Next Article in Journal
Plant-Derived Exosomes: Carriers and Cargo of Natural Bioactive Compounds: Emerging Functions and Applications in Human Health
Previous Article in Journal
Research Progress and Future Perspectives on Photonic and Optoelectronic Devices Based on p-Type Boron-Doped Diamond/n-Type Titanium Dioxide Heterojunctions: A Mini Review
Previous Article in Special Issue
First-Principles Guided Design of Pd-Decorated V2O5/Porous Silicon Composites for High-Performance NO2 Sensing at Room Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two-Dimensional Porous Beryllium Trinitride Monolayer as Multifunctional Energetic Material

1
Department of Physics, Anhui Normal University, Wuhu 241000, China
2
Xuancheng Ecological Environment Law Enforcement Monitoring Station, Xuancheng 242000, China
3
Hefei National Laboratory for Physical Sciences at Microscale, Department of Chemical Physics, University of Science and Technology of China, Hefei 230026, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(13), 1004; https://doi.org/10.3390/nano15131004
Submission received: 12 June 2025 / Revised: 24 June 2025 / Accepted: 26 June 2025 / Published: 29 June 2025
(This article belongs to the Special Issue Theoretical Calculation Study of Nanomaterials: 2nd Edition)

Abstract

Polynitrogen compounds have broad applications in the field of high-energy materials, making the exploration of two-dimensional polynitride materials with both novel properties and practical utility a highly attractive research challenge. Through global structure search methods and first-principles theoretical calculations at the Perdew–Burke–Ernzerhof (PBE) level of density functional theory (DFT), the globally minimum-energy configuration of a novel planar BeN3 monolayer (tetr-2D-BeN3) is predicted. This material exhibits a planar quasi-isotropic structure containing pentagonal, hexagonal, and dodecagonal rings, as well as “S”-shaped N6 polymeric units, exhibiting a high energy density of 3.34 kJ·g−1, excellent lattice dynamic stability and thermal stability, an indirect bandgap of 2.66 eV (HSE06), high carrier mobility, and ultraviolet light absorption capacity. In terms of mechanical properties, it shows a low in-plane Young’s stiffness of 52.3–52.9 N·m−1 and a high in-plane Poisson’s ratio of 0.55–0.56, indicating superior flexibility. Furthermore, its porous structure endows it with remarkable selectivity for hydrogen (H2) and argon (Ar) gas separation, achieving a maximum selectivity of up to 1023 (He/Ar). Therefore, the tetr-2D-BeN3 monolayer represents a multifunctional two-dimensional polynitrogen-based energetic material with potential applications in energetic materials, flexible semiconductor devices, ductile materials, ultraviolet photodetectors, and other fields, thereby expanding the design possibilities for polynitride materials.

1. Introduction

High-energy-density materials are crucial in fields such as propellants, explosives, and rocket fuels due to their ability to store exceptionally high energy [1]. Polynitrogen compounds are considered ideal candidates for such materials, primarily because the N≡N triple bond energy in nitrogen gas (N2) (954 kJ·mol−1) is significantly higher than that of N-N single bonds (~160 kJ·mol−1) and N=N double bonds (~418 kJ·mol−1) [2]. This allows for the release of enormous energy when N-N/N = N bonds break to form N≡N bonds. Various polynitrogen structures have been investigated, including the cubic gauche phase [3], layered polymeric phases [4], black phosphorus-like structures [5], adamantane-type N10 [6], P4/nbm phase [7], and ring/chain/cage configurations [8,9,10]. Alkaline (earth) metal polynitrogen compounds (e.g., MN3, MN5, MN10, MN4; M = Li, Na, K, Rb, Cs, Be, Mg, Ca, Sr, Ba, etc.) [11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26] often contain abundant polymeric nitrogen networks, such as N3/N5/N6 rings or N chains [11,14,16,26]. Among them, beryllium (Be)-based polynitrogen compounds exhibit unique advantages in energy density due to their smallest atomic mass and radius when forming compounds with polymeric nitrogen. For instance, the most stable γ-BeN4 features a chair-like nitrogen chain structure and exhibits an energy density of 3.1 kJ·g−1 [27]. High-pressure-synthesized β-BeN4 (obtained from the reaction of Be3N2 and N2 at 25.4 GPa) demonstrates an even higher energy density of 3.6 kJ·g−1 while retaining the same chair-like nitrogen chain configuration [28]. Notably, δ-BeN4 [27] and P21/c-BeN4 [29] exhibit exceptionally high energy densities of 4.7 kJ·g−1 and 6.3 kJ·g−1, respectively. These values significantly surpass those of alkaline earth metal nitrides such as P1-MgN4 (2.084 kJ·g−1) [30] and P1-CaN10 (2.438 kJ·g−1) [31] and even exceed that of the conventional explosive TNT (4.3 kJ·g−1). Furthermore, theoretical predictions suggest that α-BeN6 and β-BeN6 [32] possess high energy densities of 3.32 kJ·g−1 and 3.59 kJ·g−1, respectively. Therefore, the development of novel high-energy-density materials based on the beryllium–nitrogen (Be-N) system is a highly promising and important direction.
Two-dimensional (2D) materials have been extensively studied, exhibiting various and novel physical and chemical properties with wide applications [33,34,35,36]. In contrast, 2D polynitrogen materials is still in its early stages, that only some 2D polynitrogen materials have been reported, such as, 2D KN3 [37], h-MN2 (M = Be, Mg) [38], h-MN3 (M = Be, Ge) [39], and 2D MN4 (M = Be, Mg, Ir, Rh, Ni, Cu, Au, Pd, Pt) [40]. Similar to bulk polynitrogen compounds, these 2D materials also contain diverse polymeric nitrogen structures, such as N3 trimers (2D KN3) [37], N4 tetramers (h-MgN2) [38], N6 hexagons (h-BeN3) [39], and infinite N chains (2D MgN4) [20]. For the 2D Be-N system, many 2D N-rich beryllium polynitrogen compounds have been intensively explored, e.g., BeN4, BeN3, BeN2, Be2N3, Be3N4, BeN, and Be3N2, revealing rich properties and broad application prospects [22,38,41,42]. For example, h-BeN2 monolayer [38], featuring isolated “Y”-shaped N4 tetramers, is predicted to possess a direct bandgap, excellent carrier mobility, ultrahigh on-current in transistors, and water photocatalysis capability. Furthermore, α-2D-BeN2 consists of penta-, hexa-, and hepta-atomic rings, along with an N4 tetramer, currently the most stable 2D BeN2 structure, theoretically exhibits a direct bandgap of 1.82 eV, and demonstrates outstanding performance in oxygen reduction/evolution reaction catalysis and potassium-ion storage [43]. A benzene-like N6 hexagonal honeycomb monolayer material, h-BeN3, shows high stability, excellent carrier mobility, and n-type semiconductor properties, making it promising for nanoelectronic device applications [39]. In addition, recent studies have achieved multiscale correlation from atomic mechanisms to mesoscopic morphology in group III nitride nanosystems by coupling DFT with phase-field modeling, establishing a theoretical foundation for predicting broader semiconductor nanostructures and advancing the development of novel nitrogen-containing compounds [44]. Experimentally, the successful synthesis of a 2D BeN4 monolayer with parallel armchair-like infinite [N]n chains possesses anisotropic characteristics and potential applications in ion battery storage, CO2 capture, and hydrogen storage [22]. Nevertheless, searching for more practical 2D polynitrogen materials with novel, advanced properties is in the ascendant.
In this work, using the artificial bee colony (ABC) algorithm for 2D global structure search, a novel tetragonal porous BeN3 monolayer (tetr-2D-BeN3), consisting of Be atoms and “S”-shaped N6 tetramer segments, is systematically investigated. Compared with h-BeN3 [39], it has lower energy while maintaining excellent lattice dynamics and thermal stability, along with a high energy density of 3.34 kJ·g−1. This new structure exhibits an indirect bandgap (2.66 eV) with high carrier mobilities, strong ultraviolet light absorption, and a large-pore structure that enables superior gas separation performance. It is expected to serve as a highly stable multifunctional material with broad application prospects in various fields.

