Next Article in Journal
Leather Dyeing by Plant-Derived Colorants in the Presence of Natural Additives
Next Article in Special Issue
Assessment of Strains Produced by Thermal Expansion in Printed Circuit Boards
Previous Article in Journal
The Influence of Solar Sintering on Copper Heat Exchanger Parts with Controlled 3D-Printed Morphology
Previous Article in Special Issue
Influence of Texture on the Mechanical Properties of a Mg-6Al-1Zn-0.9Sn Alloy Processed by ECAP
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Overview on the Evaluation of the Elastic Properties of Non-Carbon Nanotubes by Theoretical Approaches

by
Jorge M. Antunes
1,2,*,
André F. G. Pereira
1 and
Nataliya A. Sakharova
1
1
Centre for Mechanical Engineering, Materials and Processes (CEMMPRE), Department of Mechanical Engineering, University of Coimbra, Rua Luís Reis Santos, Pinhal de Marrocos, 3030-788 Coimbra, Portugal
2
Abrantes High School of Technology, Polytechnic Institute of Tomar, Quinta do Contador, Estrada da Serra, 2300-313 Tomar, Portugal
*
Author to whom correspondence should be addressed.
Materials 2022, 15(9), 3325; https://doi.org/10.3390/ma15093325
Submission received: 31 March 2022 / Revised: 29 April 2022 / Accepted: 30 April 2022 / Published: 5 May 2022
(This article belongs to the Collection Materials Investigations in Mechanical Systems)

Abstract

:
Low-dimensional structures, such as nanotubes, have been the focus of research interest for approximately three decades due to their potential for use in numerous applications in engineering and technology. In addition to extensive investigation of carbon nanotubes, those composed of elements other than carbon, the so-called non-carbon nanotubes, have also begun to be studied, since they can be more suitable for electronic and optical nano-devices than their carbon counterparts. As in the case of carbon nanotubes, theoretical (numerical and analytical) approaches have been established predominantly to study non-carbon nanotubes. So far, most of work has dealt with the investigation of the structural and electrical properties of non-carbon nanotubes, paying less attention to the evaluation of their mechanical properties. As the understanding of the mechanical behaviour of the constituents is fundamental to ensure the effective performance of nanotube-based devices, this overview aims to analyse and systematize the literature results on the elastic properties of inorganic non-carbon nanotubes.

1. Introduction

The discovery of carbon nanotubes (CNTs) gave rise to studies of the prediction and synthesis of new graphene-like structures based on other elements and chemical compounds, such as nitrides, phosphides and carbides, among others. Compounds of elements of the 12th–15th groups of the periodic table are able to establish a honeycomb diatomic arrangement, forming nanotubes (NTs) with a graphene-like hexagonal lattice. Boron nitride (BN), aluminum nitride (AlN), gallium nitride (GaN), indium nitride (InN), boron phosphide (BP), aluminum phosphide (AlP), gallium phosphide (GaP), indium phosphide (InP) and silicon carbide (SiC) nanotubes are examples of nano-tubular structures beyond the CNTs. These non-carbon nanotubes (N-CNTs), being wide band gap semiconductors or dielectrics, have promising applications in nano-devices for optoelectronics and electronics, such as light emitting diodes (LEDs) [1,2,3], field effect transistors [4,5], sensors and bio-detectors [6,7,8], high-frequency transistors [9] and tips in scanning probe microscopy [10,11]. A number of these non-carbon NTs have already been successfully synthesized, as in the cases of BNNTs [12], AlNNTs [13,14,15], GaNNTs [16,17], GaPNTs [18], InPNTs [19] and SiCNTs [20]; others were predicted theoretically, such as InNNTs [1,21], BPNTs [22,23], and AlPNTs [24].
Until now, work on N-CNTs has been mainly devoted to the study of their stability, structural and electronic properties (see, for example [1,22,23,24,25,26,27,28,29,30]). Regarding the investigation of their mechanical behaviour, studies are at a relatively early stage, due to the complexity of accurately measuring the mechanical properties, whose study needs to be in-depth and diversified. With the exception of certain works dealing with the evaluation of the mechanical properties of boron nitride NTs (see, for example [31,32,33,34,35,36]), the other inorganic NTs have received noticeably less research attention [29,37,38,39,40,41,42,43]. This lack of knowledge about the mechanical behaviour of N-CNTs is associated with difficulties in building robust and effective nanotube-based devices.
Since the experimental procedures for materials characterization at the nanoscale are costly and resource-intensive, the investigation of the mechanical behaviour of non-carbon nanotubes is carried out in a more theoretical way, with recourse to analytical and numerical techniques, as in the case of carbon nanotubes. Similar to studies on CNTs, there are three classes of theoretical methodologies that are used for the mechanical characterization of N-CNTs, namely: the atomistic approach, comprising, ab initio, molecular dynamics (MD) and tight-binding molecular dynamics (TBMD); the continuum mechanics (CM) approach; and the nanoscale continuum modelling (NCM) or molecular structural mechanics (MSM) approach. A comprehensive review of theoretical approaches for modelling and characterizing the mechanical behaviour of CNTs can be found in the literature [44,45].
With regard to atomistic approaches, ab initio simulation was used by Kochaev [37] to assess the surface Young’s modulus (product of Young’s modulus by the nanotube wall thickness) and Poisson’s ratio of BNNTs, AlNNTs, GaNNTs, AlPNTs and GaPNTs. Baumeier et al. [46] applied ab initio density functional theory (DFT) within self-interaction-corrected (SIC) pseudopotentials to calculate the surface Young’s modulus of BNNTs and SiCNTs. Hao et al. [47] employed ab initio DFT calculations coupled with linear combination of atomic orbitals (LCAO) to study size-dependent mechanical behaviour of the AlNNTs and to calculate their Young’s modulus. The existing MD simulations, performed to evaluate the mechanical properties of N-CNTs, used analytical or empirical potential functions to describe the interactions between atoms, which form a diatomic graphene-like hexagonal lattice. The Tersoff–type potential was used by Kang and Hwang [40] in their MD simulation study to describe the mechanical behaviour of BNNTs, AlNNTs and GaNNTs under compressive loading and to assess their Young’s modulus. Chandra et al. [48], using the Tersoff–type potential, also analysed the thermal vibrational characteristics of BNNTs. In two of their works Jeng et al. [38,49] adopted MD simulation with Tersoff many-body potential to evaluate the mechanical properties of GaNNTs under tension [38,49] and fatigue [38] loadings. Xiong and Tian [50] carried out a comprehensive study on the torsional properties of BNNTs, using MD simulation with Tersoff potential and, compared the force approach and the energy approach for calculation of BNNTs’ shear modulus. Moon et al. [51], Setoodeh et al. [52], Pan and Si [53] and Zhou et al. [54], using MD simulations based on Tersoff potentials, studied mechanical properties, mechanical behaviour under axial compression, tensile behaviour, and elastic and melting properties of the SiCNTs, respectively. Choyal et al. [31], Verma et al. [55], Tao et al. [56] and Ansary and Ajori [57] used MD with Tersoff–Brenner (TB) potential to describe the interactions between boron (B) and nitride (N) atoms, to clarify the effect of the aspect ratio on the Young’s modulus, to evaluate the Young’s and shear moduli and the Poisson’s ratio, to calculate the Young’s modulus, and to study vibrational behaviour of BNNTs, respectively. Wang et al. [58] employed a MD approach, describing the interactions between gallium (Ga) and nitride (N) atoms by Stilliger-Weber potential, to model the mechanical behaviour of GaNNTs under combined tension and torsion and study their failure. Vijayaraghavan and Zhang [35] used an empirical reactive bond order (REBO) to describe the atomic interactions in the BNNTs and studied their mechanical behaviour under tensile loading. Santosh et al. [59] adopted a force—constant approach for describing the B—N interactions under axial compression in the MD simulation study to calculate the Young’s and shear moduli of the BNNTs. Le [43], in MD simulation with harmonic force fields, obtained explicit expressions for the Young’s modulus of the BNNTs and SiCNTs. Regarding other atomistic approaches, Hernandez et al. [60] calculated the Young’s modulus and Poisson’s ratio of BNNTs using TBMD, while Pinhal et al. [29] coupled periodic DFT calculations with the functional B3LYP (Becke, 3-parameter, Lee–Yang–Parr) to assess the elastic constants of AlNNTs and GaNNTs. Zhang et al. [61] used the density-functional-based tight-binding (DFTB) model in combination with MD to evaluate the Young’s and shear moduli of BNNTs.
Works employing the CM approach, which consider the nanotube as a continuum structure, are relatively scarce in the literature to our knowledge. Oh [62] calculated the Young’s modulus and Poisson’s ratio of BNNTs, using a continuum lattice (CL) analytical thermodynamic approach, together with the TB potential. Panchal et al. [63] modelled the single-walled BNNT as a thin walled tube with thickness and investigated the vibrational response of the BNNTs with attached mass at the free nanotube end. Song et al. [64] developed a finite-deformation shell model to study the mechanical behaviour of BNNTs under tension, compression and torsional loads.
The NCM/MSM approach explores the link between molecular arrangement in nanotubes and solid mechanics and considers the bonds between two atoms in the diatomic structure as elements (for example, beams or springs) well described by elasticity theory. With regard to boron nitride NTs, which have received the most research attention up to now, Li and Chou [34], Salavati et al. [65], Ansari et al. [66], Sakharova et al. [36] and Panchal et al. [67] used the beam element to replace the B-N bond within the framework of the NCM/MSM approach, to study the elastic and dynamic properties, elastic moduli and Poisson’s ratio, electromechanical properties, size-dependent elastic properties and vibrational properties of BNNTs, respectively. In a recent study, Zakaria [68] modelled bonds between B and N atoms by two-sectioned beam elements to evaluate the elastic and vibrational properties of the BNNTs. Giannopoulos et al. [69], instead of the beam elements, used spring-like elements in the modelling of the B-N bond to investigate vibrational behaviour of BNNTs. Yan and Liew [70] considered a representative cell assembled by a boron atom connected to three neighbouring nitride atoms by B-N covalent bonds, to model the BNNTs under NCM/MSM approach. Yan et al. [71] assessed fundamental frequencies and elastic moduli by modelling longitudinal and torsional free vibrations of BNNTs, based on the NCM/MSM approach combined with an Euler beam model. Moreover, there are studies using the NCM/MSM approach, which deal not only with BNNTs, but also with other N-CNTs. Genoese et al. [41] combined the NCM/MSM and CM approaches to perform a numerical simulation study of the mechanical behaviour of BNNTs and SiCNTs under tensile, bending and torsional tests, using a “stick-and-spring” model involving Morse and cosine potential functions. In their work Genoese et al. [41] evaluate the Young’s and shear moduli, and Poisson’s ratio of BN and SiC nanotubes adopting a linkage between the “stick-and-spring” and continuum thin shell Donnell models. Jiang and Guo [39] proposed closed-formed analytical solutions based on the “stick-and-spring” model, to study the buckling and assess the surface Young’s modulus and Poisson’s ratio of BN, AlN, GaN, BP, GaP, InP and SiC nanotubes.
The important issue in modelling the N-CNTs’ mechanical behaviour under the NCM/MSM approach is to properly select the force field constants, regarding bond stretching, kr, bond bending, kθ, and torsional resistance, kτ, which are necessary for calculating the elastic properties of elements representing bonds between two atoms in the hexagonal diatomic structure. In the case of the N-CNTs, the computation of the kr, kθ and kτ force constants, used as input for theoretical (numerical or analytical) models, is questionable and not as explicit as for CNTs. In the work by Sakharova et al. [36], the influence of input parameters, obtained based on force field constants assessed by different calculation methods, on the results of finite element (FE) modelling of the mechanical behaviour of BNNTs was studied. A significant scattering of elastic properties was reported. Among the most well-established and commonly used methods for calculation of the force field constants of N-CNTs are those based on UFF (Universal Force Fields) [72], DREIDING force field [73] and ab initio DFT computations, combined with the analytical expressions resulting from molecular mechanics (MM) models [32,33,41]. Jiang and Guo [39] used this methodology, based on the combination of ab initio DFT computations and MM models, to calculate the out-of-plane torsion force constant, which is a component of the torsional resistance force constant, for a wide class of N-CNTs, including BN, AlN, GaN, BP, GaP, InP and SiCNTs.
Although almost all the developments in the mechanical characterization of N-CNTs have been achieved with the help of theoretical approaches, there are still some studies involving experimental evaluation of the elastic properties of N-CNTs. Arenal et al. [74] calculated the Young’s modulus of single-walled boron nitride nanotubes (SWBNNTs) from results of in situ uniaxial compression tests carried out by high-resolution transmission-electron microscopy (HRTEM) and atomic force microscopy (AFM). Chopra and Zettl [75] and Suryavanshi et al. [76] used a transmission electron microscope (TEM) to measure the Young’s modulus of multi-walled boron nitride nanotubes (MWBNNTs) from the thermal vibrational amplitude of a cantilevered nanotube and by the electric-field-induced resonance method, respectively. Golberg et al. [77] and Ghassemi et al. [78], with the help of AFM set up within TEM, performed in situ bending and cycling bending experiments, respectively, to evaluate the Young’s modulus of the MWBNNTs. Tanur et al. [79] studied the mechanical properties of MWBNNTs using a three-point bending technique in AFM and calculated their Young’s and shear moduli. Zhou et al. [80] measured the Young’s modulus of MWBNNTs using a high-order resonance technique within HRTEM. Chen et al. [81] studied the in situ mechanical behaviour of MWBNNTs under axial compression, using a TEM set up with a force transducer holder, and calculated the Young’s modulus of the MWBNNT, based on the directly measured critical compressive force. Hung et al. [16,82] studied the mechanical behaviour of single-walled gallium nitride nanotubes (SWGaNNTs) in compression using the nano-indentation technique and assessed their Young’s modulus and Poisson’s ratio. Stan et al. [15] performed experimental measurements of the Young’s modulus of faceted AlNNTs with triangular cross-section by contact resonance atomic force microscopy (CR-AFM).
The present review is focused on collecting and systematizing recent accomplishments in the mechanical characterization of inorganic non-carbon NTs, by numerical and analytical approaches. The outcomes achieved in the evaluation of the elastic constants (Young’s and shear moduli and Poisson’s ratio) and vibrational properties of the N-CNTs are examined.

