural and mechanical properties make it a highly promising candidate for use as a 2D functional material.

3.2. Electronic Structure

Figure 3a,b shows the electronic band structures and density of states (DOS) with partial density of states (PDOS) of the penta−NiPN monolayer. The electronic band structures are presented using both the standard GGA–PBE (solid gray line) and hybrid functional HSE06 (solid red line). For the DOS and PDOS, only the results at the HSE06 level are presented. The calculated band morphologies based on GGA–PBE and HSE06 are almost identical. Unlike the penta−NiN2 and −NiP2 monolayers, which exhibit direct bandgap features, the penta−NiPN monolayer is an indirect bandgap semiconductor with a bandgap value of 0.518/1.237 eV at the PBE/HSE06 level. As summarized in Table 1, the bandgap of the penta−NiPN monolayer is larger than that of the penta−NiN2 (1.10 eV) and −NiP2 (0.81 eV) monolayers. The valence band maximum (VBM) and conduction band minimum (CBM) of the penta−NiPN monolayer are both located near the S−point along the S−X direction, whereas for the penta−NiN2 and −NiP2 monolayers, the VBM and CBM are both located at the S−point. It should be noted that, although the penta−NiPN monolayer is an indirect bandgap semiconductor, the energy difference (ΔE) between its direct and indirect bandgaps is only 0.0013/0.0449 eV at the PBE/HSE06 level, making it a quasi−direct bandgap semiconductor. The combination of a suitable bandgap and quasi−direct band features in the penta−NiPN monolayer may have potential applications in optoelectronic devices.
As shown in Figure 3b, the PDOS results indicate that there is strong hybridization between the Ni 3d orbits and P 3p and N 2p orbits. Furthermore, the orbital contributions to VBM and CBM are dominated by Ni 3d orbits, which is further confirmed by the spatial distribution of the wave−functions corresponding to the VBM and CBM of the penta−NiPN monolayer at the GGA–PBE level (see Figure 3c). The spatial distribution of the wave functions of the VBM is mainly located around Ni atoms, while that of CBM is located around Ni atoms and N atoms. Regarding bonding properties, as shown in Figure 3d, the electron localization function (ELF) [41,42] reveals that Ni−N and Ni−P are typical ionic bonds with the electron mainly located around the non−metal element. The N−P bond is mainly an ionic bond with some covalent bond characteristics. The results of Bader charge analysis [43] indicate that both Ni and P lose electrons during the bonding process of the penta−NiPN monolayer, with −0.6848 e/atom and −1.5487 e/atom, respectively, while N gains +2.2335 e/atom (“+” and “−” representing gained and lost electrons, respectively). The charge transfer during the bonding process is mainly determined by the electronegativity of each element. The corresponding electronegativities of Ni, P, and N are 1.91, 2.19, and 3.04, respectively. As a result, N gains the most electrons during the bonding process, as expected.
To evaluate the potential of the penta−NiPN monolayer for electronic devices, we calculated its acoustic phonon−limited carrier mobility, a critical parameter in this regard. Since the band morphologies of the monolayer are almost identical at the GGA–PBE and HSE06 levels, we used the GGA–PBE results directly to evaluate the carrier mobility here. The method used is based on the deformation potential theory (DPT) proposed by Bardeen and Shockley [44], with further details described in numerous works [45,46], which we do not repeat here. To calculate the carrier mobility, we evaluated the carrier effective mass m*, the deformation potential constant |Eil|, and the elastic modulus C2D of the penta−NiPN monolayer, as summarized in Table 2. In addition, Figure 4 shows more details for the calculation of C2D and Eil. Our calculations show that the effective masses m* of electrons and holes in the penta−NiPN monolayer along the a/b−direction are 0.38/0.36 me and 0.22/0.24 me, respectively, which are higher than those in the penta−NiP2 monolayer (0.106/0.140 me and 0.119/0.170 me along the a/b−direction, respectively) [22]. Meanwhile, the deformation potential constant of the penta−NiPN monolayer is much smaller than that of penta−NiP2 (2.10/0.85 eV vs. 5.23/5.23 eV; 0.74/0.99 eV vs. 1.53/1.53 eV), whereas the elastic constant of the penta−NiPN monolayer is larger than that of the penta−NiP2 monolayer (147.68/146.24 N m−1 vs. 118.19/118.19 N m−1). By combining the effective mass, deformation potential constant, and elastic constant, we calculated the carrier mobility of penta−NiPN monolayer. The obtained electron mobilities are 0.51 and 3.24 × 104 cm2V−1 s−1 along the a− and b−directions, respectively, whereas the hole mobility is even higher, reaching up to 11.36 and 5.76 × 104 cm2V−1s−1 along a− and b−directions, respectively. These values are much higher than those of some typical 2D materials, such as MoS2 (~49 to 200 cm2V−1s−1) [47], GaPS4 (~14 to 1306 cm2V−1s−1) [40], and GeP3 (~14 to 190 cm2V−1s−1) [48], and comparable to that of phosphorene (~104 cm2V−1s−1) [49] and penta−MX2 (M = Ni, Pd, Pt; X = N, P, As, Sb; 103~105 cm2V−1s−1) [22,23,24]. The suitable bandgap and ultra−high carrier mobility in the penta−NiPN monolayer make it promising for nanoelectronics and microelectronics.

