Next Article in Journal
Recent Progress of Squaraine-Based Fluorescent Materials and Their Biomedical Applications
Previous Article in Journal
Scattering Properties of an Acoustic Anti-Parity-Time-Symmetric System and Related Fabry–Perot Resonance Mode
Previous Article in Special Issue
QSPR Models for the Molar Refraction, Polarizability and Refractive Index of Aliphatic Carboxylic Acids Using the ZEP Topological Index
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural Stability and Electronic Properties of Boron Phosphide Nanotubes: A Density Functional Theory Perspective

by
Dolores García-Toral
1,*,
Raúl Mendoza-Báez
2,
Ernesto Chigo-Anota
1,
Antonio Flores-Riveros
3,
Víctor M. Vázquez-Báez
4,
Gregorio Hernández Cocoletzi
3 and
Juan Francisco Rivas-Silva
3
1
Facultad de Ingeniería Química, Benemérita Universidad Autónoma de Puebla, Av. San Claudio y 18 Sur S/N, San Manuel, Puebla 72570, Mexico
2
Departamento de Química, Centro de Investigación y de Estudios Avanzados del IPN (Cinvestav), Av. IPN 2508, Col. San Pedro Zacatenco, Mexico City 07360, Mexico
3
Instituto de Física, Benemérita Universidad Autónoma de Puebla, Av. San Claudio y Blvd. 18 Sur, Col. San Manuel, Puebla 72570, Mexico
4
Facultad de Ingeniería, Benemérita Universidad Autónoma de Puebla, Puebla 72570, Mexico
*
Author to whom correspondence should be addressed.
Symmetry 2022, 14(5), 964; https://doi.org/10.3390/sym14050964
Submission received: 12 April 2022 / Revised: 30 April 2022 / Accepted: 6 May 2022 / Published: 9 May 2022

Abstract

:
Based on the Density Functional Theory (DFT) calculations, we analyze the structural and electronic properties of boron phosphide nanotubes (BPNTs) as functions of chirality. The DFT calculations are performed using the M06-2X method in conjunction with the 6-31G(d) divided valence basis set. All nanostructures, (n,0) BPNT (n = 5–8, 10, 12, 14) and (n,n) BPNT (n = 3–11), were optimized minimizing the total energy, assuming a non-magnetic nature and a total charge neutrality. Results show that the BPNT diameter size increases linearly with the chiral index “n” for both chiralities. According to the global molecular descriptors, the (3,3) BPNT is the most stable structure provided that it shows the largest global hardness value. The low chirality (5,0) BPNT has a strong electrophilic character, and it is the most conductive system due to the small |HOMO-LUMO| energy gap. The chemical potential and electrophilicity index in the zigzag-type BPNTs show remarkable chirality-dependent behavior. The increase in diameter/chirality causes a gradual decrease in the |HOMO-LUMO| energy gap for the zigzag BPNTs; however, in the armchair-type BPNTs, a phase transition is generated from a semiconductor to a conductor system. Therefore, the nanostructures investigated in this work may be suggested for both electrical and biophysical applications.

1. Introduction

Since the discovery of the single-walled carbon nanotubes (SWCNT) in 1991 [1], numerous investigations have been conducted on their structural, chemical, physical, mechanical and electronic properties, which have proven to be extraordinary for certain applications. There exists an increased interest in fabricating novel inorganic analogues to CNTs made up of group III and V elements of the periodic table (groups neighboring the carbon group), such as nitrides and phosphides of boron, aluminum, gallium and indium (BN, AlN, GaN, InN, BP, AlP, GaP and InP, respectively). Among them, boron nitride nanotubes (BNNTs), as predicted in 1994 [2] and synthesized in 1995 [3], have been quite attractive and extensively investigated. Despite their structural similarity, CNT electronic properties show a strong dependence on the chirality [4,5,6,7], while the properties of BNNTs with large diameters remain almost constant [2,8], which in turn indicate chirality independence. The boron composed nanotubes have been more studied than those of phosphorus. Although BPNTs have not been synthesized yet, theoretical research during the last decade suggests that BPNTs may be good adsorbent materials for some molecules of environmental, industrial, pharmaceutical and catalytic interest, such as peroxide, phenol, imidazole, thiazole, carbon monoxide and even some anions such as cyanates and thiocyanates [9,10,11,12,13,14,15]. Studies reveal that the band gap of BPNTs is similar to that of their silicon carbide analogs (SiCNTs) [16,17]. Furthermore, BPNTs doped with carbon, germanium, gallium, silicon and palladium atoms show a drastic change in their electronic properties as compared to the pristine state [10,17,18,19,20,21,22,23]. However, not much information is available about investigations focused on the dependence of the structural and electronic properties of BP nanotubes as a function of their diameter/chirality. For example, Srivastava et al. [24] describe that the value of the band gap of the zigzag-type BPNTs (n,0) (n = 4–15) ranges from a metallic to a semiconductor behavior, while Azizi et al. [25] show values corresponding to semiconductor for both armchair- (n,n) and zigzag-type (n,0) BPNTs.
The interest of this work is to study the structural and electronic properties of armchair- (n,n) (n = 3–11) and zigzag-type (n,0) (n = 5–8,10,12,14) BPNTs as functions of chirality. The nanostructures were optimized and verified by calculating the vibrational frequencies. Based on the DFT, the reactivity and chemical stability of the BPNTs are studied through cohesion energy ( E c o h ) and global molecular descriptors such as chemical potential (μ), global hardness (η), electrophilicity index (ω), energy gap (Eg), ionization potential (I) and electronic affinity (A).

