Next Article in Journal
Study on Load Transfer Mechanism of Local Curved Prestressed Hollow-Core Slab Bridge
Previous Article in Journal
Carbon Peaking and Carbon Neutrality in the Cement-Based Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Size- and Chirality-Dependent Structural and Mechanical Properties of Single-Walled Phenine Nanotubes

1
Faculty of Civil Engineering and Mechanics, Jiangsu University, Zhenjiang 212013, China
2
Zienkiewicz Centre for Computational Engineering, Faculty of Science and Engineering, Bay Campus, Swansea University, Swansea SA1 8EN, Wales, UK
3
Key Laboratory for Intelligent Nano Materials and Devices of Ministry of Education, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(13), 4706; https://doi.org/10.3390/ma16134706
Submission received: 17 May 2023 / Revised: 19 June 2023 / Accepted: 22 June 2023 / Published: 29 June 2023
(This article belongs to the Section Carbon Materials)

Abstract

:
Phenine nanotubes (PNTs) have recently been synthesized as a promising new one-dimensional material for high-performance electronics. The periodically distributed vacancy defects in PNTs result in novel semiconducting properties, but may also compromise their mechanical properties. However, the role of these defects in modifying the structural and mechanical properties is not yet well understood. To address this, we conducted systematic molecular dynamics simulations investigating the structural evolution and mechanical responses of PNTs under various conditions. Our results demonstrated that the twisting of linear carbon chains in both armchair and zigzag PNTs led to interesting structural transitions, which were sensitive to chiralities and diameters. Additionally, when subjected to tensile and compressive loading, PNTs’ cross-sectional geometry and untwisting of linear carbon chains resulted in distinct mechanical properties compared to carbon nanotubes. Our findings provide comprehensive insights into the fundamental properties of these new structures while uncovering a new mechanism for modifying the mechanical properties of one-dimensional nanostructures through the twisting–untwisting of linear carbon chains.

1. Introduction

One-dimensional (1D) tubular structures have garnered considerable interest from the scientific community due to their remarkable features that make them useful in various applications, including the reinforcement of composites [1,2,3,4], electronic devices [5,6,7,8], drug delivery [2,9], among others. Specifically, carbon nanotubes (CNTs) possess unique properties such as a narrow and seamless crystal structure, exceptional electronic transport properties, and chemical stability [1,10,11,12,13,14,15,16], which make them promising candidates for electronic devices. However, fully exploiting the potential of CNTs for the semiconducting industry remains uncertain, largely due to their band gap being dependent on their chiralities. Therefore, synthesized CNT arrays often consist of both metallic and semiconducting tubes. While some strategies have been developed to address this issue, such as breaking down metallic CNTs [17,18,19,20,21] or synthesizing CNTs with specific chiralities [22,23,24,25,26], these methods usually result in low yields and are therefore unsuitable for industrialization. Recently, Sun et al. succeeded in synthesizing a novel type of tubular carbon called phenine nanotubes (PNTs). PNTs can be thought of as porous graphene layers rolled into tubes [27]. One advantage of this member of the carbon allotrope family is that it possesses a finite band gap that is well-suited for semiconducting applications.
While the promising electronic properties of PNTs have been well studied, less is known about their mechanical properties in comparison to perfect CNTs or those with localized defects. It is widely accepted that the presence of defects in PNTs will lead to a deterioration in their ability to withstand external loads, but the unique porous nature of this new structure and its impact on the structural, elastic, and failure behaviors of PNTs remains unclear. CNTs are known for their extremely high strength and can undergo plastic behavior at high temperatures [3,11,28,29,30] and through various types of defect formation such as Stone–Wales transformation. However, vacancy defects typically reduce the strength of CNTs and do not facilitate ductile deformation [31,32,33]. PNTs with regularly patterned pores can be considered as a specific type of CNTs with periodically arranged vacancies, suggesting that the mechanical strength of PNTs is likely to be degraded; this is confirmed by recent simulation results [34]. It is important to note that defects in PNTs are not localized, resulting in a weakened tubular skeleton along the transverse direction. Recent calculations by Yu et al. [35] showed that the cross-sectional geometry of PNTs does not retain a circular shape for some cases, thereby causing the collapse of bundles of small-sized PNTs to flattened structures. However, there has been no systematic investigation into how these geometric changes affect the mechanical properties of PNTs. Such features are likely correlated to the sizes and chiralities of the PNTs and require comprehensive studies into their atomistic details.
In this study, we have examined the structural and mechanical properties of armchair and zigzag periodic nanotubes (PNTs) using molecular dynamics simulations. Our results have shown that PNTs exhibit unique structural features based on their diameter and chirality. For armchair PNTs, the cross-sectional shape changes from a circular to a polygonal shape as the diameter increases. When the diameter is greater than 50 Å, complex structures are observed. On the other hand, for zigzag PNTs, the circular cross-sectional shape remains until the diameter reaches around 50 Å, beyond which it transforms into an oval shape. At diameters larger than 150 Å, it tends to collapse towards a flattened geometry. Such anisotropic structural features lead to distinct mechanical properties of PNTs under tensile and compressive loads, which are common loads in mechanical devices. We have investigated the elastic properties and failure modes of PNTs, which were found to show size dependency. Our study provides fundamental insights into how periodic defects modify the structural and mechanical properties of 1D nanotube systems. This allows us to establish a comprehensive database for further studies when selecting this novel structure for functional nano devices.