3.3. Gas Adsorption

Today, one of the most significant uses of 2D materials is in gas sensors. Previous studies have shown that 2D materials with pentagonal structures are good gas−sensitive materials, such as penta−graphene [50], penta−BCN [51], and penta−BeP2 [52]. On the other hand, penta−NiN2 [21] and penta−PdAs2 [27], which belong to the same family as the penta−NiPN proposed in this work, have proved to be very good gas−sensitive materials as well. Therefore, we believe that the gas−sensitive properties of the penta−NiPN monolayer are worth exploring. In this section, we focus on evaluating the potential applications of the penta−NiPN monolayer in gas sensors by exploring its gas adsorption properties. First, there are eight unequal adsorption sites in the penta−NiPN monolayer, as labeled in Figure 1f. These eight adsorption sites comprise the top site of the Ni/P/N atom (site 1/2/3), a bridge site located along the Ni−P/Ni−N/N−P bond (site 4/5/6), and a hollow site present in the Ni2P2N/Ni2PN2 pentagon (site 7/8). For gases, we selected 12 typical gas molecules as the study objects, i.e., CO, CO2, CH4, H2, H2O, H2S, N2, NO, NO2, NH3, SO2, and O2. Our main reason for choosing H2O and O2 was to evaluate the moisture sensitivity and oxidation resistance of the penta−NiPN monolayer, respectively. The remaining gases included greenhouse gases, toxic gases, or gases commonly found in the air.
To identify the optimal adsorption sites of the 12 gas molecules on the penta−NiPN monolayer, we constructed 3 × 3 supercells (area: 14.98 Å × 15.03 Å). Figure 5 illustrates the most favorable adsorption configurations for the selected 12 molecules on the penta−NiPN monolayer via top and side views. For most gas molecules, the preferred adsorption site was the top site of the Ni atom, including CO, H2, H2S, NO, NO2, NH3, and SO2. For CO2 and N2, the top site of the N atom was more favorable. CH4 tended to be adsorbed at the bridge site along the Ni−N bond, while H2O was adsorbed at the hollow site of the Ni2PN2 pentagon. The most unusual case was during the adsorption of O2 on the penta−NiPN monolayer, whereby it reacted directly with the substrate, leading to the dissociation of O2 molecules into O atoms and ultimately forming a new material. However, for a reusable gas−sensitive material, it is necessary to have the ability to both adsorb and release gases. According to the results of O2 adsorption on the penta−NiPN monolayer, this process is irreversible. Therefore, our findings indicate that the penta−NiPN monolayer may require an oxygen−free environment if used as a medium material for a gas sensor. We do not discuss the case of O2 in the subsequent studies, considering the strong reactivity during the adsorption of oxygen and the penta−NiPN monolayer.
We begin our analysis of gas adsorption on the penta−NiPN monolayer with the evaluation of the adsorption energy (Ea) and adsorption distance (d). The Ea of gas adsorption on the penta−NiPN monolayer is defined as follows: E a = E NiPN gas E NiPN E gas , where E NiPN gas , E NiPN , and E gas are the total energy of the NiPN monolayer with gas adsorption, a pristine NiPN monolayer, and a single gas molecule, respectively. By definition, a negative Ea implies that gas adsorption is an exothermic process and can be spontaneous. Conversely, if the Ea value is positive, the process is endothermic and non−spontaneous. The magnitude of the absolute value determines the likelihood of the reaction. The adsorption distance d refers to the minimum distance between the gas molecule and the substrate at the optimal adsorption site. We conducted a statistical analysis of Ea and d for 11 gases adsorbed on the penta−NiPN monolayer, as shown in Figure 6 and Table 3. For the 11 gases studied, the adsorption energy on the NiPN monolayer is negative (−1.011 to −0.072 eV), indicating that all adsorption could be spontaneous. Furthermore, the absolute values of adsorption energy are in the order of | E a NO 2 | > | E a NO | > | E a CO | > | E a NH 3 | > | E a SO 2 | > | E a H 2 S | > | E a H 2 O | > | E a CO 2 | > | E a CH 4 | > | E a N 2 | > | E a H 2 | . The largest was | E a NO 2 | (1.011 eV), followed by | E a NO |   = 0.751 eV, indicating that the penta−NiPN monolayer is an excellent trapping material for these two gases. On the other hand, H2 and N2 exhibited very small adsorption energy values (0.072 eV and 0.100 eV), suggesting that they are challenging to capture in normal environments. Regarding the adsorption distances, the values ranged from 1.834 Å (CO) to 3.117 Å (N2). The adsorption distances for NO and NO2 are 1.862 Å and 2.065 Å, respectively, which are shorter than the values observed for NO and NO2 absorption on the penta−NiN2 monolayer (2.190 Å and 2.124 Å) [21]. Adsorption energy and distance can characterize the strength or weakness of interactions between gas molecules and host materials. Our results show that six gas molecules, including CO, H2S, NO, NO2, NH3, and SO2, had relatively strong interactions with the penta−NiPN monolayer.