2. Materials and Methods

2.1. Materials

The artificial bee colony (ABC) algorithm within the CALYPSO code [45] is employed to conduct a global structural search for two-dimensional (2D) compounds with a Be:N = 1:3 (BeN3) stoichiometric ratio. This study focuses on the lowest-energy structure from these 1800 candidates, a 2D planar BeN3 structure with P4m 2D space group (wallpaper group).

2.2. Calculation Details

First-principles simulations based on density functional theory (DFT) [46,47] are performed using the VASP (vesion 6) software package [48]. The electron–ion interactions are described by the projector augmented wave (PAW) method [49], with a plane-wave basis set energy cutoff of 520 eV. The exchange–correlation effects are treated using the Perdew–Burke–Ernzerhof (PBE) correlation–exchange (XC) functional. For geometric optimization and electronic property calculations, the Γ-centered k-point mesh in the Brillouin zone is set to 6 × 6 × 1. The energy convergence threshold is set to 10−5 eV, and the atomic force convergence threshold is 0.01 eV/Å. All intrinsic pristine structures are optimized using the ISIF = 3 tag in VASP, where stress convergence is automatically controlled by the force convergence criterion. The more accurate electronic structure is obtained using the Heyd–Scuseria–Ernzerhof (HSE06) hybrid functional [50]. The lattice dynamical stability of the material is verified through phonon spectra calculations using the finite displacement method [51]. Ab initio molecular dynamics (AIMD) simulations are conducted in the NVT ensemble with a time step of 1 fs for a total duration of 5 ps to evaluate the structural stability at elevated temperatures. The structural search within the CALYPSO code [45] settings included parallel calculations for unit cells with varying numbers of atoms, specifically examining 2:6, 3:9, and 4:12 configurations. All calculations use a population size of 20 and 30 generations of evolution, generating a total of 1800 candidate structures.
The formation energy of tetr-2D-BeN3 is calculated using the following equation:
E f = E B e N 3 E B e 3 E N / 4
Here, E B e and E N represents the energy of per metal atoms in the Be bulk metal phase and the energy of per atoms in the N2 gas.
The vacancy formation energies (EF-vac) are calculated using the following equation:
E F-vac = E V B e / V N @ B e N 3 + E B e- a t o m / N- a t o m E B e N 3
Here, E B e- a t o m and E N- a t o m represents the energy of a single Be and N atom.
The in-plane Young’s stiffness Y ( θ ) and Poisson’s ratio v ( θ ) are calculated as functions of θ based on the equations as listed below [52]:
Y ( θ ) = C 11 C 22 C 12 2 C 11 sin 4 θ + C 22 cos 4 θ + ( C 11 C 22 C 12 2 C 66 2 C 12 ) cos 2 θ sin 2 θ
v ( θ ) = ( C 11 + C 12 C 11 C 22 C 12 2 C 66 ) cos 2 θ sin 2 θ C 12 sin 4 θ C 12 cos 4 θ C 11 sin 4 θ + C 22 cos 4 θ + ( C 11 C 22 C 12 2 C 66 2 C 12 ) cos 2 θ sin 2 θ
According to the DP theory [53], the carrier mobility of a 2D structure is calculated as
μ 2 D = e 3 C 2 D k B T m * m d E 1 2
where e , , and k B are the electron charge, reduced Planck constant, and Boltzmann constant, respectively. C 2 D are the elastic moduli, T is the temperature (300 K), m * is the effective mass along the transport direction, which can be obtained by fitting the band structure at the CBM and VBM, and m d = m x * m y * is the average effective mass. E l is the deformation potential constant, which is the key to the magnitude of mobility, defined by E l = E e d g e / ε , where E e d g e is the energy of the band edge and ε = l / l 0 . The strain range is from −0.6% to +0.6% with an interval of 0.2%.
According to the frequency-dependent permittivity of ε ω = ε 1 ω + i ε 2 ω , the absorption coefficient is obtained based on the following equation [54,55]:
α ω = 2 ω c ε 1 2 ω + ε 2 2 ( ω ) ε 1 ( ω ) 1 / 2
The selectivity S g a s 1 / g a s 2 is an important indicator of the efficiency of gas separation and can be assessed with the Arrhenius equation [56]:
S g a s 1 / g a s 2 = r g a s 1 r g a s 2 = A g a s 1 A g a s 2 e E g a s 1 / R T e E g a s 2 / R T
where r is the diffusion rate, A is the diffusion prefactor (assuming that the A of gas molecules are the same as 1011) [57], E is the diffusion energy barrier, R is the gas constant, and T is the temperature.

