Hong
Zheng
*a,
Zhenhan
Zhao
a,
Jiasheng
Chen
a,
Hanwen
Qin
a,
Jun
Li
a and
Xiang
Zhao
*b
aState Key Laboratory of Electrical Insulation and Power Equipment, Center of Nanomaterials for Renewable Energy, School of Electrical Engineering, Xi'an Jiaotong University, Xi'an 710049, China. E-mail: zhenghong666@xjtu.edu.cn
bInstitute for Chemical Physics, School of Chemistry, MOE Key Laboratory for Nonequilibrium Synthesis and Modulation of Condensed Matter, Xi'an Jiaotong University, Xi'an 710049, China. E-mail: xzhao@mail.xjtu.edu.cn
First published on 31st March 2025
Designing new auxetic materials is important for flexible electronics. The micromechanism of auxeticity in two-dimensional (2D) materials has attracted significant attention but the main factors of auxeticity are case-dependent and their connection with the geometrical/electronic features of 2D materials still requires systematic exploration. In this work, two new phases of 2D group-VA materials, namely, X10-2_2 and X12-14 (X = P, As, Sb, and Bi), were predicted. All these structures were proven to be stable from the aspects of their thermodynamic and kinetic stabilities. Investigations of their electronic properties revealed that these structures were all semiconductors with high anisotropic mobility. Mechanical analyses showed that all the X12-14 phases exhibited auxeticity, whereas the Poisson's ratios of the X10-2_2 series presented a strong dependence on the group-VA elements. The P10-2_2 structure exhibited auxeticity under both tensile and compression strain. However, As10-2_2 exhibited a peculiar half-auxeticity and Sb10-2_2 and Bi10-2_2 were non-auxetic. Evaluations of the interatomic interaction forces revealed that it is the unique folded structures and the changes of the atomic interactions that induce the different mechanical properties of the X10-2_2 structures. This work uncovers the special relationship between the auxeticity of 2D group-VA materials and the inherent natures of the group-VA elements from the viewpoint of the interatomic interaction force, which will assist future research on auxetic materials.
Similar to phosphorous, other group-VA elements, such as arsenic, stibium, and bismuth, can also form 2D materials with similar lattice structures as phosphorenes as their sp2 hybridization leads to analogous bonding features to phosphorous. Experimentally synthesized arsenene, antimonene, and bismuthene were reported to possess structural similarities to β-phosphorene, all exhibiting a non-planar, buckled configuration.29–32 Concurrently, there exist many theoretically stable allotropes of these two-dimensional materials that have yet to be experimentally synthesized. These unexplored allotropic forms are predicted to possess a diverse range of superior electronic and mechanical properties.31 Despite the similar valence electronic configurations, the chemical properties of group-VA elements varies with the increasing tendencies of their atomic size and metallicity from P to Bi, which would induce different chemical and physical characteristics of 2D group-VA materials based on the same phase. Nevertheless, the influence of group-VA elements toward the electronic/mechanical properties of 2D group-VA materials has not yet been illustrated, and still requires comprehensive investigation and comparison between 2D group-VA materials based on the same conformation.
Owing to their atomic scale of thickness, the mechanical property becomes important for 2D materials and many of them have been found to exhibit auxeticity.33 Auxeticity is reflected in the Poisson's ratio (ν), which is defined as the ratio of transverse contraction strain to longitudinal tensile strain and herein measures the fundamental mechanical response of a solid to external loads. Most materials exhibit positive Poisson's ratios, indicating that they will thin out (or thicken) laterally when subjected to longitudinal tension (or compression). By contrast, materials with negative Poisson's ratios, known as auxetic materials,34 exhibit counterintuitive properties compared to traditional materials. Auxeticity endows nanomaterials with excellent impact resistance, superior elasticity and shear resistance, and sound/vibration absorption, leading to diverse potential applications in fields such as biomedicine, bulletproof vests, sensors, and aerospace.
