Next Article in Journal
Novel Sol-Gel Synthesis of TiO2 Spherical Porous Nanoparticles Assemblies with Photocatalytic Activity
Previous Article in Journal
3D Magnonic Conduits by Direct Write Nanofabrication
Previous Article in Special Issue
First-Principles Study on Mechanical, Electronic, and Magnetic Properties of Room Temperature Ferromagnetic Half-Metal MnNCl Monolayer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principles Study of B16N16 Cluster-Assembled Porous Nanomaterials

School of Physical Science and Technology, Inner Mongolia University, Hohhot 010021, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(13), 1927; https://doi.org/10.3390/nano13131927
Submission received: 30 May 2023 / Revised: 21 June 2023 / Accepted: 22 June 2023 / Published: 24 June 2023
(This article belongs to the Special Issue First-Principles Investigations of Low-Dimensional Nanomaterials)

Abstract

:
Owing to the similar valence electron structures between the B-N bond and the C-C bond, boron nitride, similar to carbon, can form abundant polymorphs with different frameworks, which possess rich mechanical and electronic properties. Using the hollow, cage-like B16N16 cluster as building blocks, here, we established three new BN polymorphs with low-density porous structures, termed Cub-B16N16, Tet-B16N16, and Ort-B16N16, which have cubic ( P 4 ¯ 3 m ), tetragonal (P4/nbm), and orthomorphic (Imma) symmetries, respectively. Our density functional theory (DFT) calculations indicated that the existence of porous structure Cub-B16N16, Tet-B16N16, and Ort-B16N16 were not only energetically, dynamically, thermally and mechanically stable, they were even more stable than some known phases, such as sc-B12N12 and Hp-BN. The obtained Pugh’s ratio showed that the Cub-B16N16 and Tet-B16N16 structures were brittle materials, but Ort-B16N16 was ductile. The analysis of ideal strength, Young’s moduli, and shear moduli revealed that the proposed new phases all exhibited sizable mechanical anisotropy. Additionally, the calculation of electronic band structures and density of states showed that they were all semiconducting with a wide, indirect band gap (~3 eV). The results obtained in this work not only identified three stable BN polymorphs, they also highlighted a bottom-up way to obtain the desired materials with the clusters serving as building blocks.

1. Introduction

The similar valance electronic configuration between the B-N bond and the C-C bond makes boron nitrides have many polymorphs, similar to carbon, including zero-dimensional (0D) clusters, one-dimensional (1D) nanotubes and nanoribbons, two-dimensional (2D) nanosheets, and three-dimensional (3D) crystalline or amorphous BN [1,2,3,4,5]. As a new level of material structure, the study of clusters is helpful to understand the evolution law of matter from microscopic atoms and molecules to macroscopic condensed matter; it also provides an ideal platform to explore the novel physic phenomena of the 0D nano-system.
In recent years, BkNk cage clusters have attracted numerous attentions, due to their potential application in materials, energy, environment, and other fields [6,7,8,9,10,11,12,13,14,15,16,17]. From previous works, one can know that there are two main types of BkNk cage structures: one is constructed from alternating B-N bonds and consists entirely of four-membered and six-membered rings. The other is a fullerene-like structure based on combinations of pentagons and hexagons, with N-N and B-B bonds [17,18,19]. Oku et al. [6,7,8,9] synthesized and detected BkNk (k = 12, 24–60) nanocages by laser desorption time-of-flight mass spectrometry and found that the BN clusters consisting of 4-, 6- and 8-membered BN rings satisfied the isolated tetragonal rule, which was optimized by molecular orbital calculation. Stéphan et al. [10] presented experimental evidence for the formation of small BN cage-like molecules by an electron-irradiation experiment and observed that the diameters of the smallest and most observed cages were from 0.4 to 0.7 nm, close to those of octahedron-like structures of B12N12, B16N16, and B28N28 cages, which were predicted to be stable magic number clusters by electronic structural calculations [11]. These studies provided experimental evidence for the stable existence of boron nitride clusters.
In fact, early theoretical studies proved that the hollow cage clusters (e.g., B12N12 [12], B16N16 [13,14], B24N24 [15], and B36N36 [16]) are quite stable. For the specific atomic structure, Strout [17] compared the two classes of boron nitrides: fullerene-like structures consisting of pentagons and hexagons and alternant structures consisting of squares and hexagons. These two classes were compared for B13N13, B14N14, and B16N16 by theoretical calculations using the Hartree–Fock theory and the density functional theory (B3LYP and LDA). The major result was that the alternative structures were more stable than the fullerene-based cage structure.
Over the past few decades, the bottom-up approach, with stable clusters as building blocks, has been considered a promising way to design new materials with desired properties [20,21]. Xiong et al. [22] studied the stabilities and electronic structures of two boron nitride crystals, Pm3n B12N12 and sc-B12N12, assembled from the experimental synthesized B12N12 cluster with an alternative structure by first-principles calculations. They found that the two structures were stable, and both of them were wide-band gap insulators. In our recent previous work, we established a new sp3-hybridized BN allotrope sc-B24N24, based on the cage-like alternative B24N24 cluster, which was energetically, dynamically, and mechanically stable. The analysis of the electronic and optical properties showed that sc-B24N24 was a semiconductor. Remarkably, if the sc-B24N24 framework was taken as the host for endohedral doping of magnetic impurities, a desirable magnetic material was obtained, which exhibited a ferromagnetic (FM) half-metallic ground state with complete spin polarization [5]. In all these reported BN cluster-assembled phases, the clusters could maintain their structural characteristics when interacting with neighboring clusters. This means that the clusters B12N12 and B24B24 can be served as stable assembly motifs to construct new nanoscale materials using a bottom-up way.
In 2002, Alexandre et al. [23], taking the stable stoichiometric B16N16 [14] as building blocks, proposed that B16N16 could form covalent-bound, low-density, cluster-assembled solids with large interstitial channels. However, they only considered one B16N16-assembled phase. Were the other assembled structures also stable, similar to that of the Zn16O16 cluster [24]? If so, what about their mechanical and electronic properties?
Inspired by these questions, in this work, based on the density functional theory (DFT), we selected the stable B16N16 nanocluster as the building block for constructing new possible stable structures. Our results showed that three new porous nanostructures named Cub-B16N16, Tet-B16N16, and Ort-B16N16 were mechanically, dynamically, and thermally stable. On the basis of this, we further explored their structural, mechanical, and electronic properties. As for mechanical properties, the bulk elastic modulus (B), shear modulus (G), Young’s modulus (Y), Pugh’s ratio (K = B/G), and elastic anisotropy index (AU) were calculated. The results of Pugh’s ratio suggested that Cub-B16N16 and Tet-B16N16 were brittle materials, but Ort-B16N16 was ductile. The obtained elastic anisotropy indices indicated that the proposed phases were all anisotropic, and among them, Ort-B16N16 had the highest anisotropy. Moreover, the calculations of electronic band structure and densities of states revealed that these assembled phases were all indirect band gap semiconductors.