The study of the interaction between adsorbed gas molecules and the host material was further characterized using the charge density difference (CDD). Additionally, by conducting Bader charge analysis, we obtained the value of charge transfer (Q) between gas molecules and substrates. The transferred charge values for all models are summarized in Table 3, and the CDDs are plotted in Figure 7. In general, the values of the transferred charge were arranged in the order of Q NO 2 > Q NO > Q SO 2 > Q NH 3 > Q CO > Q H 2 S > Q CO 2 > Q H 2 O > Q N 2 > Q H 2 > Q CH 4 . CO2, CH4, H2, H2O, N2, and the host material showed very little charge transfer (< 0.1 e), indicating weak interactions that can be neglected. The remaining six gas molecules can be categorized into two classes, depending on the direction of charge transfer between them and the penta−NiPN monolayer. The first class comprised electron donors, such as H2S and NH3, which donated electrons (−0.100 e and −0.103 e, respectively) to the penta−NiPN monolayer, with the corresponding Q < 0. In contrast, the second class comprised electron acceptors, such as CO, NO, NO2, and SO2. For these four gas molecules, the electrons transferred from the host material to the gas molecules during the absorption process. CO and SO2 both obtained 0.100 e and 0.187 e, respectively, which were lower than NO (0.216 e) and NO2 (0.553 e). Therefore, in comparison with other gas molecules, NO and NO2 exhibited higher adsorption energy and larger charge transfer during adsorption, suggesting that these gases are more easily adsorbed on the penta−NiPN monolayer, with stronger coupling between them and the host material. Taken together, these results indicate that penta−NiPN monolayer may be a promising material for sensing NO and NO2 gases.
Last but not least, we further investigated the electronic properties of the penta−NiPN monolayer with adsorption of various gas molecules. Spin polarization was considered in the calculation. We found that, except for NO and NO2 adsorbed penta−NiPN monolayers with residual magnetic moments (M), all the other adsorbed systems were nonmagnetic. Therefore, we present only the results of NO and NO2, considering spin polarization in the latter electronic structures. When nonmagnetic molecules, such as CO, CO2, CH4, H2, H2O, H2S, N2, NH3, and SO2, were adsorbed, the system remained a nonmagnetic semiconductor with various bandgaps (see Figure 8). The results of PDOS indicated that the orbital hybridization between the gas molecules and the host material was weak or almost non−existent after the adsorption of nine nonmagnetic molecules. Most significantly, the orbital energy levels of gas molecules were primarily in the deep valence band and were distant from the Fermi level (see Figure 8).
Notably, the adsorption of magnetic NO and NO2 molecules (1.00 µB) could transform the penta−NiPN monolayer into a magnetic semiconductor. After the adsorption of NO and NO2 on the penta−NiPN monolayer, the magnetic moment of NO and NO2 was reduced to 0.604 µB and 0.242 µB, respectively. At the same time, the magnetic moment of 0.091 µB and 0.636 µB was introduced into the host material penta−NiPN monolayer, resulting in a total magnetic moment of 0.695 µB and 0.878 µB, respectively, in each system (see Table 3). Therefore, the penta−NiPN monolayer can be electrically and magnetically sensitive to both NO and NO2 molecules. In Figure 9, we have plotted the spin−dependent PDOS for the NO and NO2 adsorbed systems with various energy ranges, respectively. For the NO adsorbed system, there were electronic states (spin−up DOS) just below the Fermi energy, which contributed to the orbital hybridization between the penta−NiPN monolayer and the NO molecule. Similarly, electron states were introduced into the NO2 adsorbed system, but unlike the NO condition, the electron state in NO2 was mainly below the conduction band. Clearly, the adsorption of NO and NO2 on the surface of the penta−NiPN monolayer induced strong coupling between the gas molecules and the host material. Due to the strong orbital coupling between them, a significant charge transfer occurred. Moreover, the strong adsorption interaction also provided a large magnitude of adsorption energy.
In general, the main gas−sensitive mechanisms of 2D materials are surface charge transfer and Schottky barrier (SB) modulation. Upon adsorption of gas molecules on the surface of the host material, the resistance of the material can be shifted. The interactions between gas molecules and 2D materials can either increase or decrease various resistive behaviors, depending on the major charge carriers of the semiconductor and the electron−donating/electron−withdrawing properties of the gas molecules. This resistance change can be further enhanced by creating SBs at the interface between the metal and the semiconductor. In our work, the absorption of NO and NO2 by the penta−NiPN monolayer introduced electronic states or impurity levels near the Fermi level or the bottom of the conduction band, directly affecting the electronic transmission properties of the system. This change in electron transport characteristics is manifested as resistance drift in the system. Therefore, gas sensing can be achieved by measuring resistance changes with and without gas adsorption. Figure 10 shows a schematic diagram of a gas sensor based on the penta−NiPN monolayer. When a bias voltage is introduced at the right/left electrode, specific recognition is achieved through the current−voltage curve, reflecting the difference in electronic transmission properties after the adsorption of different gas molecules [21]. However, as the penta−NiPN monolayer serves as the host material for gas sensing, it must be used in an oxygen−free environment. Otherwise, there may be irreversible reactions between the penta−NiPN monolayer and atmospheric oxygen.