2. Materials and Methods

Boron phosphide nanotube structural and electronic properties are investigated using first principles total energy calculations within the Density Functional Theory. The BPNT dangling bonds were passivated with hydrogen atoms to carry out a finite molecular study, that is, each nanotube length was limited through covalent bonds with hydrogen atoms at their ends. The number of atoms of each chemical species (H, B and P) that make up each nanotube is proportional to the chirality index (n). In the case of the armchair-type (n,n) BPNTs, H = 4 n and B = P = 12 n , while for the zigzag-type (n,0) BPNTs, H = 2 n and B = P = 7 n . Calculations were performed to determine the geometric optimization, the energy of the frontier molecular orbitals, HOMO and LUMO, dipole moment and the total energy. Studies account for the gas phase within the M06-2X/6-31G(d) approach implemented in the Gaussian 16 code [26]. The M06-2X functional is suitable to be applied in medium-sized systems [27]. Furthermore, results show that the M06-2X functional is suitable for unraveling non-covalent interactions, which allows more reliable results for later functionalization studies [28,29].
The stabilized nanotubes showed a singlet multiplicity ( M = 1 ) with a neutral charge state ( Q = 0 ). Vibrational frequency calculations were performed to verify that the structures relax with local minima. The reactivity of the systems was determined using global quantum molecular descriptors such as chemical potential (μ), global hardness (η) and electrophilicity index (ω), obtained from the HOMO and LUMO energies through the quantities obtained using Koopmans’ theorem [30]:
I = −EHOMO,
A = −ELUMO,
where I is the ionization potential and A is the electron affinity. This is true for density functionals that correctly describe the derivative discontinuity at the integer particle numbers. Since M06-2X includes 54% of the HF exchange, we expect this functional to yield accurate estimates. Therefore, the chemical potential (μ) and global hardness (η) can be calculated with the following equations:
μ = ( E N ) ν ( r ) = I + A 2
η = ( 2 E N 2 ) ν ( r ) = I A 2
where E is the total energy, N is the number of electrons and ν(r) is the external potential of the system [31]. The global electrophilicity index ( ω ) is described by Parr [32], which uses the electronic chemical potential and is given by:
ω = μ 2 2 η
This descriptor measures the tendency of the chemical species to accept electrons. Small values of ω indicate that the chemical species behaves as an electron donor (nucleophile), while high values of ω characterize the electron acceptor (electrophile).
The stability of all the nanotubes was validated by calculating the cohesion energy, which is expressed as:
E c o h = E B P N T ( n B E B o r o n + n P E P h o s p + n H E H y d r o ) n B + n P + n H
where E B P N T is the total energy of the nanotube, while n B , n P , n H , E B o r o n , E P h o s p and E H y d r o represent the number of atoms and total energies of the isolated boron, phosphorus and hydrogen atoms, respectively. The E c o h is expressed in e V / a t o m . A stable geometry will be characterized by a negative cohesion energy.
Solvation energies ( E S o l v ) of each BPNT were obtained based on the total energy of the optimized nanotube in vacuum ( E B P N T ) and water ( E B P N T C P C M ) (via Conductor-like Polarizable Continuum Model), as indicated by the following equation:
E S o l v = E B P N T C P C M E B P N T
The more negative the E S o l v value is, the greater the degree of solubility of the system is.

3. Results and Discussion

3.1. Structural Properties

The diameter size trend with respect to the chirality in each nanotube is shown in Figure 1. It is evident that for both armchair- and zigzag-type nanotubes, as the chiral index increases, the diameter increases almost linearly, whose coefficients of determination (termed as R2) are 0.9984 and 0.9996 for the (n,n) BPNT and (n,0) BPNT, respectively. These R2 values indicate a very good linear fit since values are close to unity. Table 1 shows in detail the diameter size corresponding to each nanotube.
The axial length of each nanotube shows no considerable variation with respect to the chirality, with an average value of 19.94 Ȧ and 17.98 Ȧ for armchair- and zigzag-type nanotubes, respectively. Table 2 and Table 3 summarize the relaxed bond lengths and angles. In both cases, the B-P bond length decreases as the chirality index increases. The average length of the B-P bond for both chiralities is ~1.88 Ȧ. In addition, the boron–hydrogen bond length is shorter than that of the phosphorus–hydrogen bond ( B H < P H ), because of the electronegativity difference in the B-H, which is greater than in the P-H, inducing a more strengthened bond. The average value of the angle formed by the P-B-P atoms is 120.08° and 120.45° for armchair- and zigzag-type BPNTs, respectively. The angles of the H-P-B and H-B-P bonds at the ends of armchair and zigzag nanotubes increase slightly with increasing chirality. This small increase occurs because of chirality increase, as there is a higher density of hydrogen atoms at the end of the tube, causing a repulsion effect between them, which in turn generates an increase in the formed angle. The cohesion energy values are reported in Table 4 (armchair type) and Table 5 (zigzag type). Negative values were obtained for each nanotube, with an average value of 2.96   eV / atom for armchair-type and 2.95   eV / atom for zigzag-type BPNTs, indicating that they are all stable structures. These values suggest that pristine BPNTs are less stable than CNTs and BNNTs, since the latter have cohesion energies of −8.72 and −7.27 eV, respectively [33]. For both types of chirality, the cohesion energy decreases as the chiral index ( n ) increases, indicating that the greater the diameter is, the better the stability of the system is. However, the decrease in E c o h is very small, since the energy difference between the maximum and the minimum value is only 0.13   eV / atom , both in the armchair- and zigzag-type. Vibrational frequency calculations showed that all nanotubes have non-imaginary frequencies.
The calculated IR spectra of all systems are shown in Figure 2 and Figure 3 for the armchair- and zigzag-type boron phosphide nanotubes, respectively. In both cases, a pair of sets of characteristic peaks are noted in the frequency region of >2400 cm−1, corresponding to the P-H stretch vibration (2503–2521 cm−1) and the B-H stretch vibration (~2725 cm−1). The highest ε intensity peaks occur in the frequency range between 850–1100 cm−1, again for both cases, attributed to a P-B-P stretch vibration (900–955 cm−1) and to the B-H bending vibration (1031–1040 cm−1). It is noted that the P-B-P stretch vibration mode is affected by a blue shift when the chiral index increases (+10 cm−1/n, Figure 2b and Figure 3b). The armchair- and zigzag-type BPNTs display a very similar IR spectra; however, the armchair-type nanotubes exhibit a characteristic small peak in the frequency range of 820–840 cm−1 corresponding to the P-H bond bending mode. This characteristic signal allows differentiating between both chiralities. In general, the peak intensity increases with the chiral index, as induced by the number of bonds present in the structure.

