Next Article in Journal
Recent Developments in Photoluminescent, Photothermal and Photocatalytic Nanomaterials
Next Article in Special Issue
Combination of CNTs with Classical Drugs for Treatment in Human Colorectal Adenocarcinoma (HT-29) Cell Line
Previous Article in Journal
Cu(II) and Mn(II) Anchored on Functionalized Mesoporous Silica with Schiff Bases: Effects of Supports and Metal–Ligand Interactions on Catalytic Activity
Previous Article in Special Issue
Carbon Nanotube-Supported Dummy Template Molecularly Imprinted Polymers for Selective Adsorption of Amide Herbicides in Aquatic Products
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Prediction of Temperature-Dependent Mechanical Properties for SWCNT/Cu Nanocomposite Metamaterials: A Molecular Dynamics Study

School of Aeronautics and Astronautics, Shanghai Jiao Tong University, Shanghai 200240, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(12), 1885; https://doi.org/10.3390/nano13121885
Submission received: 19 May 2023 / Revised: 7 June 2023 / Accepted: 11 June 2023 / Published: 19 June 2023
(This article belongs to the Special Issue Current State-of-the-Art of SWCNT, MWCNT, and Mixed CNT)

Abstract

:
Single-walled carbon nanotube (SWCNT) is a promising candidate for strengthening nanocomposite. As the matrix of nanocomposite, a single crystal of copper is designed to be in-plane auxetic along the crystal orientation [1 1 0]. In that way, the nanocomposite could also be auxetic when enhanced by (7, 2) a single-walled carbon nanotube with relatively small in-plane Poisson’s ratio. A series of molecular dynamics (MD) models of the nanocomposite metamaterial are then established to study mechanical behaviors of the nanocomposite. In the modelling, the gap between copper and SWCNT is determined following the principle of crystal stability. The enhanced effect for different content and temperature in different directions is discussed in detail. This study provides a complete set of mechanical parameters of nanocomposite including thermal expansion coefficients (TECs) from 300 K to 800 K for five weight fractions, which is essential for a wide range of applications of auxetic nanocomposites in the future.

1. Introduction

Carbon nanotubes (CNTs) [1,2] have outstanding physical properties, such as superior mechanical behaviors, good electrical conductivity, and exceptional thermal performance, together with light weight [3]. Accordingly, for high-performance nanocomposites, CNTs are considered as ideal reinforcements to fulfill diverse requirements from high-end manufacturing [4]. Metal matrix nanocomposites (MMCs) reinforced by CNTs [5,6,7], which can serve in a high-temperature environment, have great potential in electronic, aerospace, and other industries. Experimental studies [8,9] have reported that only a small addition of CNT could result in a significant enhancement in mechanical properties of metals. Among MMCs, CNT reinforced copper (CNT/Cu) nanocomposites gain considerable attention due to their excellent electrical and thermal conductivity as well as high strength. Moreover, low TECs of CNT/Cu nanocomposites are favorable for reducing the strains caused by the TEC mismatch between electronic devices [10,11]. It is worth mentioning that the high performance of CNT/Cu nanocomposites is attributed to homogeneous distribution, improved interfacial reaction, and enhanced structural integrity of CNTs in copper matrix [12,13]. Therefore, many fabrication technologies have been developed to address these issues, such as molecular-level mixing [14,15,16,17], high-energy ball milling [18,19], electrochemical co-deposition [20], and spark plasma sintering [21]. Obviously, these experimental results are influenced by the processing method. In this case, an atomistic simulation like MD may be suitable for parametric investigations of the effects of composites and geometry on mechanical properties [22,23,24,25,26].
Most studies on mechanical behaviors of metal matrix nanocomposites, however, have not paid enough attention to a phenomenon of negative Poisson’s ratio (NPR). NPR, also known as “auxetic” [27], is an abnormal phenomenon where a material expands transversely under tension and contracts transversely under compression. NPR can be obtained through microstructural design, such as porosity and re-entrant, inside materials [28]. Hence, any material, theoretically, can be auxetic with a designed microstructure. Auxeticity can bring higher vibration absorption [29], increased indentation resistance [30], and improved plane strain fracture resistance [31]. However, there is a proverb that says you cannot have your cake and eat it too. Internal microstructure also means loss of material, inevitably resulting in degradation of stiffness and strength, but they are highly pursued in engineering. It is a contradiction that NPR and high stiffness and strength are mutually exclusive [32]. One solution is to adopt the sandwich structure that takes advantage of the unique properties of auxetic cores and composite face sheets [33]. Another solution is to design auxetic metamaterial in atomic scale to avoid porosity. In this way, the auxetic characteristics are obtained without sacrificing physical performance. For example, some metals with face-centered cubic (FCC) lattice have NPR along a non-axial direction [34,35]. Recently, based on MD simulations, Fan et al. [36,37] confirmed that single-crystal copper showed in-plane NPR of about −0.19 when subjected to uniaxial [1 1 0] loading.
It must be admitted that research on the mechanical behavior of CNT reinforced copper matrix composites has been relatively limited and the available data for further mechanical analysis are not comprehensive. In particular, the content of CNT plays a key role in strengthening the properties of composites. The main goal of this study is to design a SWCNT reinforced copper matrix (SWCNT/Cu) nanocomposite metamaterial and establish an effective corresponding molecular dynamics model to predict its mechanical properties, especially NPR, under different temperature. Additionally, the enhanced effect of SWCNT will be discussed in detail. A new design idea of auxetic materials is proposed to provide useful guidance for preparation.