3. Results and Discussion

3.1. Structural Search and Energetic Properties of BeN3

First-principles calculations reveal that the global energy-minimum configuration of BeN3 is found in the largest unit cell system (containing 4 Be atoms and 12 N atoms) (Figure 1a,b), while previously reported monolayer structure (h-Be2N6) [39] with the same stoichiometry are identified in the corresponding smallest unit cells. This result suggests that increasing the unit cell size facilitates the discovery of global energy-minimum structures for two-dimensional (2D) Be-N compounds, providing a new strategy for global structural searches of similar 2D materials.
As shown in Figure 1a, the global energy-minimum 2D-Be4N12 (here named tetr-2D-BeN3) exhibit significant thermodynamic advantages over the reported h-Be2N6 [39], with a total energy reduction of 110 meV/atom. The tetr-2D-BeN3 monolayer possesses a tetragonal orthogonal lattice with lattice constants a = b = 7.74 Å and a space group (wallpaper group in 2D) symmetry of P4m. As shown in Figure 1b, the unit cell contains four equivalent Be atoms and three types of nitrogen atoms (N1, N2, N3), where N1 adopts a Be-N-N tri-coordination mode, N2 exhibits N-N di-coordination, and N3 shows Be-Be-N tri-coordination. The tetr-2D-BeN3 planar structure consists of five-membered, eight-membered, and twelve-membered rings, featuring an “S”-shaped N6 cluster. The Be-N bond lengths range from 1.61 to 1.65 Å, while the N-N bond lengths (1.32–1.41 Å) lie between the typical single N–N bond (1.45 Å) [58] and double N=N bond (1.25 Å) values [59], suggesting its potential as an energetic material. Under ambient pressure, the exothermic decomposition of BeN3 yields the most stable Be-N compound Be3N2 ( I a 3 ¯ ) [28] and N2, as shown in the following equation:
6 B e N 3 2 B e 3 N 2 + 7 N 2 + 10.6   e V
This process releases a chemical energy of 10.6 eV, corresponding to an energy density of approximately 3.34 kJ·g−1, which is slightly lower than that of BeN4 (3.60 kJ·g−1) [28], δ-BeN4 (4.70 kJ·g−1) [27], β-BeN6 (3.59 kJ·g−1) [32], and HEDM TNT (4.30 kJ·g−1) but higher than CNO (2.2 kJ·g−1) [60], LiN5 (2.72 kJ·g−1) [61], GdN6 (1.62 kJ·g−1) [62], YN10 (3.05 kJ·g−1) [63], γ-BeN4 (3.10 kJ·g−1) [27].

3.2. Bonding Characteristics and Stability of BeN3

To further elucidate the bonding characteristics in tetr-2D-BeN3, deformation charge density, electron localization function (ELF), and Bader charge analysis are calculated (Figure 1c). The deformation charge density reveals electron accumulation at Be-N bridging sites and partial electron distribution between adjacent nitrogen atoms within the N6 chain segment, indicating significant covalent hybridization between Be-N atoms and within the N6 chains. Additionally, significant electron density is observed on the N2 surface along the direction opposite to the N1-N2-N3 angle bisector. Analysis of the ELF (where high values (>0.5) correspond to lone pairs, core electrons, or covalent bonds, low values (<0.5) represent ionic bonds, and ELF~0.5 signifies metallic bonding) shows a highly localized electron cloud (ELF ≈ 1) on the N2 surface, indicating the presence of unhybridized p-orbitals along the direction normal to the plane. The high ELF values at Be-N and N-N sites further suggest the formation of strong covalent bonds. Bader charge analysis quantifies electron transfer, revealing that Be atoms transferred 1.66 e, while nitrogen atoms exhibited significant charge variation. Specifically, the tri-coordinated N1 (Be-N-N), di-coordinated N2 (N-N), and tri-coordinated N3 (Be-Be-N) acquired charges of +0.47 e, +0.05 e, and +1.14 e, respectively. Crucially, the notably low charge acquisition and the high ELF localization on the di-coordinated N2 atom, indicative of poorly hybridized/underbonded states due to its low coordination and minimal charge transfer, are highly pertinent to the observed high energy density of tetr-2D-BeN3. The electron-rich N2 center, featuring unpaired electrons or highly localized electrons within its pz orbital with a lone pair electron perpendicular to the plane, represents a significant metastable site storing considerable strain energy. Upon decomposition or reaction initiation, these energetically frustrated sites, particularly the under-bonded and electron-deficient N2, provide a substantial thermodynamic driving force, facilitating the rapid release of stored chemical energy and thereby underpinning the material’s high energy density.
Furthermore, the stability of the novel tetr-2D-BeN3 monolayer is evaluated from three perspectives: structural stability, thermodynamic feasibility, and dynamical stability. The formation energy (Ef, referenced to Be metal and N2 molecules) is calculated as +0.02 eV/atom, suggesting that the formation of tetr-2D-BeN3 from Be metal and N2 molecules requires slight external energy input. Phonon spectrum analysis (Figure 2a) confirms the dynamical stability of tetr-2D-BeN3, as no imaginary frequencies are observed across the entire Brillouin zone. Finally, AIMD simulations verify the thermal stability of this new structure. Figure 2b displays geometric snapshots after 5 ps of simulation at temperatures of 300, 500, 1000, 1200, and 1400 K, showing that the tetr-2D-BeN3 monolayer maintains structural integrity below 1400 K, exhibits partial bond breaking at 1500 K. At 2000 K, tetr-2D-BeN3 completely decomposes into amorphous Be-N clusters and N2 molecules. This theoretical prediction of decomposition temperatures (2000 K) is higher than the experimentally measured decomposition temperatures of graphene and MoS2 monolayers (1100–1300 K) [64,65], and serves as a reference for comparative thermal stability analysis. Furthermore, the environmental stability of tetr-2D-BeN3 through AIMD simulations at 300 K under both H2O and O2 atmospheres is investigated. As shown in Figure 3, the tetr-2D-BeN3 monolayer demonstrated excellent structural stability during 5 ps AIMD simulations in H2O environment, with energy fluctuations remaining within a stable range. However, in an O2 environment, significant oxidation is observed, characterized by the formation of localized Be-O compounds and the cleavage of Be-N bonds. These results, combined with vacuum stability tests, demonstrate that while tetr-2D-BeN3 maintains thermal stability up to 1400 K in vacuum, it shows pronounced oxygen sensitivity at ambient conditions, necessitating oxygen-free environments during both preparation and application.
Considering the tendency for point vacancies to form during the synthesis of 2D materials, four different types of point vacancies on the tetr-2D-BeN3 surface are tested, labeled as VBe, VN1, VN2, and VN3 (Figure 4a). As shown in Figure 4b,c, after fully structural optimization, the VBe vacancy is found to cause structural disruption in the tetr-2D-BeN3 sheet, while the structures with VN1, VN2, and VN3 vacancies only transform the original Be-N pentagonal rings into Be-N quadrilateral rings without significant changes to the overall crystal structure. Furthermore, calculations of vacancy formation energies (EF-vac) reveal that the EF-vac values for VBe and VN1/N2/N3 are 7.413 eV and 6.861 eV, respectively, indicating that VBe is more difficult to form than VN (Figure 4d). Additionally, EF-vac exceeding 5 eV suggests that low-density vacancies are energetically unlikely to form on the tetr-2D-BeN3 surface. Furthermore, considering that many nitrides require high-pressure conditions for synthesis and that tetr-2D-BeN3 additionally possesses porous characteristics, the experimental synthesis of this material still faces significant challenges.