Over the past decades, various negative Poisson's ratio 2D materials have been theoretically predicted, synthesized, or manufactured.35–44 Black phosphorene is one of the most representative examples, which was predicted to possess an out-of-plane negative Poisson's ratio of −0.027.45 Researchers attributed such an out-of-plane negative Poisson's ratio to the folded structure of black phosphorene. Subsequently, researchers have predicted multiple 2D materials with negative Poisson's ratios.46–48 For instance, monolayer arsenene with a structure similar to black phosphorene is believed to have an out-of-plane negative Poisson's ratio of −0.09 and the magnitude becomes more negative as the number of layers increases, reaching approximately −0.12 for a four-layer arsenene.49 δ-P was also predicted to have an in-plane negative Poisson's ratio of −0.267 due to its folded geometry.50 Many 2D materials with geometries akin to δ-P have been suggested to possess negative Poisson's ratios, such as δ-PbO.51 Furthermore, some 2D materials with unique auxetic properties have been proposed, such as Ag2S with biaxial negative Poisson's ratios (in-plane and out-of-plane)52 and PdB4 exhibiting half-auxeticity.53 However, to date, the number of intrinsic 2D auxetic materials remains quite limited and the search for new 2D auxetic materials remains an urgent task in modern nanodevice research. In addition, knowledge of the rules and main factors in the Poisson's ratio of auxetic materials, which strongly depend on the geometrical structures and bonding features of 2D materials, is crucial for understanding the origin of auxeticity and can only be provided by theoretical studies. Since the phosphorus element can form various 2D materials, as mentioned above, it is possible to design auxetic 2D phosphorus materials through theoretical investigations. Moreover, other group-VA elements, such as arsenic, stibium, and bismuth, are different from phosphorus in atomic size and bonding feature, leading to possible differences in the chemical/physical properties of 2D group-VA materials, which can only be elucidated by theoretical investigations.
In this work, we first searched out a series of new phases of 2D phosphorus allotropes utilizing a structure prediction method. We then focused on two of them with unique in-plane negative Poisson's ratios and constructed analogues with other group-VA elements. The stabilities of these 2D materials were then studied from the viewpoints of cohesive energy, phonon spectrum, and ab initio molecular dynamics (AIMD). The electronic properties of these group-VA 2D structures were next analysed, including their band structures, partial density of states (PDOS), and carrier mobilities. For assessing their mechanical properties, the Poisson's ratios were calculated and the relationship between the Poisson's ratios and the changes among the group-VA elements was uncovered based on analyses of their geometrical structures and interatomic forces.
All the density functional theory (DFT) calculations were performed with the Vienna Ab initio Simulation Package (VASP).56 In the optimizations of the 2D group-VA structures, the projector augmented wave (PAW)57 method was used to describe the interaction between the valence electrons and core electrons. The generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE)58 exchange–correlation functional was used as the electronic exchange–correlation function. Convergence tests for the plane wave energy cutoff and electronic wave vector k were performed, and the appropriate wave function energy cutoff value ENCUT was set as 500 eV. The contribution of the van der Waals (vdW) interactions was considered using the DFT-D2 method developed by Grimme.59 The Brillouin zone was sampled using mesh sizes of 5 × 5 × 1 in the Γ-centered Monkhorst–Pack scheme.60 The conjugate gradient (CG) algorithm was used to optimize the lattice structure, and convergence was reached when the Hellmann–Feynman force acting on all the unconstrained atoms was less than 0.01 eV Å−1 and the energy was less than 10−8 eV per atom. For the phonon spectrum calculations, a 3 × 3 × 1 supercell was used, and the phonon spectra were obtained using the PHONOPY program.61 In the AIMD calculations, a 3 × 3 × 1 supercell was used to minimize the effect of the periodic boundary conditions. The temperature was set as room temperature (300 K) and the time step was set to 2 fs with 3000 steps, resulting in a total simulation time of 6 ps in the NVT ensemble. For the calculations of the band structures and density of states, a Gaussian broadening of 0.05 eV and the first Brillouin zone integration path were used. For orthogonal lattices, the path was Γ(0.0,0.0,0.0) → X(0.5,0.0,0.0) → S(0.5,0.5,0.0) → Y(0.0,0.5,0.0) → Γ(0.0,0.0,0.0). Beside the PBE functional, the HSE06 hybrid functional62 was also used for comparison in the calculations of the band structures.