3.2. Electronic Properties

The highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) are discussed in this section. Figure 4 shows the |HOMO-LUMO| energy gaps as functions of the chirality as obtained in this work and compared with those obtained in the previously mentioned reports (see Introduction Section). The energy gap of the (n,n) BPNTs decreases as the index (diameter) increases, reaching stabilization when n ≥ 10, (10,10) BPNT. This characteristic has also been reported for BNNT and SiCNT, where starting at a certain diameter the band gap energy is chirality independent [16,24]. On the other hand, zigzag nanotubes transform from a metallic character (5,0) BPNT with 0.08 eV to semiconductor, with (10,0) BPNT being the one with a larger energy gap value (0.98 eV). Thus, low chirality zigzag BPNTs can be used in electronic and energy transport devices. It can be noted that, in the case of zigzag BPNTs, our values have a similar trend and are close to those reported by Srivastava et al. [24], particularly with those calculated by the LDA/PZ/DZP method. However, our results differ largely from those described by Azizi et al. [25] for both armchair- and zigzag-type nanotubes. Other authors have also reported the value of the energy gap for some chiralities in studies about the functionalization of boron phosphide nanotubes. Table 6 summarizes these values reported by different authors [9,10,11,12,14,15,17,18,20,21,22,34,35,36,37] for the nanotubes (5,0), (6,0), (7,0), (8,0) and (4,4), comparing them with the values obtained in this work. Considerable different values of molecular gap are noted for the same chirality, however, all the values reported by other authors are calculated with the B3LYP or BLYP functional. This suggests that there is an underestimation or overestimation when calculating the HOMO-LUMO gap when the functionals B3LYP/BLYP and the M06-2X are used, respectively. Another factor that could be responsible for this difference in the molecular gap is the number of atoms that make up the nanotube, that is, its molecular formula. For all the cases shown in Table 6, our models of BPNTs have a larger number of atoms and, therefore, a greater length, because chirality determines the diameter of the nanotube but not its length. This would be a great challenge to synthesize BPNTs in a controlled way, since their electronic properties would present a strong dependence on the length of the nanotube.
The distribution of the HOMO and LUMO orbitals through the nanotubes are graphically depicted in Figure 5 and Figure 6 for the armchair and zigzag BPNTs, respectively. It is evident that the HOMO orbital is uniformly distributed throughout the entire nanotube in the armchair type, however, different features display the zigzag type, since this is concentrated at one end of the nanotube. The LUMO orbital, in the zigzag-type BPNTs, is uniformly distributed until chirality (8,0). At larger index, this is concentrated at one of the extremes. In the armchair-type BPNTs, analogous to HOMO, the LUMO is distributed throughout the entire nanotube.
The reactivity and chemical stability of the nanotubes was measured through global quantum molecular descriptors such as chemical potential (μ), global hardness (η) and electrophilicity index (ω), energy gap (Eg), ionization potential (I) and electronic affinity (A), as summarized in Table 4 and Table 5. These parameters have been widely used for many years in computational chemistry studies [38,39,40,41] and, recently, to study the functionalization and stability of some III-V nanotubes [42,43,44,45,46]. The global hardness represents the system resistance to a charge transfer, that is, high values of η indicate greater electronic stability. Thus, the electronic stability of the armchair-type BPNTs decreases as chirality increases, with (3,3) BPNT being the most electronically stable system (see Figure 7). The zigzag-type BPNTs exhibit an arbitrary variation in the global hardness with respect to the chirality, showing a maximum energy value of ~0.492 eV in the (10,0) BPNT. In addition, the zigzag nanotubes display the highest (most negative) μ values and these decrease with increasing chirality, so that the zigzag BPNTs with the highest chirality are the least chemically reactive, since the chemical potential is a property that characterizes the tendency of electrons to escape from a system in equilibrium. No significant variation in the value of μ of the armchair-type BPNTs, −6.34 eV < μ <−6.26 eV, is found, indicating chirality independence (see Figure 8). However, the chemical potential of both chiralities tend to converge towards the same value, i.e., μ = 6.36   eV ± 0.37 % for n 10 . Comparing the global hardness and chemical potential values of boron phosphide nanotubes with their most studied structural analogues, CNTs and BNNTs, it is noted that BPNTs show η values in a range very close to that of CNTs, however, this later structure is much smaller than those reported for the BNNTs. Moreover, in the case of the chemical potential, μ , the BPNT values are similar to those of the BNNTs, being much smaller than those reported for CNTs [46,47,48]. Therefore, we can establish that η C N T s η B P N T s η B N N T s and | μ C N T s | < | μ B P N T s | | μ B N N T s | .
Regarding the electrophilicity index, the (5,0) BPNT exhibits a large value provided that ω (599.59 eV) is large. However, the electrophilic character of the zigzag BPNTs drops abruptly with increasing chirality, while in the armchair type, this parameter increases gradually (see Figure 9). The ionization potential ( I ) measures the tendency to yield electrons, while electron affinity ( A ) measures the tendency to accept electrons. Thus, high ionization potential values indicate that large energy is required to yield electrons from the HOMO orbital, while large values of electron affinity indicate high electron accepting capacity. Both the armchair- and the zigzag-type chiralities exhibit large energy values of I and A , therefore, BPNTs are systems that easily accept electrons. This justifies their great electrophilic character, as indicated by the ω index.
The dipole moment μ D is a quantitative measure of the dipole moment magnitude and allows differentiating between polar and non-polar molecules. It is considered to a certain extent an index of reactivity, which is very important to define some biological properties [49]. The dipole moment of a polyatomic system is the vector sum of the bond dipoles. Our results show that the (n,n) BPNTs exhibit nearly zero dipole moment values (i.e., non-polar systems), as induced by the structural symmetry of the armchair-type nanotubes (see Figure 10). Nearly-zero polarity is independent of the type of atoms that make up the nanotube, since other nanotubes such as GaNNT, BNNT and AlNNT also display null dipole moment values of armchair-type chirality [43,50]. On the other hand, zigzag-type nanotubes are polar structures, with the (14,0) BPNT being the one with the largest value μ D = 2.32   D , although with no dependence of polarity ( μ D ) on the chiral index ( n ). This suggests that zigzag-type boron phosphide nanotubes are more soluble than armchair-type structures in polar solvents such as water, which in turn indicates that they are suitable for applications in biological systems. This characteristic of zigzag-type nanotubes of having a higher dipole moment, and therefore greater solubility, compared to armchair type, has also been reported for BNNTs [51,52]. These results make it clear that the zigzag-type nanotubes have a higher degree of solubility, since polarity is a key property for high solubility. We can validate this through the solvation energy values calculated using Equation (7). The results show that, in general, the solvation energy in the zigzag type is higher (more negative) than in armchair nanotubes, as expected. It may be thought that Figure 11 contradicts this last assertion since, for example, the armchair-type (11,11) BPNT has higher E S o l v values than most of the zigzag-type nanotubes. However, this is not an appropriate comparison, since the (11,11) BPNT contains a much larger number of atoms that promotes intermolecular interactions which in turn favors solvation. Then, it is necessary to compare the E S o l v values between nanotubes with a similar number of atoms and having almost the same diameters, for example the (5,5) BPNT (−12.08 kcal/mol) vs. (8,0) BPNT (−17.27 kcal/mol) or the (6,6) BPNT (−13.49 kcal/mol) vs. (10,0) BPNT (−19.83 kcal/mol). There are several recent investigations that show the important role that the dipole moment plays on the solubility of some pristine and/or functionalized nanotubes [51,52,53,54,55,56].