3.3. Mechanical, Electronic, and Optical Properties of BeN3

For mechanical properties, the elastic constants of tetr-2D-BeN3 are calculated as C11 = C22 = 76.2, C12 = 42.7, and C66 = 17.0 N·m−1 satisfying the mechanical stability criteria for 2D structures (C11 > 0, C22 > 0, C11C22C12C21 > 0, C66 > 0) [66], confirming its mechanical stability. Tetr-2D-BeN3 exhibits quasi-isotropic mechanical behavior, the in-plane Young’s stiffness ( Y ( θ ) ) is in range of 52.3–52.9 N·m−1 with a tiny anisotropy of Ymax/Ymin = 1.01, indicating a relatively soft in-plane characteristic, comparable to silicene (62 N·m−1) [67] and phosphorene (21–91 N·m−1) [68] but significantly lower than graphene (330 N·m−1) [69] and h-BN (279 N·m−1) [70]. The Poisson’s ratio of tetr-2D-BeN3 is high (0.55–0.56), reflecting high ductility.
The band structure of tetr-2D-BeN3 shown in Figure 5a, reveals an indirect bandgap of 1.68 eV (PBE), with the valence band maximum (VBM) located at the M point (1/2, 1/2, 0) and the conduction band minimum (CBM) at the Γ point (0, 0, 0). HSE06 hybrid functional calculations yielded a more accurate bandgap of 2.66 eV. As shown in Figure 5b,c, projected density of states (PDOS, PBE) and partial charge analysis show that the VBM of tetr-2D-BeN3 is dominated by the pz orbitals of N1 and N3, while the CBM arises from the pz orbitals of N2 and minor contributions from Be pz orbitals.
To gain deeper insight into the carrier transport properties of tetr-2D-BeN3, deformation potential theory is employed to calculate its carrier mobility. With a tetragonal lattice structure, tetr-2D-BeN3 exhibits identical mobility along its two principal axes (a/b directions). For electrons (E) and holes (H), the deformation potentials and effective masses of E/H along these axes are 1.012/0.986 eV and 0.974/0.806 me, respectively. The calculated mobilities of E/H reach 1.6/2.5 × 103 cm2V−1s−1, which is lower than those of graphene (340 × 103 cm2V−1s−1) [71] and monolayer black phosphorus (26 × 103 cm2V−1s−1) [72], but higher than that of MoS2 monolayer (0.072–0.2 × 103 cm2V−1s−1) [73]. These findings suggest that tetr-2D-BeN3 could be a promising candidate for microelectronic applications.
The optical properties of tetr-2D-BeN3 (calculated at the HSE06 level) reveal anisotropic dielectric constants, satisfying εaa = εbbεcc. As shown in Figure 6, the tetr-2D-BeN3 exhibits an indirect bandgap of 2.66 eV and anisotropic optical absorption spectra, with significant differences between the in-plane (a/b) and out-of-plane (c) absorption coefficients. Its absorption in the visible range is weak, with only a gradual increase along the a/b direction at 2.6 eV. However, in the ultraviolet range, absorption peaks at 3.5 eV and 8 eV, reaching 3.7 and 5.5 × 105 cm−1. This unique anisotropic optical response, combined with its 2.66 eV indirect bandgap, suggests that tetr-2D-BeN3 holds potential for applications in polarization-sensitive optoelectronic devices and ultraviolet photodetection.

3.4. Gas Separation Applications of BeN3

The tetr-2D-BeN3 monolayer structure contains twelve-membered rings with a pore diameter of 4.98 Å, indicating its potential for gas separation. Further studies reveal that tetr-2D-BeN3 exhibits low energy barriers for the permeation of helium (He), neon (Ne), and hydrogen (H2), with values of 0.241 eV, 0.486 eV, and 0.564 eV, respectively. This suggests that He, Ne, and H2 can diffuse rapidly across the membrane over a wide temperature range, as illustrated in Figure 7a,b. Moreover, tetr-2D-BeN3 demonstrates high selectivity for Ar and H2 compared with other noble gases and common atmospheric gases (Figure 7c). Specifically, the selectivity ratios for He/Ar, Ne/Ar, H2/O2, H2/H2O, and H2/N2 reach 1023, 1019, 106, 108, and 1013, respectively. These gas separation capabilities are comparable to those of porous phosphorene and porous silicene [74,75]. Therefore, tetr-2D-BeN3 holds significant application value and broad prospects in the field of gas separation.

4. Conclusions

In summary, based on first-principles calculations, the energetic properties, structural stability, mechanical properties, electronic properties, optical absorption characteristics, and gas separation performance of the porous tetr-2D-BeN3 monolayer are systematically investigated. The tetr-2D-BeN3 is a global energy-minimum 2D structure searched by the artificial bee colony (ABC) algorithm. This monolayer material exhibits high energy density of 3.34 kJ·g−1, excellent lattice dynamic stability and maintains its fundamental structural framework with outstanding thermal stability up to approximately 1400 K. It demonstrates in-plane quasi-isotropic mechanical, electronic, and optical properties, including a Young’s stiffness of 52.3–52.9 N·m−1, a large Poisson’s ratio of 0.55–0.56, a moderate indirect bandgap of 2.66 eV, along with high carrier mobilities of 1.6/2.5 × 103 cm2V−1s−1 and strong ultraviolet light absorption capability. The porous nature of the material enables highly efficient gas separation for hydrogen and argon, achieving maximum selectivity values of 1013 (H2/N2) and 1023 (He/Ar), respectively. Owing to these novel characteristics and exceptional performance, the tetr-2D-BeN3 monolayer shows potential applications in energetic materials, flexible semiconductor devices, ductile materials, ultraviolet photodetectors, and other fields, thereby expanding the design possibilities for polynitride materials.