The calculations of the carrier mobilities according to the deformation potential theory63 were carried out using the following formula:64,65
![]() | (1) |
In the analyses of the mechanical properties, the Young's modulus E(θ) and Poisson's ratio ν(θ) were calculated using the following equations:66–69
![]() | (2) |
![]() | (3) |
In Fig. 1, the top and side views of the two phases of the group-VA monolayers together with the Wigner–Seitz cell are shown. The X10-2_2 and X12-14 phases had the space groups P and P21/c, respectively. Herein, 10 and 12 represent the number of atoms contained in the unit cell, and 2 and 14 represent the space group serial number. The ‘_2’ ordinal number is used to distinguish the structure of multiple identical space groups (another two phases named X10-2_1 and X10-2_3 have also been discovered). The geometrical structures of the X10-2_2 and X12-14 phases could be described as a combination of two layers of pentagonal phosphorous rings along the z axis. Specifically, in the X10-2_2 structures, each pentagonal phosphorous ring of the upper layer is connected with another neighboring ring of the lower layer by three P–P bonds, two of which form a tetragonal ring. The pentagonal phosphorous rings X12-14 structures adopt a similar layer connection through P–P bonds. For each group-VA element, the ranges of the bond length and bond angle vary little between the X10-2_2 and X12-14 phases (Table 1). As the atomic radius increases from P to Bi, the bond lengths and layer thicknesses present increasing tendencies.
![]() | ||
Fig. 1 (a–h) Top and side views of the X10-2_2 and X12-14 monolayers. The lower atom layers are marked with darker colors. |
Structure | a/Å | b/Å | γ/° | h/Å | d/Å | Y/° |
---|---|---|---|---|---|---|
P10-2_2 | 5.00 | 7.11 | 90 | 3.98 | 2.23–2.28 | 85.48–114.76 |
As10-2_2 | 5.51 | 7.40 | 90 | 4.51 | 2.49–2.52 | 82.62–113.92 |
Sb10-2_2 | 6.28 | 8.04 | 90 | 5.28 | 2.86–2.93 | 80.37–112.39 |
Bi10-2_2 | 6.61 | 8.16 | 90 | 5.64 | 3.01–3.11 | 79.74–111.37 |
P12-14 | 4.87 | 9.71 | 90 | 3.71 | 2.25–2.27 | 81.86–113.32 |
As12-14 | 5.37 | 10.15 | 90 | 4.18 | 2.49–2.52 | 80.00–111.73 |
Sb12-14 | 6.13 | 10.80 | 90 | 4.90 | 2.86–2.91 | 78.36–111.21 |
Bi12-14 | 6.47 | 10.98 | 90 | 4.95 | 3.00–3.07 | 77.59–108.26 |
To investigate the thermodynamic stability of the X10-2_2 and X12-14 structures, we first calculated their cohesive energies and compared them with other reported 2D group-VA allotropes (α and β phases), as shown in Table 2. For these 2D elemental structures, the wrinkled α phase, known as black phosphorus, was found to be the lowest-energy phase, consistent with previous studies.70 The bulk materials of arsenene, stibnene, and bismuthene were gray arsenic, gray antimony, and β-type bismuth, respectively, corresponding to the β phase. Our calculations showed that among all the 2D group-VA structures, the Ec values of the α and β phases were rather similar (ΔEc < 0.05 eV), indicating they had similar thermodynamic stabilities.