2. Simulation Details

We conducted all simulations using the large-scale atomic/molecular massively parallel simulator (LAMMPS) package [36]. Visualization was carried out using the OVITO package [37]. The construction of the PNT structures is described in Section 3. To describe interatomic interactions, we adopted the AIREBO potential [38], which has proven successful in studying the structural and mechanical properties of carbon nanomaterials. To avoid a nonphysical description of mechanical properties by the original AIREBO potential, we modified the cutoff distance to 2 Å [39,40,41]. Geometry optimization was first performed using the conjugate gradient method with periodic boundary conditions. The system was then relaxed in the NVT ensemble with a time step of 1 fs until it reached an equilibrium state; this time step is commonly used in studying nano carbon systems [34,42]. It is worth noting that the exact values of the mechanical properties could be different if one choose other force fields, such as the reax force field. As the focus of this work is to uncover a correlation between structural evolution and mechanical responses, we did not concentrate our discussions on this aspect.
For investigating the mechanical properties of the structure, it was uniformly stretched or compressed along its axial direction with a strain rate of 0.1/ps until failure, which occurred at 1 K in the NVT ensemble. The results from this strain rate are comparable to those obtained by Yu et al. [35]. This allowed us to obtain insights into the behavior of the structure under tensile and compressive loading, and provided valuable information on its elastic properties and failure modes.