4. Conclusions

In conclusion, we present a novel ternary penta−NiPN monolayer via first−principles calculations. Our results show that the penta−NiPN monolayer possesses high robust stability, including kinetic, thermodynamic, and mechanical stabilities. We confirmed its quasi−direct bandgap feature with a bandgap value of 1.237 eV based on the hybrid functional HSE06. Its moderate Young’s modulus and Poisson’s ratio indicate that the penta−NiPN monolayer is a promising candidate for flexible electronics. Furthermore, we predicted a hole mobility as high as 1.14 × 105 cm2V−1s−1 in the penta−NiPN monolayer based on deformation potential theory. Moreover, we systematically studied the adsorption properties of 12 common gas molecules (CO, CO2, CH4, H2, H2O, H2S, N2, NO, NO2, NH3, O2, and SO2) on the surface of the penta−NiPN monolayer. Our results show that the penta−NiPN monolayer exhibits good selectivity for NO and NO2, and has the potential to be used as a sensor for these two gases. Overall, our findings suggest that the penta−NiPN monolayer is a desirable candidate for high−performance electronic devices, as well as NO and NO2 gas sensors.

Author Contributions

Conceptualization, H.W.; software, H.W., J.-H.Y. and J.W.; formal analysis, H.W., G.L., J.-H.Y., P.Z. and Y.S.; resources, H.W., G.L., J.-H.Y., J.W., P.Z. and Y.S.; data curation, H.W.; writing—original draft, H.W.; writing—review and editing, G.L., J.-H.Y., J.W., P.Z. and Y.S.; supervision, G.L., P.Z. and Y.S.; funding acquisition, Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Foundation of Hubei Province (Grant No. 2022CFB935).

Data Availability Statement

All data needed to evaluate the conclusions in the paper are present in the paper.

Acknowledgments

This work was supported by the National Science Foundation of Hubei Province (Grant No. 2022CFB935).