2. Computational Methods

The local structure optimization and electronic properties calculations of cluster assembly materials were performed using the DFT method in the Vienna ab initio package (VASP) [25]. The Perdew–Burke–Ernzerhof (PBE) form of the generalized gradient functional (GGA) was used to solve the exchange correlation energy [26,27]. The plane-wave base vector was based on the projector-augmented wave (PAW) method [28]. The Grimme method was adopted to correct the Van der Waals interaction [29]. A cutoff energy of 500 eV was used for the plane-wave basis, and the energy and force constant convergences were set to 10−6 eV and 10−3 eV Å−1, respectively. The Monkhorst-pack k-point mesh with uniform spacing was adopted in the Brillouin zones, which were 7 × 7 × 7 for the Cub-B16N16 structure, 6 × 6 × 7 for the Tet-B16N16 structure, and 7 × 6 × 7 for the Ort-B16N16 structure. The phonon dispersion spectra and phonon state densities of the systems were calculated by the density-functional perturbation theory implemented in the PHONOPY package [30] to evaluate the dynamic stability. In addition, molecular dynamics simulations using NVT canonical ensemble at a 300 K temperature were performed to investigate the initial decomposition mechanism and thermal stability of the supercells. The time step was set to 1 fs, and the total simulation time was 10 ps. The elastic constant Cij, the bulk modulus (B, the average of BV, and BR), and shear modulus (G, the average of GV, and GR) of the new structures were calculated according to the Voigt–Reuss–Hill (VRH) approximation [31]. Elastic constants were defined by means of the stress–strain method [32,33].

3. Results and Discussion

3.1. Structural Properties

B16N16 cage cluster, a stable cluster with magic number characteristics [10], is an octahedral structure, whose basic units are four-membered rings and six-membered rings, among which six four-membered rings are independently separated by twelve six-membered rings [23]. Based on their structural properties and energy stability, we made it the building blocks and considered three possible cluster–cluster interactions, i.e., six-membered ring facing six-membered ring (H), four-membered ring facing four-membered ring (C), and B-N edge-to-edge (S) connection modes. Then, three new periodic 3D solids were obtained and named Cub-B16N16, Tet-B16N16, and Ort-B16N16, according to their crystallographic systems and primitive names. Their optimized atomic structures, including the coordination polymerization mode of each building block and the primitive cells of every phase, are displayed in Figure 1. The Cub-B16N16, Tet-B16N16, and Ort-B16N16 structures had P 4 ¯ 3m, P4/nbm, and Imma symmetries, respectively.
From Figure 1, one can see that the B16N16 hollow-cage cluster structures were very stable, with one B atom and four N atoms forming sp3 hybridization, maintaining its structural integrity in the three assembled new crystalline phases and showing an excellent “element” role. The structural optimization parameters (space groups, lattice constants, unit atomic volumes, and equilibrium densities) of Cub-B16N16, Tet-B16N16, and Ort-B16N16 and several considered structures (e.g., c-BN, d-BN, Hp-BN, Pm3n-BN, and sc-B12N12) are listed in Table 1. For Cub-B16N16, Tet-B16N16, and Ort-B16N16 structures, the average bond lengths were 1.519 Å, 1.530 Å, and 1.544 Å, respectively; the average bond angles of the four-membered rings were 93.6°, 89.8°, and 90.0°; and the six-membered rings were 119.2°, 117.1°, and 112.8°, respectively. One can see that the mass densities of Cub-B16N16, Tet-B16N16, and Ort-B16N16 were 2.124 g/cm3, 2.379 g/cm3, and 2.163 g/cm3, respectively, which were much lower than that of c-BN (3.472 g/cm3) [34], Hp-BN (3.633 g/cm3) [35], and Pm3n-BN (2.849 g/cm3) [22], due to the existence of hollow holes. The optimized, non-equivalent atomic coordinates of Cub-B16N16, Tet-B16N16, and Ort-B16N16 structures are shown in Table S1. Due to the characteristics of low density and nanopores, the proposed B16N16-assembled phases might be promising for future applications in heterogeneous catalysis, molecular transport, and other fields [36,37].