3. Results and Discussion

PNTs can be viewed as seamless rolled-up structures of graphene sheets, analogous to CNTs. As shown in Figure 1, rolling up along different directions leads to PNTs with different chiralities, such as armchair and zigzag types. For the ease of explanation, if the sheet is rolled up along x direction (corresponding to armchair edge) in the left panel of Figure 1, an armchair PNT is produced, as shown in the top right of Figure 1, while if rolled up vertically along the y direction (corresponding to zigzag edge), a zigzag PNT is obtained, as presented in the lower right panel of Figure 1. Due to the presence of defects, PNTs exhibited distinct structural properties compared to perfect CNTs. Our MD simulations revealed these differences for both armchair and zigzag PNTs, as seen in Figure 2, Figure 3 and Figure 4. To ensure equilibrium states were reached, energy minimization was performed first followed by relaxation at 1 K for 100 ps, or up to 2500 ps for larger models.
In the case of armchair PNTs shown in Figure 2, an interesting phenomenon was observed during energy minimization: the cross-sectional geometry transitions to a polygonal type when the diameter is small (less than 20.6 Å), while no sudden changes were observed during structural relaxation at a finite temperature for the duration of the simulations. This transition is attributed to the twisting of the linear carbon chain along zigzag edges, which reduces the PNT’s energy by about 0.33 eV/atom (Figure S2). The schematic twisting behavior is presented in Figure 3a, where a C-C bond linked to 2 benzene rings can be rotated along the bond direction; a series of rotations along the zigzag chain leads to notable reduction in energy. As a result, the type of polygon formed corresponds exactly to the number of linear zigzag chains in the PNT structure. As the diameter of the PNT increases, the trend varies, as depicted in Figure 2. For example, for a PNT with diameter D = 24.75 Å and 12 linear zigzag chains, the energy-minimized geometry exhibits a hexagonal cross-sectional structure that deforms further after finite-temperature relaxation. Consequently, the twisting angles of different linear carbon chains vary, indicating that the deformation modes become inhomogeneous and localized. The details of twisting angle variations can be found in Figure 3. With further diameter increases, the cross-section becomes more severely deformed and even collapses into a flattened shape at D = 53.6 Å. This effect arises because the linear carbon chain along the axis of the PNT is no longer able to resist radial deformation, and the van der Waals interaction between opposite walls plays a dominant role in reducing the total energy of the PNT system. From the potential energy plot and selected geometry evolution shown in Figure S2, it can be observed that for small-diametered PNTs, energy reduction during the relaxation process primarily comes from the twisting of carbon chains. On the other hand, for larger-diametered PNTs, the energy decreases gradually initially, suggesting that the contribution from twisting is less significant, while the flattening of the tube wall leads to a more significant decrease in potential energy (see, e.g., Figure S2c), consistent with the above analysis.
The evolution of the cross sectional shape in zigzag PNTs displays a different trend, owing to the fact that there is an alignment of the carbon chains either circularly or as a 30° helix along the axis. The high axial symmetry resulting from these alignments prevents any major transformation to a polygon type. Nonetheless, the twisting of the zigzag carbon chains causes an increase in wall thickness, as explicated in Figure 4. In small diametered zigzag PNTs, the twisting angle is notably strong compared to that seen in armchair PNTs. Larger PNTs, however, do not exhibit evenly distributed twisting. Check Figure S3 for further details. Due to the helical and circular alignment of carbon chains, PNTs show a slower collapse trend with an increasing diameter. It was not until D = 157.2 Å that the completely flattened structure was observed in zigzag PNTs, almost thrice that of their armchair counterparts. Nonetheless, there are certain overall similarities in the size-dependent correlation between the energy release mechanism and geometry revolution, as seen in Figure S4. Thus, we shall skip any detailed discussions of these shared premises here.
We next proceed to investigate the mechanical behavior of PNTs under various strain conditions. Figure 5 displays the tensile responses of selected armchair and zigzag PNTs. It is evident from the stress–strain relationship that both PNT types are brittle materials and undergo structural failure after reaching peak stress values, resulting in separated segments. This behavior is consistent with covalently bonded crystals such as CNTs and graphene. Note here that some simulations using the reax force field have reported a ductile deformation mode in nanocarbon materials [43]; this is different from the approach used in the present study. By computing the slopes of the stress–strain curves, our results reveal that armchair PNTs possess a higher Young’s modulus than zigzag PNTs. The average value for armchair PNTs is roughly 214 GPa, while it is approximately 154 GPa for zigzag PNTs (see Figure S5 in the Supplementary Materials). We attribute this difference to the alignment of the carbon chains. In armchair PNTs, the chains are aligned with the PNT axis, whereas