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Feng, S.; Farha, F.; Li, Q.; Wan, Y.; Xu, Y.; Zhang, T.; Ning, H. Review on Smart Gas Sensing Technology. Sensors 2019, 19, 3760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Mamom, J.; Ratanadecho, P.; Mingmalairak, C.; Rungroungdouyboon, B. Humidity−Sensing Mattress for Long−Term Bedridden Patients with Incontinence−Associated Dermatitis. Micromachines 2023, 14, 1178. [Google Scholar] [CrossRef] [PubMed]
  3. Marti, E.; De Miguel, M.A.; Garcia, F.; Perez, J. A Review of Sensor Technologies for Perception in Automated Driving. IEEE Intell. Transp. Syst. Mag. 2019, 11, 94–108. [Google Scholar] [CrossRef] [Green Version]
  4. Tyagi, D.; Wang, H.; Huang, W.; Hu, L.; Tang, Y.; Guo, Z.; Ouyang, Z.; Zhang, H. Recent Advances in Two−Dimensional−Material−Based Sensing Technology toward Health and Environmental Monitoring Applications. Nanoscale 2020, 12, 3535–3559. [Google Scholar] [CrossRef]
  5. Zhang, L.; Khan, K.; Zou, J.; Zhang, H.; Li, Y. Recent Advances in Emerging 2D Material-Based Gas Sensors: Potential in Disease Diagnosis. Adv. Mater. Interfaces 2019, 6, 1901329. [Google Scholar] [CrossRef]
  6. Buckley, D.J.; Black, N.C.G.; Castanon, E.G.; Melios, C.; Hardman, M.; Kazakova, O. Frontiers of Graphene and 2D Material−Based Gas Sensors for Environmental Monitoring. 2D Mater. 2020, 7, 032002. [Google Scholar] [CrossRef]
  7. Yang, S.; Jiang, C.; Wei, S. Gas Sensing in 2D Materials. Appl. Phys. Rev. 2017, 4, 021304. [Google Scholar] [CrossRef]
  8. Ou, J.Z.; Ge, W.; Carey, B.; Daeneke, T.; Rotbart, A.; Shan, W.; Wang, Y.; Fu, Z.; Chrimes, A.F.; Wlodarski, W.; et al. Physisorption−Based Charge Transfer in Two−Dimensional SnS2 for Selective and Reversible NO2 Gas Sensing. ACS Nano 2015, 9, 10313–10323. [Google Scholar] [CrossRef]
  9. Qin, Z.; Song, X.; Wang, J.; Li, X.; Wu, C.; Wang, X.; Yin, X.; Zeng, D. Development of Flexible Paper Substrate Sensor Based on 2D WS2 with S Defects for Room−Temperature NH3 Gas Sensing. Appl. Surf. Sci. 2022, 573, 151535. [Google Scholar] [CrossRef]
  10. Liu, X.; Ma, T.; Pinna, N.; Zhang, J. Two−Dimensional Nanostructured Materials for Gas Sensing. Adv. Funct. Mater. 2017, 27, 1702168. [Google Scholar] [CrossRef]
  11. Thomas, S.; Asle Zaeem, M. Superior Sensing Performance of Two−Dimensional Ruthenium Carbide (2D−RuC) in Detection of NO, NO2 and NH3 Gas Molecules. Appl. Surf. Sci. 2021, 563, 150232. [Google Scholar] [CrossRef]
  12. Hakimi Raad, N.; Manavizadeh, N.; Frank, I.; Nadimi, E. Gas Sensing Properties of a Two−Dimensional Graphene/h−BN Multi−Heterostructure toward H2O, NH3 and NO2: A First Principles Study. Appl. Surf. Sci. 2021, 565, 150454. [Google Scholar] [CrossRef]
  13. Aasi, A.; Mortazavi, B.; Panchapakesan, B. Two−Dimensional PdPS and PdPSe Nanosheets: Novel Promising Sensing Platforms for Harmful Gas Molecules. Appl. Surf. Sci. 2022, 579, 152115. [Google Scholar] [CrossRef]
  14. Bykov, M.; Bykova, E.; Ponomareva, A.V.; Tasnádi, F.; Chariton, S.; Prakapenka, V.B.; Glazyrin, K.; Smith, J.S.; Mahmood, M.F.; Abrikosov, I.A.; et al. Realization of an Ideal Cairo Tessellation in Nickel Diazenide NiN2: High−Pressure Route to Pentagonal 2D Materials. ACS Nano 2021, 15, 13539–13546. [Google Scholar] [CrossRef]
  15. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [Green Version]
  16. Li, L.; Yu, Y.; Ye, G.J.; Ge, Q.; Ou, X.; Wu, H.; Feng, D.; Chen, X.H.; Zhang, Y. Black Phosphorus Field−Effect Transistors. Nat. Nanotechnol. 2014, 9, 372–377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Yuan, J.-H.; Song, Y.-Q.; Chen, Q.; Xue, K.-H.; Miao, X.-S. Single−Layer Planar Penta−X2N4 (X = Ni, Pd and Pt) as Direct−Bandgap Semiconductors from First Principle Calculations. Appl. Surf. Sci. 2019, 469, 456–462. [Google Scholar] [CrossRef]
  18. Mortazavi, B.; Zhuang, X.; Rabczuk, T.; Shapeev, A.V. Outstanding Thermal Conductivity and Mechanical Properties in the Direct Gap Semiconducting Penta−NiN2 Monolayer Confirmed by First−Principles. Phys. E Low−Dimens. Syst. Nanostructures 2022, 140, 115221. [Google Scholar] [CrossRef]
  19. Hu, Y.; Zhao, X.; Yang, Y.; Xiao, W.; Zhou, X.; Wang, D.; Wang, G.; Bi, J.; Luo, Z.; Liu, X. Coordination Engineering on Novel 2D Pentagonal NiN2 for Bifunctional Oxygen Electrocatalysts. Appl. Surf. Sci. 2023, 614, 156256. [Google Scholar] [CrossRef]
  20. Zhang, C.; Sun, J.; Shen, Y.; Kang, W.; Wang, Q. Effect of High Order Phonon Scattering on the Thermal Conductivity and Its Response to Strain of a Penta−NiN2 Sheet. J. Phys. Chem. Lett. 2022, 13, 5734–5741. [Google Scholar] [CrossRef] [PubMed]
  21. Li, M.; Wang, X.-F. Adsorption Behaviors of Small Molecules on Two−Dimensional Penta−NiN2 Layers: Implications for NO and NO2 Gas Sensors. ACS Appl. Nano Mater. 2023, 6, 6151–6160. [Google Scholar] [CrossRef]