3.2. Stabilities

Were the Cub-B16N16, Tet-B16N16, Ort-B16N16 polymorphs stable? To explore this, we evaluated their energies and their mechanical, dynamic, and thermal stabilities. Firstly, the total energies of the three assembled materials as functions of volume at a temperature of zero were calculated to determine the energy stability. For comparison, five related boron nitride isomerized materials, including c-BN, d-BN, Hp-BN, Pm3n-BN, and sc-B12N12 polymorphs, were also considered, as shown in Figure 2a. The results showed that although the equilibrium total energies of these three materials were higher than that of c-BN and d-BN, they were energetically more stable than that of Hp-BN and sc-B12N12, based on the results, by fitting the third-order Birch–Murnaghan equation of state (EOS) [42]. Among the assembled phases, the Ort-B16N16 phase had the lowest energy, implying that it was the most stable phase.
The parameters of enthalpy and pressure were obtained from the equation H = Et + PV. The enthalpy pressure relationships of cluster-assembled crystal phases Cub-B16N16, Tet-B16N16, and Ort-B16N16 and a series of boron nitride isomeric phases within the range of 0~10 GPa are shown in Figure 2b to confirm the stabilities of three assembled materials, with respect to the five synthesized phases (e.g., c-BN, d-BN, Hp-BN, Pm3n-BN, and sc-B12N12) in different ranges of pressure. A more stable phase will generally have a lower enthalpy for a given pressure. It can be seen from Figure 2b that Cub-B16N16, Tet-B16N16, and Ort-B16N16 were all more stable than sc-B12N12 and Hp-BN in the whole considered ranges of pressure, indicating their good mechanical stabilities with respect to sc-B12N12 and Hp-BN.
To assess the dynamical stability of the proposed phases, we further calculated the phonon dispersion and the corresponding phonon density of states along highly symmetric paths throughout the Brillouin zone, as shown in Figure S1. The numbers of atoms per unit cell of Cub-B16N16, Tet-B16N16, and Ort-B16N16 were 32, 64 and 64, respectively, indicating that there were 96, 192, and 192 branches of the dispersion spectrum. Since there were no imaginary frequencies in the dispersion spectrum, the proposed Cub-B16N16, Tet-B16N16, and Ort-B16N16 should be dynamically stable at T = 0 K. Moreover, the densities of phonon states of the Cub-B16N16, Tet-B16N16, and Ort-B16N16 phases illustrated that the vibrational modes in the low-frequency region were mainly contributed by N atoms, while those in the higher-frequency region were mainly contributed by B atoms, due to its relatively smaller atomic mass.
However, the above discussion could not guarantee the stabilities of the three phases at elevated temperatures. In this regard, further exploration of the thermal stabilities of Cub-B16N16, Tet-B16N16, and Ort-B16N16 at room temperature was necessary. By building a 2 × 2 × 1 supercell with 128 atoms for the Cub-B16N16 structure, 256 atoms for the Tet-B16N16 structure, and a 2 × 1 × 1 supercell with 128 atoms for the Ort-B16N16 structure, we performed the ab initio molecular dynamics (AIMD) simulations at 300 K with a Nosé–Hoover thermostat. Figure S2 shows the potential energy and temperature fluctuations of the three systems as a function of simulation times. Throughout the simulation, the potential energy was almost constant, with small variations due to thermal fluctuations for all the assembled crystal phases. Correspondingly, the structures maintained their original structures without damage. Therefore, the simulation results confirmed that all the assembled structures were thermally stable and could survive, at least at room temperature.
To examine the mechanical stability of the assembled phases, we further calculated their elastic constants. The results are listed in Table 2. According to Born’s mechanical stability criterion, a material with mechanical stability should follow the related mechanical stability criteria:
For the cubic system,
C 11 > 0 , C 44 > 0 , C 11 C 12 > 0 , C 11 + 2 C 12 > 0 .
For the tetragonal system,
C 11 > 0 , C 33 > 0 , C 44 > 0 , C 66 > 0 , C 11 C 12 > 0 , C 11 + C 33 2 C 13 > 0 , 2 ( C 11 + C 12 ) + C 33 + 4 C 13 > 0 .
For the orthorhombic system,
C 11 > 0 , C 22 > 0 , C 33 > 0 , C 44 > 0 , C 55 > 0 , C 66 > 0 , C 11 + C 22 2 C 12 > 0 , C 11 + C 33 2 C 13 > 0 , C 22 + C 33 2 C 23 > 0 , C 11 + C 22 + C 33 + 2 ( C 12 + C 13 + C 23 ) > 0 .
It can be seen that all of the Cij for Cub-B16N16, Tet-B16N16, and Ort-B16N16 met the Born stability criterion [43]; this means that they were all mechanically stable.
Electron localization function (ELF) was an effective method for analyzing the types of chemical bonds, which can accurately characterize the distribution characteristics of electron delocalization in both molecules and solids [45]. Values of 1.00 and 0.50 indicated complete localization and delocalization of electrons, respectively, while 0.00 indicated very low electron density. The electronic local functions of Cub-B16N16, Tet-B16N16, and Ort-B16N16 along the four-numbered ring in the [001] direction was calculated to investigate the local characteristics of the assembled materials. Figure 3 illustrates a 2D contour map of ELF along the [001] direction of Cub-B16N16, Tet-B16N16, and Ort-B16N16. It can be seen that the ELF value in the middle of the B-N bond was close to 1.0, implying that the electrons were highly localized in this region. In other words, the B-N bonds in the assembled materials had strong covalent properties. This should be the reason that Cub-B16N16, Tet-B16N16, and Ort-B16N16 had superb energy and dynamical, thermal, and mechanical stability. Near the N and B atoms, the ELF values were about 0.5 and 0.25, respectively. This means that electrons were more likely to be localized around the N atom, while the densities of electrons near the B atom were very low.