Author Contributions

Conceptualization, J.J. and W.W.; methodology, validation, formal analysis, investigation, data curation, writing—original draft preparation, and visualization, J.J.; writing—review editing, Q.H.; supervision, writing—review editing, and project administration; software, and resources, H.G. and W.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

We thank Hefei Advanced Computing Center.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhang, C.; Sun, C.; Hu, B.; Yu, C.; Lu, M. Synthesis and characterization of the pentazolate anion cyclo-N5ˉ in (N5)6(H3O)3(NH4)4Cl. Science 2017, 355, 374–376. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, J.; Shreeve, J.N.M. 3,3’-Dinitroamino-4,4’-azoxyfurazan and Its Derivatives: An Assembly of Diverse N–O Building Blocks for High-Performance Energetic Materials. J. Am. Chem. Soc. 2014, 136, 4437–4445. [Google Scholar] [CrossRef] [PubMed]
  3. Mailhiot, C.; Yang, L.H.; McMahan, A.K. Polymeric nitrogen. Phys. Rev. B 1992, 46, 14419–14435. [Google Scholar] [CrossRef]
  4. Laniel, D.; Geneste, G.; Weck, G.; Mezouar, M.; Loubeyre, P. Hexagonal Layered Polymeric Nitrogen Phase Synthesized near 250 GPa. Phys. Rev. Lett. 2019, 122, 066001. [Google Scholar] [CrossRef]
  5. McMahan, A.K.; LeSar, R. Pressure Dissociation of Solid Nitrogen under 1 Mbar. Phys. Rev. Lett. 1985, 54, 1929–1932. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, X.; Wang, Y.; Miao, M.; Zhong, X.; Lv, J.; Cui, T.; Li, J.; Chen, L.; Pickard, C.J.; Ma, Y. Cagelike Diamondoid Nitrogen at High Pressures. Phys. Rev. Lett. 2012, 109, 175502. [Google Scholar] [CrossRef]
  7. Sun, J.; Martinez-Canales, M.; Klug, D.D.; Pickard, C.J.; Needs, R.J. Stable All-Nitrogen Metallic Salt at Terapascal Pressures. Phys. Rev. Lett. 2013, 111, 175502. [Google Scholar] [CrossRef]
  8. Glukhovtsev, M.N.; Jiao, H.; Schleyer, P.v.R. Besides N(2), What Is the Most Stable Molecule Composed Only of Nitrogen Atoms? Inorg Chem 1996, 35, 7124–7133. [Google Scholar] [CrossRef]
  9. Zhou, H.; Wong, N.-B.; Zhou, G.; Tian, A. Theoretical Study on “Multilayer” Nitrogen Cages. J. Phys. Chem. A 2006, 110, 3845–3852. [Google Scholar] [CrossRef]
  10. Thompson, M.D.; Bledson, T.M.; Strout, D.L. Dissociation Barriers for Odd-Numbered Acyclic Nitrogen Molecules N9 and N11. J. Phys. Chem. A 2002, 106, 6880–6882. [Google Scholar] [CrossRef]
  11. Wang, X.; Li, J.; Botana, J.; Zhang, M.; Zhu, H.; Chen, L.; Liu, H.; Cui, T.; Miao, M. Polymerization of nitrogen in lithium azide. J. Chem. Phys. 2013, 139, 164710. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, M.; Yin, K.; Zhang, X.; Wang, H.; Li, Q.; Wu, Z. Structural and electronic properties of sodium azide at high pressure: A first principles study. Solid State Commun. 2013, 161, 13–18. [Google Scholar] [CrossRef]
  13. Steele, B.A.; Oleynik, I.I. Novel Potassium Polynitrides at High Pressures. J. Phys. Chem. A 2017, 121, 8955–8961. [Google Scholar] [CrossRef] [PubMed]
  14. Peng, F.; Han, Y.; Liu, H.; Yao, Y. Exotic stable cesium polynitrides at high pressure. Sci. Rep. 2015, 5, 16902. [Google Scholar] [CrossRef]
  15. Wei, S.; Li, D.; Liu, Z.; Li, X.; Tian, F.; Duan, D.; Liu, B.; Cui, T. Alkaline-earth metal (Mg) polynitrides at high pressure as possible high-energy materials. Phys. Chem. Chem. Phys. 2017, 19, 9246–9252. [Google Scholar] [CrossRef]
  16. Zhu, S.; Peng, F.; Liu, H.; Majumdar, A.; Gao, T.; Yao, Y. Stable Calcium Nitrides at Ambient and High Pressures. Inorg Chem 2016, 55, 7550–7555. [Google Scholar] [CrossRef]
  17. Shen, Y.; Oganov, A.R.; Qian, G.; Zhang, J.; Dong, H.; Zhu, Q.; Zhou, Z. Novel lithium-nitrogen compounds at ambient and high pressures. Sci. Rep. 2015, 5, 14204. [Google Scholar] [CrossRef]
  18. Steele, B.A.; Oleynik, I.I. Sodium pentazolate: A nitrogen rich high energy density material. Chem. Phys. Lett. 2016, 643, 21–26. [Google Scholar] [CrossRef]
  19. Williams, A.S.; Steele, B.A.; Oleynik, I.I. Novel rubidium poly-nitrogen materials at high pressure. J. Chem. Phys. 2017, 147, 234701. [Google Scholar] [CrossRef]
  20. Xia, K.; Zheng, X.; Yuan, J.; Liu, C.; Gao, H.; Wu, Q.; Sun, J. Pressure-Stabilized High-Energy-Density Alkaline-Earth-Metal Pentazolate Salts. J. Phys. Chem. C 2019, 123, 10205–10211. [Google Scholar] [CrossRef]
  21. Yi, W.; Zhao, L.; Liu, X.; Chen, X.; Zheng, Y.; Miao, M. Packing high-energy together: Binding the power of pentazolate and high-valence metals with strong bonds. Mater. Des. 2020, 193, 108820. [Google Scholar] [CrossRef]
  22. Bykov, M.; Fedotenko, T.; Chariton, S.; Laniel, D.; Glazyrin, K.; Hanfland, M.; Smith, J.S.; Prakapenka, V.B.; Mahmood, M.F.; Goncharov, A.F.; et al. High-Pressure Synthesis of Dirac Materials: Layered van der Waals Bonded BeN4 Polymorph. Phys. Rev. Lett. 2021, 126, 175501. [Google Scholar] [CrossRef] [PubMed]