2. Atomic Structure of N-CNTs

Figure 1 shows how the parameters characterizing the atomic structure of non-carbon nanotubes are defined, namely the chiral indices, (n, m), the chiral vector, Ch, and the chiral angle, θ, taking as an example a gallium phosphide (GaP) honeycomb lattice. Rolling up a hexagonal diatomic sheet into a cylinder results in the formation of non-carbon NT.
The chiral vector, Ch, and the chiral angle, θ, are expressed in terms of the chiral indices, n and m as follows:
C h = n a 1 + m a 2 ,
θ = sin 1 3 2 m n 2 + nm + m 2 ,
where a 1 and a 2 are the unit vectors of the hexagonal diatomic lattice, constituting atoms A1. and A2 and the chiral indices n and m are integers. The length of the unit vector a is defined as a = 3 a A 1 A 2 , where a A 1 A 2 is the equilibrium bond length. The bond length values for several diatomic nanostructures, which have been reported in the literature, are shown in Table 1. As can be seen, there is no agreement among the research community with regard to the bond length values for N-CNTs.
Considering the bond length, a A 1 A 2 , the N-CNT diameter, D n is defined as follows:
D n = a A 1 A 2 3 ( n 2 + nm + m 2 ) π .
Three main symmetry groups of N-CNTs are defined through the chiral angles, whose magnitudes are within the range of 0° to 30°, as follows:
  • zigzag NTs (n, 0) when θ = 0° and m = 0;
  • armchair NTs (n, n) when θ = 30° and n = m;
  • chiral NTs (n, m) when 0° < θ < 30° and n ≠ m.
The two limiting configurations in terms of the chiral angle, zigzag (θ = 0°) and armchair (θ = 30°), are designated as non-chiral nanotubes. Segments of selected N-CNTs, representing three fundamental groups of symmetry, armchair, zigzag and chiral, are shown in Figure 2.