3.3. Mechanical Properties

As is known, many reported polymorphs of boron nitride have excellent mechanical properties. This naturally poses a question: can the new low-density porous boron nitride polymorphs preserve their intrinsic configuration under external stress? Using the VRH method [31], we calculated the bulk elastic modulus (B), shear modulus (G), Young’s modulus (Y), and Pugh’s ratio (K = B/G) of the three low-density assembled phases on the basis of the obtained elastic constants. The data are listed in Table 2. The moduli of Cub-B16N16, Tet-B16N16, and Ort-B16N16 were relatively lower, compared to that of the super-hard boron nitride c-BN. However, considering their low densities, they still had good elastic properties. To prove this, we further calculated the Vickers hardness (Hv) using the empirical formula [37]:
H ν = 2 ( G 2 B 2 ) 0.585 3
The calculated results of Vickers hardness are shown in Table 2. One can see that the hardnesses of Cub-B16N16, Tet-B16N16, and Ort-B16N16 were less than 40 GPa, which was the critical value of Vickers hardness to distinguish super-hard materials from ordinary materials. Although they were not super-hard materials, Cub-B16N16, Tet-B16N16, and Ort-B16N16 were still hard, even superior to some metal nitrides and carbides, such as TiC, TiN, and WC [46]. In addition, the ratio of the bulk elastic modulus to the shear modulus (i.e., B/G) is often used to distinguish ductile and brittle materials. If the B/G ratio is greater than 1.75, the material is ductile; otherwise, it is brittle. The B/G ratios of Cub-B16N16, Tet-B16N16, and Ort-B16N16 were calculated to be 1.61, 1.60, and 2.07, respectively. There was no doubt that Cub-B16N16 and Tet-B16N16 were brittle materials, but Ort-B16N16 was ductile.
Figure 4 shows the ideal tensile strength as a function of strain for Cub-B16N16, Tet-B16N16, and Ort-B16N16 along the three directions of [100], [110], and [111]. It can be seen that Cub-B16N16, Tet-B16N16, and Ort-B16N16 could withstand a certain tensile strain before the structure broke. Specifically, along the [100], [110], and [111] directions, the ideal tensile strength and the corresponding maximum strain withstood at the fracture point of the three structures are shown in Table 3.
Moreover, from Figure 4, one can conclude that Cub-B16N16, Tet-B16N16, and Ort-B16N16 were mechanically anisotropic. Microcracks and lattice deformations of materials are important factors for reflecting the elastic anisotropy, which also plays a key role in enhancing the mechanical durability of materials. In order to characterize the anisotropy degree of materials, Ranganathan proposed a universal elastic anisotropy index AU for each crystal phase, based on the mean values of Reuss and Voigt [47]:
A U = 5 G V G R + B V B R 6
where GV, GR, BV, and BR are the shear moduli and bulk moduli of Voigt and Reuss approximations, respectively.
The material is isotropic when the value of AU is zero; otherwise, it is anisotropic. The degree of anisotropy of the material can be reflected by the magnitude of the AU deviation from zero. The greater the deviation of AU, the stronger the anisotropy. We calculated the universal elastic anisotropy indices, AU, of Cub-B16N16, Tet-B16N16, and Ort-B16N16 to provide an effective perspective on mechanical anisotropy. The results showed that the anisotropy indices of Young’s moduli for Cub-B16N16, Tet-B16N16, and Ort-B16N16 were 1.30, 1.32, and 4.82, respectively, and the anisotropy indices of shear moduli were 1.37, 1.39, and 4.62, respectively. Therefore, Cub-B16N16, Tet-B16N16, and Ort-B16N16 were all anisotropic. Among the proposed phases, Ort-B16N16 had the strongest anisotropy, due to its largest AU. Figure 5 shows the three-dimensional diagrams of Young’s moduli and shear moduli, which can provide more intuitive physical images for the anisotropic elastic characteristics of the three assembly structures.

3.4. Electronic Properties

The electronic band structures along the high-symmetry k-points were calculated, and the Monkhorst-pack k-point meshes increased to 15 × 15 × 15 for the Cub-B16N16 structure, 11 × 11 × 15 for the Tet-B16N16 structure, and 12 × 15 × 14 for the Ort-B16N16 structure, as shown in Figure 6. Similar to most of the BN polymorphs, one can see that Cub-B16N16, Tet-B16N16, and Ort-B16N16 were semiconducting with wide energy band gaps, which were 2.94, 2.80, and 3.34 eV, respectively. Since the valence band maximum (VBM) and conduction band minimum (CBM) were not at the same high-symmetry k-points for Cub-B16N16 and Ort-B16N16 structures, they belonged to indirect band gap semiconductors. However, the Tet-B16N16 structure was a direct bandgap semiconductor, since VBM and CBM were at the same high-symmetry k-points. As is usual for PBE calculations, the absolute band gaps were systematically underestimated, whereas the relative magnitudes provided a gauge of the electronic change with respect to the change in bulk topology [48]. We calculated the relative band gap magnitude of each structure to the most stable c-BN (Eg/Eg,c), which is shown in Table 1.
The total and partial densities of states (DOS) of Cub-B16N16, Tet-B16N16, and Ort-B16N16 are shown in Figure 7. There was obvious hybridization between the N atomic orbital and the B orbital, due to the fact that the B-N bonds of the assembled crystal phases were strong covalent bonds. According to the partial densities of states, the densities of states of the VBM and CBM were dominated by B-2p and N-2p orbitals, respectively. In the vicinity of the Fermi level, the states of the valence were mainly contributed by the N-2p orbital, and the states of the conduction band mainly came from the B-2p orbital.