  23. Huang, B.; Frapper, G. Barium–Nitrogen Phases Under Pressure: Emergence of Structural Diversity and Nitrogen-Rich Compounds. Chem. Mater. 2018, 30, 7623–7636. [Google Scholar] [CrossRef]
  24. Wei, S.; Lian, L.; Liu, Y.; Li, D.; Liu, Z.; Cui, T. Pressure-stabilized polymerization of nitrogen in alkaline-earth-metal strontium nitrides. Phys. Chem. Chem. Phys. 2020, 22, 5242–5248. [Google Scholar] [CrossRef] [PubMed]
  25. Bhadram, V.S.; Kim, D.Y.; Strobel, T.A. High-Pressure Synthesis and Characterization of Incompressible Titanium Pernitride. Chem. Mater. 2016, 28, 1616–1620. [Google Scholar] [CrossRef]
  26. Niwa, K.; Suzuki, K.; Muto, S.; Tatsumi, K.; Soda, K.; Kikegawa, T.; Hasegawa, M. Discovery of the Last Remaining Binary Platinum-Group Pernitride RuN2. Chem. A Eur. J. 2014, 20, 13885–13888. [Google Scholar] [CrossRef]
  27. Zhang, X.; Xie, X.; Dong, H.; Zhang, X.; Wu, F.; Mu, Z.; Wen, M. Pressure-Induced High-Energy-Density BeN4 Materials with Nitrogen Chains: First-Principles Study. J. Phys. Chem. C 2021, 125, 25376–25382. [Google Scholar] [CrossRef]
  28. Zhang, S.; Zhao, Z.; Liu, L.; Yang, G. Pressure-induced stable BeN4 as a high-energy density material. J. Power Sources 2017, 365, 155–161. [Google Scholar] [CrossRef]
  29. Wei, S.; Li, D.; Liu, Z.; Wang, W.; Tian, F.; Bao, K.; Duan, D.; Liu, B.; Cui, T. A Novel Polymerization of Nitrogen in Beryllium Tetranitride at High Pressure. J. Phys. Chem. C 2017, 121, 9766–9772. [Google Scholar] [CrossRef]
  30. Yu, S.; Huang, B.; Zeng, Q.; Oganov, A.R.; Zhang, L.; Frapper, G. Emergence of Novel Polynitrogen Molecule-like Species, Covalent Chains, and Layers in Magnesium–Nitrogen MgxNy Phases under High Pressure. J. Phys. Chem. C 2017, 121, 11037–11046. [Google Scholar] [CrossRef]
  31. Yuan, J.; Xia, K.; Wu, J.; Sun, J. High-energy-density pentazolate salts: CaN10 and BaN10. Sci. China Phys. Mech. Astron. 2020, 64, 218211. [Google Scholar] [CrossRef]
  32. Zhang, X.; Dong, H.; Huang, L.; Long, H.; Zhang, X.; Wu, F.; Mu, Z.; Wen, M. Pressure-Induced High-Energy-Density BeN6 Materials: First-Principles study. Comput. Mater. Sci. 2024, 244, 113272. [Google Scholar] [CrossRef]
  33. 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] [PubMed]
  34. Carvalho, A.; Wang, M.; Zhu, X.; Rodin, A.S.; Su, H.; Castro Neto, A.H. Phosphorene: From theory to applications. Nat. Rev. Mater. 2016, 1, 16061. [Google Scholar] [CrossRef]
  35. Wang, Z.-Q.; Lü, T.-Y.; Wang, H.-Q.; Feng, Y.P.; Zheng, J.-C. Review of borophene and its potential applications. Front. Phys. 2019, 14, 33403. [Google Scholar] [CrossRef]
  36. Tiwari, S.K.; Sahoo, S.; Wang, N.; Huczko, A. Graphene research and their outputs: Status and prospect. J. Sci. Adv. Mater. Devices 2020, 5, 10–29. [Google Scholar] [CrossRef]
  37. Yang, Q.; Zhao, Y.; Jiang, X.; Wang, B.; Zhao, J. Unconventional stoichiometric two-dimensional potassium nitrides with anion-driven metallicity and superconductivity. J. Mater. Chem. C 2024, 12, 103–109. [Google Scholar] [CrossRef]
  38. Wei, Y.; Ma, Y.; Wei, W.; Li, M.; Huang, B.; Dai, Y. Promising Photocatalysts for Water Splitting in BeN2 and MgN2 Monolayers. J. Phys. Chem. C 2018, 122, 8102–8108. [Google Scholar] [CrossRef]
  39. Li, X.; Zhang, S.; Zhang, C.; Wang, Q. Stabilizing benzene-like planar N6 rings to form a single atomic honeycomb BeN3 sheet with high carrier mobility. Nanoscale 2018, 10, 949–957. [Google Scholar] [CrossRef]
  40. Mortazavi, B.; Shojaei, F.; Zhuang, X. Ultrahigh stiffness and anisotropic Dirac cones in BeN4 and MgN4 monolayers: A first-principles study. Mater. Today Nano 2021, 15, 100125. [Google Scholar] [CrossRef]
  41. Zhou, W.; Guo, S.; Zeng, H.; Zhang, S. High-Performance Monolayer BeN2 Transistors With Ultrahigh On-State Current: A DFT Coupled With NEGF Study. IEEE Trans. Electron. Devices 2022, 69, 4501–4506. [Google Scholar] [CrossRef]
  42. Ding, Y.-M.; Ji, Y.; Dong, H.; Rujisamphan, N.; Li, Y. Electronic properties and oxygen reduction reaction catalytic activity of h-BeN2 and MgN2 by first-principles calculations. Nanotechnology 2019, 30, 465202. [Google Scholar] [CrossRef] [PubMed]
  43. Ni, S.; Jiang, J.; Wang, W.; Wu, X.; Zhuo, Z.; Wang, Z. Beryllium dinitride monolayer: A multifunctional direct bandgap anisotropic semiconductor containing polymeric nitrogen with oxygen reduction catalysis and potassium-ion storage capability. J. Mater. Chem. A 2025, 13, 10214–10223. [Google Scholar] [CrossRef]
  44. Filho, M.A.M.; Farmer, W.; Hsiao, C.-L.; dos Santos, R.B.; Hultman, L.; Birch, J.; Ankit, K.; Gueorguiev, G.K. Density Functional Theory-Fed Phase Field Model for Semiconductor Nanostructures: The Case of Self-Induced Core–Shell InAlN Nanorods. Cryst. Growth Des. 2024, 24, 4717–4727. [Google Scholar] [CrossRef]
  45. Wang, Y.; Lv, J.; Zhu, L.; Ma, Y. Crystal structure prediction via particle-swarm optimization. Phys. Rev. B 2010, 82, 094116. [Google Scholar] [CrossRef]
  46. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar] [CrossRef]