2. Materials and Methods

In this study, an MD model of SWCNT/Cu nanocomposite is created along the x[1 1 0], y[1 1 ¯ 0], and z[0 0 1] crystal orientations of single-crystal copper as depicted in Figure 1a. For simplicity, we utilize the numbers 1, 2, and 3 to represent the directions of x, y, and z, respectively. Thus, a new representative cell for copper matrix along the crystal orientations is shown in Figure 1b. To ensure that the nanocomposite has obvious auxetic properties, the Poisson’s ratio of SWCNT, according to rule of mixture, should be selected as small as possible. Furthermore, the minimum diameter for synthetic CNT is greater than 6 Å in the laboratory [38]. Therefore, the SWCNT with chiral indices (7, 2) is selected as a reinforcement in this article. The chiral angle θ and diameter d of (7, 2) SWCNT are 12.22° and 6.41 Å, respectively. According to our previous work [39], Poisson’s ratios for some chiral SWCNTs with variable chiral indices are listed in Table 1.
There are three potentials for atom pairs, i.e., Cu-Cu, C-C, and Cu-C, to be determined in this SWCNT/Cu nanocomposite model. The interaction between each pair of copper atoms is described by an embedded-atom method (EAM) potential [40,41], which is presented as:
E Cu = F α j i ρ β r   ij + 1 2 j i φ   α β ( r   ij ) ,
where ECu is the total energy for a pair of Cu atoms, F is the embedding energy which is a function of the atomic electron density ρ, φ is a pair potential interaction between atoms i and j separated by a distance rij, and subscripts α and β are the element types of atoms i and j. Conventionally, we use 3.615 Å and 63.546 as lattice constant and atomic weight of copper, respectively.
The adaptive intermolecular reactive empirical bond order (AIREBO) [42] potential with a cutoff radius (r = 3 Å) is adopted for the force field between carbon atoms, which is widely employed to study CNTs [22,23,24]. The cohesive energy of this potential consists of three terms:
E CNTs = 1 2 i j i E ij REBO + E ij LJ + k i , j g i , j , k E kijg TORSION ,
in which EREBO, ELJ, and ETORSION are potential energy from short-ranged reactive empirical bond order (REBO) potential, standard Lennard-Jones (LJ) potential that adds longer-ranged interactions and the torsional interactions term that describes various dihedral angle preferences in hydrocarbon configurations. The EREBO term in the AIREBO potential gives the model its reactive capabilities and only describes short-ranged interactions whose cutoff radii are less than 2 Å. The EREBO term is given as:
E ij REBO = f ( r ij ) ( V ij R + B ij V ij A ) ,
where Bij, V ij R , and V ij A are the bond-dependent parameters which weigh the bond order, repulsive and attractive pair energy terms. f (rij) is the cut-off function of the distance rij between atoms i and j [42,43]. The C-C bond length used is a commonly accepted value of 1.42 Å, and the atom weight of C is 12.011.
To estimate the non-bonded interaction between copper and carbon atoms, we employ the LJ potential which has adequately predicted the Cu-C interaction and has been extensively used in the existing literature [22,23,24,25] on the MD study of carbon/copper molecular structures. The LJ potential energy is a function of the distance rij between two atoms:
E ij LJ = 4 ε σ r ij 12 σ r ij 6 ,
in which ε and σ are the potential well depth and equilibrium interatomic distance, respectively. It is apparent that these two parameters have important consequences in the MD results. Here, we assign ε = 0.01996 eV and σ = 3.225 Å, which are also adopted in studies [44,45,46].
In this study, simulations are carried out using a large-scale atomic/molecular massively parallel simulator (LAMMPS) [47] and results are visualized by the Open Visualization Tool (OVITO) [48]. Periodic boundary conditions are set in the three directions to consider Poisson’s effect of material. Generally, there are two ways to present content of CNT in nanocomposites: volume fraction (vol.%) and weight fraction (wt.%). However, it might be inaccurate to measure SWCNT content by using volume fraction because determination of effective wall thickness of SWCNT is inevitable in this method but controversy truly exists [39]. Alternatively, the weight fraction of SWCNTs is adopted, which is obtained directly from the calculation based on the number of Cu and C atoms and their atomic mass. It should be noted that the weight fraction of CNT has an upper limitation to ensure stability of the whole system. Moreover, when designing MD models for SWCNT/Cu nanocomposites shown in Figure 1c, we use the common multiple of Cu lattice parameters and length of (7, 2) SWCNT to reduce the effect of mismatch in the longitudinal direction (i.e., x-axis) which may additionally cause the internal stress at the periodic interface. For instance, Figure 1c,d illustrates the three-dimensional structure and yoz section of the SWCNT/Cu nanocomposite with 8.29 wt.%. The simulation box containing 1681 Cu atoms and 804 C atoms is 104.8 × 15.3 × 18.1 Å3 in dimension. It should be pointed out that interface between Cu and SWCNT plays a key role in some mechanical properties, such as in-plane shear modulus. In this paper, determination of the gap will be discussed in detail in the next section. The temperature varying from 300 K to 800 K is also taken into account to evaluate the thermal effect on mechanical properties of nanocomposites. The duration of each timestep is set at 1 fs, and temperature and pressure of the system are both controlled by employing the Nosé–Hoover thermostat. The minimized structure is then equilibrated for a period of 5 × 104 timesteps within the context of an isothermal-isobaric (NPT) ensemble.
After equilibration, the tensile and shear behaviors controlled by strain are applied to MD models to study in-plane and out-of-plane mechanical properties, containing Young’s modulus E, shear modulus G, and Poisson’s ratio υ. For tensile simulation, the specified constant engineering strain rate is 10−5/ps, and a 1% stretching strain is performed after 106 timesteps in the NPT ensemble. Then, we can obtain Young’s moduli and Poisson’s ratios as:
E ii = σ i ε i ,
υ ij = ε j ε i ,
where subscripts i and j refer to 1, 2, and 3 directions (i.e., x, y, and z). Hence, there are three Young’s moduli and six Poisson’s ratios that need to be determined.
For shear testing, we also control the engineering strain rate to achieve the 1% shear strain. The calculation method of three shear moduli (G12, G13, and G23) is similar to Equation (5). It is emphasized that we set two “one-layer” of Cu and SWCNT sections, as shown in Figure 1d, in each MD model as “rigid” to restrain sliding of SWCNT during shear deformation. The sliding phenomenon means that only the interaction between Cu and C atoms works in resistance to shear deformation. In other words, the MD results only reflect the strength of the SWCNT/Cu interface without any limitations. On the other hand, the “one-layer” of “rigid” part should be as thin as possible to avoid unnecessary disturbance. Therefore, we utilize two “rigid” parts of thickness measuring 1.8 Å in the longitudinal direction (half of the copper lattice structure). Each part consists of 16 Cu atoms (marked as purple in Figure 1d). Furthermore, the thermal expansion coefficient (TEC) α and density ρ of nanocomposites can be obtained from the equilibration process as:
α ii = 1 L i 0 d L i d T ,
ρ = n C   M C + n Cu   M Cu V   N A ,
in which the unit of α and ρ are K−1 and g/cm3. i represents the 1, 2, and 3 directions, respectively. n, Li, and V, respectively, denote number of atoms, length in the i direction and volume of SWCNT/Cu nanocomposites. Correspondingly, Li0 represents the initial length of SWCNT/Cu nanocomposites. T is temperature with the unit of Kelvin and NA is the Avogadro constant.
According to the density formula of nanocomposites (Equation (8)), the relationship between vol.% and wt.% of CNTs can be derived as:
vol . % = wt . % wt . % + ρ C N T ρ C u ρ C N T ρ C u   wt . % .
where ρCNT and ρCu stand for the density of CNTs and Cu in nanocomposites. Generally, ρCNT takes a value of 1.8–2.1 g/cm3 and ρCu is chosen as 8.9–8.96 g/cm3 [14,49].

3. Results and Discussion

3.1. Determination of Gap

The interface between SWCNT and copper plays a fundamental role in some mechanical performance of nanocomposites [50], such as shear modulus. The selection criterion of minimum gap (distance) between copper atom and carbon atom is following the principle of a stable crystal system after relaxation. Three representative models with various gaps (i.e., 1.9 Å, 2.3 Å, and 3.0 Å) and their simulation results after relaxation at room temperature (300 K) are shown in Figure 2. They are the same size, 104.8 × 17.9 × 18.1 Å3, but have different weight fractions (6.5 wt.%, 6.8 wt.%, and 7.3 wt.%). Figure 2d–f shows the atomic stress distributions of the three models. In Figure 2d, the highest atom stress in the atomic structure with 1.9 Å gap is up to 520 GPa⸱Å3 after energy minimization. In contrast, models for 2.3 Å and 3.0 Å gaps have the distribution of atom stress at a lower level. Furthermore, although the cross-section of SWCNT in Figure 2f is well maintained in a circular shape, the nanocomposite with 3.0 Å gap does not have symmetric stress distribution in the z-direction. After equilibration, the structures with 1.9 Å and 3.0 Å gap both exhibit lattice deformation, shown in Figure 2g,i. One reason may be the performance difference due to the different lattice constants of [1 1 0] and [0 0 1] Cu crystal orientations. Therefore, 2.3 Å gap seems the best choice among the three from the view of system stability.