  22. Yuan, J.-H.; Zhang, B.; Song, Y.-Q.; Wang, J.-F.; Xue, K.-H.; Miao, X.-S. Planar Penta−Transition Metal Phosphide and Arsenide as Narrow−Gap Semiconductors with Ultrahigh Carrier Mobility. J. Mater. Sci. 2019, 54, 7035–7047. [Google Scholar] [CrossRef] [Green Version]
  23. Qian, S.; Sheng, X.; Xu, X.; Wu, Y.; Lu, N.; Qin, Z.; Wang, J.; Zhang, C.; Feng, E.; Huang, W.; et al. Penta−MX2 (M = Ni, Pd and Pt; X = P and As) Monolayers: Direct Band−Gap Semiconductors with High Carrier Mobility. J. Mater. Chem. C 2019, 7, 3569–3575. [Google Scholar] [CrossRef]
  24. Liu, Z.; Wang, H.; Sun, J.; Sun, R.; Wang, Z.F.; Yang, J. Penta−Pt2N4: An Ideal Two−Dimensional Material for Nanoelectronics. Nanoscale 2018, 10, 16169–16177. [Google Scholar] [CrossRef] [PubMed]
  25. Shao, X.; Liu, X.; Zhao, X.; Wang, J.; Zhang, X.; Zhao, M. Electronic Properties of a π−Conjugated Cairo Pentagonal Lattice: Direct Band Gap, Ultrahigh Carrier Mobility, and Slanted Dirac Cones. Phys. Rev. B 2018, 98, 085437. [Google Scholar] [CrossRef] [Green Version]
  26. Shao, X.; Sun, L.; Ma, X.; Feng, X.; Gao, H.; Ding, C.; Zhao, M. Multiple Dirac Cones and Lifshitz Transition in a Two−Dimensional Cairo Lattice as a Hawking Evaporation Analogue. J. Phys. Condens. Matter 2021, 33, 365001. [Google Scholar] [CrossRef]
  27. Raval, D.; Gupta, S.K.; Gajjar, P.N. Detection of H2S, HF and H2 Pollutant Gases on the Surface of Penta−PdAs2 Monolayer Using DFT Approach. Sci. Rep. 2023, 13, 699. [Google Scholar] [CrossRef] [PubMed]
  28. He, Y.; Xiong, S.; Xia, F.; Shao, Z.; Zhao, J.; Zhang, X.; Jie, J.; Zhang, X. Tuning the Electronic Transport Anisotropy in α−Phase Phosphorene through Superlattice Design. Phys. Rev. B 2018, 97, 085119. [Google Scholar] [CrossRef]
  29. Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total−Energy Calculations Using a Plane−Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  30. 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]
  31. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar] [CrossRef] [Green Version]
  32. Kohn, W.; Sham, L.J. Self−Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef] [Green Version]
  33. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Krukau, A.V.; Vydrov, O.A.; Izmaylov, A.F.; Scuseria, G.E. Influence of the Exchange Screening Parameter on the Performance of Screened Hybrid Functionals. J. Chem. Phys. 2006, 125, 224106. [Google Scholar] [CrossRef] [PubMed]
  35. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin−Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  36. Grimme, S. Semiempirical GGA−Type Density Functional Constructed with a Long−Range Dispersion Correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  37. Togo, A.; Oba, F.; Tanaka, I. First−Principles Calculations of the Ferroelastic Transition between Rutile−Type and CaCl2−Type SiO2 at High Pressures. Phys. Rev. B 2008, 78, 134106. [Google Scholar] [CrossRef] [Green Version]
  38. Zhang, Y.; Wu, Z.-F.; Gao, P.-F.; Fang, D.-Q.; Zhang, E.-H.; Zhang, S.-L. Structural, Elastic, Electronic, and Optical Properties of the Tricycle−like Phosphorene. Phys. Chem. Chem. Phys. 2017, 19, 2245–2251. [Google Scholar] [CrossRef]
  39. Born, M.; Huang, K. Dynamical Theory of Crystal Lattices; Clarendon Press: Oxford, UK, 1954. [Google Scholar]
  40. Yuan, J.-H.; Xue, K.-H.; Wang, J.-F.; Miao, X.-S. Gallium Thiophosphate: An Emerging Bidirectional Auxetic Two−Dimensional Crystal with Wide Direct Band Gap. J. Phys. Chem. Lett. 2019, 10, 4455–4462. [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.; Jepsen, O.; Flad, J.; Andersen, O.K.; Preuss, H.; von Schnering, H.G. Electron Localization in Solid−State Structures of the Elements: The Diamond Structure. Angew. Chem. Int. Ed. Engl. 1992, 31, 187–188. [Google Scholar] [CrossRef]
  43. Tang, W.; Sanville, E.; Henkelman, G. A Grid−Based Bader Analysis Algorithm without Lattice Bias. J. Phys. Condens. Matter 2009, 21, 084204. [Google Scholar] [CrossRef]
  44. Bardeen, J.; Shockley, W. Deformation Potentials and Mobilities in Non−Polar Crystals. Phys. Rev. 1950, 80, 72–80. [Google Scholar] [CrossRef]
  45. Yuan, J.-H.; Xue, K.-H.; Miao, X. Two−Dimensional ABC3 (A = Sc, Y; B = Al, Ga, In; C = S, Se, Te) with Intrinsic Electric Field for Photocatalytic Water Splitting. Int. J. Hydrog. Energy 2023, 48, 5929–5939. [Google Scholar] [CrossRef]
  46. Yuan, J.-H.; Xue, K.-H.; Wang, J.; Miao, X. Designing Stable 2D Materials Solely from VIA Elements. Appl. Phys. Lett. 2021, 119, 223101. [Google Scholar] [CrossRef]
  47. Cai, Y.; Zhang, G.; Zhang, Y.-W. Polarity−Reversed Robust Carrier Mobility in Monolayer MoS2 Nanoribbons. J. Am. Chem. Soc. 2014, 136, 6269–6275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Jing, Y.; Ma, Y.; Li, Y.; Heine, T. GeP3: A Small Indirect Band Gap 2D Crystal with High Carrier Mobility and Strong Interlayer Quantum Confinement. Nano Lett. 2017, 17, 1833–1838. [Google Scholar] [CrossRef] [Green Version]