3. Analysis of the Literature Results

3.1. Elastic Constants of N-CNTs

3.1.1. Young’s and Shear Moduli

Table 2 summarises theoretical results from the literature on the Young’s, E, and shear, G, moduli of different non-carbon NTs. Experimental results on the Young’s modulus evaluation of BNNTs and GaNNTs are also shown. Arenal et al. [74] calculated the Young’s modulus of SWBNNTs from the force—displacement curve obtained in the in situ uniaxial compression test of an individual SWBNNT, carried out in HRTEM and AFM set-ups. Hung et al. [16,82] also computed the Young’s modulus of SWGaNNTs with two different nanotube lengths, Ln, from the force—displacement curve obtained in the uniaxial compression test performed using the Nano-indentation System (NS) with Berkovich indenter. Chopra and Zettl [75] examined a cantilevered MWBNNT by TEM and measured its Young’s modulus from the thermal vibrational amplitude, while Suryavanshi et al. [76] used, for the same propose, the electric-field-induced resonance method inside TEM. Zhou at al. [80] measured the Young’s modulus of the MWBNNTs using high-order resonance modes induced by electric fields, within HRTEM. Chen et al. [81], Golberg et al. [77] and Ghassemi et al. [78] calculated the MWBNNTs Young’s modulus in bending from experimentally obtained critical buckling force [81] and the force—displacement curves [77,78]. Tanur et al. [79] calculated the MWBNNTs Young’s and shear moduli from the results of the in situ three-point bending test, by AFM. The methodology of Stan et al. [15] to evaluate the Young’s modulus of the AlNNTs consisted of determining the contact stiffness between the AFM tip and the nanotube, combined with finite element analysis (FEA) calculations to take into account realistic contact area and, finally, to assess the modulus value from contact stiffness.
The elastic properties of the multi-walled (MW) nanotubes, i.e., the structures constituted by several single-walled NTs (layers) with the diameter of the outer layer, D out , were investigated in a few experimental works [75,76,77,78,79,80,81] in the case of MWBNNTs. It can be seen from this table that studies dealing with the Young’s modulus of N-CNTs are considerably more frequent than those on the evaluation of their shear modulus. The elastic moduli reported by analytical and numerical studies from the literature are obtained for the case of non-chiral (zigzag and armchair) N-CNTs. In the case of SWBNNTs, only Yan et al. [71] reported shear modulus values and Sakharova et al. [36] reported Young’s and shear moduli values. Pinhal et al. [29] calculated Young’s modulus for chiral SWAlNNTs and SWGaNNTs (see Table 2).
Since the calculation of the N-CNTs’ elastic moduli, E and G, almost always requires reliable knowledge of the value of the nanotube wall thickness, tn, several authors have provided the results concerning the surface Young’s ( E s = Et n ) and shear ( G s = Gt n ) moduli [33], this is taken into account in Table 2.
Although the InNNTs have been predicted theoretically [1,21], to our knowledge results regarding their mechanical properties are not available in the literature.
Table 2 shows that the Young’s modulus values obtained by theoretical methods for the BNNTs are in the range E ≈ 0.72–1.1 TPa, which is in reasonable agreement with the experimental results. Such resemblance of the elastic properties of the BNNTs to those of the CNTs (E ≈ 1.0 TPa) makes BNNTs suitable candidates either to replace CNTs in different technological applications [91,92], or to generate novel hybrid structures, based on carbon and boron-nitride NTs.
Moreover, it is evident from Table 2 that the studies to determine the shear modulus, G, of BNNTs are less common than those dealing with the evaluation of their Young’s modulus. The value of G is reported as half of that calculated for E in the respective studies [34,36,61,70,71]. Tanur et al. [79] in the only experimental study on this topic, as far we know, evaluated the shear modulus of MWBNNTs to be about 250 times smaller than their Young’s modulus. This low value was explained by the occurrence of shear between adjacent layers in the structure of the nanotube, due to the geometry and dimensions of the MWBNNTs under study.
With respect to the Young’s modulus, E, of other non-carbon NTs, only a few results were reported for AlNNTs and GaNNTs. The Young’s modulus of the AlNNTs, whose value is about half of that calculated for the BNNTs, was evaluated in the works of Hao et al. [47] (E ≈ 0.350 TPa), Pinhal et al. [29] (E ≈ 0.390 TPa), and Kang and Hwang [40] (E = 0.453 TPa). The results of Hao et al. [47] and Pinhal et al. [29] are in satisfactory agreement with those experimentally obtained for AlNNTs in the work of Stan et al. [15] (E = 0.325 TPa). A more considerable scatter of the Young’s modulus results can be observed for the GaNNTs. While Kang and Hwang [40] and Jeng et al. [38] obtained identical E values in their MD studies, equal to 0.796 TPa and 0.793 TPa, respectively, for (5, 5) GaNNT, Pinhal et al. [29] estimated by DFT calculations the Young’s modulus of (20, 20) GaNNT, equal to 0.383 GPa, and Hung et al. [16,82] in their experimental study using a nano-indentation test obtained the E values of 0.484 TPa and 0.223 TPa for the Young’s modulus of the SWGaNNTs with two different lengths.
In several works, only the surface Young’s modulus, Es, of the non-carbon nanotubes was evaluated. Based on the results of Baumeier et al. [46], Le [43], Genoese et al. [41] and Jiang and Guo [39], it can be concluded that the Es value for the SiCNTs is about 40% lower than that calculated in the respective works for BNNTs. Kochaev [37] found that the surface Young’s moduli of AlNNTs, GaNNTs, AlPNTs and GaPNTs are approximately 30%, 40%, 50% and 70%, respectively, lower than the Es value of BNNTs. Jiang and Guo [39] evaluated the surface Young’s moduli of AlNNTs, GaNNTs and BPNTs to be about 60% lower than that of BNNTs, and those of GaPNTs and InPNTs to be nearly 80% smaller than that of BNNTs. Although the studies concerning the evaluation of the mechanical properties of the N-CNTs other than BNNTs are rare, the preliminary results mentioned above point out that SiC, AlN, GaN, BP and, especially, AlP, GaP and InP nanotubes have low mechanical strength when compared with boron-nitride and carbon NTs. This should be taken into account when designing NT-based devices and hybrid nanostructures, where weaker N-CNTs are combined with other N-CNTs with high mechanical strength or CNTs.
In order to better analyse the evolutions of the Young’s modulus of the N-CNTs, the values of E in Table 2 were plotted, whenever it was possible, as a function of the nanotube diameter, D n , in Figure 3a,b for the case of BNNTs (Figure 3a) and for the case of AlNNTs and SiCNTs (Figure 3b).
Discrepancies in the Young’s modulus values of BNNTs and in their evolutions can be observed in Figure 3. The Young’s modulus results for the BNNTs from Figure 3a, show three trends in the evolution of the value of E with the nanotube diameter, D n :
  • at first, the Young’s modulus decreases and then becomes almost stable for D n > 1.5 nm [36];
  • the E value is nearly constant over the range of nanotube diameters [34,60,62,66,71];
  • at first, the Young’s modulus increases and then becomes almost stable for 0.8 < D n < 1.5 nm [56,59,61].
It is worth noting that the Young’s modulus results obtained by Oh [62] for zigzag (n, 0) BNNTs, are in remarkable consonance with those evaluated by Yan et al. [71] for armchair (n, n), zigzag (n, 0) and chiral (n, m) BNNTs, over a wide range of D n , despite different modelling approaches: CM by Oh [62] and longitudinal vibrations within NCM/MSM by Yan et al. [71]. Both Young’s modulus evolutions reported in the works of Oh and Yan et al. reasonably coincide with the E evolution obtained by Sakharova et al. [36] for armchair (n, n), zigzag (n, 0) and chiral (n, m) BNNTs with D n ≥ 1.715 nm. Satisfactory agreement is observed when comparing the Young’s modulus evolutions assessed by Zhang et al. [61] in their MD study and Ansari et al. [66], who used NCM/MSM approach, for (n, n) and (n, 0) BNNTs with D n ≥ 0.702 nm. A reasonable correspondence is observed between the Young’s modulus values evaluated by Tao et al. [56], who used the MD simulation, and Li and Chou [34], who employed the FE beam model under the NCM/MSM approach, for zigzag (n, 0) BNNTs. In most studies, the Young’s modulus evolution with nanotube diameter for (n, n) armchair BNNTs is separated from that for (n, 0) zigzag BNNTs [34,54,56,59,62,66], with the E values for armchair NTs being higher than those for zigzag NTs. Santos et al. [59] reported the same Young’s modulus evolution for both (n, n) and (n, 0) BNNTs. Yan et al. [71] and Sakharova et al. [36] found that the evolutions of E with D n are described by a unique trend for armchair, zigzag and chiral nanotubes.
As can be seen from Figure 3b, the evolutions of the Young’s modulus as a function of the nanotube diameter, D n , reported by Moon et al. [51] for the SiCNTs, shows a unique trend for (n, n) and (n, 0) zigzag nanotubes, where E increases for small D n and then tends to a nearly constant value. On the contrary, the E value of the monocrystalline SiCNTs slightly decreases at the beginning and then becomes stable with increasing D n [54]. In the case of AlNNTs, the Young’s modulus evolutions in nanotube diameter can be separated for (n, n) armchair and (n, 0) zigzag NTs, as reported by Hao at al. [47] (see Figure 3b). The E value of (n, n) AlNNTs is almost constant for all range of D n , while E of (n, 0) AlNNTs substantially increases for D n   1.554 nm and after that it stabilizes. The E values obtained by Hao at al. [47] for (n, n) AlNNTs are about 9% lower than those calculated by Pinhal et al. [29] for individual (20, 20), (20, 0) and (20, 10) AlNNTs.
To further analyse, the Young’s modulus results of the BNNTs in Table 2, were plotted as a function of the nanotube aspect ratio, L n / D n , and the nanotube diameter, D n (see Figure 4).
/ D n [31,36,65,68,69].
The E evolutions as a function of L n / D n reported by Sakharova et al. [36] for (10, 10) and (18, 0) BNNTs are in satisfactory agreement with those obtained by Choyal et al. [31] for (10, 10) and (17, 0) BNNTs and by Salvati et al. [65] for (20, 0) BNNT, in case of nanotube aspect ratios L n / D n > 8. Zakaria [68] and Giannopoulos et al. [69], in their numerical simulation studies within NCM/MSM approach, obtained for (8, 8) and (21, 0) BNNTs Young’s modulus values similar in all range of the nanotube aspect ratio (see, Figure 4), with differences of less than about 8% in the E values. Note that, despite using dissimilar methods to model the B–N covalent bond, Giannopoulos et al. [69] used spring elements and Zakaria [68] represented the interatomic bond as two-section beam, in both works the elastic properties of the elements to be used as input for the numerical simulation were computed based on the same values of the constants kr, kθ and kτ.
Figure 5 shows the evolution of the surface Young’s modulus, E s , with the nanotube diameter, D n , for BNNTs (Figure 5a), SiCNTs (Figure 5b), AlN, GaNNTs (Figure 5c), and BP, AlP, GaP, InPNTs (Figure 5d). Regarding Figure 5a,c,d, it can be noted that there are two characteristic trends of the evolution of the surface Young’s modulus, E s , with the nanotube diameter, D n , one for the nitrides (BN, AlN, GaN) and another for phosphides (BP, AlP, GaP, InP). In the first case, the E s value slightly increases with increasing of D n , and then becomes nearly constant for nanotube diameters D n   0.7 nm [39,41,43,46]. The second trend is reported by Kochaev [37] for BN, AlN, GaN, AlP, GaPNTs, consists of a significant increase of the surface Young’s modulus for D n   1.2 nm (armchair NTs) and D n   0.9 nm (zigzag NTs), and then a strong decrease of E s after reaching the maximum.
In the case of SiC, the only representative of carbide NTs in the present analysis, the E s evolution follows a trend in which its value increases for small nanotube diameters, D n , of less than 0.7 nm, and remains approximately constant for large diameters (Figure 5b).
Concerning the BNNTs, Le [43], Baumeier et al. [46] and Jiang and Guo [39] provided obtained comparable E s results (see, Figure 5a), Genoese et al. [41], employing “stick-and-spring” model similar to that used by Jiang and Guo [39], evaluated the values of E s at about 6% lower than those reported by the latter. Such difference is probably related to the approaches used for computation of the force-field constants. The maximum value of the surface Young’s modulus obtained by Kochaev [37] for BNNTs is notably higher than the values calculated by other authors [39,41,43,46]. The same is true for the surface Young’s modulus results of AlN, GaN and GaP nanotubes, when compared with the values reported by Kochaev [37] and Jiang and Guo [39] (see, Figure 5c,d). Jiang and Guo [39] found that the E s values of armchair and zigzag structures are approximately identical for AlN, GaN, BP, GaP, and InPNTs. Regarding the surface Young’s modulus of SiCNTs, there is a good agreement between the results reported by Baumeier et al. [46] and Jiang and Guo [39]. The values of E s calculated by Genoese et al. [41] and Le [43] are, respectively, about 7% and 11% lower than those of Baumeier et al. [46] and Jiang and Guo [39], while the E s value assessed by Setoodeh et al. [52] is about 8% higher (see, Figure 5b).
The results of Table 2 concerning the evolutions of the shear, G, and surface shear, G s , moduli with the nanotube diameter, D n , are shown in Figure 6a,b, respectively. With an exception of the G s values for SiCNTs reported by Genoese et al. [41], all other shear and surface shear moduli results plotted in Figure 6 are related to BNNTs. As in the case of the Young’s modulus, three trends of the evolution of G and G s with D n can be found in the literature:
  • at the beginning, the shear modulus decreases along with increase in the diameter of the BNNTs and then tends to stabilize at high values of D n [36,55];
  • the shear modulus almost does not vary within the entire range of BNNT diameters [61,70,71];
  • at the beginning, the shear modulus slightly increases and then becomes nearly constant for high values of D n [33,34,41,50,59].
Most authors, such as Li and Chou [34], Sakharova et al. [36], Zhang et al. [61], Yan and Liew [70] and Yan et al. [71] pointed to the value of 0.426 ± 0.06 TPa as the shear modulus of BNNTs (see, Figure 6a). In these works [34,36,61,70,71], the G values for (n, n) armchair, (n, 0) zigzag and (n, m) chiral structures (when those are reported, as for example [36,71]) are also practically identical for nanotube diameters D n > 0.8 nm.
Santos et al. [59] and Xiong and Tian [50] reported similar G values for BNNTs, especially for nanotube diameters D n > 2 nm (see, Figure 6a). The shear modulus values obtained by Santos et al. [59] and Xiong and Tian [50] are about 25% lower than those presented [34,36,61,70,71] (see, Figure 6b).
With regard to surface shear modulus, Genoese et al. [41] reported the value of G s of SiCNTs at about 40% lower than that of BNNTs. In the work of Jiang and Guo [39], the surface shear modulus of the zigzag structure of SiCNTs and BNNTs are higher when compared to those of the armchair structure for D n   1.95 nm. The G s values evaluated by Jiang and Guo [39] are lower than those obtained by Genoese et al. [41]. These studies share the modelling approach, but the methods for calculating the force-field constants and the surface shear modulus are different, which could explain the discrepancy of the results. Jiang and Guo [39] introduced the out-of-plane inversion force constant to describe the bond torsion, which is essential to calculate the shear modulus, while Genoese et al. [41] assumed a continuum thin shell model to assess G s .
It is worth noting that the scattering observed in the values of the elastic modulus of the N-CNTs can be attributed to the different modelling and calculation methods used. The main limitation of the NCM/MSM approach is related to accurate determination of the force field constants. Although knowledge of these constants is not necessary to employ the atomistic approach, the latter is generally time consuming and computationally expensive. Another restriction to the calculation of the elastic moduli is associated with the value of the nanotube wall thickness, t n . Except for the case of BNNTs, there is an apparent lack of information in the literature about the value of t n for other N-CNTs. For this reason, the surface elastic moduli of N-CNTs, different from BNNTs, were mainly assessed, which makes it difficult to compare the available results.