4. Conclusions

In this study, the structural and electronic properties of boron phosphide nanotubes in the zigzag and armchair chiralities were investigated. The study was conducted under the Density Functional Theory approach with the M06-2X/6-31G(d) method. The total energy minimization, assuming a non-magnetic nature and a total charge neutrality, yielded the ground state of all nanostructures, (n,0) BPNT (n = 5–8, 10, 12, 14) and (n,n) BPNT (n = 3–11). In both chiralities, results indicate that the BPNT diameter grows linearly with the chiral index (n). Global molecular descriptors establish that (3,3) BPNT is the most stable nanotube because it has the largest global hardness value. The (5,0) BPNT has a strong electrophilic character, acting as an excellent electron acceptor. The chemical potential and the electrophilicity index in the zigzag-type BPNTs show remarkable chirality-dependent behavior. Based on the characteristic behavior of the |HOMO-LUMO| energy gap, the zigzag-type BPNTs transform from metallic to semiconductor as their chirality increases. In contrast, the armchair-type nanotubes show a semiconductor behavior with the energy gap decreasing gradually with the increase in the chiral index. Finally, in this study of nanotubes as finite molecules, it is observed that armchair BPNTs are non-polar systems, unlike the remarkable polar behavior of zigzag-type BPNTs regardless of chirality. In general, studying the variation in the structural and electronic properties of BPNTs allows proposing these systems for possible applications. Therefore, the set of nanotubes studied in this work may be proposed for both electrical and biophysical applications.

Author Contributions

Conceptualization, D.G.-T. and R.M.-B.; methodology, J.F.R.-S.; software, E.C.-A.; validation, D.G.-T. and J.F.R.-S.; formal analysis, G.H.C.; investigation, V.M.V.-B.; resources, V.M.V.-B.; data curation, A.F.-R.; writing—original draft preparation, J.F.R.-S.; writing—review and editing, R.M.-B.; visualization, V.M.V.-B.; supervision, D.G.-T.; project administration, E.C.-A.; funding acquisition, V.M.V.-B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thankfully acknowledge the computer resources, technical expertise and support provided by the Laboratorio Nacional de Supercómputo del Sureste de México, CONACYT member of the network of national laboratories.

Conflicts of Interest

The authors declare no conflict of interest. On behalf of all authors, the corresponding author states that there is no conflict of interest.