3.2. Effect of SWCNT Content

The correspondence between the present work and experiment results [14,15,16,17,18,19,21] is shown in Figure 3, where Young’s modulus of CNT/Cu composites synthesized by experiments are compared. In the figure, volume fractions of CNTs are converted to weight fractions by Equation (9) (i.e., 5 vol.% ≈ 1.1 wt.%; 10 vol.% ≈ 2.2 wt.%) for comparison. Although a difference does exist for various material systems, the Young’s moduli from the present work are close to those from mainstream experimental methods [14,15,16,17,18,19,21]. The figure also verifies the effectiveness of the present MD model.
To study the effect of SWCNT weight fraction on mechanical properties of SWCNT/Cu nanocomposite, 100 MD models are built and divided into four groups according to their lengths in the y- and z-directions (Ly and Lz), respectively. All the models are simulated at room temperature. Figure 4a,d,g show that Young’s modulus E11 drops when length (Ly or Lz) increases, which means the SWCNT weight fraction decreases simultaneously. The nanocomposite with 11.3 wt.% obtains a maximum E11 at about 399.3 GPa. For the group where Lz = 14.46 Å in Figure 4a, E11 changes from 399.3 GPa to 260.5 GPa within the interval of 20 Å, which dropped by 35%. In the next interval, E11 reduces to 203.9 GPa, which dropped by 22%. The rate of decline slows slightly in the interval 50~70 Å. This trend is similar to the effect of Lz in Figure 4d. In other words, the influence of size increase in any direction on E11 is similar. In contrast, E22 is improved as lengths (Ly and Lz) enlarge or wt.% decreases. It is apparent, in Figure 4e, that the sensitivity of E22 to lengths diminishes when lengths exceed 45 Å and the role of SWCNT turns from weakened to enhanced. After this, E22 remains at around 102.5 GPa, a little higher than copper (101.1 GPa). Unlike E11 and E22, both having a monotone relation to weight fraction, there is a maximum value of out-of-plane modulus E33 existing with the change of weight fraction. This trend was also observed in some experimental studies [16,21,49]. The composites with higher wt.% have lower relative density, which can be attributed to the original clustering of CNTs [51]. The maximum value is about 68.5 GPa at 3.3% weight fraction which is independent on Ly. Thus, it is clear that the enhanced effect of the SWCNT to the auxetic copper is different in the three directions.
There are two cases displayed in Figure 5a to study the effect of SWCNT size: (1) nanocomposites with same length but owning SWCNTs of larger diameter; (2) nanocomposites with same weight fraction but owning SWCNTs of larger diameter. In can be seen in Figure 5c that E11 enlarges with increasing wt.%, while E22 and E33 decline. However, E11 of nanocomposite with 10.5 wt.% (28, 8) SWCNT is only 267.8 GPa, about 60% E11 of nanocomposite with 11.3 wt.% (7, 2). The reason is that SWCNTs of small diameter possess higher modulus [39] and thus greater enhancement effects. As for the case of nanocomposites with the same weight fraction, all three moduli have been slightly reduced. In the meantime, negative Poisson’s ratios have not been well improved, considering the much larger size of the copper. Moreover, it can be observed that the larger the diameter, the more irregular the cross-section of SWCNT becomes after equilibration. This means that the overall structure is more prone to failure and therefore exhibit low strength. In conclusion, smaller diameter carbon tubes have greater enhancement potential [52].

3.3. Effect of Temperature

We simulate single-crystal copper and SWCNT/Cu nanocomposites with five SWCNT weight fractions of 1.37%, 3.35%, 4.96%, 6.78%, and 8.29% from 300 K to 800 K to predict temperature-dependent mechanical properties. The thickness to width ratio (Lz/Ly) is limited between 0.8 and 1.2 to eliminate size effect. The MD results of in-plane moduli are compared in Table 2 and plotted in Figure 6. It can be seen that the modulus of nanocomposites in the x-direction is significantly higher than that of pure copper. As expected, SWCNT/Cu nanocomposite with 8.29 wt.% has the largest E11 about 344.1 GPa, increased by 240% compared with pure copper (0 wt.%) at room temperature (300 K). When the temperature reaches 800 K, the enhanced effect achieves about 304%, which is consistent with the observation in [25]. Moreover, all three Young’s moduli decrease with the increase in temperature. It is easy to understand the weakened loading bearing capacity of SWCNT/Cu nanocomposites caused by the expansion of distances of atoms and softened crystalline structure due to raised temperature [24]. It is easy to find that the enhanced effect in the x direction is more remarkable than those in the other two directions from Table 2, that is because of the mechanical characteristics of the reinforcement SWCNT. In other words, contribution of SWCNT to E11 of nanocomposite is more than that of Cu. In addition, SWCNT contribution proportion is improved with higher content. For the cases of E22 and E33, however, contributions of Cu t are dominant. Furthermore, from Table 2, the degradation of mechanical property is sensitive to temperature for copper. However, it is reported that Young’s modulus of SWCNT in the longitudinal direction is only reduced by 3% when temperature is raised from 300 K to 700 K [53], which means longitudinal modulus of SWCNT is almost temperature independent. Hence, this explains well why Young’s modulus of SWCNT/Cu is almost temperature-independent in the x-direction but temperature-dependent in other directions.
Due to the symmetrical lattice structure, E11 of copper is approximately equal to E22. However, we observe that the enhanced properties of SWCNT/Cu nanocomposites in the x-direction comes at the cost of a decrease in two/three-directional performance. For example, in the case of 8.29 wt.% SWCNT, E11 is almost four times that of E22 at room temperature. It is believed that SWCNT is a typical one-dimensional single-crystal nanomaterial and had greater excellent mechanical performance in the longitudinal direction. Figure 7 illustrates the effect of temperature and weight fraction on in-plane shear modulus G12 and out-of-plane shear moduli (G13 and G23). It was found that G12 is weakened by adding SWCNT reinforcement. This is because without intensive treatment the strength of the copper–carbon interface is very low due to poor wettability and interaction [12], which is precisely the decisive factor in resisting shear deformation. Therefore, the high strength of CNTs is not fully exploited [54]. G12 of 8.29 wt.% nanocomposite decreases from 13.5 GPa to 9.8 GPa when temperature increases to 800 K. Both in-plane and out-of-plane moduli of copper and SWCNT/Cu nanocomposites are provided in Table 2. It can be inferred that the performances of SWCNT/Cu nanocomposites and copper are close in the thickness direction while out-of-plane shear moduli of SWCNT/Cu nanocomposites are all lower than copper.
To compare the effect of SWCNT and defect (hole) on in-plane properties, five cases listed in Table 3 are investigated. The MD results in Table 3 reveal that the defect can significantly decrease the in-plane moduli and Poisson’s ratios of copper, which is in agreement with findings Zhao et al. [55] reported. It is apparent that there is no enhancement on E22 and G12 of nanocomposite with SWCNT. Inversely, these properties are degraded somehow compared with pure copper (Case 1). In other words, the effect of SWCNT in certain directions is similar to that of defect, although is not as obvious as that.
In order to study the auxetic behavior of the SWCNT/Cu nanocomposites, in-plane Poisson’s ratios (υ12 and υ21) are depicted in Figure 8 and presented in Table 4. Single crystal copper, as a typical FCC metal, exhibits an in-plane auxetic behavior when the load is applied in the [1 1 0] direction. It is apparent that auxeticity is more significant with the increase in temperature. This phenomenon is also observed from the SWCNT/Cu nanocomposites. Furthermore, we find that the increase in SWCNT content may diminish the auxetic nature of SWCNT/Cu nanocomposites, although the Poisson’s ratio of SWCNT is positive but relatively small [39]. As listed in Table 3, the presence of nano-hole leads to NPRs υ12 and υ21 decreases by −19.4% and −37.0%, respectively. SWCNTs slightly improve the NPR υ12 but further weaken NPR υ21. In general, the NPRs SWCNT/Cu of composites are lower than that of the copper. Table 4 also presents the results of out-of-plane Poisson’s ratios from 300 K to 800 K. It should be pointed out that the limit of Poisson’s ratio from −1.0 to 0.5, which follows from stability conditions for isotropic materials, does not apply [37]. In addition, results in Table 2 and Table 4 reflect the remarkable anisotropy for the SWCNT/Cu nanocomposites, which is caused by the addition of SWCNTs.
The TECs and density of (7, 2) SWCNT, copper, and the SWCNT/Cu nanocomposites varying from 300 K to 800 K are shown in Figure 9 and Table 5. Although TEC α11 of copper in Figure 9a is gradually enlarged as temperature rises, TEC α11 of SWCNT/Cu nanocomposites seems to be insensitive to temperature changes. It is evident that α11 is more sensitive to the SWCNT content, and for 8.29 wt.% at room temperature, α11 is just one third of that of copper. We find that α22 in Figure 9b is also insensitive to temperature changes but is larger than α11. In contrast, α33 shows a clear increasing trend as the temperature rises.
In order to further analyze the structural changes of SWCNT/Cu nanocomposites in the temperature field, the model has been divided into three components as depicted in Figure 10a: the copper part (LCu), the embedded SWCNT (dCNT), and the gap between copper and SWCNT (Lgap). In fact, the sum of these three is the thickness Lz of the model and the same is true in the y-direction. The detailed α of three components evaluated by Equation (7) under various temperatures are plotted in Figure 10. It is clear that the Cu part accounts for most of the variation in α22 and α33. However, it is gradually weakened with increased weight fractions. Correspondingly, the effect of the gap between copper and SWCNT becomes progressively greater, even over the Cu part results. The larger distance between the copper and SWCNTs leads to a weaker interface, which accounts for the diminished capacity of mechanical loading transfer for SWCNT/Cu nanocomposites. It is worth noting that α22 caused by SWCNT is negative, while α33 caused by SWCNT is positive, which means that the SWCNT expands in the z-direction but contracts in the y-direction.