  47. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef]
  48. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  49. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef]
  50. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef]
  51. Togo, A.; Tanaka, I. First principles phonon calculations in materials science. Scr. Mater. 2015, 108, 1–5. [Google Scholar] [CrossRef]
  52. Cadelano, E.; Palla, P.L.; Giordano, S.; Colombo, L. Elastic properties of hydrogenated graphene. Phys. Rev. B 2010, 82, 235414. [Google Scholar] [CrossRef]
  53. Bardeen, J.; Shockley, W. Deformation Potentials and Mobilities in Non-Polar Crystals. Phys. Rev. 1950, 80, 72–80. [Google Scholar] [CrossRef]
  54. Peng, B.; Zhang, H.; Shao, H.; Xu, Y.; Zhang, R.; Zhu, H. The electronic, optical, and thermodynamic properties of borophene from first-principles calculations. J. Mater. Chem. C 2016, 4, 3592–3598. [Google Scholar] [CrossRef]
  55. Saha, S.; Sinha, T.P.; Mookerjee, A. Electronic structure, chemical bonding, and optical properties of paraelectricBaTiO3. Phys. Rev. B 2000, 62, 8828–8834. [Google Scholar] [CrossRef]
  56. Jiang, D.-e.; Cooper, V.R.; Dai, S. Porous Graphene as the Ultimate Membrane for Gas Separation. Nano Lett. 2009, 9, 4019–4024. [Google Scholar] [CrossRef]
  57. Blankenburg, S.; Bieri, M.; Fasel, R.; Müllen, K.; Pignedoli, C.A.; Passerone, D. Porous Graphene as an Atmospheric Nanofilter. Small 2010, 6, 2266–2271. [Google Scholar] [CrossRef]
  58. Carlotti, M.; Johns, J.W.C.; Trombetti, A. The ν5 Fundamental Bands of N2H2 and N2D2. Can. J. Phys. 1974, 52, 340–344. [Google Scholar] [CrossRef]
  59. Morino, Y.; Iijima, T.; Murata, Y. An Electron Diffraction Investigation on the Molecular Structure of Hydrazine. Bull. Chem. Soc. Jpn. 1960, 33, 46–48. [Google Scholar] [CrossRef]
  60. Raza, Z.; Pickard, C.J.; Pinilla, C.; Saitta, A.M. High Energy Density Mixed Polymeric Phase from Carbon Monoxide and Nitrogen. Phys. Rev. Lett. 2013, 111, 235501. [Google Scholar] [CrossRef] [PubMed]
  61. Peng, F.; Yao, Y.; Liu, H.; Ma, Y. Crystalline LiN5 Predicted from First-Principles as a Possible High-Energy Material. J. Phys. Chem. Lett. 2015, 6, 2363–2366. [Google Scholar] [CrossRef] [PubMed]
  62. Liu, L.; Wang, D.; Zhang, S.; Zhang, H. Pressure-stabilized GdN6 with an armchair–antiarmchair structure as a high energy density material. J. Mater. Chem. A 2021, 9, 16751–16758. [Google Scholar] [CrossRef]
  63. Lu, W.; Hao, K.; Liu, S.; Lv, J.; Zhou, M.; Gao, P. Pressure-stabilized high-energy-density material YN10. J. Phys. Condens. Matter 2022, 34, 135403. [Google Scholar] [CrossRef]
  64. Liu, L.; Tang, J.; Wang, Q.-Q.; Shi, D.-X.; Zhang, G.-Y. Thermal stability of MoS2 encapsulated by graphene. Acta Phys. Sin. 2018, 67, 226501. [Google Scholar] [CrossRef]
  65. Liu, F.; Wang, M.; Chen, Y.; Gao, J. Thermal stability of graphene in inert atmosphere at high temperature. J. Solid State Chem. 2019, 276, 100–103. [Google Scholar] [CrossRef]
  66. Ding, Y.; Wang, Y. Density Functional Theory Study of the Silicene-like SiX and XSi3 (X = B, C, N, Al, P) Honeycomb Lattices: The Various Buckled Structures and Versatile Electronic Properties. J. Phys. Chem. C 2013, 117, 18266–18278. [Google Scholar] [CrossRef]
  67. Zhuo, Z.; Wu, X.; Yang, J. Two-dimensional silicon crystals with sizable band gaps and ultrahigh carrier mobility. Nanoscale 2018, 10, 1265–1271. [Google Scholar] [CrossRef]
  68. Zhuo, Z.; Wu, X.; Yang, J. Two-Dimensional Phosphorus Porous Polymorphs with Tunable Band Gaps. J. Am. Chem. Soc. 2016, 138, 7091–7098. [Google Scholar] [CrossRef]
  69. Li, Y.; Liao, Y.; Chen, Z. Inside Back Cover: Be2C Monolayer with Quasi-Planar Hexacoordinate Carbons: A Global Minimum Structure. Angew. Chem. Int. Ed. 2014, 53, 7369. [Google Scholar] [CrossRef]
  70. Roy, S.; Zhang, X.; Puthirath, A.B.; Meiyazhagan, A.; Bhattacharyya, S.; Rahman, M.M.; Babu, G.; Susarla, S.; Saju, S.K.; Tran, M.K.; et al. Structure, Properties and Applications of Two-Dimensional Hexagonal Boron Nitride. Adv. Mater. 2021, 33, 2101589. [Google Scholar] [CrossRef]
  71. Chen, J.; Xi, J.; Wang, D.; Shuai, Z. Carrier Mobility in Graphyne Should Be Even Larger than That in Graphene: A Theoretical Prediction. J. Phys. Chem. Lett. 2013, 4, 1443–1448. [Google Scholar] [CrossRef] [PubMed]
  72. Qiao, J.; Kong, X.; Hu, Z.-X.; Yang, F.; Ji, W. High-mobility transport anisotropy and linear dichroism in few-layer black phosphorus. Nat. Commun. 2014, 5, 4475. [Google Scholar] [CrossRef]
  73. 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]
  74. Hu, W.; Wu, X.; Li, Z.; Yang, J. Helium separation via porous silicene based ultimate membrane. Nanoscale 2013, 5, 9062–9066. [Google Scholar] [CrossRef] [PubMed]
  75. Hu, W.; Wu, X.; Li, Z.; Yang, J. Porous silicene as a hydrogen purification membrane. Phys. Chem. Chem. Phys. 2013, 15, 5753–5757. [Google Scholar] [CrossRef]