References

  1. Iijima, S. Helical microtubules of graphitic carbon. Nature 1991, 354, 56–58. [Google Scholar] [CrossRef]
  2. Rubio, A.; Corkill, J.L.; Cohen, M.L. Theory of graphitic boron nitride nanotubes. Phys. Rev. B Condens. Matter 1994, 49, 5081–5084. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Chopra, N.G.; Luyken, R.J.; Cherrey, K.; Crespi, V.H.; Cohen, M.L.; Louie, S.G.; Zettl, A. Boron Nitride Nanotubes. Science 1995, 269, 966–967. [Google Scholar] [CrossRef] [PubMed]
  4. Odom, T.W.; Huang, J.-L.; Kim, P.; Lieber, C.M. Structure and Electronic Properties of Carbon Nanotubes. J. Phys. Chem. B. 2000, 104, 2794–2809. [Google Scholar] [CrossRef]
  5. Sfeir, M.Y. Optical Spectroscopy of Individual Single-Walled Carbon Nanotubes of Defined Chiral Structure. Science 2006, 312, 554–556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Marana, N.L.; Albuquerque, A.R.; La Porta, F.A.; Longo, E.; Sambrano, J.R. Periodic density functional theory study of structural and electronic properties of single-walled zinc oxide and carbon nanotubes. J. Solid State Chem. 2016, 237, 36–47. [Google Scholar] [CrossRef]
  7. Dass, D.; Vaid, R. Chirality dependence of electronic band structure and density of states in single-walled carbon nanotubes. Afr. Rev. Phys. 2018, 12, 104–113. [Google Scholar]
  8. Blase, X.; Rubio, A.; Louie, S.G.; Cohen, M.L. Stability and Band Gap Constancy of Boron Nitride Nanotubes. Europhys. Lett. (EPL) 1994, 28, 335–340. [Google Scholar] [CrossRef] [Green Version]
  9. Baei, M.T.; Moradi, A.V.; Torabi, P.; Moghimi, M. Adsorption properties of H2O2 trapped inside a boron phosphide nanotube. Monatsh. Chem. 2011, 143, 37–41. [Google Scholar] [CrossRef]
  10. Peyghan, A.A.; Baei, M.T.; Moghimi, M.; Hashemian, S. Theoretical Study of Phenol Adsorption on Pristine, Ga-Doped, and Pd-Decorated (6,0) Zigzag Single-Walled Boron Phosphide Nanotubes. J. Clust. Sci. 2012, 24, 49–60. [Google Scholar] [CrossRef]
  11. Sayyad-Alangi, S.Z.; Hashemian, S.; Baei, M.T. Covalent Functionalization of Pristine and Ga-Doped Boron Phosphide Nanotubes with Imidazole. Phosphorus Sulfur Silicon Relat. Elem. 2014, 189, 453–464. [Google Scholar] [CrossRef]
  12. Sayyad-Alangi, S.Z.; Baei, M.T.; Hashemian, S. Adsorption and electronic structure study of thiazole on the (6,0)zigzagsingle-walled boron phosphide nanotube. J. Sulphur Chem. 2013, 34, 407–420. [Google Scholar] [CrossRef]
  13. Beheshtian, J.; Baei, M.T.; Peyghan, A.A. Theoretical study of CO adsorption on the surface of BN, AlN, BP and AlP nanotubes. Surf. Sci. 2012, 606, 981–985. [Google Scholar] [CrossRef]
  14. Kanani, Y.; Baei, M.T.; Moradi, A.V.; Soltani, A. Adsorption mechanism of single OCN− and SCN− upon single-walled BP nanotubes. Phys. E Low Dimens. Syst. Nanostruct. 2014, 59, 66–74. [Google Scholar] [CrossRef]
  15. Soltani, A.; Taghartapeh, M.R.; Mighani, H.; Pahlevani, A.A.; Mashkoor, R. A first-principles study of the SCN− chemisorption on the surface of AlN, AlP, and BP nanotubes. Appl. Surf. Sci. 2012, 259, 637–642. [Google Scholar] [CrossRef]
  16. Wu, I.J.; Guo, G.Y. Optical properties of SiC nanotubes: An ab initio study. Phys. Rev. B Condens. Matter 2007, 76, 035343. [Google Scholar] [CrossRef] [Green Version]
  17. Mirzaei, M. Carbon doped boron phosphide nanotubes: A computational study. J. Mol. Model. 2010, 17, 89–96. [Google Scholar] [CrossRef]
  18. Esrafili, M.D. Carbon-Doped (6,0) Single-Walled Boron-Phosphide Nanotubes: A DFT Investigation of Electronic Structure, Surface Electrostatic Potential and QTAIM Analysis. Fuller. Nanotub. Carbon Nanostruct. 2014, 23, 142–147. [Google Scholar] [CrossRef]
  19. Baei, M.T.; Moghimi, M.; Torabi, P.; Moradi, A.V. The Ge-doped (6,0) zigzag single-walled boron phosphide nanotubes: A computational study. Comput. Theor. Chem. 2011, 972, 14–19. [Google Scholar] [CrossRef]
  20. Rezaei-Sameti, M. Gallium doped in armchair and zigzag models of boron phosphide nanotubes (BPNTs): A NMR study. Phys. B Condens. Matter 2012, 407, 3717–3721. [Google Scholar] [CrossRef]
  21. Baei, M.T.; Peyghan, A.A.; Moghimi, M. Electronic structure study of Si-doped (4,4) armchair single-walled boron phosphide nanotube as a semiconductor. Monatsh. Chem. 2012, 143, 1627–1635. [Google Scholar] [CrossRef]
  22. Baei, M.T.; Sayyad-Alangi, S.Z.; Moradi, A.V.; Torabi, P. NMR and NQR parameters of the SiC-doped on the (4,4) armchair single-walled BPNT: A computational study. J. Mol. Model. 2011, 18, 881–889. [Google Scholar] [CrossRef] [PubMed]
  23. Baei, M.T. Ge-doped (4,4) armchair single-walles boron phosphide nanotube as a semiconductor: A computational study. Monatsh. Chem. 2012, 6, 881–889. [Google Scholar] [CrossRef]
  24. Srivastava, A.; Sharma, M.; Tyagi, N.; Kothari, S.L. Diameter Dependent Electronic Properties of Zigzag Single Wall BX (X = N, P, As) Nanotubes: Ab-Initio Study. J. Comput. Theor. Nanosci. 2012, 9, 1693–1699. [Google Scholar] [CrossRef]
  25. Salabat, K.; Azizi, K. A Computational Study on the Effects of Size and Chirality on the Electronic and Structural Properties of BP Nanotubes. In Proceedings of the 16th Iranian Physical Chemistry Conference, Mazandaran, Iran, 29–31 October 2013; Volume 16, pp. 867–869. Available online: https://www.sid.ir/FileServer/SE/247E20131694.pdf (accessed on 9 May 2021).
  26. Frish, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Montgomery, J.A., Jr.; Vreven, T.; Kudin, K.N.; Burant, J.C.; et al. G09W®; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  27. Hou, N.; Wu, Y.Y.; Wei, Q.; Liu, X.L.; Ma, X.J.; Zhang, M.; Wu, H.S. A comparative theoretical study on the electrical and nonlinear optical properties of Li atom adsorbed on AlN and BN single-walled nanotubes. J. Mol. Model. 2017, 23, 1–8. [Google Scholar] [CrossRef] [PubMed]
  28. Mardirossian, N.; Head-Gordon, M. How Accurate Are the Minnesota Density Functionals for Noncovalent Interactions, Isomerization Energies, Thermochemistry, and Barrier Heights Involving Molecules Composed of Main-Group Elements? J. Chem. Theory Comput. 2016, 12, 4303–4325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2007, 120, 215–241. [Google Scholar]
  30. Koopmans, T. Über die Zuordnung von Wellenfunktionen und Eigenwerten zu den Einzelnen Elektronen Eines Atoms. Physica 1934, 1, 104–113. [Google Scholar] [CrossRef]
  31. Parr, R.G.; Pearson, R.G. Absolute hardness: Companion parameter to absolute electronegativity. J. Am. Chem. Soc. 1983, 105, 7512–7516. [Google Scholar] [CrossRef]
  32. Parr, R.G.; Szentpály, L.v.; Liu, S. Electrophilicity Index. J. Am. Chem. Soc. 1999, 121, 1922–1924. [Google Scholar] [CrossRef]
  33. Juárez, A.R.; Anota, E.C.; Cocoletzi, H.H.; Ramírez, J.S.; Castro, M. Stability and electronic properties of armchair boron nitride/carbon nanotubes. Fuller. Nanotub. Carbon Nanostruct. 2017, 25, 716–725. [Google Scholar] [CrossRef]
  34. Beheshtian, J.; Soleymanabadi, H.; Kamfiroozi, M.; Ahmadi, A. The H2 dissociation on the BN, AlN, BP and AlP nanotubes: A comparative study. J. Mol. Model. 2011, 18, 2343–2348. [Google Scholar] [CrossRef] [PubMed]
  35. Mirzaei, M.; Giahi, M. Computational studies on boron nitride and boron phosphide nanotubes: Density functional calculations of boron-11 electric field gradient tensors. Phys. E Low Dimens. Syst. Nanostruc. 2010, 42, 1667–1669. [Google Scholar] [CrossRef]
  36. Arshadi, S.; Bekhradnia, A.R.; Alipour, F.; Abedini, S. Pure and carbon-doped boron phosphide (6,0) zigzag nanotube: A computational NMR study. Phys. B Condens. Matter 2015, 477, 1–7. [Google Scholar] [CrossRef]
  37. Rezaei-Sameti, M.; Yaghoobi, S. Theoretical study of adsorption of CO gas on pristine and AsGa-doped (4, 4) armchair models of BPNTs. Comput. Condens. Matter 2015, 3, 21–29. [Google Scholar] [CrossRef] [Green Version]