4. Conclusions

An MD study on mechanical properties and TECs of a chiral SWCNT reinforced auxetic copper nanocomposite was carried out in this paper. In the modelling, an SWCNT with indices (7, 2) and FCC crystals of copper which have auxeticity in the [1 1 0] crystallographic orientation, were selected as the reinforcement and the matrix, respectively. The main conclusions can be drawn as follows:
  • To sustain the stability in subsequent mechanical behavior, it is most important to determine the gap between Cu and SWCNT according to minimization of energy. The comparison between three gaps on stress distribution and crystal deformation after relaxation shows that the gap of 2.3 Å is optimum in the present case. The enhanced effect of SWCNT on the considered nanocomposite in various direction is different.
  • As expected, elastic modulus E11 is significantly raised as it is along the length direction of SWCNT. In our simulation, E11 can reach about 399.3 GPa with 11.3 wt.% SWCNT, which is about four times that of the copper matrix. Whereas, E22 of nanocomposite is decreased because the effect of SWCNT in the y-direction is similar to a defect in this case. Unlike E11 and E22, E33 has an “inverse” trend with increase in SWCNT weight fraction, and the peak value of E33 is at 3.3 wt.% of around 68.5 GPa. These indicate that the considered nanocomposite is anisotropic in mechanics.
  • In addition, we discovered that smaller diameter CNTs have a better enhanced effect and its nanocomposite has more remarkable auxeticity. Although the SWCNT with positive Poisson’s ratio weakens the auxeticity of copper, υ12 of SWCNT/Cu can still reach about −0.2.
  • Finally, thermal effect is considered as temperature rises from 300 K to 800 K. With temperature increase, all elastic moduli and shear moduli are degraded due to the weakening of their own chemical bonds and interactions between SWCNT and Cu. However, at the same time, it is noted that the enhanced effect of SWCNT is relatively more obvious at a high temperature owing to the property degradation of SWCNT being less. On the whole, TECs α11 and α22 of nanocomposite are in between those of SWCNT and copper. However, an abnormal phenomenon is that TEC α33 is higher than that of any component material. Thus, we discussed the contributions of the three components (SWCNT, copper, and the gap) to the TEC in that direction to explain this phenomenon.
The MD model and material data of the nanocomposite metamaterial proposed in this paper can be referenced for further mechanical analysis.

Author Contributions

Conceptualization, Y.F. and H.-S.S.; methodology, H.-N.Z., Y.F. and H.-S.S.; software, H.-N.Z.; data curation, H.-N.Z. and Y.F.; writing—original draft preparation, H.-N.Z. and Y.F.; writing—review and editing, H.-N.Z., Y.F. and H.-S.S.; funding acquisition, Y.F. All authors have read and agreed to the published version of the manuscript.

Funding

The work is sponsored by the National Natural Science Foundation of China (NSFC) under Grant No. 12102255 and Natural Science Foundation of Shanghai under Grant No. 22ZR1429500.

Data Availability Statement

Any data included in the manuscript is adequately referenced and available.