Structures | E c (eV per atom) | ΔEc (eV per atom) |
---|---|---|
a E c = (Etotal − n × Eatom)/n, where Etotal and Eatom are the total energy of the 2D group-VA allotropes and the energy of a free single atom, respectively. | ||
α-P | −3.608 | 0 |
β-P | −3.573 | +0.035 |
P10-2_2 | −3.482 | +0.126 |
P12-14 | −3.484 | +0.124 |
α-As | −3.133 | 0 |
β-As | −3.128 | +0.005 |
As10-2_2 | −3.028 | +0.105 |
As12-14 | −3.039 | +0.094 |
α-Sb | −2.879 | 0 |
β-Sb | −2.855 | +0.024 |
Sb10-2_2 | −2.799 | +0.080 |
Sb12-14 | −2.807 | +0.072 |
α-Bi | −2.811 | 0 |
β-Bi | −2.764 | +0.047 |
Bi10-2_2 | −2.744 | +0.067 |
Bi12-14 | −2.761 | +0.050 |
For the X10-2_2 and X12-14 phases, their cohesive energies were higher than that of the α phase with only moderate differences (<0.15 eV). These results indicate that although both phases are metastable compared to the α and β phases, they are possible new group-VA allotropes that can exist stably. Moreover, it was revealed that the ΔEc vales for the X10-2_2 or X12-14 phase presented a decreasing tendency going from P to Bi, suggesting the increasing tendency of their relative thermodynamic stabilities. Therefore, it is likely that the X10-2_2 and X12-14 phases could be synthesized experimentally, possibly through chemical techniques, such as epitaxial growth, colloidal suspension, or other unconventional methods.
To further validate the dynamical stability of these 2D structures, we calculated the phonon spectra for the X10-2_2 and X12-14 structures, as shown in Fig. S1 and S2.† All the structures exhibited no imaginary frequencies throughout the entire Brillouin zone, indicating that these structures are dynamically stable and can exist stably. The highest frequency lines and phonon band gaps of the four structures within the same phase decreased as the atomic electronegativity weakened, consistent with the trend observed in the cohesive energy changes.
Furthermore, we employed first-principles molecular dynamics (AIMD) simulations to investigate the thermal stability of these structures at room temperature. The energy variations of these structures at room temperature (300 K) are depicted in Fig. S3 and S4.† After a 6 ps simulation at room temperature, the total energies of these systems fluctuated within a very small range (<0.05 eV per atom) and they maintained their equilibrium structures almost unchanged. These results indicate that these systems can stably exist at room temperature.
Structures | PBE (eV) | PBE + SOC (eV) | HSE06 (eV) | HSE06 + SOC (eV) |
---|---|---|---|---|
P10-2_2 | 0.99 | 0.98 | 1.73 | 1.66 |
As10-2_2 | 0.64 | 0.62 | 1.12 | 1.23 |
Sb10-2_2 | 0.38 | 0.33 | 0.68 | 0.61 |
Bi10-2_2 | 0.21 | 0.09 | 0.42 | 0.18 |
P12-14 | 1.40 | 1.39 | 2.14 | 2.15 |
As12-14 | 1.38 | 1.36 | 2.14 | 2.18 |
Sb12-14 | 0.95 | 0.90 | 1.38 | 1.33 |
Bi12-14 | 0.40 | 0.19 | 0.68 | 0.37 |
In addition, the partial density of states (PDOS) was calculated for all the structures (Fig. 3), and the results were consistent with the electronic band structure results. It was revealed that the PDOS near the Fermi level for these 2D group-VA structures was mainly contributed by p orbitals due to the classic sp3 hybridization of group-VA atoms. Specially, the VBM was mainly contributed by the pz orbital for the P and As elements, whereas it was primarily contributed by the py orbital for the Sb and Bi elements.