  38. Geerlings, P.; De Proft, F.; Langenaeker, W. Conceptual Density Functional Theory. Chem. Rev. 2003, 103, 1793–1874. [Google Scholar] [CrossRef]
  39. Parr, R.G. Density Functional Theory of Atoms and Molecules. Horiz. Quantum Chem. 1980, 3, 5–15. [Google Scholar]
  40. Chermette, H. Chemical reactivity indexes in density functional theory. J. Comput. Chem. 1999, 20, 129–154. [Google Scholar] [CrossRef]
  41. Roos, G.; Geerlings, P.; Messens, J. Enzymatic Catalysis: The Emerging Role of Conceptual Density Functional Theory. J. Phys. Chem. B 2009, 113, 13465–13475. [Google Scholar] [CrossRef]
  42. Hosseinzadeh, B.; Salimi Beni, A.; Eskandari, R.; Karami, M.; Khorram, M. Interaction of propylthiouracil, an anti-thyroid drug with boron nitride nanotube: A DFT study. Adsorption 2020, 26, 1385–1396. [Google Scholar] [CrossRef]
  43. Makiabadi, B.; Zakarianezhad, M.; Hosseini, S.S. Investigation and comparison of pristine/doped BN, AlN, and CN nanotubes as drug delivery systems for Tegafur drug: A theoretical study. Struct. Chem. 2021, 32, 1019–1037. [Google Scholar] [CrossRef]
  44. Chen, K.; Zeng, Q.; He, G.; Sarkar, A. Tyrosine amino acid as a sustainable material for chemical functionalization of single-wall BC2N nanotubes: Quantum chemical study. Struct. Chem. 2021, 32, 1197–1203. [Google Scholar] [CrossRef]
  45. Muz, İ.; Kurban, H.; Kurban, M. A DFT study on stability and electronic structure of AlN nanotubes. Mater. Today Commun. 2021, 26, 102118. [Google Scholar] [CrossRef]
  46. García-Toral, D.; Báez, R.M.; Sánchez, S.J.I.; Flores-Riveros, A.; Cocoletzi, G.H.; Rivas-Silva, J.F. Encapsulation of Pollutant Gaseous Molecules by Adsorption on Boron Nitride Nanotubes: A Quantum Chemistry Study. ACS Omega 2021, 6, 14824–14837. [Google Scholar] [CrossRef] [PubMed]
  47. Beheshtian, J.; Peyghan, A.A.; Bagheri, Z. Carbon nanotube functionalization with carboxylic derivatives: A DFT study. J. Mol. Model. 2013, 19, 391–396. [Google Scholar] [CrossRef]
  48. Chermahini, A.N.; Teimouri, A.; Farrokhpour, H. A DFT-D study on the interaction between lactic acid and single-wall carbon nanotubes. RSC Adv. 2015, 5, 97724–97733. [Google Scholar] [CrossRef]
  49. Mendoza-Wilson, A.M.; Glossman-Mitnik, D. CHIH-DFT study of the electronic properties and chemical reactivity of quercetin. J. Mol. Struct. THEOCHEM 2005, 716, 67–72. [Google Scholar] [CrossRef]
  50. Erkoç, Ş.; Malcioğlu, O.B.; Taşci, E. Structural and electronic properties of single-wall GaN nanotubes: Semi-empirical SCF-MO calculations. J. Mol. Struct. THEOCHEM 2004, 674, 1–5. [Google Scholar] [CrossRef]
  51. Kaur, J.; Singla, P.; Goel, N. Adsorption of oxazole and isoxazole on BNNT surface: A DFT study. Appl. Surf. Sci. 2015, 328, 632–640. [Google Scholar] [CrossRef]
  52. Sánchez, S.J.I.; Rivas-Silva, J.F.; García-Toral, D. Study of weak interactions of boron nitride nanotubes with anticancer drug by quantum chemical calculations. Theor. Chem. Acc. 2020, 139, 1–9. [Google Scholar] [CrossRef]
  53. Makiabadi, B.; Zakarianezhad, M.; Nasab, S.S. Theoretical study of interaction of NH2X (X = H, CH3, CH2OCH3, and CH2COOH) molecules with AlN and AlP nanotubes. Phosphorus Sulfur Silicon Relat. Elem. 2017, 192, 81–87. [Google Scholar] [CrossRef]
  54. Mananghaya, M.R.; Santos, G.N.; Yu, D.N. Solubility of amide functionalized single wall carbon nanotubes: A quantum mechanical study. J. Mol. Liq. 2017, 242, 1208–1214. [Google Scholar] [CrossRef]
  55. Beheshtian, J.; Peyghan, A.A.; Tabar, M.B.; Bagheri, Z. DFT study on the functionalization of a BN nanotube with sulfamide. Appl. Surf. Sci. 2013, 266, 182–187. [Google Scholar] [CrossRef]
  56. Mananghaya, M. Modeling of single-walled carbon nanotubes functionalized with carboxylic and amide groups towards its solubilization in water. J. Mol. Liq. 2015, 212, 592–596. [Google Scholar] [CrossRef]
Figure 1. Diameter length (Ȧ) as a function of chiral index (n). The black line corresponds to the armchair-type nanotubes, while the red line to the zigzag type. Atom colors: orange—phosphorus; pink—boron; white—hydrogen.
Figure 1. Diameter length (Ȧ) as a function of chiral index (n). The black line corresponds to the armchair-type nanotubes, while the red line to the zigzag type. Atom colors: orange—phosphorus; pink—boron; white—hydrogen.
Symmetry 14 00964 g001
Figure 2. (a) Theoretical IR spectra of armchair-type BPNTs; (b) close-up in range from 825 cm−1 to 1075 cm−1. The signal of the peaks was shifted by 350 ε units relative to each other.
Figure 2. (a) Theoretical IR spectra of armchair-type BPNTs; (b) close-up in range from 825 cm−1 to 1075 cm−1. The signal of the peaks was shifted by 350 ε units relative to each other.
Symmetry 14 00964 g002
Figure 3. (a) Theoretical IR spectra of zigzag-type BPNTs; (b) close-up in range from 700 cm−1 to 1100 cm−1. The signal of the peaks was shifted by 350 ε units relative to each other.
Figure 3. (a) Theoretical IR spectra of zigzag-type BPNTs; (b) close-up in range from 700 cm−1 to 1100 cm−1. The signal of the peaks was shifted by 350 ε units relative to each other.
Symmetry 14 00964 g003
Figure 4. |HOMO-LUMO| energy gaps as functions of the chiral index (n) for armchair- (black line) and zigzag-type (red line) BPNTs. Comparison of values calculated in: a this work (unfilled squares), and in Refs; b [25] (filled circles) and c [24] (filled/unfilled triangles).
Figure 4. |HOMO-LUMO| energy gaps as functions of the chiral index (n) for armchair- (black line) and zigzag-type (red line) BPNTs. Comparison of values calculated in: a this work (unfilled squares), and in Refs; b [25] (filled circles) and c [24] (filled/unfilled triangles).
Symmetry 14 00964 g004
Figure 5. Spatial distribution of the HOMO and LUMO orbitals through the armchair-type BP nanotubes. Atom colors: orange—phosphorus; pink—boron; white—hydrogen.
Figure 5. Spatial distribution of the HOMO and LUMO orbitals through the armchair-type BP nanotubes. Atom colors: orange—phosphorus; pink—boron; white—hydrogen.
Symmetry 14 00964 g005
Figure 6. Spatial distribution of the HOMO and LUMO orbitals through the zigzag-type BP nanotubes. Atom colors: orange—phosphorus; pink—boron; white—hydrogen.
Figure 6. Spatial distribution of the HOMO and LUMO orbitals through the zigzag-type BP nanotubes. Atom colors: orange—phosphorus; pink—boron; white—hydrogen.
Symmetry 14 00964 g006
Figure 7. |HOMO-LUMO| gap and global hardness as functions of the chiral index for armchair-type (black line) and zigzag-type (red line) BPNTs. Inset: equation of the relationship between |HOMO-LUMO| gap and global hardness (bottom); (3,3) and (10,0) BPNTs (top).
Figure 7. |HOMO-LUMO| gap and global hardness as functions of the chiral index for armchair-type (black line) and zigzag-type (red line) BPNTs. Inset: equation of the relationship between |HOMO-LUMO| gap and global hardness (bottom); (3,3) and (10,0) BPNTs (top).
Symmetry 14 00964 g007
Figure 8. Chemical potential behavior as a function of the chiral index for armchair-type (black line) and zigzag-type (red line) BPNTs.
Figure 8. Chemical potential behavior as a function of the chiral index for armchair-type (black line) and zigzag-type (red line) BPNTs.
Symmetry 14 00964 g008
Figure 9. Electrophilicity index behavior as a function of the chiral index for armchair-type (black line) and zigzag-type (red line) BPNTs.
Figure 9. Electrophilicity index behavior as a function of the chiral index for armchair-type (black line) and zigzag-type (red line) BPNTs.
Symmetry 14 00964 g009
Figure 10. Dipole moment behavior as a function of the chiral index for armchair-type (black line) and zigzag-type (red line) BPNTs.
Figure 10. Dipole moment behavior as a function of the chiral index for armchair-type (black line) and zigzag-type (red line) BPNTs.