Acknowledgments

The computations in this work were run on the ARM cluster supported by the Center for High Performance Computing at Shanghai Jiao Tong University.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Iijima, S. Helical microtubules of graphitic carbon. Nature 1991, 354, 56–58. [Google Scholar] [CrossRef]
  2. Iijima, S.; Ichihashi, T. Single-shell carbon nanotubes of 1-nm diameter. Nature 1993, 363, 603–604. [Google Scholar] [CrossRef]
  3. Zhang, R.F.; Zhang, Y.Y.; Wei, F. Controlled Synthesis of Ultralong Carbon Nanotubes with Perfect Structures and Extraordinary Properties. Acc. Chem. Res. 2017, 50, 179–189. [Google Scholar] [CrossRef] [PubMed]
  4. Chiba, T.; Yabuki, H.; Takashiri, M. High thermoelectric performance of flexible nanocomposite films based on Bi2Te3 nanoplates and carbon nanotubes selected using ultracentrifugation. Sci. Rep. 2023, 13, 3010. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, G.J.; Cai, Y.P.; Ma, Y.J.; Tang, S.C.; Syed, J.A.; Cao, Z.H.; Meng, X.K. Ultrastrong and Stiff Carbon Nanotube/Aluminum-Copper Nanocomposite via Enhancing Friction between Carbon Nanotubes. Nano Lett. 2019, 19, 6255–6262. [Google Scholar] [CrossRef]
  6. Cao, L.; Chen, B.; Wan, J.; Kondoh, K.; Guo, B.; Shen, J.; Li, J.S. Superior high-temperature tensile properties of aluminum matrix composites reinforced with carbon nanotubes. Carbon 2022, 191, 403–414. [Google Scholar] [CrossRef]
  7. Liu, K.Y.; Li, J.S.; Wan, J.; Yan, Q.; Kondoh, K.; Shen, J.; Li, S.; Chen, B. Sintering-free fabrication of high-strength titanium matrix composites reinforced with carbon nanotubes. Carbon 2022, 197, 412–424. [Google Scholar] [CrossRef]
  8. Stein, J.; Lenczowski, B.; Frety, N.; Anglaret, E. Mechanical reinforcement of a high-performance aluminium alloy AA5083 with homogeneously dispersed multi-walled carbon nanotubes. Carbon 2012, 50, 2264–2272. [Google Scholar] [CrossRef]
  9. Pan, Q.Y.; Yu, X.P.; Yu, F.J. Creep and mechanical properties of aluminium A356 composites reinforced with coated and un-coated MWCNTs fabricated using the stir casting method. Mater. Sci. Technol. 2023, 39, 1100–1112. [Google Scholar] [CrossRef]
  10. Xu, L.S.; Chen, X.H.; Liu, X.J. Preparation and properties of carbon nanotubes reinforced Cu matrix composites for electronic packaging application. Appl. Mech. Mater. 2013, 275–277, 1789–1793. [Google Scholar] [CrossRef]
  11. Ishibe, T.; Kaneko, T.; Uematsu, Y.; Sato-Akaba, H.; Komura, M.; Iyoda, T.; Nakamura, Y. Tunable Thermal Switch via Order-Order Transition in Liquid Crystalline Block Copolymer. Nano Lett. 2022, 22, 6105–6111. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, D.; Yan, A.; Liu, Y.; Wu, Z.; Gan, X.; Li, F.; Tao, J.; Li, C.; Yi, J. Interfacial Bonding Improvement through Nickel Decoration on Carbon Nanotubes in Carbon Nanotubes/Cu Composite Foams Reinforced Copper Matrix Composites. Nanomaterials 2022, 12, 2548. [Google Scholar] [CrossRef]
  13. Tian, D.; Liu, Y.; Yu, J.; Zhao, Q.; Tao, J.; Wu, Z.; Zhang, J.; Fan, Y.; Liu, Y.; Li, C.; et al. A Study of Silver Decoration on Carbon Nanotubes via Ultrasonic Chemical Synthesis and Their Reinforced Copper Matrix Composites. Nanomaterials 2023, 13, 887. [Google Scholar] [CrossRef] [PubMed]
  14. Cha, S.I.; Kim, K.T.; Arshad, S.N.; Mo, C.B.; Hong, S.H. Extraordinary strengthening effect of carbon nanotubes in metal-matrix nanocomposites processed by molecular-level mixing. Adv. Mater. 2005, 17, 1377–1381. [Google Scholar] [CrossRef] [PubMed]
  15. Kim, K.T.; Il Cha, S.; Gemming, T.; Eckert, J.; Hong, S.H. The Role of Interfacial Oxygen Atoms in the Enhanced Mechanical Properties of Carbon-Nanotube-Reinforced Metal Matrix Nanocomposites. Small 2008, 4, 1936–1940. [Google Scholar] [CrossRef] [PubMed]
  16. Lim, B.K.; Mo, C.B.; Nam, D.H.; Hong, S.H. Mechanical and Electrical Properties of Carbon Nanotube/Cu Nanocomposites by Molecular-Level Mixing and Controlled Oxidation Process. J. Nanosci. Nanotechnol. 2010, 10, 78–84. [Google Scholar] [CrossRef]
  17. Xue, Z.W.; Wang, L.D.; Zhao, P.T.; Xu, S.C.; Qi, J.L.; Fei, W.D. Microstructures and tensile behavior of carbon nanotubes reinforced Cu matrix composites with molecular-level dispersion. Mater. Des. 2012, 34, 298–301. [Google Scholar] [CrossRef]
  18. Kim, K.T.; Il Cha, S.; Hong, S.H.; Hong, S.H. Microstructures and tensile behavior of carbon nanotube reinforced Cu matrix nanocomposites. Mat. Sci. Eng. A-Struct. 2006, 430, 27–33. [Google Scholar] [CrossRef]
  19. Liu, J.; Xiong, D.-B.; Tan, Z.; Fan, G.; Guo, Q.; Su, Y.; Li, Z.; Zhang, D. Enhanced mechanical properties and high electrical conductivity in multiwalled carbon nanotubes reinforced copper matrix nanolaminated composites. Mat. Sci. Eng. A-Struct. 2018, 729, 452–457. [Google Scholar] [CrossRef]
  20. Chai, G.; Sun, Y.; Sun, J.J.; Chen, Q. Mechanical properties of carbon nanotube-copper nanocomposites. J. Micromech. Microeng. 2008, 18, 035013. [Google Scholar] [CrossRef]
  21. Daoush, W.M.; Lim, B.K.; Mo, C.B.; Nam, D.H.; Hong, S.H. Electrical and mechanical properties of carbon nanotube reinforced copper nanocomposites fabricated by electroless deposition process. Mater. Sci. Eng. A 2009, 513–514, 247–253. [Google Scholar] [CrossRef]
  22. Duan, K.; Li, L.; Hu, Y.J.; Wang, X.L. Damping characteristic of Ni-coated carbon nanotube/copper composite. Mater. Des. 2017, 133, 455–463. [Google Scholar] [CrossRef]
  23. Faria, B.; Guarda, C.; Silvestre, N.; Lopes, J.N.C.; Galhofo, D. Strength and failure mechanisms of cnt-reinforced copper nanocomposite. Compos. Part B Eng. 2018, 145, 108–120. [Google Scholar] [CrossRef]
  24. Wang, P.J.; Cao, Q.; Wang, H.P.; Nie, Y.T.; Liu, S.; Peng, Q. Fivefold enhancement of yield and toughness of copper nanowires via coating carbon nanotubes. Nanotechnology 2020, 31, 115703. [Google Scholar] [CrossRef]
  25. Yan, Y.P.; Lei, Y.X.; Liu, S. Tensile responses of carbon nanotubes-reinforced copper nanocomposites: Molecular dynamics simulation. Comput. Mater. Sci. 2018, 151, 273–277. [Google Scholar] [CrossRef]
  26. Shim, J.S.; Beom, H.G. Atomistic Investigation on the Blocking Phenomenon of Crack Propagation in Cu Substrate Reinforced by CNT. Nanomaterials 2023, 13, 575. [Google Scholar] [CrossRef]