4. Conclusions

In summary, based on the bottom-up approach, we predicted three new low-density boron nitride polymorphs, Cub-B16N16, Tet-B16N16. and Ort-B16N16, which can be considered three-dimensional structures assembled from B16N16 cage clusters. Based on the density functional theory modified by Van der Waals, the following interesting features of Cub-B16N16, Tet-B16N16, and Ort-B16N16 were characterized: (i) they were low-density (2.124 g/cm3, 2.379 g/cm3, and 2.163 g/cm3, respectively) porous materials, due to the existence of boron nitride hollow cages B16N16; (ii) Cub-B16N16, Tet-B16N16, and Ort-B16N16 exhibited good energy and dynamic, thermal, mechanical, and chemical stability, due to the strong covalence interaction between B and N atoms, which was proven by our electron localization function analysis; (iii) the results of elastic properties calculations showed that Cub-B16N16 and Tet-B16N16 were brittle materials, but Ort-B16N16 was ductile; the Young’s moduli and shear moduli of the three assembled materials harbored strong anisotropy; (iv) the electronic band structure showed that the three assembled crystal phases were all indirect wide-band gap semiconductors. Our results not only highlighted some novel low-density boron nitride polymorphs, they also provided a bottom-up way to design new solid materials by using the clusters as building blocks.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13131927/s1, Table S1: The optimized non-equivalent atomic coordinates of Cub-B16N16, Tet-B16N16, and Ort-B16N16 structures. Figure S1: Phonon band structures and phonon density of states of Cub-B16N16, Tet-B16N16, and Ort-B16N16. Figure S2: The fluctuation of potential energy and temperature of Cub-B16N16, Tet-B16N16, and Ort-B16N16 as a function of molecular dynamics simulation time at room temperature. Reference [49] is cited in the Supplementary Materials.

Author Contributions

Conceptualization, X.C.; methodology, T.S.; software, T.S.; validation, Z.L. and X.C.; formal analysis, X.W. and X.Z.; investigation, X.W. and X.Z.; resources, X.C.; data curation, X.W. and L.L.; writing—original draft preparation, X.W.; writing—review and editing, X.C., Z.L. and X.Z.; visualization, X.W.; supervision, X.C.; project administration, X.C.; funding acquisition, X.C. and Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (12264033, 11964023) and the Natural Science Foundation of Inner Mongolia Autonomous Region (No. 2021JQ-001).