Structure | Direc-tion | C 2D (J m−2) | E e1 (eV) | E h1 (eV) | μ e (cm2 V−1 s−1) | μ h (cm2 V−1 s−1) | ||
---|---|---|---|---|---|---|---|---|
P10-2_2 | x | 89.89 | 10.41 | 2.15 | 0.16 | 0.55 | 228 | 930 |
y | 14.26 | 0.31 | 3.96 | 0.66 | 0.53 | 9786 | 45 | |
As10-2_2 | x | 63.75 | 7.50 | 1.59 | 0.13 | 0.45 | 470 | 2043 |
y | 20.92 | 0.51 | 4.46 | 0.57 | 0.34 | 7272 | 114 | |
Sb10-2_2 | x | 43.65 | 6.25 | 2.13 | 0.13 | 0.31 | 429 | 1630 |
y | 29.07 | 0.28 | 4.26 | 0.60 | 0.23 | 31![]() |
361 | |
Bi10-2_2 | x | 37.43 | 4.32 | 2.34 | 0.10 | 0.39 | 919 | 401 |
y | 24.47 | 0.53 | 3.04 | 0.93 | 1.00 | 4405 | 61 | |
P12-14 | x | 76.64 | 1.31 | 1.17 | 0.26 | 3.45 | 6447 | 95 |
y | 8.37 | 1.40 | 1.22 | 0.57 | 1.70 | 279 | 19 | |
As12-14 | x | 55.29 | 6.14 | 1.74 | 0.15 | 0.74 | 396 | 343 |
y | 7.43 | 0.34 | 1.62 | 0.82 | 1.41 | 3228 | 28 | |
Sb12-14 | x | 38.25 | 4.84 | 3.55 | 0.12 | 0.20 | 645 | 357 |
y | 5.65 | 0.36 | 1.02 | 0.68 | 1.73 | 3171 | 76 | |
Bi12-14 | x | 29.76 | 3.53 | 2.07 | 0.10 | 0.15 | 1894 | 1993 |
y | 6.66 | 1.11 | 0.73 | 0.34 | 0.69 | 1247 | 793 |
It should be noted that the deformation potential theory generally overestimates the carrier mobility because several important parameters (especially the E1 term) are obtained by fitting and the error becomes relatively large. Also, the influence of DFT functionals toward the deformation potential constant and carrier effective mass has not been adequately analysed and discussed yet. Meanwhile, in experiments, the influence of the substrate and layer number would lead to a large deviation from an ideal 2D material. Despite all this, the absolute values of carrier mobility have poor meaning, whereas their relative values compared with those of a semiconductor material that has been proven to have high carrier conductivity (such as black phosphorus) are helpful for screening nanoelectronic devices. Therefore, in this study, we mainly compared our results with previous theoretical studies based on the same computational method.
![]() | ||
Fig. 4 Stress–strain curves along the x and y directions for the X10-2_2 phase (a–d) and the X12-14 phase (e–h). |
Based on the numerical values of the elastic constants, the Young's modulus, Poisson's ratio, and shear modulus in all directions within the 2D plane for each structure were calculated to further elucidate the mechanical properties of these single-layer structures. The angular-dependent Young's modulus E(θ) and Poisson's ratio ν(θ) are shown in Fig. 5 and 6. The changes in the two-dimensional polar coordinate curves indicated that the mechanical properties of all these structures were highly anisotropic. For the X10-2_2 phase, all the structures reached their maximum values in the x direction, while the maximum Young's modulus decreased in the order of P > As > Sb > Bi (Fig. 5), suggesting a weakening tendency of anisotropy. Specially, the Young's modulus of P10-2_2 reached a maximum value of 90 N m−1 in the x direction, comparable to the maximum Young's modulus of black phosphorus. The Young's modulus of P10-2_2 reached a minimum value of 15 N m−1 in the y direction, smaller than the minimum value of black phosphorus, indicating the stronger anisotropy of this structure. The X12-14 phase structures all reached their maximum values in the x direction and showed their minimum values in the y direction. Similar to X10-2_2, the four X12-14 structures also exhibited a weakening tendency of anisotropy from P to Bi.
These results also revealed that the Young's modulus values of these structures were significantly lower than those of h-BN (271 N m−1) and graphene (341 N m−1), and slightly lower than the maximum Young's modulus of MoS2 (128 N m−1), suggesting that they are relatively flexible materials. For the shear modulus, all the structures reach their minimum values at 45°, 135°, 225°, and 315° directions and maximum values in the x and y directions. Similar to the Young's modulus, the anisotropy of the shear modulus in the X10-2_2 phase weakened from P to Bi. The shear modulus values also indicated that these eight structures are relatively soft materials.