  27. Evans, K.E. Auxetic polymers: A new range of materials. Endeavour 1991, 15, 170–174. [Google Scholar] [CrossRef]
  28. Saxena, K.K.; Das, R.; Calius, E.P. Three Decades of Auxetics Research—Materials with Negative Poisson’s Ratio: A Review. Adv. Eng. Mater. 2016, 18, 1847–1870. [Google Scholar] [CrossRef]
  29. Ma, C.; Lei, H.; Liang, J.; Wu, W.; Wang, T.; Fang, D. Macroscopic mechanical response of chiral-type cylindrical metastructures under axial compression loading. Mater. Des. 2018, 158, 198–212. [Google Scholar] [CrossRef]
  30. Fan, Y.; Wang, Y. The effect of negative Poisson’s ratio on the low-velocity impact response of an auxetic nanocomposite laminate beam. Int. J. Mech. Mater. Des. 2021, 17, 153–169. [Google Scholar] [CrossRef]
  31. Zhou, C.; Bai, L.; Gao, Y.; Liang, X.; Sun, J.; Yang, S.; Wen, M.; Wu, F.; Dong, H. Negative Poisson’s ratio effect of P2/m phosphine. J. Appl. Phys. 2022, 132, 205106. [Google Scholar] [CrossRef]
  32. Ren, X.; Das, R.; Phuong, T.; Tuan Duc, N.; Xie, Y.M. Auxetic metamaterials and structures: A review. Smart Mater. Struct. 2018, 27, 023001. [Google Scholar] [CrossRef]
  33. Li, C.; Shen, H.-S.; Yang, J. Design and nonlinear dynamics of FG curved sandwich beams with self-adapted auxetic 3D double-V meta-lattice core. Eng. Struct. 2022, 272, 115023. [Google Scholar] [CrossRef]
  34. Milstein, F.; Huang, K. Existence of a negative Poisson ratio in fcc crystals. Phys. Rev. B 1979, 19, 2030–2033. [Google Scholar] [CrossRef]
  35. Baughman, R.H.; Shacklette, J.M.; Zakhidov, A.A.; Stafstrom, S. Negative Poisson’s ratios as a common feature of cubic metals. Nature 1998, 392, 362–365. [Google Scholar] [CrossRef]
  36. Fan, Y.; Xiang, Y.; Shen, H.-S. Temperature-Dependent Mechanical Properties of Graphene/Cu Nanocomposites with In-Plane Negative Poisson’s Ratios. Research 2020, 2020, 5618021. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Fan, Y.; Shen, H.-S. Non-symmetric stiffness of origami-graphene metamaterial plates. Compos. Struct. 2022, 297, 115974. [Google Scholar] [CrossRef]
  38. Su, W.; Li, X.; Li, L.; Yang, D.; Wang, F.; Wei, X.; Zhou, W.; Kataura, H.; Xie, S.; Liu, H. Chirality-dependent electrical transport properties of carbon nanotubes obtained by experimental measurement. Nat. Commun. 2023, 14, 1672. [Google Scholar] [CrossRef] [PubMed]
  39. Zhang, H.-N.; Fan, Y.; Shen, H.-S. Chirality-Dependent and Intrinsic Auxeticity for Single-Walled Carbon Nanotubes. Materials 2022, 15, 8720. [Google Scholar] [CrossRef]
  40. Daw, M.S.; Baskes, M.I. Embedded-atom method: Derivation and application to impurities, surfaces, and other defects in metals. Phys. Rev. B 1984, 29, 6443–6453. [Google Scholar] [CrossRef] [Green Version]
  41. Foiles, S.M.; Baskes, M.I.; Daw, M.S. Embedded-atom-method functions for the fcc metals Cu, Ag, Au, Ni, Pd, Pt, and their alloys. Phys. Rev. B 1986, 33, 7983–7991. [Google Scholar] [CrossRef] [PubMed]
  42. Brenner, D.W.; Shenderova, O.A.; Harrison, J.A.; Stuart, S.J.; Ni, B.; Sinnott, S.B. A second-generation reactive empirical bond order (REBO) potential energy expression for hydrocarbons. J. Phys. Condens. Matter 2002, 14, 783–802. [Google Scholar] [CrossRef] [Green Version]
  43. Qian, C.; McLean, B.; Hedman, D.; Ding, F. A comprehensive assessment of empirical potentials for carbon materials. APL Mater. 2021, 9, 061102. [Google Scholar] [CrossRef]
  44. He, Y.Z.; Huang, F.; Li, H.; Sui, Y.W.; Wei, F.X.; Meng, Q.K.; Yang, W.M.; Qi, J.Q. Tensile mechanical properties of nano-layered copper/graphene composite. Phys. E 2017, 87, 233–236. [Google Scholar] [CrossRef]
  45. Zhang, C.; Lu, C.; Pei, L.Q.; Li, J.Q.; Wang, R.; Tieu, K. The negative Poisson’s ratio and strengthening mechanism of nanolayered graphene/Cu composites. Carbon 2019, 143, 125–137. [Google Scholar] [CrossRef]
  46. Huang, C.-W.; Chang, M.-P.; Fang, T.-H. Effects of temperature and repeat layer spacing on mechanical properties of graphene/polycrystalline copper nanolaminated composites under shear loading. Beilstein J. Nanotechnol. 2021, 12, 863–877. [Google Scholar] [CrossRef] [PubMed]
  47. Thompson, A.P.; Aktulga, H.M.; Berger, R.; Bolintineanu, D.S.; Brown, W.M.; Crozier, P.S.; Veld, P.J.i.t.; Kohlmeyer, A.; Moore, S.G.; Nguyen, T.D.; et al. LAMMPS-a flexible simulation tool for particle-based materials modeling at the atomic, meso, and continuum scales. Comput. Phys. Commun. 2022, 271, 108171. [Google Scholar] [CrossRef]
  48. Stukowski, A. Visualization and analysis of atomistic simulation data with OVITO-the Open Visualization Tool. Modelling Simul. Mater. Sci. Eng. 2010, 18, 015012. [Google Scholar] [CrossRef]
  49. Koti, V.; George, R.; Shakiba, A.; Murthy, K.V.S. Mechanical properties of Copper Nanocomposites Reinforced with Uncoated and Nickel Coated Carbon Nanotubes. FME Transactions 2018, 46, 623–630. [Google Scholar] [CrossRef]
  50. Duan, K.; Li, L.; Wang, F.; Meng, W.; Hu, Y.; Wang, X. Importance of Interface in the Coarse-Grained Model of CNT/Epoxy Nanocomposites. Nanomaterials 2019, 9, 1479. [Google Scholar] [CrossRef] [Green Version]
  51. Chu, K.; Jia, C.-c.; Li, W.-s.; Wang, P. Mechanical and electrical properties of carbon-nanotube-reinforced CuTi alloy matrix composites. Phys. Status Solidi A 2013, 210, 594–599. [Google Scholar] [CrossRef]
  52. Sun, Y.; Chen, Q. Diameter dependent strength of carbon nanotube reinforced composite. Appl. Phys. Lett. 2009, 95, 021901. [Google Scholar] [CrossRef] [Green Version]
  53. Shen, H.-S.; Zhang, C.-L. Thermal buckling and postbuckling behavior of functionally graded carbon nanotube-reinforced composite plates. Mater. Des. 2010, 31, 3403–3411. [Google Scholar] [CrossRef]
  54. Liu, J.; Fan, G.; Tan, Z.; Guo, Q.; Su, Y.; Li, Z.; Xiong, D.-B. Mechanical properties and failure mechanisms at high temperature in carbon nanotube reinforced copper matrix nanolaminated composite. Compos. Part A Appl. Sci. Manuf. 2019, 116, 54–61. [Google Scholar] [CrossRef]
  55. Zhao, K.J.; Chen, C.Q.; Shen, Y.P.; Lu, T.J. Molecular dynamics study on the nano-void growth in face-centered cubic single crystal copper. Comput. Mater. Sci. 2009, 46, 749–754. [Google Scholar] [CrossRef]