Symmetry 14 00964 g010
Figure 11. Solvation energy as a function of the chiral index for the armchair-type (black line) and zigzag-type (red line) BPNTs. Values were calculated using the Conductor-like Polarizable Continuum Model (CPCM), as implemented in the software package Gaussian16.
Figure 11. Solvation energy as a function of the chiral index for the armchair-type (black line) and zigzag-type (red line) BPNTs. Values were calculated using the Conductor-like Polarizable Continuum Model (CPCM), as implemented in the software package Gaussian16.
Symmetry 14 00964 g011
Table 1. Diameter length (Ȧ) for the geometrically optimized armchair- (n,n) and zigzag-type (n,0) BP nanotubes as calculated with the M06-2X/6-31G(d) method.
Table 1. Diameter length (Ȧ) for the geometrically optimized armchair- (n,n) and zigzag-type (n,0) BP nanotubes as calculated with the M06-2X/6-31G(d) method.
Chirality/Indexn = 345678910111214
(n,0)--5.566.467.438.44-10.42-12.5514.40
(n, n)5.196.858.159.9912.1113.9115.5517.3819.31--
Table 2. Bond length (Ȧ) and bond angle (°) values of armchair-type (n,n) BP nanotubes (where n = 3–11) optimized via DFT/M06-2X/6-31G(d).
Table 2. Bond length (Ȧ) and bond angle (°) values of armchair-type (n,n) BP nanotubes (where n = 3–11) optimized via DFT/M06-2X/6-31G(d).
NanotubeBond Lengths (Ȧ)Angle (°)
ArmchairAxial LengthB-PP-HB-HP-B-PB-P-BH-P-BH-B-P
(3,3)19.88941.91291.42591.1861122.2794109.5253108.55118.1869
(4,4)19.90631.89241.42091.1859119.2242114.2072109.4626118.777
(5,5)19.92371.88621.41871.1856120.7065115.1849111.0487119.6057
(6,6)19.93761.88141.41661.1857119.691116.4117111.9162119.8233
(7,7)19.95031.87981.41521.1856119.6188117.8321112.3506119.6171
(8,8)19.96181.87711.41431.1856119.4768118.3511112.997119.7843
(9,9)19.97081.87561.41341.1855119.8691118.3118113.6048119.9248
(10,10)19.98051.87471.41271.1856120.1455118.9187114.2501120.085
(11,11)19.98611.8751.41221.1854119.7613119.2771114.7068120.1182
Table 3. Bond length (Ȧ) and bond angle (°) values of zigzag-type (n,0) BP nanotubes (where n = 5–8, 10, 12, 14) optimized via DFT/M06-2X/6-31G(d).
Table 3. Bond length (Ȧ) and bond angle (°) values of zigzag-type (n,0) BP nanotubes (where n = 5–8, 10, 12, 14) optimized via DFT/M06-2X/6-31G(d).
NanotubeBond Lengths (Ȧ)Angle (°)
ZigzagAxial LengthB-PP-HB-HP-B-PB-P-BH-P-BH-B-P
(5,0)17.86141.91351.43511.1861120.8479105.0648101.9944117.7055
(6,0)17.93461.89951.43161.1858121.7117110.9808102.8412117.5638
(7,0)17.98061.89391.42941.1856120.3019111.5138104.9868118.5491
(8,0)17.99661.88511.42791.1854120.3811114.968105.9255118.8265
(10,0)18.02591.88191.42551.1852119.7837115.6483107.5051119.1812
(12,0)18.04351.87781.4241.1848119.8862116.4566108.7472119.3955
(14,0)18.05371.87751.42281.185120.2846118.1968109.7213119.513
Table 4. Optimized total energy ( E T o t a l ), energy of the highest occupied molecular orbital ( E H O M O ), energy of the lowest unoccupied molecular orbital ( E L U M O ), energies of the quantum molecular descriptors ( η , μ , ω ) and cohesion energy ( E c o h ) for armchair-type BPNTs. All values are given in eV units and performed via DFT/M06-2X/6-31G(d).
Table 4. Optimized total energy ( E T o t a l ), energy of the highest occupied molecular orbital ( E H O M O ), energy of the lowest unoccupied molecular orbital ( E L U M O ), energies of the quantum molecular descriptors ( η , μ , ω ) and cohesion energy ( E c o h ) for armchair-type BPNTs. All values are given in eV units and performed via DFT/M06-2X/6-31G(d).
Armchair BPNT Quantum   Molecular   Descriptors   and   E c o h   for   ( n , n )   BPNTs
Descriptors(3,3)(4,4)(5,5)(6,6)(7,7)(8,8)(9,9)(10,10)(11,11)
E T O T A L −358,734.516−478,319.133−597,903.132−717,486.631−837,069.831−956,652.817−1,076,235.658−1,195,818.389−1,315,401.058
E H O M O −6.8995−6.7725−6.6888−6.6428−6.6145−6.5971−6.5857−6.5764−6.5712
E L U M O −5.6244−5.7531−5.9084−5.9922−6.0406−6.0694−6.0874−6.0988−6.1061
Eg Gap1.27511.01950.78040.65060.57390.52770.49830.47760.4651
I = E H O M O 6.89966.77256.68886.64286.61456.59716.58576.57646.5712
A = E L U M O 5.62445.75315.90845.99226.04066.06946.08746.09886.1061
η = (IA)/20.63760.50970.39020.32530.28700.26380.24920.23880.2326
µ = −(I + A)/2−6.2620−6.2628−6.2986−6.3175−6.3275−6.3332−6.3365−6.3376−6.3387
ω = μ 2 / 2 η 30.751638.474150.837561.341869.761876.012080.576184.092386.3841
E c o h −2.8618−2.9193−2.9494−2.9665−2.9772−2.9843−2.9892−2.9927−2.9954
Table 5. Optimized total energy ( E T o t a l ), energy of the highest occupied molecular orbital ( E H O M O ), energy of the lowest unoccupied molecular orbital ( E L U M O ), energies of the quantum molecular descriptors ( η , μ , ω ) and cohesion energy ( E c o h ) for zigzag-type BPNTs. All values are given in eV and as calculated via DFT/M06-2X/6-31G(d).
Table 5. Optimized total energy ( E T o t a l ), energy of the highest occupied molecular orbital ( E H O M O ), energy of the lowest unoccupied molecular orbital ( E L U M O ), energies of the quantum molecular descriptors ( η , μ , ω ) and cohesion energy ( E c o h ) for zigzag-type BPNTs. All values are given in eV and as calculated via DFT/M06-2X/6-31G(d).
Zigzag BPNT Quantum   Molecular   Descriptors   and   E c o h   for   ( n , 0 )   BPNTs
Descriptors(5,0)(6,0)(7,0)(8,0)(10,0)(12,0)(14,0)
E T O T A L −348,742.4988−418,495.3783−488,247.8623−558,000.0492−697,503.8110−837,007.0566−976,509.9845
E H O M O −6.9882−6.9210−7.0652−6.8106−6.8770−6.7910−6.7140
E L U M O −6.9077−6.7092−6.4298−6.2049−5.8934−5.9500−6.0411
Eg Gap0.08050.21190.63540.60570.98360.84100.6729
I = E H O M O 6.98826.92107.06526.81066.87706.79106.7140
A = E L U M O 6.90776.70926.42986.20495.89345.95006.0411
η = (IA)/20.04030.10590.31770.30290.49180.42050.3365
µ = −(I + A)/2−6.9480−6.8151−6.7475−6.5077−6.3852−6.3705−6.3776
ω = μ 2 / 2 η 599.5909219.198571.654769.915141.452648.254860.4427
E c o h −2.8633−2.9090−2.9380−2.9575−2.9809−2.9939−3.0017
Table 6. Review of the HOMO-LUMO gap values (eV) reported in the literature for different boron phosphide nanotubes. Values calculated by this work (Gaussian16/M06-2X/6-31G(d)) and references.
Table 6. Review of the HOMO-LUMO gap values (eV) reported in the literature for different boron phosphide nanotubes. Values calculated by this work (Gaussian16/M06-2X/6-31G(d)) and references.
Summary of HOMO-LUMO Gap Values and Molecular Formula Reported in the Literature
ChiralitySoftware + MethodHOMO-LUMO GapMolecular FormulaReference
(5,0)GAMESS/B3LYP/6-31G(d,p)1.62B27P27H10[34]
(5,0)Gaussian16/M06-2X/6-31G(d)0.0805B35P35H10This work
(6,0)Gaussian03/B3LYP/6-31G(d)2.27B24P24H12[12]
(6,0)Gaussian03/B3LYP/6-31G(d)2.43B24P24H12[18]
(6,0)GAMESS/B3LYP/6-31G(d)1.93B36P36H12[14]
(6,0)Gaussian98/BLYP/6-31G(d)1.09B24P24H12[35]
(6,0)GAMESS/B3LYP/6-31G(d)2.06B30P30H12[11]
(6,0)Gaussian98/BLYP/6-31G(d)1.09B24P24H12[17]
(6,0)Gaussian03/B3LYP/6-31G(d)2.06B30P30H12[10]
(6,0)Gaussian98/B3LYP/6-311G(d,p)2.25B24P24H12[36]
(6,0)Gaussian98/B3LYP/6-31G(d)2.27B24P24H12[15]
(6,0)Gaussian16/M06-2X/6-31G(d)0.2118B42P42H12This work
(7,0)GAMESS/B3LYP/6-31G(d)2.22B42P42H14[14]
(7,0)Gaussian16/M06-2X/6-31G(d)0.6353B49P49H14This work
(8,0)Gaussian03/B3LYP/6-31G(d)2.57B32P32H16[20]
(8,0)Gaussian16/M06-2X/6-31G(d)0.6057B56P56H16This work
(4,4)Gaussian98/BLYP/6-31G(d)1.77B28P28H16[17]
(4,4)Gaussian03/B3LYP/6-31G(d)2.95B28P28H16[9]
(4,4)Gaussian98/BLYP/6-31G(d)1.77B28P28H16[35]
(4,4)GAMESS/B3LYP/6-31G(d)2.95B28P28H16[14]
(4,4)Gaussian03/BLYP/6-31G(d)1.75B28P28H16[22]
(4,4)Gaussian03/B3LYP/6-31G(d)2.95B28P28H16[22]
(4,4)Gaussian03/B3LYP/6-31G(d)2.95B28P28H16[21]
(4,4)Gaussian03/B3LYP/6-31G(d)2.95B32P32H16[37]
(4,4)Gaussian16/M06-2X/6-31G(d)1.0194B48P48H16This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