3.1.2. Poisson’s Ratio

The results of the Poisson’s ratio of N-CNTs available in the literature are summarized in Table 3. The only experimental value in the table was obtained by Hung et al. [16,82] in the uniaxial compression test performed with the help of the nano-indentation system (NS) for SWGaNNTs.
As can be seen in Table 3, the Poisson’s ratio values for GaNNTs evaluated by Jeng et al. [38], in their MD study, and by Jiang and Guo [39], using the analytical model under NCM/MSM approach, are in acceptable agreement with the ν value calculated by Hung et al. [16] from the experimental results of the nano-indentation test. The difference between the values of ν reported by Jeng et al. [39] and Jiang and Guo [39] and those experimentally obtained [16] is around 8% and 15%, respectively.
To facilitate the analysis of the literature results from Table 3, the Poisson’s ratio of N-CNTs was plotted as a function on the nanotube diameter, D n , as shown in Figure 7. The stabilized values of Poisson’s ratio found in the literature for BNNTs can be separated into two groups (Figure 7a). Jiang and Guo [39], Genoese et al. [41] and Ansari et al. [66] obtained the values of ν in the range of 0.217 to 0.239, for the nanotube diameter D n   1.2 nm. The correspondent evolutions of the ν value with D n show that the Poisson’s ratio decreases for small BNNT diameters and tends to stabilize with the increase of D n , the evolutions of ν reported by Jiang and Guo [39] for (n, n) armchair, and Ansari et al. [66] for (n, n) armchair and (n, 0) zigzag BNNTs, being in especially good agreement (see, Figure 7a). For all the above mentioned results, the ν value for (n, 0) nanotubes is higher than that obtained for (n, n) nanotubes.
Sakharova et al. [36], Oh [62] and Verma et al. [55] evaluated the BNNT Poisson’s ratio in the range of 0.14–0.17 for nanotubes with diameter D n   1.2 nm (see, Figure 7a). Regarding trends in the evolution of ν with D n , Sakharova et al. [36] reported that for D n < 1.5 nm the Poisson’s ratio value decreases for the cases of (n, n) armchair and (n, m) chiral BNNTs, but increases in the case of (n, 0) zigzag BNNTs. When the nanotube diameter becomes higher than 1.5 nm, ν tends to a nearly constant value for (n, n), (n, 0) and (n, m) BNNTs.
Thus, according to Sakharova et al. [36], for BNNT diameters D n < 1.5 nm, the value of ν is clearly influenced by the chiral angle of the nanotube and increases from zigzag structure (θ = 0°) to armchair (θ = 30°). Nevertheless, Oh [62] found that the Poisson’s ratio of both BNNTs, (n, n) armchair and (n, 0) zigzag, decreases for D n < 1.5 nm and afterwards becomes constant, the value of ν being greater for the (n, 0) structures.
Concerning N-CNTs different from BNNTs, it can be concluded from Figure 7b,c that the evolutions of the Poisson’s ratio, ν, with nanotube diameter, D n , follow a similar trend, for which the ν value decreases for small nanotube diameters and then tends to stabilize with increasing D n (cases of SiC, AlN, GaN, BP, GaP and InPNTs). For all the above mentioned evolutions of the ν value with D n , ν is higher for (n, 0) zigzag NTs. Jiang and Guo [39] established, in their study for AlN, GaN, BP, GaP and InPNTs, that the greater the value of the bond length, a A 1 A 2 , of the diatomic hexagonal nanostructure (see, Table 1), the greater the value of the nanotube diameter, D n , for which the Poisson’s ratio becomes stable (see, Figure 7b,c).
Genoese et al. [41] reported the value of ν for SiCNTs to be about 3.5 times higher than that calculated by Jiang and Guo [39] (see, Figure 7b). Such a large scatter in the ν values cannot be attributed solely to the different methods for computation of the force field constants, under a similar modelling approach in both studies, because the results of the Poisson’s ratio for BNNTs, which were also obtained in these studies, do not show such considerable dissimilarity (see, Figure 7a). In fact, Jiang and Guo [39] calculated the force field constants for BN and SiC nanotubes with the same calculation method. Genoese et al. [41] took into account the assumption of the so-called buckled surface of SiC nanotube, which occurs due to the fact that Si and C atoms form two coaxial cylinders [46,93,94]. This hypothesis modified the way of calculating the force field constants for SiCNTs compared to BNNTs.
The results of Jiang and Guo [39] obtained for other N-CNTs can be compared only for the cases of GaNNTs and GaPNTs, reported by Jeng et al. [38] and Kochaev [37], respectively. Jeng et al. [38] obtained the Poisson’s ratio values for (5, 5) and (9, 0) GaNNTs 12% and 29% lower, respectively, than those evaluated by Jiang and Guo [39] for GaNNTs with comparable diameters. In the work by Kochaev [37], the Poisson’s ratio of (10, 10) and (10, 0) GaPNTs is 19% and 17% higher, respectively, than the value of ν estimated by Jiang and Guo [39], for GaP nanotubes with comparable diameters.
It can be noted that there is large scattering of the ν values reported in the literature for N-CNTs, regardless of the modelling and calculation approaches used for this purpose.

3.2. Vibtational Properties of N-CNTs

The developing of innovative nanoelectromechanical devices, such as oscillators, amplifiers, mass/charge detectors ultra-high frequency resonators and resonant-based nano-mechanical sensors [95,96,97], as well their top-notch performance, requires constituents with appropriate vibrational characteristics. The N-CNTs, in particular the BNNTs, offer the necessary properties for nanoscale electromechanical applications, specifically sensing ones. In this context, understanding the dynamical properties of the N-CNTs is crucial for creating novel nano-devices with desirable functioning.
The studies reported so far on the dynamic properties of N-CNTs mainly addressed the theoretical analysis of the vibrational behaviour of BNNTs [34,48,57,63,67,68,69,70,71,98,99] and SiCNTs [100,101].
Table 4 summarizes the available theoretical results on the vibrational properties of N-CNTs. Since in general the works deal with the calculation of the fundamental frequencies of nanotubes, the first fundamental frequency, fn1, is herein chosen for the comparison purpose. The vibration analysis of NTs is typically performed assuming different support constraints, depending on their potential applications, i.e., simply supported (SS), clamped-free (CF) and clamped-clamped (CC) boundary conditions, which are indicated in Table 4.
In order to simplify the understanding of the results from Table 4, the first fundamental frequency, fn1, was plotted as a function of the NT aspect ratio, L n / D n , and NT diameter, D n , in Figure 8 and Figure 9, respectively. It can be concluded from Figure 8a,b and Table 4 that the values of fn1 are higher for the case of clamped-clamped (CC) boundary conditions than for the case of clamped-free (CF) boundary conditions [34,57,67,69,71,100]. The first fundamental frequency decreases with the nanotube aspect ratio and the decreasing rate is higher when the clamped-clamped boundary conditions is applied. Regarding the fn1 value of BNNTs, the results reported by Li and Chou [34] and Panchal et al. [67] for clamped-free support, and by Zakaria [68] for simply supported NTs are in a particularly good agreement for all range of the aspect ratio, L n / D n , considered (see Figure 8a).
The values of fn1 obtained by Giannopoulos et al. [69] under clamped-free boundary conditions are about 30% higher when compared with fn1 calculated in the abovementioned studies [34,67,68] for L n / D n ≥ 10. The fn1 values reported by Li and Chou [34], Panchal et al. [67] and Giannopoulos et al. [69], considering clamped-clamped boundary conditions, are in satisfactory agreement with each other for L n / D n ≥ 15. For the rest of the results from Table 4 and Figure 8, there is a considerable scattering of the fn1 values for both BNNTs and SiCNTs. Although the representation of fn1 as a function of the nanotube diameter, D n , is less frequent, some results for the BNNTs can be found in the literature, which are shown in Figure 9.
The main trend of the evolution of the first fundamental frequency with the nanotube diameter is the decrease of fn1 when D n increases, whatever the boundary conditions [34,68,69]. According to another trend, reported by Yan et al. [71], the fn1 value slightly increases for nanotube diameters D n < 0.7 nm and then is almost constant with increasing Dn.
It is worth mentioning that most of the reported results on dynamic properties of the N-CNTs are limited to the vibrational analysis of some BNNTs and the determination of their first fundamental frequency. That is, the vibrational properties of N-CNTs are less commonly investigated than their elastic moduli and Poisson ratio. In this context, considering forthcoming applications, systematic studies are needed on the vibrational behaviour of a wider set of N-CNTs with a wide range of chiral indices and diameters.

4. Conclusions

The present review collects accomplishments in the development of the modelling of elastic properties of N-CNTs. Works in the literature, where the static (Young’s and shear moduli, and Poisson’s ratio) and dynamic (first fundamental frequency) elastic properties were evaluated, are comprehensively analysed. Despite considerable achievements in predicting the elastic properties of BNNTs by numerical simulation and analytical approaches, the other N-CNTs have received much less research attention. Theoretical studies show scattering in the values of elastic constants and the first fundamental frequency due to different modelling and calculation approaches. Most of the results on the elastic properties of N-CNTs were obtained using the atomistic and NCM/MSM approaches. Although the NCM/MSM approach proves to be cost-effective for modelling the mechanical behaviour of N-CNTs, the main challenge in its application remains the suitable choice of the force field constants to simulate a bond between two atoms in the diatomic nanotube structure. If for the BNNTs this choice is ambiguous, the methods for computing the force field constants for other N-CNTs are scarce and not well explored.
In this context, future lines of research on the evaluation of the elastic properties of N-CNTs should include the development of the NCM/MSM approach, focusing on the correct computation of the force field constants. It is expected that the upcoming investigation will provide a benchmark regarding the determination of the mechanical properties of a wide group of N-CNTs by theoretical (numerical and analytical) methods.