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Berisha, A. Unraveling the electronic influence and nature of covalent bonding of aryl and alkyl radicals on the B12N12 nanocage cluster. Sci. Rep. 2023, 13, 752. [Google Scholar] [CrossRef] [PubMed]
  2. Xiang, H.J.; Yang, J.; Hou, J.G.; Zhu, Q. First-principles study of small-radius single-walled BN nanotubes. Phys. Rev. B 2003, 68, 035427. [Google Scholar] [CrossRef]
  3. Hao, J.; Li, L.; Gao, P.; Jiang, X.; Ban, C.; Shi, N. Boron nitride nanoribbons grown by chemical vapor deposition for VUV applications. Micromachines 2022, 13, 1372. [Google Scholar] [CrossRef] [PubMed]
  4. Li, L.H.; Cervenka, J.; Watanabe, K.; Taniguchi, T.; Chen, Y. Strong oxidation resistance of atomically thin boron nitride nanosheets. ACS Nano 2014, 8, 1457–1462. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, X.; Liu, Z.; Song, T.; Cui, X. Sc-B24N24: A new low-density allotrope of BN. J. Phys. Chem. C 2022, 126, 12836–12844. [Google Scholar] [CrossRef]
  6. Oku, T.; Kuno, M.; Kitahara, H.; Narita, I. Formation, atomic structures and properties of boron nitride and carbon nanocage fullerene materials. Int. J. Inorg. Mater. 2001, 3, 597–612. [Google Scholar] [CrossRef]
  7. Oku, T.; Nishiwaki, A.; Narita, I.; Gonda, M. Formation and structure of B24N24 clusters. Chem. Phys. Lett. 2003, 380, 620–623. [Google Scholar] [CrossRef]
  8. Oku, T.; Nishiwaki, A.; Narita, I. Formation and atomic structure of B12N12 nanocage clusters studied by mass spectrometry and cluster calculation. Sci. Technol. Adv. Mater. 2004, 5, 635–638. [Google Scholar] [CrossRef]
  9. Oku, T.; Nishiwaki, A.; Narita, I. Formation and atomic structures of BnNn (n = 24–60) clusters studied by mass spectrometry, high-resolution electron microscopy and molecular orbital calculations. Phys. B 2004, 351, 184–190. [Google Scholar] [CrossRef]
  10. Stéphan, O.; Bando, Y.; Loiseau, A.; Willaime, F.; Shramchenko, N.; Tamiya, T.; Sato, T. Formation of small single-layer and nested BN cages under electron irradiation of nanotubes and bulk material. Appl. Phys. A 1998, 67, 107–111. [Google Scholar] [CrossRef]
  11. Seifert, G.; Fowler, P.W.; Mitchell, D.; Porezag, D.; Frauenheim, T. Boron-nitrogen analogues of the fullerenes: Electronic and structural properties. Chem. Phys. Lett. 1997, 268, 352. [Google Scholar] [CrossRef]
  12. Li, J.L.; He, T.; Yang, G.W. An all-purpose building block: B12N12 fullerene. Nanoscale 2012, 4, 1665–1670. [Google Scholar] [CrossRef]
  13. Yin, B.; Wang, G.; Sa, N.; Huang, Y. Bonding analysis and stability on alternant B16N16 cage and its dimers. J. Mol. Model. 2008, 14, 789–795. [Google Scholar] [CrossRef]
  14. Alexandre, S.S.; Chacham, H.; Nunes, R.W. Structure and energetics of boron nitride fullerenes: The role of stoichiometry. Phys. Rev. B 2001, 63, 045402. [Google Scholar] [CrossRef]
  15. Lan, Y.Z.; Cheng, W.D.; Wu, D.S.; Li, X.D.; Zhang, H.; Gong, Y.J.; Shen, J.; Li, F.F. Theoretical studies of third-order nonlinear optical response for B12N12, B24N24 and B36N36 clusters. J. Mol. Struc.-Theochem. 2005, 730, 9–15. [Google Scholar] [CrossRef]
  16. Wu, H.S.; Xu, X.H.; Strout, D.L.; Jiao, H. The structure and stability of B36N36 cages: A computational study. J. Mol. Model. 2006, 12, 1–8. [Google Scholar] [CrossRef]
  17. Strout, D.L. Fullerene-like cages versus alternant cages: Isomer stability of B13N13, B14N14, and B16N16. Chem. Phys. Lett. 2004, 383, 95–98. [Google Scholar] [CrossRef]
  18. Rogers, K.M.; Fowler, P.W.; Seifert, G. Chemical versus steric frustration in boron nitride heterofullerene polyhedral. Chem. Phys. Lett. 2000, 332, 43–50. [Google Scholar] [CrossRef]
  19. Xia, X.; Jelski, D.A.; Bowser, J.R.; George, T.F. MNDO Study of Boron-Nitrogen Analogues of Buckminsterfullerene. J. Am. Chem. Soc. 1992, 114, 6493–6496. [Google Scholar] [CrossRef]
  20. Khanna, S.N.; Jena, P. Assembling Crystals from Clusters. Phys. Rev. Lett. 1992, 69, 1664–1667. [Google Scholar] [CrossRef]
  21. Claridge, S.A.; Castleman, A.W., Jr.; Khanna, S.N.; Murray, C.B.; Sen, A.; Weiss, P.S. Cluster-Assembled Materials. ACS Nano 2009, 3, 244–255. [Google Scholar] [CrossRef] [PubMed]
  22. Xiong, C.; Shi, J.; Zhou, A.; Cai, Y. A comparative investigation of sp3-hybridized Pm3n-BN and sc-B12N12 based on density functional theory (DFT). Mater. Today Commun. 2020, 25, 101582. [Google Scholar] [CrossRef]
  23. Alexandre, S.S.; Nunes, R.W.; Chacham, H. Energetics of the formation of dimers and solids of boron nitride fullerenes. Phys. Rev. B 2002, 66, 085406. [Google Scholar] [CrossRef]
  24. Liu, Z.; Wang, X.; Cai, J.; Liu, G.; Zhou, P.; Wang, K.; Zhu, H. From the ZnO Hollow Cage Clusters to ZnO Nanoporous Phases: A First-Principles Bottom-Up Prediction. J. Phys. Chem. C 2013, 117, 17633–17643. [Google Scholar] [CrossRef]
  25. Kresse, G.; Furthmüller, J. Effificient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef]
  26. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  27. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671–6687. [Google Scholar] [CrossRef]
  28. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [Green Version]
  29. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  30. Togo, A.; Tanaka, I. First principles phonon calculations in materials science. Scr. Mater. 2015, 108, 1–5. [Google Scholar] [CrossRef] [Green Version]