in zigzag PNTs, they form angles with the axis, making them less capable of resisting tensile loading. Notably, the stress–strain plot for the 11.91 Å zigzag PNT exhibits a much lower slope. As a result, its modulus is much smaller (around 114 GPa) than that of the larger-sized zigzag PNTs (160–170 GPa). Upon analyzing the structure, we found that for zigzag PNTs, the length contracts significantly with a decreasing diameter, similar to the bending Poisson effect observed by Liu et al. [44]. (Here the contract ratio is measured by (L0 − L)/L0, where L0 is the length of the original model after the rolling up of the defected graphene sheet, and L is the length after geometry optimization). Consequently, when subjected to tensile loading, the response of the PNT is slower. Although armchair PNTs also exhibit longitudinal contraction, the contraction ratio is not sensitive to their diameters (see Figure S6 in the Supplementary Materials), resulting in a weaker diameter dependence of the Young’s modulus.
Comparing PNTs with CNTs and graphene, it can be concluded that the ultimate stress and Young’s modulus of PNTs are significantly lower, approximately 30 GPa and 184 GPa, respectively. In contrast, the corresponding values for CNTs are approximately 200 GPa and 1 TPa. This is attributable to the fact that PNTs fall under the category of defective CNTs, impeding their capacity to carry external loads. The overall mechanical characters investigated in this study are summarized in Table 1 and Table 2.
Figure 5c,d display the extracted ultimate stress and strain from the stress–strain curves, revealing that the aspect ratio has no considerable impact on the ultimate stress and strain of the PNTs. The diameter, on the other hand, plays a crucial role. The ultimate strain demonstrates a decreasing trend with an increasing diameter until it converges, with the ultimate stress displaying similar behavior. This reason is due to two aspects: firstly, as discussed earlier, PNTs contract at small diameters, facilitating deformation under tension; secondly, carbon chain twisting sensitivity varies according to the tube diameter. For small-diametered PNTs, the degree of twisting is stronger (up to 80°), leading to the untwisting of carbon chains under tensile loading. Figure 6 represents a typical result obtained from a D = 16.50 Å PNT. The twist angle steadily drops from 80° to 55° as the tensile strain rises to 12%. It then remains relatively stable before the breaking point when a failure transpires at a strain level of 21%, demonstrating an almost complete fracture of the PNTs since the twist angles fluctuate and return to values near the original ones. The immediately fractured structures of both armchair and zigzag PNTs are shown in Figure 7, where brittle failure mode is clearly seen.
Figure 8 illustrates the compressive behavior of PNTs featuring various chiralities and diameters. It is noteworthy that the D = 16.5 Å PNT exhibits a substantially higher stress level than the other PNTs studied in our studies, showing a peak stress value of 9.4 GPa. In contrast, the peak values for other armchair PNTs are lower than 3 GPa. This unusual behavior arises from the progressive collapse of cross-sectional geometries, meaning that larger diametered PNTs are less capable of withstanding compressive loadings. On the other hand, for zigzag PNTs, the critical stress shows a slightly different trend regarding the PNT diameter, i.e., it gradually decreases with the diameter. As discussed earlier, this is partly due to the relatively well-preserved circular cross-sectional geometry in zigzag PNTs. Additionally, as the linear carbon chain does not align along the axial direction, the maximum compressive stress that zigzag PNTs can withstand is smaller than that of armchair PNTs. The inverse correlation between PNT diameter and the critical buckling stress aligns with previous CNT buckling theories [45].
Interestingly, the critical buckling strain for zigzag PNTs is substantially higher than that for armchair PNTs. The underlying mechanisms can be better comprehended by investigating the structural evolution of two representative PNTs, as illustrated in Figure 9. Upon observation, it becomes evident that for the armchair PNT, the compressive stress is primarily experienced by the linear carbon chains along the axial direction, rapidly reaching a critical buckling strain similar to that in a rod. Meanwhile, for zigzag PNTs, the load transfer pathway is not along the axial direction, and the deformation mode in the linear carbon chains is not exclusively compression but is also bending-induced bond angle change. Therefore, the strain energy stored in the zigzag PNT is significantly smaller at the same level of strain, thereby leading to delayed buckling events. This scenario is further corroborated by energy storage density simulations performed in our experiments, as shown in Figure 10. Here, the energy storage capacity is calculated as the following:
Δ E M = Δ E n ( C ) × A r ( C ) × m ( C 12 ) 12 + n ( H ) × A r ( H ) × m ( C 12 ) 12
where Δ E represents the energy change in the relaxation process; n represents the number of atoms; A r represents the relative atomic mass; m ( C 12 ) represents the carbon-12 mass. The average energy density for armchair PNTs is approximately 30,000 KJ/Kg, while for zigzag PNTs, it is about 19,200 KJ/Kg. While an energy density lower than CNTs is expected, the density is nevertheless higher than recently predicted values of nanothread [46].