Figure 1. (a) Global structural search results for Be:N = 1:3, where the blue box indicates the previously reported 2D Be-N monolayer, and the black box marks the global energy-minimum 2D Be-N monolayer. (b) The top and side views of the calypso structure search for the lowest-energy configurations of Be:N ratios 2:6, 3:9, and 4:12. The structures are labeled as α/β/γ-2D-BeN3 in ascending order of energy. (c) The top and side views, deformation charge density, electron localization function, and Bader charge analysis of tetr-2D-BeN3.
Figure 1. (a) Global structural search results for Be:N = 1:3, where the blue box indicates the previously reported 2D Be-N monolayer, and the black box marks the global energy-minimum 2D Be-N monolayer. (b) The top and side views of the calypso structure search for the lowest-energy configurations of Be:N ratios 2:6, 3:9, and 4:12. The structures are labeled as α/β/γ-2D-BeN3 in ascending order of energy. (c) The top and side views, deformation charge density, electron localization function, and Bader charge analysis of tetr-2D-BeN3.
Nanomaterials 15 01004 g001
Figure 2. (a) Phonon spectrum of tetr-2D-BeN3, (b) the energy curves, and (c) final structural snapshots of tetr-2D-BeN3 from AIMD simulations at 300, 500, 1000, 1200, 1400, and 1500 K over 5 ps, as well as structural snapshots from AIMD simulations at 2000 K. The red circles indicate broken Be-N bonds, while the blue circles highlight N2 molecules formed upon the decomposition of tetr-2D-BeN3.
Figure 2. (a) Phonon spectrum of tetr-2D-BeN3, (b) the energy curves, and (c) final structural snapshots of tetr-2D-BeN3 from AIMD simulations at 300, 500, 1000, 1200, 1400, and 1500 K over 5 ps, as well as structural snapshots from AIMD simulations at 2000 K. The red circles indicate broken Be-N bonds, while the blue circles highlight N2 molecules formed upon the decomposition of tetr-2D-BeN3.
Nanomaterials 15 01004 g002
Figure 3. The structural snapshots (a,b) and the energy curves (c,d) of tetr-2D-BeN3 with H2O and O2 molecules from AIMD simulations at 300 K over 5 ps.
Figure 3. The structural snapshots (a,b) and the energy curves (c,d) of tetr-2D-BeN3 with H2O and O2 molecules from AIMD simulations at 300 K over 5 ps.
Nanomaterials 15 01004 g003
Figure 4. (a) Four different types of point vacancies on the tetr-2D-BeN3 surface, along with the fully optimized structures of (b) VBe@tetr-2D-BeN3 and (c) VN1/N2/N3@tetr-2D-BeN3, as well as their corresponding (d) vacancy formation energies.
Figure 4. (a) Four different types of point vacancies on the tetr-2D-BeN3 surface, along with the fully optimized structures of (b) VBe@tetr-2D-BeN3 and (c) VN1/N2/N3@tetr-2D-BeN3, as well as their corresponding (d) vacancy formation energies.
Nanomaterials 15 01004 g004
Figure 5. (a) The band structures, the black and yellow lines represent the PBE functional and HSE06 functional, respectively, (b) projected density of states (PDOS, PBE), and (c) partial electron density of band edge of tetr-2D-BeN3. The Fermi level is set to 0 eV.
Figure 5. (a) The band structures, the black and yellow lines represent the PBE functional and HSE06 functional, respectively, (b) projected density of states (PDOS, PBE), and (c) partial electron density of band edge of tetr-2D-BeN3. The Fermi level is set to 0 eV.
Nanomaterials 15 01004 g005
Figure 6. Optical absorption coefficient of tetr-2D-BeN3 computed based on the HSE06 method.
Figure 6. Optical absorption coefficient of tetr-2D-BeN3 computed based on the HSE06 method.
Nanomaterials 15 01004 g006
Figure 7. (a) Schematic diagram of the tetr-2D-BeN3 as a gas separation membrane. (b) The diffusion rate and (c) selectivity for gas molecules as a function of temperature.
Figure 7. (a) Schematic diagram of the tetr-2D-BeN3 as a gas separation membrane. (b) The diffusion rate and (c) selectivity for gas molecules as a function of temperature.
Nanomaterials 15 01004 g007
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

Jiang, J.; Hu, Q.; Wang, W.; Guo, H. Two-Dimensional Porous Beryllium Trinitride Monolayer as Multifunctional Energetic Material. Nanomaterials 2025, 15, 1004. https://doi.org/10.3390/nano15131004

AMA Style

Jiang J, Hu Q, Wang W, Guo H. Two-Dimensional Porous Beryllium Trinitride Monolayer as Multifunctional Energetic Material. Nanomaterials. 2025; 15(13):1004. https://doi.org/10.3390/nano15131004

Chicago/Turabian Style

Jiang, Jiaxin, Qifan Hu, Weiyi Wang, and Hongyan Guo. 2025. "Two-Dimensional Porous Beryllium Trinitride Monolayer as Multifunctional Energetic Material" Nanomaterials 15, no. 13: 1004. https://doi.org/10.3390/nano15131004

APA Style

Jiang, J., Hu, Q., Wang, W., & Guo, H. (2025). Two-Dimensional Porous Beryllium Trinitride Monolayer as Multifunctional Energetic Material. Nanomaterials, 15(13), 1004. https://doi.org/10.3390/nano15131004

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