  49. Jiang, J.-W.; Park, H.S. Negative Poisson’s Ratio in Single−Layer Black Phosphorus. Nat. Commun. 2014, 5, 4727. [Google Scholar] [CrossRef] [Green Version]
  50. Cheng, M.-Q.; Chen, Q.; Yang, K.; Huang, W.-Q.; Hu, W.-Y.; Huang, G.-F. Penta−Graphene as a Potential Gas Sensor for NOx Detection. Nanoscale Res. Lett. 2019, 14, 306. [Google Scholar] [CrossRef] [Green Version]
  51. Wei, M.; Dou, X.; Zhao, L.; Du, J.; Jiang, G. Monolayer Penta−BCN: A Promising Candidate for Harmful Gases Detection. Sens. Actuators Phys. 2022, 334, 113326. [Google Scholar] [CrossRef]
  52. Jiang, X.; Zhang, G.; Yi, W.; Yang, T.; Liu, X. Penta−BeP2 Monolayer: A Superior Sensor for Detecting Toxic Gases in the Air with Excellent Sensitivity, Selectivity, and Reversibility. ACS Appl. Mater. Interfaces 2022, 14, 35229–35236. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Top and side views of the penta−NiPN monolayer; the unit cell is marked by a red border. (b) Geometry of single ideal (pentagonal Cairo tiling) Ni2N3, Ni2P2N, Ni2PN2, and Ni2P3 pentagons. Numbers indicate interatomic distances (Å) and bond angles (deg). (c) The Brillouin zone of the penta−NiPN monolayer. (d) The phonon dispersion and corresponding phonon DOS of the penta−NiPN monolayer. (e) The AIMD simulation results of the penta−NiPN monolayer at 300 K; the insert is the initial and final structure of the penta−NiPN monolayer. (f) The eight adsorption sites have been labeled in the diagram.
Figure 1. (a) Top and side views of the penta−NiPN monolayer; the unit cell is marked by a red border. (b) Geometry of single ideal (pentagonal Cairo tiling) Ni2N3, Ni2P2N, Ni2PN2, and Ni2P3 pentagons. Numbers indicate interatomic distances (Å) and bond angles (deg). (c) The Brillouin zone of the penta−NiPN monolayer. (d) The phonon dispersion and corresponding phonon DOS of the penta−NiPN monolayer. (e) The AIMD simulation results of the penta−NiPN monolayer at 300 K; the insert is the initial and final structure of the penta−NiPN monolayer. (f) The eight adsorption sites have been labeled in the diagram.
Micromachines 14 01407 g001
Figure 2. Angle−dependent (a) Young’s modulus and (b) Poisson’s ratio of the penta−NiPN monolayer.
Figure 2. Angle−dependent (a) Young’s modulus and (b) Poisson’s ratio of the penta−NiPN monolayer.
Micromachines 14 01407 g002
Figure 3. Calculated (a) electronic band structures and (b) DOS, as well as PDOS, of the penta−NiPN monolayer at the HSE06 level. (c) Spatial distributions of the wave functions corresponding to the VBM and CBM of the penta−NiPN monolayer at the GGA–PBE level. The charge contour density is 0.01 e3. (d) Calculated ELF of the penta−NiPN monolayer.
Figure 3. Calculated (a) electronic band structures and (b) DOS, as well as PDOS, of the penta−NiPN monolayer at the HSE06 level. (c) Spatial distributions of the wave functions corresponding to the VBM and CBM of the penta−NiPN monolayer at the GGA–PBE level. The charge contour density is 0.01 e3. (d) Calculated ELF of the penta−NiPN monolayer.
Micromachines 14 01407 g003
Figure 4. (a) The relationship between the total energy and the applied strain (Δl/l0) along the a/b direction of the penta−NiPN monolayer. (b) The shift of VBMs and CBMs for the penta−NiPN monolayer with respect to the vacuum energy under the applied strain along the a/b direction. All the calculations are at the GGA–PBE level.
Figure 4. (a) The relationship between the total energy and the applied strain (Δl/l0) along the a/b direction of the penta−NiPN monolayer. (b) The shift of VBMs and CBMs for the penta−NiPN monolayer with respect to the vacuum energy under the applied strain along the a/b direction. All the calculations are at the GGA–PBE level.
Micromachines 14 01407 g004
Figure 5. Top and side views of the most favorable adsorption configurations for (a) CO, (b) CO2, (c) CH4, (d) H2, (e) H2O, (f) H2S, (g) N2, (h) NO, (i) NO2, (j) NH3, (k) SO2, and (l) O2 on the penta−NiPN monolayer.
Figure 5. Top and side views of the most favorable adsorption configurations for (a) CO, (b) CO2, (c) CH4, (d) H2, (e) H2O, (f) H2S, (g) N2, (h) NO, (i) NO2, (j) NH3, (k) SO2, and (l) O2 on the penta−NiPN monolayer.
Micromachines 14 01407 g005
Figure 6. The calculated adsorption energy and adsorption distance for CO, CO2, CH4, H2, H2O, H2S, N2, NO, NO2, NH3, and SO2 on the penta−NiPN monolayer.
Figure 6. The calculated adsorption energy and adsorption distance for CO, CO2, CH4, H2, H2O, H2S, N2, NO, NO2, NH3, and SO2 on the penta−NiPN monolayer.
Micromachines 14 01407 g006
Figure 7. The charge density differences between (a) CO, (b) CO2, (c) CH4, (d) H2, (e) H2O, (f) H2S, (g) N2, (h) NO, (i) NO2, (j) NH3, and (k) SO2 gas molecules and the penta−NiPN monolayer. The equivalent surface was 0.012 e3, and the electron accumulation (loss) is represented by yellow (blue). In addition, the direction of charge transfer (represented by arrows) and the amount of charge transfer are marked.