Figure 1. The schematic of the (a) single-crystal copper cell, (b) representative cell (green) along [1 1 0] crystal orientation in the SWCNT/Cu nanocomposite, (c) SWCNT/Cu nanocomposite (Cu atoms are in blue, and C atoms are in red), (d) front view of the nanocomposite (purple Cu atoms are set fixed after equilibration when simulating shear deformation).
Figure 1. The schematic of the (a) single-crystal copper cell, (b) representative cell (green) along [1 1 0] crystal orientation in the SWCNT/Cu nanocomposite, (c) SWCNT/Cu nanocomposite (Cu atoms are in blue, and C atoms are in red), (d) front view of the nanocomposite (purple Cu atoms are set fixed after equilibration when simulating shear deformation).
Nanomaterials 13 01885 g001
Figure 2. Different options of gap between Cu and SWCNT. (ac) initial construction (green shade annulus represents gap area), (df) atomic stress distribution after energy minimization, and (gi) crystal lattice structure after equilibration.
Figure 2. Different options of gap between Cu and SWCNT. (ac) initial construction (green shade annulus represents gap area), (df) atomic stress distribution after energy minimization, and (gi) crystal lattice structure after equilibration.
Nanomaterials 13 01885 g002
Figure 3. Comparison of Young’s modulus of CNT/Cu composites with varied CNT weight fractions at 300 K. MLM: molecular-level mixing; HEBM: high-energy ball milling; SPS: spark plasma sintering; ED: electroless deposition; HP: hot pressing; HR: hot rolling [14,15,16,17,18,19,21].
Figure 3. Comparison of Young’s modulus of CNT/Cu composites with varied CNT weight fractions at 300 K. MLM: molecular-level mixing; HEBM: high-energy ball milling; SPS: spark plasma sintering; ED: electroless deposition; HP: hot pressing; HR: hot rolling [14,15,16,17,18,19,21].
Nanomaterials 13 01885 g003
Figure 4. The variation of Young’s moduli E11 (a,d,g), E22 (b,e,h), and E33 (c,f,i) of SWCNT/Cu nanocomposites at 300 K with different lengths in y- and z-direction and corresponding SWCNT weight fractions (The green arrows denote the direction of size change).
Figure 4. The variation of Young’s moduli E11 (a,d,g), E22 (b,e,h), and E33 (c,f,i) of SWCNT/Cu nanocomposites at 300 K with different lengths in y- and z-direction and corresponding SWCNT weight fractions (The green arrows denote the direction of size change).
Nanomaterials 13 01885 g004
Figure 5. The variation of Young’s moduli and Poisson’s ratios of SWCNT/Cu nanocomposites at 300 K with different diameter and weight fraction of SWCNTs. (a) MD models for the nanocomposites with same size but different SWCNT or similar weight fraction but different size; (b) Young’s moduli and (c) in-plane Poisson’s ratios of the nanocomposites corresponding to the 4 models on the top in (a); (d) Young’s moduli and (e) in-plane Poisson’s ratios of the nanocomposites corresponding to the first model on the top left and 3 models on the bottom in (a).
Figure 5. The variation of Young’s moduli and Poisson’s ratios of SWCNT/Cu nanocomposites at 300 K with different diameter and weight fraction of SWCNTs. (a) MD models for the nanocomposites with same size but different SWCNT or similar weight fraction but different size; (b) Young’s moduli and (c) in-plane Poisson’s ratios of the nanocomposites corresponding to the 4 models on the top in (a); (d) Young’s moduli and (e) in-plane Poisson’s ratios of the nanocomposites corresponding to the first model on the top left and 3 models on the bottom in (a).
Nanomaterials 13 01885 g005
Figure 6. The variation of Young’s moduli: (a) E11, (b) E22, and (c) E33 of SWCNT/Cu nanocomposites with different SWCNT weight fractions under a thermal environment.
Figure 6. The variation of Young’s moduli: (a) E11, (b) E22, and (c) E33 of SWCNT/Cu nanocomposites with different SWCNT weight fractions under a thermal environment.
Nanomaterials 13 01885 g006
Figure 7. The variation of shear moduli: (a) G12, (b) G13, and (c) G23 of SWCNT/Cu nanocomposites with different SWCNT weight fractions under a thermal environment.
Figure 7. The variation of shear moduli: (a) G12, (b) G13, and (c) G23 of SWCNT/Cu nanocomposites with different SWCNT weight fractions under a thermal environment.
Nanomaterials 13 01885 g007
Figure 8. Thermal effect on in-plane Poisson’s ratios: (a) υ12 and (b) υ21 of nanocomposites with different SWCNT weight fractions.
Figure 8. Thermal effect on in-plane Poisson’s ratios: (a) υ12 and (b) υ21 of nanocomposites with different SWCNT weight fractions.
Nanomaterials 13 01885 g008
Figure 9. Thermal effect on TECs: (a) α11, (b) α22, (c) α33, and (d) ρ of SWCNT/Cu nanocomposites with different SWCNT weight fractions.
Figure 9. Thermal effect on TECs: (a) α11, (b) α22, (c) α33, and (d) ρ of SWCNT/Cu nanocomposites with different SWCNT weight fractions.
Nanomaterials 13 01885 g009
Figure 10. (a) Schematic graphic of three components for TECs of SWCNT/Cu nanocomposites. The effect of three components on TECs (bd) α22 and (eg) α33 of nanocomposites with 1 wt.%, 3 wt.%, and 5 wt.% SWCNTs.
Figure 10. (a) Schematic graphic of three components for TECs of SWCNT/Cu nanocomposites. The effect of three components on TECs (bd) α22 and (eg) α33 of nanocomposites with 1 wt.%, 3 wt.%, and 5 wt.% SWCNTs.
Nanomaterials 13 01885 g010
Table 1. Poisson’s ratios for chiral SWCNTs at 4.3 K [39] and 300 K (sorted by υ12).
Table 1. Poisson’s ratios for chiral SWCNTs at 4.3 K [39] and 300 K (sorted by υ12).
Chiral Angle θDiameter d (Å)(n, m)υ12 at 4.3 Kυ12 at 300 K
10.89°3.59(4, 1)0.03850.0437
8.95°4.36(5, 1)0.05830.0699
19.11°4.14(4, 2)0.06940.0792
7.59°5.13(6, 1)0.08160.0952
16.10°4.89(5, 2)0.08730.0988
6.59°5.91(7, 1)0.10610.1106
13.90°5.65(6, 2)0.10800.1143
5.82°6.69(8, 1)0.12820.1301
12.22°6.41(7, 2)0.12880.1312
Table 2. Temperature-dependent elastic moduli and shear moduli of SWCNT/Cu nanocomposites.
Table 2. Temperature-dependent elastic moduli and shear moduli of SWCNT/Cu nanocomposites.
wt.%T (K)E11 (GPa)E22 (GPa)E33 (GPa)G12 (GPa)G13 (GPa)G23 (GPa)
0%300101.10101.1455.89018.56268.97568.946
40096.80796.71253.23517.53766.54866.555
50091.80992.03950.31816.63564.19864.150
60087.79287.41447.65515.65961.73761.746
70082.77782.59044.67314.76259.31859.303
80077.52477.28741.41013.77256.72256.683
1%300160.47101.0265.47817.21858.51760.179
400156.5197.33362.92316.34656.73257.740