Author Contributions

Conceptualization, N.A.S. and J.M.A.; methodology, N.A.S., J.M.A. and A.F.G.P.; writing—original draft preparation, N.A.S.; writing—original manuscript, J.M.A. and N.A.S.; writing—review and editing, all the authors; supervision, N.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research is sponsored by FEDER funds through the program COMPETE—Programa Operacional Factores de Competitividade—and by national funds through FCT, Fundação para a Ciência e a Tecnologia, under the project UIDB/00285/2020.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author after obtaining permission of authorized person.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Qian, Z.; Hou, S.; Zhang, J.; Li, R.; Shen, Z.; Zhao, X.; Xue, Z. Stability and electronic structure of single-walled InN nanotubes. Physica E 2005, 30, 81–85. [Google Scholar] [CrossRef]
  2. Ambacher, O. Growth and applications of group III-nitrides. J. Phys. D 1998, 31, 2653–2710. [Google Scholar] [CrossRef]
  3. Orton, J.W.; Foxon, C.T. Group III nitride semiconductors for short wavelength light-emitting devices. Rep. Prog. Phys. 1998, 61, 1–75. [Google Scholar] [CrossRef]
  4. Bachtold, A.; Hadley, P.; Nakanishi, T.; Dekker, C. Logic circuits with carbon nanotube transistors. Science 2001, 294, 1317–1320. [Google Scholar] [CrossRef] [PubMed]
  5. Bhuiyan, A.G.; Hashimoto, A.; Yamamoto, A. Indium nitride (InN): A review ongrowth, characterization, and properties. J. Appl. Phys. 2003, 94, 2779–2808. [Google Scholar] [CrossRef]
  6. Roudbari, M.A.; Ansari, R. Single-walled boron nitride nanotube as nano-sensor. Contin. Mech. Thermodyn. 2020, 32, 729–748. [Google Scholar] [CrossRef]
  7. Kong, J.; Franklin, N.R.; Zhou, C.; Chapline, M.G.; Peng, S.; Cho, K.; Dai, H. Nanotube molecular wires as chemical sensors. Science 2000, 287, 622–625. [Google Scholar] [CrossRef]
  8. Cohen, M.L.; Zettl, A. The physics of boron nitride nanotubes. Phys. Today 2010, 63, 34–38. [Google Scholar] [CrossRef] [Green Version]
  9. Zhuang, H.L.; Singh, A.K.; Hennig, R.G. Computational discovery of single-layer III–V materials. Phys. Rev. B 2013, 87, 165415–165418. [Google Scholar] [CrossRef]
  10. Tenne, R. Inorganic Nanotubes and Fullerene-Like Materials. Chem. Eur. J. 2002, 8, 5296–5304. [Google Scholar] [CrossRef]
  11. Wong, S.; Joselevich, E.; Woolley, A.; Cheung, C.; Lieber, C.M. Covalently functionalized nanotubes as nanometre- sized probes in chemistry and biology. Nature 1998, 394, 52–55. [Google Scholar] [CrossRef] [Green Version]
  12. 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]
  13. Fan, Y. Formation of crystalline AlN nanotubes by a roll-up approach. Mater. Lett. 2011, 65, 1900–1902. [Google Scholar] [CrossRef]
  14. Wu, Q.; Hu, Z.; Wang, X.; Lu, Y.; Chen, X.; Xu, H.; Chen, Y. Synthesis and characterization of faceted hexagonal aluminum nitride nanotubes. J. Am. Chem. Soc. 2003, 125, 10176–10177. [Google Scholar] [CrossRef]
  15. Stan, G.; Ciobanu, C.V.; Thayer, T.P.; Wang, G.T.; Creighton, J.R.; Purushotham, K.P.; Bendersky, L.A.; Cook, R.F. Elastic moduli of faceted aluminum nitride nanotubes measured by contact resonance atomic force microscopy. Nanotechnology 2009, 20, 035706. [Google Scholar] [CrossRef] [Green Version]
  16. Hung, S.C.; Su, Y.K.; Fang, T.H.; Chang, S.J.; Ji, L.W. Buckling instabilities in GaN nanotubes under uniaxial compression. Nanotechnology 2005, 16, 2203–2208. [Google Scholar] [CrossRef]
  17. Goldberger, J.; He, R.; Zhang, Y.; Lee, S.; Yan, H.; Choi, H.-J.; Peidong, Y. Single-crystal gallium nitride nanotubes. Nature 2003, 422, 599–602. [Google Scholar] [CrossRef]
  18. Wu, Q.; Hu, Z.; Liu, C.; Wang, X.; Chen, Y.; Lu, Y. Synthesis and optical properties of gallium phosphide nanotubes. J. Phys. Chem. B 2005, 109, 19719–19722. [Google Scholar] [CrossRef]
  19. Bakkers, E.P.A.M.; Verheijen, M.A. Synthesis of InP Nanotubes. J. Am. Chem. Soc. 2003, 125, 3440–3441. [Google Scholar] [CrossRef]
  20. Sun, X.H.; Li, C.P.; Wong, W.K.; Wong, N.B.; Lee, C.S.; Lee, S.T.; Teo, B.K. Formation of silicon carbide nanotubes and nanowires via reaction of silicon (from disproportionation of silicon monoxide) with carbon nanotubes. J. Am. Chem. Soc. 2002, 124, 14464–14471. [Google Scholar] [CrossRef]
  21. Peng, Q.; Sun, X.; Wang, H.; Yang, Y.; Wene, X.; Huang, C.; Liu, S.; De, S. Theoretical prediction of a graphene-like structure of indium nitride: A promising excellent material for optoelectronics. Appl. Mater. Today 2017, 7, 169–178. [Google Scholar] [CrossRef]
  22. Mirzaei, M.; Giahi, M. Computational studies on boron nitride and boron phosphide nanotubes: Density functional calculations of boron-11 electric field gradient tensors. Physica E 2010, 42, 1667–1669. [Google Scholar] [CrossRef]
  23. Mirzaei, M.; Meskinfam, M. Computational studies of effects of tubular lengths on the NMR properties of pristine and carbon decorated boron phosphide nanotubes. Solid State Sci. 2011, 13, 1926–1930. [Google Scholar] [CrossRef]
  24. Mirzaei, M.; Mirzaei, M. Aluminum phosphide nanotubes: Density functional calculations of aluminum-27 and phosphorus-31 chemical shielding parameters. J. Mol. Struct. Theochem. 2010, 951, 69–71. [Google Scholar] [CrossRef]
  25. Erkoç, S. Semi-empirical SCF-MO calculations for the structural and electronic properties of single-wall InP nanotubes. J. Mol. Struct. Theochem. 2004, 676, 109–113. [Google Scholar] [CrossRef]
  26. Mirzaei, M.; Mirzaei, M. A computational study of gallium phosphide nanotubes. Physica E 2011, 43, 1343–1345. [Google Scholar] [CrossRef]
  27. Mirzaei, M.; Yousefi, M.; Meskinfam, M. Studying (n, 0) and (m,m) GaP nanotubes (n = 3−10 and m = 2−6) through DFT calculations of Ga-69 quadrupole coupling constants. Solid State Sci. 2012, 14, 801–804. [Google Scholar] [CrossRef]
  28. Guo, Y.; Yan, X.; Yang, Y. First-principles study of narrow single-walled GaN nanotubes. Phys. Lett. A 2009, 373, 367–370. [Google Scholar] [CrossRef]
  29. Pinal, G.B.; Marana, N.L.; Fabris, G.S.L.; Sambrano, J.R. Structural, electronic and mechanical properties of single-walled AlN and GaN nanotubes via DFT/B3LYP. Theor. Chem. Acc. 2019, 138, 1–11. [Google Scholar] [CrossRef]
  30. Marana, N.L.; Pinhal, G.B.; Laranjeira, J.A.S.; Buzolin, P.G.C.; Longo, E.; Sambrano, J.R. Strain-induced novel properties of alloy nitride nanotubes. Comput. Mater. Sci. 2020, 177, 109589–109598. [Google Scholar] [CrossRef]
  31. Choyal, V.K.; Choyal, V.; Nevhal, S.; Bergaley, A.; Kundalwal, S.I. Effect of aspects ratio on Young’s modulus of boron nitride nanotubes: A molecular dynamics study the continuum mechanics. Mater. Today Proc. 2020, 26, 1–4. [Google Scholar] [CrossRef]
  32. Ansari, R.; Rouhi, S.; Mirnezhad, M.; Aryayi, M. Stability characteristics of single-walled boron nitride nanotubes. Arch. Civ. Mech. Eng. 2015, 15, 162–170. [Google Scholar] [CrossRef]
  33. Jiang, L.; Guo, W. A molecular mechanics study on size-dependent elastic properties of single-walled boron nitride nanotubes. J. Mech. Phys. Solids 2011, 59, 1204–1213. [Google Scholar] [CrossRef]
  34. Li, C.; Chou, T.W. Static and dynamic properties of single-walled boron nitride nanotubes. J. Nanosci. Nanotechnol. 2006, 6, 54–60. [Google Scholar] [CrossRef]
  35. Vijayaraghavan, V.; Zhang, L. Consistent computational modeling of mechanical properties of carbon and boron nitride nanotubes. JOM 2020, 72, 3968–3976. [Google Scholar] [CrossRef]
  36. Sakharova, N.A.; Antunes, J.M.; Pereira, A.F.G.; Chaparro, B.M.; Fernandes, J.V. On the determination of elastic properties of single-walled boron nitride nanotubes by numerical simulation. Materials 2021, 14, 3183–3210. [Google Scholar] [CrossRef]
  37. Kochaev, A. Elastic properties of noncarbon nanotubes as compared to carbon nanotubes. Phys. Rev. B 2017, 96, 155428–155435. [Google Scholar] [CrossRef]
  38. Jeng, Y.R.; Tsai, P.C.; Fang, T.H. Molecular dynamics investigation of the mechanical properties of gallium nitride nanotubes under tension and fatigue. Nanotechnology 2004, 15, 1737–1744. [Google Scholar] [CrossRef]
  39. Jiang, L.; Guo, W. Analytical solutions for elastic binary nanotubes of arbitrary chirality. Acta Mech. Sin. 2016, 32, 1046–1057. [Google Scholar] [CrossRef]
  40. Kang, J.W.; Hwang, H.J. Atomistic study of III-nitride nanotubes. Comput. Mater. Sci. 2004, 31, 237–246. [Google Scholar] [CrossRef]
  41. Genoese, A.; Genoese, A.; Salerno, G. On the nanoscale behaviour of single-wall C, BN and SiC nanotubes. Acta Mech. 2019, 230, 1105–1128. [Google Scholar] [CrossRef]
  42. Mercan, K.; Civalek, O. Buckling analysis of Silicon carbide nanotubes (SiCNTs) with surface effect and nonlocal elasticity using the method of HDQ. Comp. Part B Eng. 2017, 114, 34–45. [Google Scholar] [CrossRef]
  43. Le, M.Q. Young’s modulus prediction of hexagonal nanosheets and nanotubes based on dimensional analysis and atomistic simulations. Meccanica 2014, 49, 1709–1719. [Google Scholar] [CrossRef]