The most interesting mechanical feature of these 2D structures is their Poisson's ratio. Here, the uniaxial strains ranged from −5% to +5%. Our calculations show that the thickness of the layer (strain along z) decreased with increasing the in-plane strain (along x and y), as shown in Fig. S5 and S6,† indicating that the out-of-plane Poisson's ratio for all the structures was positive. In the explorations of the in-plane Poisson's ratio, Fig. 5 and 6 reveal that the P10-2_2 and X12-14 structures possessed negative Poisson's ratios along the 35°, 145°, 215°, and 325° directions, ranging from −0.65 to −0.34. These NTR values are more negative than those of penta-B2N4 (−0.03) and α-P (−0.03). Surprisingly, the Poisson's ratio of the X10-2_2 phase exhibited a strong dependence on group-VA elements: whereby the negative Poisson's ratio disappeared in As10-2_2, Sb10-2_2, and Bi10-2_2 and the anisotropy was weakened as the atomic number of the group-VA element increased. All the X10-2_2 and X12-14 structures had positive Poisson's ratios along the x direction, whereas the Poisson's ratios of P10-2_2 and X12-14 along the y direction were close to 0.
Next, the uniaxial strain, defined as νxy = −εx/εy, was analyzed for X10-2_2 and X12-14. Here, all the structures were deformed uniaxially along the x direction and the changes along the y direction were monitored. For the P10-2_2 and X12-14 phases, the Poisson's ratio along the y direction was slightly negative. Therefore, the expansion along the x direction resulted in no significant strain in the y direction. This is consistent with the Poisson's ratio structures obtained from the elastic constants calculations above.
![]() | ||
Fig. 7 Supercell (marked with red lines) used for calculating the NPR of (a) the X10-2_2 phase and (b) the X12-14 phase along the diagonal direction. |
The corresponding strain responses along the a/b directions for the X10-2_2 and X12-14 structures are shown in Fig. 8, respectively. Due to the structural destruction in the Bi12-14 structure when applying a −5% compressive strain along the a direction, the data for this strain were omitted. The calculated strain responses were consistent with the elastic constants calculations above. All the X12-14 structures exhibited negative Poisson's ratios along both the a and b directions. The P10-2_2 structure exhibited negative Poisson's ratios along the a and b directions, while Sb10-2_2 and Bi10-2_2 exhibited positive Poisson's ratios. Specially, being similar with the case of the PdB4 monolayer, the As10-2_2 structure exhibited a half-auxeticity effect; whereby it presented negative Poisson's ratios during stretching but positive Poisson's ratios during compressing. Apparently, there existed a novel relationship between the Poisson's ratio and group-VA elements in the X10-2_2 phase, which requires deeper analysis of their structures.
As mentioned above, the mechanical properties of X10-2_2 (X = As, Sb, and Bi) were different from those of P10-2_2, indicating the different atomic displacements of X10-2_2 (X = As, Sb, and Bi) compared with P10-2_2. For the As10-2_2 structure, we studied its strain response and the atomic displacement changes after it experienced tensile or compressive strain along the a direction. As shown in Fig. 9(f–j), when As10-2_2 underwent tensile strain along the a direction, the atomic displacements were similar with those of P10-2_2, resulting in a negative Poisson's ratio (Fig. 9(f)). However, when As10-2_2 experienced compressive strain along the a direction, the atomic displacements became quite different from in the case of P10-2_2 (Fig. 9(g)). The b direction component in the moving directions for As1, As2, As3, and As7 were along the negative b direction, whereas those for As5, As6, and As4 were along the positive b direction. Moreover, the θ angle increased under tensile strain, similar with the case of P10-2_2, but changed little under compressive strain. The atomic displacements of Sb10-2_2 and Bi10-2_2 were different from those of P10-2_2 under both tensile and compressive strain (Fig. 10).