García-Toral, D.; Mendoza-Báez, R.; Chigo-Anota, E.; Flores-Riveros, A.; Vázquez-Báez, V.M.; Cocoletzi, G.H.; Rivas-Silva, J.F. Structural Stability and Electronic Properties of Boron Phosphide Nanotubes: A Density Functional Theory Perspective. Symmetry 2022, 14, 964. https://doi.org/10.3390/sym14050964

AMA Style

García-Toral D, Mendoza-Báez R, Chigo-Anota E, Flores-Riveros A, Vázquez-Báez VM, Cocoletzi GH, Rivas-Silva JF. Structural Stability and Electronic Properties of Boron Phosphide Nanotubes: A Density Functional Theory Perspective. Symmetry. 2022; 14(5):964. https://doi.org/10.3390/sym14050964

Chicago/Turabian Style

García-Toral, Dolores, Raúl Mendoza-Báez, Ernesto Chigo-Anota, Antonio Flores-Riveros, Víctor M. Vázquez-Báez, Gregorio Hernández Cocoletzi, and Juan Francisco Rivas-Silva. 2022. "Structural Stability and Electronic Properties of Boron Phosphide Nanotubes: A Density Functional Theory Perspective" Symmetry 14, no. 5: 964. https://doi.org/10.3390/sym14050964

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