  44. Sakharova, N.A.; Antunes, J.M.; Pereira, A.F.G.; Fernandes, J.V. Developments in the evaluation of elastic properties of carbon nanotubes and their heterojunctions by numerical simulation. AIMS Mater. Sci. 2017, 4, 706–737. [Google Scholar] [CrossRef]
  45. Rafiee, R.; Moghadam, R.M. On the modelling of carbon nanotubes: A critical review. Compos. Part B Eng. 2014, 56, 435–449. [Google Scholar] [CrossRef]
  46. Baumeier, B.; Kruger, P.; Pollmann, J. Structural, elastic, and electronic properties of SiC, BN, and BeO nanotubes. Phys. Rev. B 2007, 76, 85407–85417. [Google Scholar] [CrossRef]
  47. Hao, J.H.; Wang, Y.F.; Yin, Y.H.; Jiang, R.; Wang, Y.F.; Jin, Q.H. An ab initio study of the size-dependent mechanical behavior of single-walled AlN nanotubes. Solid State Sci. 2015, 45, 30–34. [Google Scholar] [CrossRef]
  48. Chandra, A.; Patra, P.K.; Bhattacharya, B. Thermal vibration characteristics of armchair boron-nitride nanotubes. J. Appl. Phys. 2015, 118, 234503–234508. [Google Scholar] [CrossRef]
  49. Jeng, Y.R.; Tsai, P.C.; Fang, T.H. Tensile deformation of tubular structures of nitride-based nanotubes: Brittle and weak behavior. Tamkang J. Sci. Eng. 2005, 8, 191–195. [Google Scholar]
  50. Xiong, Q.; Tian, X.G. Torsional properties of hexagonal boron nitride nanotubes, carbon nanotubes and their hybrid structures: A molecular dynamics study. AIP Adv. 2015, 5, 107215–107227. [Google Scholar] [CrossRef]
  51. Moon, W.H.; Ham, J.K.; Hwang, H.J. Mechanical properties of SiC nanotubes. In In Technical Proceedings of the 2003 Nanotechnology Conference and Trade Show, San Francisco, CA, USA, 23–27 February 2003. [Google Scholar]
  52. Setoodeh, A.R.; Jahanshahi, M.; Attariani, H. Atomistic simulations of the buckling behavior of perfect and defective silicon carbide nanotubes. Comput. Mater. Sci. 2009, 47, 388–397. [Google Scholar] [CrossRef]
  53. Pan, H.; Si, X. Molecular dynamics simulations of diameter dependence tensile behavior of silicon carbide nanotubes. Physica B 2009, 404, 1809–1812. [Google Scholar] [CrossRef]
  54. Zhou, R.L.; Wang, L.; Pan, B.C. Elastic and Melting Properties of Crystalline SiC Nanotubes. J. Phys. Chem. C 2010, 114, 8199–8205. [Google Scholar] [CrossRef]
  55. Verma, V.; Jindal, V.K.; Dharamvir, K. Elastic moduli of a boron nitride nanotube. Nanotechnology. 2007, 18, 435711–435717. [Google Scholar] [CrossRef]
  56. Tao, J.; Xu, G.; Sun, Y. Elastic properties of boron-nitride nanotubes through an atomic simulation method. Math. Prob. Eng. 2015, 2015, 240547. [Google Scholar] [CrossRef]
  57. Ansari, R.; Ajori, S. A molecular dynamics study on the vibration of carbon and boron nitride double-walled hybrid nanotubes. Appl. Phys. A 2015, 120, 1399–1406. [Google Scholar] [CrossRef]
  58. Wang, Z.; Zu, X.; Gao, F.; Weber, W.J. Mechanical behavior of gallium nitride nanotubes under combined tension-torsion: An atomistic simulation. J. Appl. Phys. 2008, 103, 13505–13509. [Google Scholar] [CrossRef]
  59. Santosh, M.; Maiti, P.K.; Sood, A.K. Elastic properties of boron nitride nanotubes and their comparison with carbon nanotubes. J. Nanosci. Nanotechnol. 2009, 9, 1–6. [Google Scholar] [CrossRef]
  60. Hernandez, E.; Goze, C.; Bernier, P.; Rubio, A. Elastic properties of C and BxCyNz composite nanotubes. Phys. Rev. Lett. 1998, 80, 4502–4505. [Google Scholar] [CrossRef] [Green Version]
  61. Zhang, D.B.; Akatyeva, E.; Dumitrica, T. Helical BN and ZnO nanotubes with intrinsic twisting: An objective molecular dynamics study. Phys. Rev. B 2011, 84, 115431–115438. [Google Scholar] [CrossRef] [Green Version]
  62. Oh, E.S. Elastic properties of boron-nitride nanotubes through the continuum lattice approach. Mater. Lett. 2010, 64, 859–862. [Google Scholar] [CrossRef]
  63. Panchal, M.B.; Upadhyay, S.H.; Harsha, S.P. Vibration analysis of single walled boron nitride nanotube based nanoresonators. J. Nanotechnol. Eng. Med. 2012, 3, 31004–31009. [Google Scholar] [CrossRef]
  64. Song, J.; Wu, J.; Huang, Y.; Hwang, K.C. Continuum modeling of boron nitride nanotubes. Nanotechnol. 2008, 19, 445705–445715. [Google Scholar] [CrossRef]
  65. Salavati, M.; Ghasemi, H.; Rabczuk, T. Electromechanical properties of boron nitride nanotube: Atomistic bond potential and equivalent mechanical energy approach. Comp. Mater. Sci. 2018, 149, 460–465. [Google Scholar] [CrossRef]
  66. Ansari, R.; Mirnezhad, M.; Sahmani, S. Prediction of chirality- and size-dependent elastic properties of single-walled boron nitride nanotubes based on an accurate molecular mechanics model. Superlattice Microst. 2015, 80, 196–205. [Google Scholar] [CrossRef]
  67. Panchal, M.B.; Upadhyay, S.H.; Harsha, S.P. An efficient finite element model for analysis of single walled boron nitride nanotube-based resonant nanomechanical sensors. NANO Brief Rep. Rev. 2013, 8, 1350011–1350027. [Google Scholar] [CrossRef]
  68. Zakaria, A.Z. A two-section beam element to model the B-N covalent bonds in boron nitride nanotubes. Mater. Res. Bull. 2022, 145, 111533–111540. [Google Scholar] [CrossRef]
  69. Giannopoulos, G.I.; Kontoni, D.-P.N.; Georgantzinos, S.K. Efficient FEM simulation of static and free vibration behaviour of single walled boron nitride nanotubes. Superlattice Microst. 2016, 96, 111–120. [Google Scholar] [CrossRef]
  70. Yan, J.W.; Liew, K.M. Predicting elastic properties of single-walled boron nitride nanotubes and nanocones using an atomistic-continuum approach. Compos. Struct. 2015, 125, 489–498. [Google Scholar] [CrossRef]
  71. Yan, J.W.; He, J.B.; Tong, L.H. Longitudinal and torsional vibration characteristics of boron nitride nanotubes. J. Vib. Eng. Technol. 2019, 7, 205–215. [Google Scholar] [CrossRef]
  72. Rappé, A.K.; Casewit, C.J.; Colwell, K.S.; Goddard, W.A.; Skid, W.M. UFF, a full periodic table force field for molecular mechanics and molecular dynamics simulations. J. Am. Chem. Soc. 1992, 114, 10024–10039. [Google Scholar] [CrossRef]
  73. Mayo, S.L.; Barry, D.; Olafson, B.D.; Goddard, W.A. DREIDING: A generic force field for molecular simulations. J. Phys. Chem. 1990, 94, 8897–8909. [Google Scholar] [CrossRef]
  74. Arenal, R.; Wang, M.S.; Xu, Z.; Loiseau, A.; Golberg, D. Young modulus, mechanical and electrical properties of isolated individual and bundled single-walled boron nitride nanotubes. Nanotechnology 2011, 22, 265704–265710. [Google Scholar] [CrossRef]
  75. Chopra, N.G.; Zettl, A. Measurement of the elastic modulus of a multi-wall boron nitride nanotube. Solid State Commun. 1998, 105, 297–300. [Google Scholar] [CrossRef]
  76. Suryavanshi, A.P.; Yu, M.F.; Wen, J.; Tang, C.; Bando, Y. Elastic modulus and resonance behavior of boron nitride nanotubes. Appl. Phys. Lett. 2004, 84, 2527–2529. [Google Scholar] [CrossRef] [Green Version]
  77. Golberg, D.; Costa, P.M.F.J.; Lourie, O.; Mitome, M.; Bai, X.; Kurashima, K.; Zhi, C.; Tang, C.; Bando, Y. Direct force measurements and kinking under elastic deformation of individual multiwalled boron nitride nanotubes. Nano Lett. 2007, 7, 2146–2151. [Google Scholar] [CrossRef]
  78. Ghassemi, H.M.; Lee, C.H.; Yap, Y.K.; Yassar, R.S. Real-time fracture detection of individual boron nitride nanotubes in severe cyclic deformation processes. J. Appl. Phys. 2010, 108, 24314–24318. [Google Scholar] [CrossRef] [Green Version]
  79. Tanur, A.E.; Wang, J.; Reddy, A.L.M.; Lamont, D.N.; Yap, Y.K.; Walker, G.C. Diameter-dependent bending modulus of individual multiwall boron nitride nanotubes. J. Phys. Chem. B 2013, 117, 4618–4625. [Google Scholar] [CrossRef]
  80. Zhou, X.; Tang, D.-M.; Mitome, M.; Bando, Y.; Sasaki, T.; Golberg, D. Intrinsic and defect-related elastic moduli of boron nitride nanotubes as revealed by in situ Transmission Electron Microscopy. Nano Lett. 2019, 19, 4974–4980. [Google Scholar] [CrossRef]
  81. Chen, G.; Lu, H.; Cui, J.; Yu, H.; Wang, B.; Liu, Y.; Li, H.; Jiang, N. In situ real-time study buckling behavior of boron nitride nanotubes with axial compression by TEM. Chin. Chem. Lett. 2019, 30, 1401–1404. [Google Scholar] [CrossRef]
  82. Hung, S.C.; Su, Y.K.; Fang, T.H.; Liang, T.C. Elastic modulus investigation of gallium nitride nanotubes. In Proceedings of the 1st IEEE International Conference on Nano/Micro Engineered and Molecular Systems, Zhuhai, China, 18–21 January 2006. [Google Scholar]
  83. Tapia, A.; Cab, C.; Hernández-Pérez, A.; Villanueva, C.; Peñuñuri, F.; Avilés, F. The bond force constants and elastic properties of boron nitride nanosheets and nanoribbons using a hierarchical modeling approach. Physica E 2017, 89, 183–193. [Google Scholar] [CrossRef]
  84. Şahin, H.; Cahangirov, S.; Topsakal, M.; Bekaroglu, E.; Akturk, E.; Senger, R.T.; Ciraci, S. Monolayer honeycomb structures of group-IV elements and III-V binary compounds: First-principles calculations. Phys. Rev. B 2009, 80, 155453–155465. [Google Scholar] [CrossRef] [Green Version]
  85. Menon, M.; Srivastava, D. Structure of boron nitride nanotubes: Tube closing versus chirality. Chem. Phys. Lett. 1999, 307, 407–412. [Google Scholar] [CrossRef]
  86. Huber, K.P.; Hertzberg, G. Molecular Spectra and Molecular Siructure: IV. Constants of Diatomic Molecules, 1st ed.; Van Nostrand Reinhold Company: New York, NY, USA, 1979. [Google Scholar]