Figure 7. The charge density differences between (a) CO, (b) CO2, (c) CH4, (d) H2, (e) H2O, (f) H2S, (g) N2, (h) NO, (i) NO2, (j) NH3, and (k) SO2 gas molecules and the penta−NiPN monolayer. The equivalent surface was 0.012 e3, and the electron accumulation (loss) is represented by yellow (blue). In addition, the direction of charge transfer (represented by arrows) and the amount of charge transfer are marked.
Micromachines 14 01407 g007
Figure 8. The PDOS (based on HSE06 functional) from the gas molecules are plotted together for the penta−NiPN monolayer with the adsorbed molecules (a) CO, (b) CO2, (c) CH4, (d) H2, (e) H2O, (f) H2S, (g) N2, (h) NH3, and (i) SO2. Only the spin−up DOS is shown in spin–degenerate systems. The X represents nitrogen atoms.
Figure 8. The PDOS (based on HSE06 functional) from the gas molecules are plotted together for the penta−NiPN monolayer with the adsorbed molecules (a) CO, (b) CO2, (c) CH4, (d) H2, (e) H2O, (f) H2S, (g) N2, (h) NH3, and (i) SO2. Only the spin−up DOS is shown in spin–degenerate systems. The X represents nitrogen atoms.
Micromachines 14 01407 g008
Figure 9. The PDOS (based on HSE06 functional) from the gas molecules are plotted for the penta−NiPN monolayer with the adsorbed molecules (a) NO and (b) NO2. The corresponding enlarged PDOS results of (c) NO and (d) NO2 are given as well. The spin−up and spin−down DOS are shown as positive and negative, respectively. The X represents nitrogen atoms.
Figure 9. The PDOS (based on HSE06 functional) from the gas molecules are plotted for the penta−NiPN monolayer with the adsorbed molecules (a) NO and (b) NO2. The corresponding enlarged PDOS results of (c) NO and (d) NO2 are given as well. The spin−up and spin−down DOS are shown as positive and negative, respectively. The X represents nitrogen atoms.
Micromachines 14 01407 g009
Figure 10. Schematic diagram of gas sensor based on the penta−NiPN monolayer.
Figure 10. Schematic diagram of gas sensor based on the penta−NiPN monolayer.
Micromachines 14 01407 g010
Table 1. Calculated lattice constant a/b, bond length l, cohesive energy Ecoh, and bandgap Eg (based on HSE06 functional) of the penta−NiPN monolayer.
Table 1. Calculated lattice constant a/b, bond length l, cohesive energy Ecoh, and bandgap Eg (based on HSE06 functional) of the penta−NiPN monolayer.
Materialsa (Å)b (Å)lNi−N (Å)lNi−P (Å)lN−N/lN−P/lP−P (Å)Ecoh (eV)Eg (eV)
NiPN
NiN2 [17]
NiP2 [22]
4.995
4.53
5.55
5.011
4.53
5.55
1.929, 1.910
1.88
−−
2.125, 2.107
−−
2.16
1.605
1.24
2.11
4.55
4.98
4.09
1.237
1.10
0.81
Table 2. Calculated m* (unit: me), |Eil| (unit: eV), C2D (unit: N m−1), and μ2D (unit: 104 cm2V−1 s−1) for the penta−NiPN monolayer along the a and b directions.
Table 2. Calculated m* (unit: me), |Eil| (unit: eV), C2D (unit: N m−1), and μ2D (unit: 104 cm2V−1 s−1) for the penta−NiPN monolayer along the a and b directions.
MaterialsCarrier Typema*mb*|Ela||Elb|Ca2DCb2Dμa2Dμb2D
NiPNElectron0.380.362.100.85147.68146.240.513.24
Hole0.220.240.740.99147.68146.2411.365.76
NiP2 [22]Electron0.1060.1405.235.23118.19118.190.710.54
Hole0.1190.1701.531.53118.19118.196.354.45
Table 3. Calculated adsorption energy (Ea), adsorption distance (d), magnetic moment (M), and charge transfer (Q) between the gas molecules and the penta−NiPN monolayer. Here, “+” and “−” represent gained and lost electrons, respectively.
Table 3. Calculated adsorption energy (Ea), adsorption distance (d), magnetic moment (M), and charge transfer (Q) between the gas molecules and the penta−NiPN monolayer. Here, “+” and “−” represent gained and lost electrons, respectively.
Gas MoleculesEa (eV)d (Å) M ( μ B Q (e)
CO
CO2
CH4
H2
H2O
H2S
N2
NO
NO2
NH3
SO2
−0.640
−0.184
−0.162
−0.072
−0.272
−0.316
−0.100
−0.751
−1.011
−0.545
−0.445
1.834
3.054
2.661
2.613
2.248
2.210
3.117
1.862
2.065
2.119
2.573
0
0
0
0
0
0
0
0.695
0.878
0
0
+0.100
+0.026
+0.010
+0.011
+0.025
−0.100
+0.015
+0.216
+0.553
−0.103
+0.187
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

Wang, H.; Li, G.; Yuan, J.-H.; Wang, J.; Zhang, P.; Shan, Y. Two−Dimensional Planar Penta−NiPN with Ultrahigh Carrier Mobility and Its Potential Application in NO and NO2 Gas Sensing. Micromachines 2023, 14, 1407. https://doi.org/10.3390/mi14071407

AMA Style

Wang H, Li G, Yuan J-H, Wang J, Zhang P, Shan Y. Two−Dimensional Planar Penta−NiPN with Ultrahigh Carrier Mobility and Its Potential Application in NO and NO2 Gas Sensing. Micromachines. 2023; 14(7):1407. https://doi.org/10.3390/mi14071407

Chicago/Turabian Style

Wang, Hao, Gang Li, Jun-Hui Yuan, Jiafu Wang, Pan Zhang, and Yahui Shan. 2023. "Two−Dimensional Planar Penta−NiPN with Ultrahigh Carrier Mobility and Its Potential Application in NO and NO2 Gas Sensing" Micromachines 14, no. 7: 1407. https://doi.org/10.3390/mi14071407

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