  31. Hill, R. The Elastic Behaviour of a Crystalline Aggregate. Proc. Phys. Soc. A 1952, 65, 349–354. [Google Scholar] [CrossRef]
  32. Zhao, J.J.; Winey, J.M.; Gupta, Y.M. First-principles calculations of second- and third-order elastic constants for single crystals of arbitrary symmetry. Phys. Rev. B 2007, 75, 094105. [Google Scholar] [CrossRef]
  33. Ravindran, P.; Fast, L.; Korzhavyi, P.A.; Johansson, B.; Wills, J.; Eriksson, O. Density functional theory for calculation of elastic properties of orthorhombic crystals: Application to TiSi2. J. Appl. Phys. 1998, 84, 4891–4904. [Google Scholar] [CrossRef]
  34. Brugger, K. Thermodynamic Definition of Higher Order Elastic Coefficients. Phys. Rev. 1964, 133, A1611–A1612. [Google Scholar] [CrossRef]
  35. Xu, S.; Wang, L.; Qiao, X.; Xu, X.; Cai, Y. Novel BN polymorphs transformed from the smallest BNNTs under high pressure. Comput. Mater. Sci. 2015, 110, 241–246. [Google Scholar] [CrossRef]
  36. Yu, J.; Qu, L.; Veen, E.; Katsnelson, M.I.; Yuan, S. Hyperhoneycomb boron nitride with anisotropic mechanical, electronic, and optical properties. Phys. Rev. Mater. 2017, 1, 045001. [Google Scholar] [CrossRef] [Green Version]
  37. Cai, Y.; Zeng, L.; Zhang, Y.; Xu, X. Multiporous sp2-hybridized boron nitride (d-BN): Stability, mechanical properties, lattice thermal conductivity and promising application in energy storage. Phys. Chem. Chem. Phys. 2018, 20, 20726–20731. [Google Scholar] [CrossRef]
  38. Tian, Y.; Xu, B.; Zhao, Z. Microscopic theory of hardness and design of novel superhard crystals. In. J. Refract. Met. Hard Mater. 2012, 33, 93–106. [Google Scholar] [CrossRef]
  39. Ohba, N.; Miwa, K.; Nagasako, N.; Fukumoto, A. First-principles study on structural, dielectric, and dynamical properties for three BN polytypes. Phys. Rev. B 2001, 63, 115207. [Google Scholar] [CrossRef]
  40. Peng, Q.; Ji, W.; De, S. Mechanical properties of the hexagonal boron nitride monolayer: Ab initio study. Comp. Mater. Sci. 2012, 56, 11–17. [Google Scholar] [CrossRef] [Green Version]
  41. Nag, A.; Raidongia, K.; Hembram, K.P.S.S.; Datta, R.; Waghmare, U.V.; Rao, C.N.R. Graphene analogues of BN: Novel synthesis and properties. ACS Nano 2010, 4, 1539–1544. [Google Scholar] [CrossRef] [PubMed]
  42. Soma, T.; Sawaoka, S.; Saito, S. Characterization of wurtzite type boron nitride synthesized by shock compression. Mater. Res. Bull. 1974, 9, 755–762. [Google Scholar] [CrossRef]
  43. Mouhat, F.; Coudert, F.X. Necessary and sufficient elastic stability conditions in various crystal systems. Phys. Rev. B 2014, 90, 224104. [Google Scholar] [CrossRef] [Green Version]
  44. Wu, Z.J.; Zhao, E.J.; Xiang, H.P.; Hao, X.F.; Liu, X.J.; Meng, J. Crystal structures and elastic properties of superhard IrN2 and IN3 from first principles. Phys. Rev. B 2007, 76, 054115. [Google Scholar] [CrossRef]
  45. He, C.; Sun, L.; Zhang, C.; Peng, X.; Zhang, K.; Zhong, J. Z-BN: A novel superhard boron nitride phase. Phys. Chem. Chem. Phys. 2012, 14, 10967–10971. [Google Scholar] [CrossRef]
  46. Silvi, B.; Savin, A. Classification of chemical bonds based on topological analysis of electron localization functions. Nature 1994, 371, 683–686. [Google Scholar] [CrossRef]
  47. Šimůnek, A.; Vackář, J. Hardness of covalent and ionic crystals: First-principle calculations. Phys. Rev. Lett. 2006, 96, 085501. [Google Scholar] [CrossRef] [Green Version]
  48. Ranganathan, S.I.; Ostoja-Starzewski, M. Universal elastic anisotropy index. Phys. Rev. Lett. 2008, 101, 055504. [Google Scholar] [CrossRef] [Green Version]
  49. Carrasco, J.; Illas, F.; Bromley, S.T. Ultralow-density nanocage-based metal-oxide polymorphs. Phys. Rev. Lett. 2007, 99, 235502. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The coordination polymerization mode of each B16N16 cage in three boron nitride cluster-assembled materials and the primitive cell of the three considered crystal phases. (Cub-B16N16, Tet-B16N16, and Ort-B16N16 were named according to the abbreviation of the crystal system and the name of the construction primitive.)
Figure 1. The coordination polymerization mode of each B16N16 cage in three boron nitride cluster-assembled materials and the primitive cell of the three considered crystal phases. (Cub-B16N16, Tet-B16N16, and Ort-B16N16 were named according to the abbreviation of the crystal system and the name of the construction primitive.)
Nanomaterials 13 01927 g001
Figure 2. (a) The relation curve of total energy as a function of volume per atom for different BN polymorphs. (b) The enthalpies varied with pressure per atom for BN cluster-assembled phases, compared with a few BN polymorphs.
Figure 2. (a) The relation curve of total energy as a function of volume per atom for different BN polymorphs. (b) The enthalpies varied with pressure per atom for BN cluster-assembled phases, compared with a few BN polymorphs.
Nanomaterials 13 01927 g002
Figure 3. The 2D contour plots of the electron localization functions (ELF) of (a) Cub-B16N16, (b) Tet-B16N16, and (c) Ort-B16N16 along the [001] direction. The reference bar for the ELF value is provided on the right.