  87. Zhou, Z.; Zhao, J.; Chen, Y.; Schleyer, P.R.; Chen, Z. Energetics and electronic structures of AlN nanotubes/wires and their potential application as ammonia sensors. Nanotechnology 2007, 18, 424023–424030. [Google Scholar] [CrossRef]
  88. Lee, S.M.; Lee, Y.H.; Hwang, Y.G.; Elsner, J.; Porezag, D.; Frauenheim, T. Stability and electronic structure of GaN nanotubes from density-functional calculations. Phys. Rev. B 1999, 60, 7788–7792. [Google Scholar] [CrossRef]
  89. Lisenkov, S.V.; Vinogradov, G.A.; Lebedev, N.G. New class of non-carbon AlP nanotubes: Structure and electronic properties. J. Exp. Theor. Phys. Lett. 2005, 81, 185–189. [Google Scholar] [CrossRef]
  90. Mirzaei, M.; Mirzaei, M. DFT calculations of NMR properties for GaP nanotubes. Monatsh. Chem. 2011, 142, 111–114. [Google Scholar] [CrossRef]
  91. Chowdhury, R.; Adhikari, S. Boron-nitride nanotubes as zeptogram-scale bionanosensors: Theoretical investigations. IEEE Trans. Nanotechnol. 2011, 10, 659–667. [Google Scholar] [CrossRef]
  92. Shannon, M.A.; Bohn, P.W.; Elimelech, M.; Georgiadis, J.G.; Marinas, B.J.; Mayes, A.M. Science and technology for water purification in the coming decades. Nature 2008, 452, 301–310. [Google Scholar] [CrossRef]
  93. Zhao, M.; Xia, Y.; Li, F.; Zhang, R.Q.; Lee, S.T. Strain energy and electronic structures of silicon carbide nanotubes: Density functional calculations. Phys. Rev. B 2005, 71, 85312–85318. [Google Scholar] [CrossRef]
  94. Alam, K.M.; Ray, A.K. Hybrid density functional study of armchair SiC nanotubes. Phys. Rev. B 2008, 77, 35436–35446. [Google Scholar] [CrossRef]
  95. Pelesko, J.A.; Bernstein, D.H. Modeling MEMS and NEMS, 1st ed.; Chapman & Hal/CRC: Boca Raton, FL, USA, 2002. [Google Scholar]
  96. Ekinci, K.L. Electromechanical transducers at the nanoscale: Actuation and sensing of motion in Nanoelectromechanical Systems (NEMS). Small 2005, 1, 786–797. [Google Scholar] [CrossRef]
  97. Peng, H.B.; Chang, C.W.; Aloni, S.; Yuzvinsky, T.D.; Zettl, A. Ultrahigh frequency nanotube resonators. Phys. Rev. Lett. 2006, 97, 87203–87207. [Google Scholar] [CrossRef] [Green Version]
  98. Chowdhury, R.; Wang, C.Y.; Adhikari, S.; Scarpa, F. Vibration and symmetry-breaking of boron nitride nanotubes. Nanotechnolohy 2010, 21, 365702–365711. [Google Scholar] [CrossRef] [Green Version]
  99. Panchal, M.B.; Upadhyay, S.H. Single walled boron nitride nanotube-based biosensor: An atomistic finite element modelling approach. IET Nanobiotechnol. 2014, 8, 149–156. [Google Scholar] [CrossRef]
  100. Khani, N.; Fakhrabadi, M.M.S.; Vahabi, M.; Kamkari, B. Modal analysis of silicon carbide nanotubes using structural mechanics. Appl. Phys. A 2014, 116, 1687–1694. [Google Scholar] [CrossRef]
  101. Ansari, R.; Rouhi, S. Vibrational analysis of single-layered silicon carbide nanosheets and single-walled silicon carbide nanotubes using nanoscale finite element method. Proc. Inst. Mech. Eng. C J. Mech. Eng. Sci. 2017, 231, 3455–3461. [Google Scholar] [CrossRef]
Figure 1. Representation of the hexagonal GaP lattice with designation of n and m, Ch and θ. P atoms have a smaller radius and are represented in pale green and Ga atoms are in bright green.
Figure 1. Representation of the hexagonal GaP lattice with designation of n and m, Ch and θ. P atoms have a smaller radius and are represented in pale green and Ga atoms are in bright green.
Materials 15 03325 g001
Figure 2. Illustration of the three main symmetry groups for AlN, GaN, AlP and GaP nanotubes with comparable diameters, constructed using the Nanotube Modeler© software. Al and Ga atoms are represented in red and N and P atoms are in green.
Figure 2. Illustration of the three main symmetry groups for AlN, GaN, AlP and GaP nanotubes with comparable diameters, constructed using the Nanotube Modeler© software. Al and Ga atoms are represented in red and N and P atoms are in green.
Materials 15 03325 g002aMaterials 15 03325 g002b
Figure 3. Young’s modulus, E, as a function of nanotube diameter, Dn, for (a) BNNTs [34,36,56,59,60,61,62,66,71] and (b) AlNNTs [29,47] and SiCNTs [51,54].
Figure 3. Young’s modulus, E, as a function of nanotube diameter, Dn, for (a) BNNTs [34,36,56,59,60,61,62,66,71] and (b) AlNNTs [29,47] and SiCNTs [51,54].
Materials 15 03325 g003aMaterials 15 03325 g003b
Figure 4. Young’s modulus, E, of BNNTs as a function of the nanotube aspect ratio, L n / D n [31,36,65,68,69].
Figure 4. Young’s modulus, E, of BNNTs as a function of the nanotube aspect ratio, L n / D n [31,36,65,68,69].
Materials 15 03325 g004
Figure 5. Surface Young’s modulus, E s , as a function of the nanotube diameter, D n , for (a) BNNTs [37,39,41,43,46], (b) SiCNTs [39,41,43,46], (c) AlN [37,39], GaNNTs [37,39], (d) AlP [37], GaP [37,39], BP [39], InPNTs [39].
Figure 5. Surface Young’s modulus, E s , as a function of the nanotube diameter, D n , for (a) BNNTs [37,39,41,43,46], (b) SiCNTs [39,41,43,46], (c) AlN [37,39], GaNNTs [37,39], (d) AlP [37], GaP [37,39], BP [39], InPNTs [39].
Materials 15 03325 g005aMaterials 15 03325 g005b
Figure 6. (a) Shear modulus, G, for BNNTs [34,36,50,55,59,61,70,71] and (b) surface shear modulus, G s , for BNNTs [33,41] and SiCNTs [41], as a function of the nanotube diameter, D n .
Figure 6. (a) Shear modulus, G, for BNNTs [34,36,50,55,59,61,70,71] and (b) surface shear modulus, G s , for BNNTs [33,41] and SiCNTs [41], as a function of the nanotube diameter, D n .
Materials 15 03325 g006
Figure 7. Poisson’s ratio, ν, as a function of the nanotube diameter, D n , for (a) BNNTs [36,39,41,55,62,66], (b) SiC [39,41], AlN [39], GaNNTs [38,39] and (c) BP [39], GaP [37,39], InPNTs [39].
Figure 7. Poisson’s ratio, ν, as a function of the nanotube diameter, D n , for (a) BNNTs [36,39,41,55,62,66], (b) SiC [39,41], AlN [39], GaNNTs [38,39] and (c) BP [39], GaP [37,39], InPNTs [39].
Materials 15 03325 g007
Figure 8. The first fundamental frequency, fn1, as a function of the NT aspect ratio, L n / D n [34,57,67,68,69,71,98,100,101]; the fn1 values of comparable magnitude are plotted on the different graphs (a,b).
Figure 8. The first fundamental frequency, fn1, as a function of the NT aspect ratio, L n / D n [34,57,67,68,69,71,98,100,101]; the fn1 values of comparable magnitude are plotted on the different graphs (a,b).
Materials 15 03325 g008
Figure 9. The first fundamental frequency, fn1, as a function of the nanotube diameter, D n , for BNNTs [34,68,69,71].
Figure 9. The first fundamental frequency, fn1, as a function of the nanotube diameter, D n , for BNNTs [34,68,69,71].
Materials 15 03325 g009
Table 1. Values of the bond length of hexagonal diatomic nanostructures available in the literature.
Table 1. Values of the bond length of hexagonal diatomic nanostructures available in the literature.
CompoundBNAlNGaNInNBPAlPGaPInPSiC
aA1–A2, nm0.1447 [83]
0.145 [84]
0.147 [37]
0.151 [85]
0.153 [33]
0.177 [37]
0.179 [84]
0.185 [40]
0.193 [86]
0.195 [87]
0.175 [88]
0.184 [37]
0.185 [84]
0.186 [40]
0.194 [86]
0.203 [1]
0.206 [84]
0.183 [84]
0.193 [86]
0.234 [89]
0.240 [37]
0.220 [37]
0.225 [84]
0.229 [90]
0.236 [86]
0.246 [84]
0.256 [86]
0.177 [84]
0.179 [51]
0.185 [86]
Table 2. Young’s and shear moduli results for non-carbon nanotubes reported in the literature.
Table 2. Young’s and shear moduli results for non-carbon nanotubes reported in the literature.
ApproachYearReferenceMethodType of NTs 1E, TPa 2Es, TPa⋅nmG, TPa 2Gs, TPa⋅nmComment
Atomistic1998Hernandez et al. [60]TBMDBN(n, n)0.894average value
(n, 0)0.896
2003Moon et al. [51]MD: Tersoff empirical potentialSiC(n, n)0.621average value
(n, 0)0.558
2004Jeng et al. [38]MD: TB many body potentialGaN(5, 5)0.793
(9, 0)0.721
2004Kang and Hwang [40]MD: Tersoff-type potentialBN(5, 5)0.870
AlN0.453
GaN0.796
2007Baumeier et al. [46]ab initio:
DFT-SIC
BN(n, n)0.278converged average value
(n, 0)0.272
SiC(n, n)0.167
(n, 0)0.162
2007Verma et al. [55]MD: TB potentialBN(n, n)1.1070.965average
value
(n, 0)1.0441.555
2009Santosh et al. [59]MD: force—constant approachBN(n, n); (n, 0)1.0170.326converged average value
2009Setoodeh et al. [52]MD: Tersoff
potential
SiC(n, n)0.182 average
value
(n, 0)0.180
2009Pan and Si [53]MD: Tersoff bond order potentialSiCsingle crystalline0.465tn = 0.30 nm
0.540tn = 0.90 nm
2010Zhou et al. [54]SiC0.641tn = 0.89 nm
0.595tn = 1.69 nm
0.582tn = 2.49 nm
2011Zhang et al. [61]MD: DFTBBN(n, n)0.8400.366converged average
value
(n, 0)0.8440.368
2014Le [43]MD: harmonic force fieldsBN(n, n)0.282converged average value
(n, 0)0.281
SiC(n, n)0.148
(n, 0)0.145
2015Hao et al. [47]ab initio:
LCAO
AlN(n, n)0.360converged average value
(n, 0)0.340
2015Xiong and Tian [50]MD, Tersoff potential: force approachBN(n, n)0.315average value
(n, 0)0.329
energy approach(n, n) 0.281
(n, 0)0.292
2015Tao et al. [56]MD: TB potential + FE modelBN(n, n)0.911converged average value
(n, 0)0.930
2017Kochaev [37]ab initioBN(n, n)0.347average value
(n, 0)0.340
AlN(n, n)0.253
(n, 0)0.247
GaN(n, n)0.207
(n, 0)0.193
AlP(n, n)0.172
(n, 0)0.159
GaP(n, n)0.131
(n, 0)0.106
2019Pinhal et al. [29]DFT + B3LYPAlN(20, 20)0.393
(20, 0)0.387