4. Conclusions

Extensive molecular dynamics simulation studies have been conducted on the structural and mechanical properties of recently synthesized phenine nanotubes (PNTs). The results demonstrate that due to distortion in their linear carbon chains, PNTs exhibit distinct cross-sectional geometry variation trends depending on their chirality and diameters. These phenomena further modify the PNTs’ mechanical responses to tensile and compressive loadings. Armchair PNTs are stronger than zigzag PNTs, as evidenced by a higher Young’s modulus and energy storage capacity, which can be attributed to the alignment of the linear carbon chains along the axial direction. However, PNTs’ overall mechanical properties are weaker than those of carbon nanotubes (CNTs), since they can be regarded as defective CNTs. The intriguing role of twisting-untwisting in tuning the mechanical properties of PNTs offers new insights into the mechanics of advanced nanostructures. It is expected that these results will inspire future studies aimed at designing new functional materials with desirable properties.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ma16134706/s1, Figure S1: Twisting behavior of linear carbon chains in PNTs; Figure S2: Structural evolution for 3 selected armchair PNTs as a function of time, the corresponding potential energy are plotted in the top panel, while the snapshots are shown in the bottom; Figure S3: Twisting of helical carbon chains in zigzag PNTs; Figure S4: Structural evolution for 3 selected zigzag PNTs as a function of time, the corresponding potential energy are plotted in the top panel, while the snapshots are shown in the bottom; Figure S5: Young’s modulus of PNTs under tensile loadings; Figure S6: Contract of PNT length when rolled up from defected graphene sheets.