500151.6993.07860.04115.69155.11355.789
600147.0189.82658.05915.01952.63953.101
700143.6586.39755.53314.03751.04351.025
800139.0182.13952.79313.15648.39148.699
3%300226.2496.91868.33615.49648.93249.838
400221.5593.17466.36214.77847.06148.416
500216.7089.18663.62914.09045.50846.357
600212.4685.02661.31113.56143.54745.030
700209.8481.31159.18512.58442.14543.688
800202.1378.09056.10011.70740.21642.060
5%300270.5890.43366.89714.91446.18943.764
400266.8487.12764.68114.03345.49342.899
500261.2583.46062.33313.39743.73241.059
600255.8980.24659.48012.72542.10039.943
700251.0276.11757.16711.87640.32239.127
800245.0472.07054.32210.80737.77637.549
7%300312.0187.27865.31414.03941.79339.616
400308.7483.87263.24113.38040.91138.842
500302.6180.25460.44812.59339.40737.000
600298.2277.51257.84311.83337.69435.759
700290.7373.51655.21810.90935.85734.479
800286.5069.37052.77010.03933.60832.890
8%300344.0984.88162.64713.53135.87234.853
400341.0081.43460.27712.85034.47333.951
500334.0078.16557.30012.24333.92433.109
600328.3075.02354.40111.43432.45431.559
700321.5771.12951.35010.44331.07230.365
800313.3266.80148.8029.792130.05028.922
Table 3. In-plane properties of copper, copper with the defect, and the SWCNT/Cu nanocomposite with 8.29 wt.% at 4.3 K.
Table 3. In-plane properties of copper, copper with the defect, and the SWCNT/Cu nanocomposite with 8.29 wt.% at 4.3 K.
CaseE11 (GPa)E22 (GPa)G12 (GPa)υ12υ21
Case 1114.99115.0721.499−0.1858−0.1855
Case 284.64953.7502.9898−0.1439−0.0876
Case 3361.1367.51710.969−0.1506−0.0287
Difference 2/1−26.4%−53.3%−86.1%−22.6%−52.8%
Difference 3/2326.6%25.6%266.9%4.7%−67.2%
Difference 3/1214.1%−41.3%−49.0%−18.9%−84.5%
Case 495.88574.5608.4298−0.1497−0.1168
Case 5360.2192.90716.061−0.1200−0.0316
Difference 4/1−16.6%−35.2%−60.8%−19.4%−37.0%
Difference 5/4275.7%24.6%90.5%−19.8%−72.9%
Difference 5/1213.3%−19.3%−25.3%−35.4%−83.0%
Case 1: (104.80 × 15.34 × 18.08) Å3; 2460 Cu atoms. Case 2: (104.80 × 15.34 × 18.08) Å3; 1517 Cu atoms and 3.0 Å diameter hole. Case 3: (104.80 × 15.34 × 18.08) Å3; 1517 Cu atoms, 804 C atoms of SWCNT and 3.0 Å gap. Case 4: (104.80 × 15.34 × 18.08) Å3; 1681 Cu atoms and 2.3 Å diameter hole. Case 5: (104.80 × 15.34 × 18.08) Å3; 1681 Cu atoms, 804 C atoms of SWCNT and 2.3 Å gap. Difference i/j = [(XiXj)/Xj] × 100%; X = E, G, υ.
Table 4. Temperature-dependent Poisson’s ratios of SWCNT reinforced Cu nanocomposites.
Table 4. Temperature-dependent Poisson’s ratios of SWCNT reinforced Cu nanocomposites.
wt.%T (K)υ12υ21υ13υ31υ23υ32
0%300−0.2053−0.20530.92160.42160.92150.4216
400−0.2125−0.21330.93300.42280.93390.4232
500−0.2217−0.22110.94760.42500.94670.4251
600−0.2304−0.23080.96030.42620.96090.4265
700−0.2409−0.24120.97630.42830.97660.4284
800−0.2547−0.25530.99640.43090.99740.4310
1%300−0.1745−0.11050.83260.29720.83700.4664
400−0.1779−0.11130.83710.29710.83930.4672
500−0.1833−0.11310.84700.29660.84600.4672
600−0.1894−0.11560.86030.29410.85010.4721
700−0.1945−0.11830.86640.29100.85270.4708
800−0.1996−0.11980.87540.28990.85620.4744
3%300−0.1537−0.06920.75080.20680.81210.4983
400−0.1574−0.06880.75720.20430.81560.5017
500−0.1637−0.06860.76770.20230.82150.5053
600−0.1695−0.07340.77140.20060.82840.5073
700−0.1762−0.07500.78330.20010.83160.5090
800−0.1814−0.07520.79840.19800.83790.5102
5%300−0.1463−0.04800.70750.16150.79750.5230
400−0.1505−0.04870.71360.15990.80110.5233
500−0.1568−0.05260.72430.15830.80660.5244
600−0.1620−0.05290.73180.15650.81050.5279
700−0.1671−0.05490.74570.15850.81870.5297
800−0.1739−0.05540.75540.15880.82650.5300
7%300−0.1356−0.04090.66670.12720.80640.5287
400−0.1405−0.04050.67270.12780.81090.5312
500−0.1474−0.04320.68100.12530.81480.5293
600−0.1551−0.04460.69050.12090.81940.5321
700−0.1616−0.04590.70970.12010.82530.5369
800−0.1691−0.04550.72880.12240.83260.5361
8%300−0.1290−0.03230.63240.10370.81890.5367
400−0.1332−0.03260.63850.10430.82350.5381
500−0.1392−0.03480.64790.10410.83010.5381
600−0.1472−0.03770.66320.10230.83580.5393
700−0.1541−0.03830.67950.10460.83880.5413
800−0.1629−0.03880.70130.10590.84510.5411
Table 5. Temperature-dependent TECs of SWCNT/Cu nanocomposites.
Table 5. Temperature-dependent TECs of SWCNT/Cu nanocomposites.
wt.%T (K)α11 (×10−6 K−1)α22 (×10−6 K−1)α33 (×10−6 K−1)ρ (g/cm3)
0%30016.65916.71516.6778.8043
40017.18317.22917.2458.7597
50017.78317.81217.8478.7138
60018.45918.46218.4868.6665
70019.20619.17419.1658.6177
80020.02419.94419.8868.5673
1%30011.43314.48325.4618.3552
40011.42014.89925.5048.3131
50011.40715.16526.0738.2703
60011.39415.33826.9968.2265
70011.38115.47628.1048.1818
80011.36815.63629.2288.1360
3%3008.336612.64928.8297.7709
4008.329612.82529.2667.7328
5008.322713.00029.9207.6940
6008.315813.17531.0037.6542
7008.308913.34932.2997.6134
8008.302013.52233.5977.5717
5%3007.274711.66830.0727.3495
4007.491911.82330.5907.3139
5007.451111.87831.3927.2778
6007.471911.74232.5447.2408
7007.253611.38934.0717.2030
8007.248411.37635.7547.1644
7%3006.504610.04230.6506.9240
4006.500410.03131.4446.8913
5006.496210.02132.4626.8580
6006.492010.01133.6986.8240
7006.487810.00135.1516.7893
8006.48359.991436.8146.7539
8%3006.08739.08430.8156.6027
4006.08369.07531.6766.5722
5006.07999.06733.1936.5408
6006.07629.05934.7446.5087
7006.07259.05136.2786.4760
8006.06889.04237.7946.4430
100%
(7, 2)
3003.81505.5715 (α22 = α33)
4003.86416.2998 (α22 = α33)
5003.91317.0267 (α22 = α33)
6003.96217.7521 (α22 = α33)
7004.01108.4758 (α22 = α33)
8004.05999.1977 (α22 = α33)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, H.-N.; Fan, Y.; Shen, H.-S. Prediction of Temperature-Dependent Mechanical Properties for SWCNT/Cu Nanocomposite Metamaterials: A Molecular Dynamics Study. Nanomaterials 2023, 13, 1885. https://doi.org/10.3390/nano13121885

AMA Style

Zhang H-N, Fan Y, Shen H-S. Prediction of Temperature-Dependent Mechanical Properties for SWCNT/Cu Nanocomposite Metamaterials: A Molecular Dynamics Study. Nanomaterials. 2023; 13(12):1885. https://doi.org/10.3390/nano13121885

Chicago/Turabian Style

Zhang, Hai-Ning, Yin Fan, and Hui-Shen Shen. 2023. "Prediction of Temperature-Dependent Mechanical Properties for SWCNT/Cu Nanocomposite Metamaterials: A Molecular Dynamics Study" Nanomaterials 13, no. 12: 1885. https://doi.org/10.3390/nano13121885

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