(20, 10)0.392
GaN(20, 20)0.383
(20, 0)0.367
(20, 10)0.370
2020Choyal et al. [31]MD: TB potentialBN(10, 10)1.053Ln ≈ 21 nm
(17, 0)1.066
2020Vijayaraghavan and Zhang [35]MD: REBOBN(10, 10)2.8tn = 0.105 nm
CM2010Oh [62]CL thermodynamic approach + TB potentialBN(n, n)0.960converged
average value
(n, 0)0.975
NCM/MSM2006Li and Chou [34]beams +
FE model
BN(n, n)0.9160.465converged
average value
(n, 0)0.9130.475
2011Jiang and Guo [33]“stick-and-spring” model + closed-form solutionBN(n, n) 0.270 0.095converged
average value
(n, 0)0.2620.088
2015Ansari et al. [66]analytical solutionBN(n, n)0.825average value
(n, 0)0.823
2015Yan and Liew [70]representative cellBN(n, n)0.9700.416converged
average value
(n, 0)0.9670.418
2016Giannopoulos et al. [69]springs + FE model: free vibrationsBN(12, 12)0.592Ln ≈ 11 nm
(21, 0)0.523
2016Jiang and Guo [39]“stick-and-spring” model + analyticalBN(n, n)0.278converged
average value
(n, 0)0.276
AlN(n, n)0.121
(n, 0)0.120
GaN(n, n)0.120
(n, 0)0.119
BP(n, n)0.118
(n, 0)0.117
GaP(n, n)0.060
(n, 0)0.059
InP(n, n)0.051
(n, 0)0.051
SiC(n, n)0.169
(n, 0)0.168
2018Salavati et al. [65]beams +
FE model
BN(n, n); (n, 0)0.928converged
average value
2019Yan et al. [71]longitudinal and torsional vibrationsBN(n, n); (n, 0); (n, m)0.972 0.418converged
average value
2019Genoese et al. [41]“stick-and-spring” model + Donnell thin shell modelBN(n, n)0.2550.092converged
average value
(n, 0)0.2500.104
SiC(n, n)0.1520.053
(n, 0)0.1490.061
2021Sakharova et al. [36]beams +
FE model
BN(n, n); (n, 0); (n, m)0.9840.486converged
average value
2022Zakaria [68]two-section beams + FE modelBN(12, 12)0.538 0.108 Ln ≈ 11 nm
(21, 0)0.4890.117
Experimental1998Chopra and Zettl [75]TEM: thermal vibrational amplitudeMWBNNTs1.220
± 0.240
2004Suryavanshi et al. [76]TEM: electric-field-induced resonanceMWBNNTs with D out = 34–94 nm0.722
± 0.140
average value for 18 MWBNNTs
2005Hung et al. [16]NS +
analytical
SWGaNNTs0.484Ln = 500 nm
0.223Ln = 300 nm
2007Goldberg et al. [77]AFM-TEM: bending + analyticalMWBNNTs with D out = 40–100 nm0.5–0.6average value
2009Stan et al. [15]CR-AFM +
FEA
faceted AlNNTs with triangular cross-section0.3252
± 0.015
inner facet
0.3050
± 0.013
outer facet
2010Ghassemi et al. [78]AFM-TEM: bending + analyticalMWBNNTs with D out = 38–51 nm0.5 ± 0.1average value for 5 NTs
2011Arenal et al. [74]HRTEM-AFM
+ analytical
SWBNNTs1.11±0.17tn = 0.07 nm
0.87±0.13tn = 0.09 nm
0.25±0.04tn = 0.34 nm
2013Tanur et al. [79]AFM: a three-point bending + analyticalMWBNNTs with
D out = 18–55 nm
0.760 ± 0.030.07
± 0.01
E in bending, average value (0.1 ± 0.02 to 1.8 ± 0.3 TPa) for 20 NTs
2019Zhou et al. [80]HRTEM: high-order resonanceMWBNNTs with D out = 28–57 nm0.906average value
2019Chen et al. [81]TEM: force transducer holder + analyticalMWBNNT with D out = 37.34 nm and 40 layers1.050–1.370E calculated from tree compression cycles
1 All theoretical results presented were obtained for single-walled NTs. 2 The Young’s, E, and shear, G, moduli values for the BNNTs were evaluated, considering the nanotube wall thickness, tn ≈ 0.34 nm; otherwise, the tn value is indicated in the column “Comments”.
Table 3. Poisson’s ratio results for non-carbon nanotubes reported in the literature.
Table 3. Poisson’s ratio results for non-carbon nanotubes reported in the literature.
ApproachYearReferenceMethodType of NTs 1νComment
Atomistic1998Hernandez et al. [60]TBMDBN(n, n)0.260average value
(n, 0)0.240
2004Jeng et al. [38]MD: TB many body potentialGaN(5, 5)0.263
(9, 0)0.221
2007Verma et al. [55]MD: TB potentialBN(n, n), (n, 0)0.140average value
2017Kochaev [37] ab initioBN(10, 10)0.560
(10, 0)0.570
AlN(10, 10)0.520
(10, 0)0.550
GaN(10, 10)0.530
(10, 0)0.550
AlP(10, 10)0.510
(10, 0)0.510
GaP(10, 10)0.510
(10, 0)0.520
CM2010Oh [62]CL thermodynamic approach + TB potentialBN(n, n)0.150converged average value
(n, 0)0.170
NCM/MSM2015Ansari et al. [66]analytical solutionBN(n, n), (n, 0)0.217average value
2016Jiang and Guo [39]“stick-and-spring” model + analyticalBN(n, n)0.216converged average value
(n, 0)0.219
AlN(n, n)0.281
(n, 0)0.287
GaN(n, n)0.285
(n, 0)0.290
BP(n, n)0.360
(n, 0)0.365
GaP(n, n)0.428
(n, 0)0.435
InP(n, n)0.455
(n, 0)0.460
SiC(n, n)0.095
(n, 0)0.100
2019Genoese et al. [41]“stick-and-spring” model + Donnell thin shell modelBN(n, n)0.239converged average value
(n, 0)0.226
SiC(n, n)0.330
(n, 0)0.331
2021Sakharova et al. [36]beams +
FE model
BN(n, n); (n, 0);
(n, m)
0.150converged average value
Experimental2005Hung et al. [16]NS +
analytical
SWGaNNTs0.242
1 All theoretical results presented were obtained for single-walled NTs.
Table 4. Vibrational properties of N-CNTs available in the literature.
Table 4. Vibrational properties of N-CNTs available in the literature.
ApproachYearReferenceMethodType of NTs Support CaseLn, nmfn1, THzComments
Atomistic2015Ansari and Ajori [57]MD: TB potentialBN(10, 10)CF60.145The BNNTs with clamped-clamped support have higher values of f n 1 than those with cantilevered support. The first fundamental frequency decreases for small nanotube length and then tends to stabilize for NT length Ln > 4 nm.
80.081
100.048
120.016
CC60.855
80.532
100.274
120.210
2015Chandra et al. [48]MD: Tersoff-type potentialBN(10, 10)CC70.52The values of fn1 were determined at T = 400K. The bigger the BNNT length, the higher the first fundamental frequency.
210.06
CM2013Panchal et al. [63]thin wall tube (outer diameter of 0.8 nm, thickness of 0.065 nm) + analyticalBNCF60.744The values of fn1 were obtained for the case of attached mass at free NT end of 10−8 fg (from the range of 10−8 to 10−2 fg). The fn1 value increases with decreasing of the attached mass and NT length.
80.419
100.268
NCM/MSM2006Li and Chou [34]beams +
FE model
BN(4, 4)CF70.044The first fundamental frequency values decrease with increased in the nanotube length and diameter. The decrease rate is higher in the case of clamped-clamped support.
90.030
110.019
130.015
CC70.289
90.178
110.122
130.089
(7, 0)CF70.044
90.030
110.019
130.015
CC70.281
90.174
110.119
130.085
2010Chowdhury et al. [98]unspecified elastic elementsBN(4, 4)CC61.647The fn1 frequency decreases with increasing NT length, Ln, and diameter, Dn.
81.253
91.118
100.941
(8, 0)61.794
81.382
101.029
120.824
2013Panchal et al. [67]beams +
FE model
BN(5, 5)CF60.077The values of fn1 de-crease with increasing Ln and Dn
The fundamental frequencies for high-er-order vibrational modes were also cal-culated.
90.037
120.021
130.016
CC60.426
90.227
120.135
130.107
2014Khani et al. [100]beams +
FE model
SiC(15, 15)CF40.025The fn1 frequency decreases with increasing SiCNT length.First five vibrational modes shapes were analysed and respective values of the fundamental frequencies were calculated. The fundamental frequencies are higher for the case of clamped-clamped support.
60.021
80.011
100.008
CC40.062
60.044
80.029
100.025
(13, 0)CF40.024
60.013
80.006
100.004
CC40.060
60.039
80.023
100.015
2016Giannopoulos et al. [69]springs + FE modelBN(12, 12)CF80.210The values of fn1 decrease with increasing NT length, Ln, and diameter, Dn. The fundamental frequencies for first three modes were calculated. The fn values decrease with increasing of the vibrational mode order. The fundamental frequencies are higher for the case of clamped-clamped support.
160.124
250.073
350.061
CC80.408
160.256
250.171
350.123
(20, 0)CF80.207
160.125
240.081
320.064
CC80.467
160.256
240.174
320.127
2017Ansari and Rouhi [101]beams +
FE model
SiC(5, 5)CF40.359The first fundamental frequency, fn1 decreases for small values of the NT lengths and then tends to stabilize for Ln > 3 nm. The values of fn1 obtained for clamped-clamped support are about two times higher than those obtained for clamped-free support.
60.236
80.154
90.103
CC40.780
60.565
80.458
90.390
(9, 0)CF40.462
60.205
80.134
90.103
CC40.878
60.565
80.390
90.390
2019Yan et al. [71]analytical solution + Euler beam theoryBN(5, 5)CF51.052The values of fn1 decrease with increasing Ln. fn1 slightly increases for small NT diameters and then tends to nearly constant value for Dn > 1.0 nm.
60.918
70.731
CC52.127
61.860
71.487
2022Zakaria [68]two-section beams + FE modelBN(12, 12)SS80.189The values of fn1 decrease with increasing NT length, Dn, and diameter, Dn.
160.054
250.022
350.012
(20, 0)80.179
160.048
240.023
320.012
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Antunes, J.M.; Pereira, A.F.G.; Sakharova, N.A. Overview on the Evaluation of the Elastic Properties of Non-Carbon Nanotubes by Theoretical Approaches. Materials 2022, 15, 3325. https://doi.org/10.3390/ma15093325

AMA Style

Antunes JM, Pereira AFG, Sakharova NA. Overview on the Evaluation of the Elastic Properties of Non-Carbon Nanotubes by Theoretical Approaches. Materials. 2022; 15(9):3325. https://doi.org/10.3390/ma15093325

Chicago/Turabian Style

Antunes, Jorge M., André F. G. Pereira, and Nataliya A. Sakharova. 2022. "Overview on the Evaluation of the Elastic Properties of Non-Carbon Nanotubes by Theoretical Approaches" Materials 15, no. 9: 3325. https://doi.org/10.3390/ma15093325

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