Author Contributions

Conceptualization, Y.L. and C.T.; Methodology, R.W., L.W. and J.X.; Software, Y.L., R.W., L.W. and J.X.; Formal analysis, R.W., L.W., J.X. and C.W.; Resources, C.T.; Data curation, Y.L.; Writing—review & editing, Y.L., R.W. and C.T.; Supervision, C.W. and C.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant Nos. 12072134, 12102151), NJ2022002 (INMD-2022M03), and the Jiangsu Province Postdoctoral Foundation (Grant No. 2021K113B).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Available upon request to corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gao, J.; Itkis, M.E.; Yu, A.; Bekyarova, E.; Zhao, B.; Haddon, R.C. Continuous spinning of a single-walled carbon nanotube−nylon composite fiber. J. Am. Chem. Soc. 2005, 127, 3847–3854. [Google Scholar] [CrossRef]
  2. Yang, X.; Tang, H.; Cao, K.; Song, H.; Sheng, W.; Wu, Q. Templated-assisted one-dimensional silica nanotubes: Synthesis and applications. J. Mater. Chem. 2011, 21, 6122–6135. [Google Scholar] [CrossRef]
  3. Thostenson, E.T.; Ren, Z.; Chou, T.-W. Advances in the science and technology of carbon nanotubes and their composites: A review. Compos. Sci. Technol. 2001, 61, 1899–1912. [Google Scholar] [CrossRef] [Green Version]
  4. Qian, D.; Wagner, G.J.; Liu, W.K.; Yu, M.-F.; Ruoff, R.S. Mechanics of carbon nanotubes. Appl. Mech. Rev. 2002, 55, 495–533. [Google Scholar] [CrossRef]
  5. Xiang, J.; Yu, X.-Y.; Paik, U. General synthesis of vanadium-based mixed metal oxides hollow nanofibers for high performance lithium-ion batteries. J. Power Sources 2016, 329, 190–196. [Google Scholar] [CrossRef]
  6. Wang, C.; Takei, K.; Takahashi, T.; Javey, A. Carbon nanotube electronics—Moving forward. Chem. Soc. Rev. 2013, 42, 2592–2609. [Google Scholar] [CrossRef]
  7. Peng, L.-M.; Zhang, Z.; Qiu, C. Carbon nanotube digital electronics. Nat. Electron. 2019, 2, 499–505. [Google Scholar] [CrossRef]
  8. Raychowdhury, A.; Roy, K. Carbon nanotube electronics: Design of high-performance and low-power digital circuits. IEEE Trans. Circuits Syst. I Regul. Pap. 2007, 54, 2391–2401. [Google Scholar] [CrossRef]
  9. Madani, S.Y.; Naderi, N.; Dissanayake, O.; Tan, A.; Seifalian, A.M. A new era of cancer treatment: Carbon nanotubes as drug delivery tools. Int. J. Nanomed. 2011, 6, 2963–2979. [Google Scholar]
  10. Choudhary, M.; Sharma, A.; Aravind Raj, S.; Sultan, M.T.H.; Hui, D.; Shah, A.U.M. Contemporary review on carbon nanotube (CNT) composites and their impact on multifarious applications. J. Am. Chem. Soc. 2022, 11, 2632–2660. [Google Scholar] [CrossRef]
  11. Vigolo, B.; Pénicaud, A.; Coulon, C.; Sauder, C.; Pailler, R.; Journet, C.; Bernier, P.; Poulin, P. Macroscopic fibers and ribbons of oriented carbon nanotubes. Science 2000, 290, 1331–1334. [Google Scholar] [CrossRef]
  12. Iijima, T.; Oshima, H.; Hayashi, Y.; Suryavanshi, U.B.; Hayashi, A.; Tanemura, M.J.D.; Materials, R. In-situ observation of carbon nanotube fiber spinning from vertically aligned carbon nanotube forest. Diam. Relat. Mater. 2012, 24, 158–160. [Google Scholar] [CrossRef]
  13. Jee, M.H.; Choi, J.U.; Park, S.H.; Jeong, Y.G.; Baik, D.H. Influences of tensile drawing on structures, mechanical, and electrical properties of wet-spun multi-walled carbon nanotube composite fiber. Macromol. Res. 2012, 20, 650–657. [Google Scholar] [CrossRef]
  14. Liu, C.; Fan, Y.Y.; Liu, M.; Cong, H.T.; Cheng, H.M.; Dresselhaus, M.S. Hydrogen Storage in Single-Walled Carbon Nanotubes at Room Temperature. Science 1999, 286, 1127–1129. [Google Scholar] [CrossRef] [Green Version]
  15. Ajayan, P.M. Nanotubes from Carbon. Chem. Rev. 1999, 99, 1787–1800. [Google Scholar] [CrossRef] [PubMed]
  16. García-Hernández, E.; Chigo-Anota, E. Structural defects on (5,5) single-walled carbon nanotubes: Impact on their electronic properties and chemical reactivity from a DFT perspective. Physica E Low Dimens. Syst. 2021, 134, 114874. [Google Scholar] [CrossRef]
  17. Collins, P.G.; Hersam, M.; Arnold, M.; Martel, R.; Avouris, P. Current Saturation and Electrical Breakdown in Multiwalled Carbon Nanotubes. Phys. Rev. Lett. 2001, 86, 3128–3131. [Google Scholar] [CrossRef] [PubMed]
  18. Collins, P.G.; Arnold, M.S.; Avouris, P. Engineering Carbon Nanotubes and Nanotube Circuits Using Electrical Breakdown. Science 2001, 292, 706–709. [Google Scholar] [CrossRef]
  19. Liao, A.; Alizadegan, R.; Ong, Z.-Y.; Dutta, S.; Xiong, F.; Hsia, K.J.; Pop, E. Thermal dissipation and variability in electrical breakdown of carbon nanotube devices. Phys. Rev. B 2010, 82, 205406. [Google Scholar] [CrossRef] [Green Version]