![]() | ||
Fig. 11 Mechanical responses of the X10-2_2 phase under strain along the a direction (black lines) and the −ICOHP between X5 and X6 (red lines). |
It was revealed that the −ICOHP value between the X5 and X6 atoms increased in the order of P < As < Sb < Bi under the case of no strain, suggesting an increasing tendency of bonding features between X5 and X6. When strain was applied, this −ICOHP value showed a decreasing tendency from the −ICOHP value between P5 and P6 from a compressive strain of −5% to a tensile strain of +5%. In P10-2_2, the −ICOHP between the P5 and P6 atoms showed a decreasing range from 1.0 eV under a strain of −5% to 0.8 eV when the strain was increased to −2%. In As10-2_2, the −ICOHP value between the As5 and As6 atoms was above 0.9 eV under compressive strain, and correspondingly, was below 0.9 eV under tensile strain, the former corresponded to negative correlations between εa and εb (positive Poisson's ratio) while the latter corresponded to positive correlations between εa and εb (negative Poisson's ratio). As for Sb10-2_2 and Bi10-2_2, the −ICOHP value stayed above 0.9 eV over the whole changing process of strain, corresponding to negative correlations between εa and εb. The −ICOHP values and the strain responses along the b direction are shown in Fig. S7,† which exhibited quite similar results to those in Fig. 11. Therefore, it could be concluded that an −ICOHP value of ∼0.9 eV is a critical point determining the auxeticity of X10-2_2. Of course this −ICOHP value is just a reference but it shows the dependence of auxeticity on the bonding between the X5 and X6 atoms. The weak bonding between X5 and X6 (such as in P10-2_2 and stretched As10-2_2) had little influence on the movement of the upper and lower pentagon rings mentioned above, resulting in negative Poisson's ratios. Nevertheless, if this interatomic interaction becomes strong, compression and stretching of the X5-X6 bond are difficult and the atomic displacements of the upper and lower pentagon rings become less flexible. Consequently, other atoms have to adjust their moving directions and the movements with opposite directions between the upper and lower pentagon rings are destroyed.
In short, the weak interatomic interaction between the X5 and X6 atoms in P10-2_2 and stretched As10-2_2 induced negative Poisson's ratios, whereas the strong interatomic interaction between them in Sb10-2_2, Bi10-2_2, and compressed As10-2_2 induced positive Poisson's ratios. The increasing tendency of this interatomic interaction from P to Bi is connected with the large extension of the valence orbitals of heavy atoms, which increases the orbital overlap between the X5 and X6 atoms. Such interatomic interaction regularly increasing from non-bonding to partial bonding along with the change of group-VA elements was not found in the X12-14 phase and herein all the four X12-14 structures had negative Poisson's ratios.
Analyses of the electronic structures of the X10-2_2 and X12-14 structures revealed that they are all indirect bandgap semiconductors. The structures with the same phase had similar band structures but the bandgaps decreased in the order of P > As > Sb > Bi. All these structures had high electron mobility, especially the Sb10-2_2 structure with an electron mobility of 31493 cm2 V−1 s−1 along the y direction, making them suitable for nanoelectronic devices.
In addition, analyses of the uniaxial tensile strain, Young's modulus, and Poisson's ratio disclosed that these materials possessed unique mechanical properties. Both X10-2_2 and X12-14 phases were stable, soft, and anisotropic materials. Analysis of the diagonal Poisson's ratio suggested that the four X12-14 structures exhibited in-plane negative Poisson's ratios. For the X10-2_2 phase, the P10-2_2 structure exhibited in-plane negative Poisson's ratios in the 35° and 145° directions, while the As10-2_2 structure exhibited half-auxeticity. The Sb10-2_2 and Bi10-2_2 structures exhibited positive Poisson's ratios. It was found that the strengthened atomic interactions between the atoms of the upper and lower pentagon rings led to positive Poisson's ratio in Sb10-2_2, Bi10-2_2, and compressed As10-2_2. Such atomic interactions for As, Sb, and Bi elements were assigned to their large extension of valence orbitals, which increases atomic orbital overlap. These results are valuable for the designation of auxetic nanomaterials and for expanding their potential applications in the fields of flexible electronics, biomedicine, protection devices, aerospace, nanosensors, et al.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4na00966e |
This journal is © The Royal Society of Chemistry 2025 |