Figure 3. The 2D contour plots of the electron localization functions (ELF) of (a) Cub-B16N16, (b) Tet-B16N16, and (c) Ort-B16N16 along the [001] direction. The reference bar for the ELF value is provided on the right.
Nanomaterials 13 01927 g003
Figure 4. Ideal tensile strength along the [100], [110], and [111] directions for (a) Cub-B16N16, (b) Tet-B16N16, and (c) Ort-B16N16.
Figure 4. Ideal tensile strength along the [100], [110], and [111] directions for (a) Cub-B16N16, (b) Tet-B16N16, and (c) Ort-B16N16.
Nanomaterials 13 01927 g004
Figure 5. The directional dependence of Young’s modulus (Y) for (a) Cub-B16N16, (c) Tet-B16N16, and (e) Ort-B16N16 and shear modulus (G) for (b) Cub-B16N16, (d) Tet-B16N16, and (f) Ort-B16N16.
Figure 5. The directional dependence of Young’s modulus (Y) for (a) Cub-B16N16, (c) Tet-B16N16, and (e) Ort-B16N16 and shear modulus (G) for (b) Cub-B16N16, (d) Tet-B16N16, and (f) Ort-B16N16.
Nanomaterials 13 01927 g005
Figure 6. Graphs (ac) are the electronic band structures of Cub-B16N16, Tet-B16N16, and Ort-B16N16, respectively.
Figure 6. Graphs (ac) are the electronic band structures of Cub-B16N16, Tet-B16N16, and Ort-B16N16, respectively.
Nanomaterials 13 01927 g006
Figure 7. Total and partial DOS for Cub-B16N16, Tet-B16N16, and Ort-B16N16 at equilibrium structure.
Figure 7. Total and partial DOS for Cub-B16N16, Tet-B16N16, and Ort-B16N16 at equilibrium structure.
Nanomaterials 13 01927 g007
Table 1. The space group (SG), lattice parameters a, b, and c (Å), volume per atom V3/atom), equilibrium density ρ (g/cm3), cohesive energy per atom Etot (eV/atom), energy gap Eg (eV), and band gap relative to ground-state c-BN phase (Eg/Eg,c) for cluster-assembled BN phases. The corresponding data of several reported BN polymorphs (c-BN, 2D h-BN, w-BN, d-BN, Hp-BN, Pm3n-BN, and sc-B12N12) were also compared.
Table 1. The space group (SG), lattice parameters a, b, and c (Å), volume per atom V3/atom), equilibrium density ρ (g/cm3), cohesive energy per atom Etot (eV/atom), energy gap Eg (eV), and band gap relative to ground-state c-BN phase (Eg/Eg,c) for cluster-assembled BN phases. The corresponding data of several reported BN polymorphs (c-BN, 2D h-BN, w-BN, d-BN, Hp-BN, Pm3n-BN, and sc-B12N12) were also compared.
Structure SGa(Å)b(Å)c(Å)VρEtot(eV)Eg(eV)Eg/Eg,c
Cub-B16N16 P 4 ¯ 3 m 6.7726.7726.7729.772.124−8.382.940.659
Tet-B16N16 P4/nbm8.9918.9916.8578.892.379−8.372.800.628
Ort-B16N16 Imma8.97512.66910.7269.512.163−8.423.340.749
c-BNThis work F 4 ¯ 3 m 3.6153.6153.6155.903.489−8.864.461.00
Cal. [22] 3.6253.6253.625 −9.37
Expt. [38] 3.6153.6153.615 3.489 6.1~6.4
h-BN(2D)This workP63/mmc2.5062.506 −8.163.950.89
Cal. [39] 2.5122.512
Expt. [40] 2.4902.490
w-BNThis workP63mc2.5492.5494.2315.922.095−8.855.201.17
Cal. [22] 2.5552.5554.225 −9.35
Expt. [41] 2.5532.5534.228
d-BNThis work F d ¯ 3 c 12.29012.29012.2909.812.101−8.724.841.09
Cal. [36] 12.29212.29212.292 2.130 4.86
Hp-BNThis workP62222.6002.6005.8115.673.633−7.933.700.83
Cal. [35] 2.6102.6105.828 −7.783.45
Pm3n-BNThis work P m 3 ¯ n 4.4284.4284.4287.232.849−8.584.551.02
Cal. [22] 4.4184.4184.418 2.868−8.334.53
sc-B12N12This work F m 3 ¯ C 11.81911.81911.8198.752.345−8.334.981.12
Cal. [22] 11.81911.81911.819 2.396−8.205.02
Table 2. Calculated elastic constants Cij (GPa), bulk modulus B (GPa), shear modulus G (GPa), Pugh’s ratio K (B/G), Young’s modulus Y (GPa), and Vickers hardness Hv (GPa) of cluster-assembled BN phases and a few typical BN phases.
Table 2. Calculated elastic constants Cij (GPa), bulk modulus B (GPa), shear modulus G (GPa), Pugh’s ratio K (B/G), Young’s modulus Y (GPa), and Vickers hardness Hv (GPa) of cluster-assembled BN phases and a few typical BN phases.
StructureC11C12C13C22C23C33C44C55C66BGKYHv
Cub-B16N16304118 127 1801121.6127915.19
Tet-B16N1637684122 314123 1341971231.6030516.30
Ort-B16N162257354480112200762676141692.071788.17
c-BN797175 456 3823910.9887563.37
Cal. [44] 780173 444 376382 62.82
d-BN300175 120 216932.322437.38
Cal. [36] 2521112.17
Hp-BN873154 360 3843661.0483255.97
Cal. [35]892166 363 375
Pm3n-BN71290 195 2972351.2655833.74
Cal. [22]781116 218 337218~332
sc-B12N12452127 163 2321620.6939117.28
Cal. [22]483160 190 268162~190
Table 3. Calculated ideal tensile strengths and corresponding maximum tensile strains withstood at the fracture point.
Table 3. Calculated ideal tensile strengths and corresponding maximum tensile strains withstood at the fracture point.
StructureDirectionIdeal Tensile Strength (GPa)Maximum Strain (%)
Cub-B16N16[100]22.398
[110]14.296
[111]18.567
Tet-B16N16[100]25.0912
[110]18.478
[111]18.697
Ort-B16N16[100]12.127
[110]18.919
[111]12.457
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, X.; Zhang, X.; Liu, L.; Song, T.; Liu, Z.; Cui, X. First-Principles Study of B16N16 Cluster-Assembled Porous Nanomaterials. Nanomaterials 2023, 13, 1927. https://doi.org/10.3390/nano13131927

AMA Style

Wang X, Zhang X, Liu L, Song T, Liu Z, Cui X. First-Principles Study of B16N16 Cluster-Assembled Porous Nanomaterials. Nanomaterials. 2023; 13(13):1927. https://doi.org/10.3390/nano13131927

Chicago/Turabian Style

Wang, Xin, Xiaoyue Zhang, Liwei Liu, Tielei Song, Zhifeng Liu, and Xin Cui. 2023. "First-Principles Study of B16N16 Cluster-Assembled Porous Nanomaterials" Nanomaterials 13, no. 13: 1927. https://doi.org/10.3390/nano13131927

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