  20. Jaehyun, C.; Kyong-Hoon, L.; Junghoon, L.; Diego, T.; George, C.S. Multi-walled carbon nanotubes experiencing electrical breakdown as gas sensors. Nanotechnology 2004, 15, 1596. [Google Scholar]
  21. Buh, G.H.; Hwang, J.H.; Jeon, E.K.; So, H.M.; Lee, J.O.; Kong, K.J.; Chang, H. On-Chip Electrical Breakdown of Metallic Nanotubes for Mass Fabrication of Carbon-Nanotube-Based Electronic Devices. IEEE Trans. Nanotechnol. 2008, 7, 624–627. [Google Scholar]
  22. Liu, J.; Wang, C.; Tu, X.; Liu, B.; Chen, L.; Zheng, M.; Zhou, C. Chirality-controlled synthesis of single-wall carbon nanotubes using vapour-phase epitaxy. Nat. Commun. 2012, 3, 1199. [Google Scholar] [CrossRef] [Green Version]
  23. Sanchez-Valencia, J.R.; Dienel, T.; Gröning, O.; Shorubalko, I.; Mueller, A.; Jansen, M.; Amsharov, K.; Ruffieux, P.; Fasel, R. Controlled synthesis of single-chirality carbon nanotubes. Nature 2014, 512, 61–64. [Google Scholar] [CrossRef] [Green Version]
  24. Liu, B.; Wu, F.; Gui, H.; Zheng, M.; Zhou, C. Chirality-Controlled Synthesis and Applications of Single-Wall Carbon Nanotubes. ACS Nano 2017, 11, 31–53. [Google Scholar] [CrossRef]
  25. Tomada, J.; Dienel, T.; Hampel, F.; Fasel, R.; Amsharov, K. Combinatorial design of molecular seeds for chirality-controlled synthesis of single-walled carbon nanotubes. Nat. Commun. 2019, 10, 3278. [Google Scholar] [CrossRef] [Green Version]
  26. Liu, B.; Wang, C.; Liu, J.; Che, Y.; Zhou, C. Aligned carbon nanotubes: From controlled synthesis to electronic applications. Nanoscale 2013, 5, 9483–9502. [Google Scholar] [CrossRef]
  27. Sun, Z.; Ikemoto, K.; Fukunaga, T.M.; Koretsune, T. Finite phenine nanotubes with periodic vacancy defects. Science 2019, 363, 151–155. [Google Scholar] [CrossRef]
  28. Treacy, M.M.J.; Ebbesen, T.W.; Gibson, J.M. Exceptionally high Young’s modulus observed for individual carbon nanotubes. Nature 1996, 381, 678–680. [Google Scholar] [CrossRef]
  29. Yu, M.-F.; Files, B.S.; Arepalli, S.; Ruoff, R.S. Tensile Loading of Ropes of Single Wall Carbon Nanotubes and their Mechanical Properties. Phys. Rev. Lett. 2000, 84, 5552–5555. [Google Scholar] [CrossRef] [Green Version]
  30. Tang, C.; Guo, W.; Chen, C. Mechanism for Superelongation of Carbon Nanotubes at High Temperatures. Phys. Rev. Lett. 2008, 100, 175501. [Google Scholar] [CrossRef]
  31. Hao, X.; Qiang, H.; Xiaohu, Y. Buckling of defective single-walled and double-walled carbon nanotubes under axial compression by molecular dynamics simulation. Compos. Sci. Technol. 2008, 68, 1809–1814. [Google Scholar] [CrossRef]
  32. Xin, H.; Han, Q.; Yao, X.-H. Buckling and axially compressive properties of perfect and defective single-walled carbon nanotubes. Carbon 2007, 45, 2486–2495. [Google Scholar] [CrossRef]
  33. Xin, H.; Han, Q. The Strain Rate Effect of Perfect and Defective Single-Walled Carbon Nanotubes Under Axial Compression. J. Comput. Theor. Nanosci. 2012, 9, 371–378. [Google Scholar] [CrossRef]
  34. Faria, B.; Silvestre, N. Mechanical properties of phenine nanotubes. Extreme Mech. Lett. 2022, 56, 101893. [Google Scholar] [CrossRef]
  35. Yu, H.T.; Yang, M.; Zhu, W.; Chang, T.; Jiang, J.W. Diameter-dependent polygonal cross section for holey phenine nanotubes. Nanotechnology 2019, 31, 085702. [Google Scholar] [CrossRef]
  36. Plimpton, S. Fast parallel algorithms for short-range molecular dynamics. J. Comput. Phys. 1995, 117, 1–19. [Google Scholar] [CrossRef] [Green Version]
  37. Stukowski, A. Visualization and analysis of atomistic simulation data with OVITO–the Open Visualization Tool. Model. Simul. Mat. Sci. Eng. 2009, 18, 015012. [Google Scholar] [CrossRef]
  38. Stuart, S.J.; Tutein, A.B.; Harrison, J.A. A reactive potential for hydrocarbons with intermolecular interactions. J. Chem. Phys. 2000, 112, 6472–6486. [Google Scholar] [CrossRef] [Green Version]
  39. Zhan, H.; Zhang, G.; Tan, V.B.C.; Cheng, Y.; Bell, J.M.; Zhang, Y.-W.; Gu, Y. From brittle to ductile: A structure dependent ductility of diamond nanothread. Nanoscale 2016, 8, 11177–11184. [Google Scholar] [CrossRef] [Green Version]
  40. Zhao, H.; Min, K.; Aluru, N.R. Size and chirality dependent elastic properties of graphene nanoribbons under uniaxial tension. Nano Lett. 2009, 9, 3012–3015. [Google Scholar] [CrossRef]
  41. Carpenter, C.; Maroudas, D.; Ramasubramaniam, A. Mechanical properties of irradiated single-layer graphene. Appl. Phys. Lett. 2013, 103, 013102. [Google Scholar] [CrossRef]
  42. Zhou, K.; Wang, L.; Wang, R.; Wang, C.; Tang, C. One Dimensional Twisted Van der Waals Structures Constructed by Self-Assembling Graphene Nanoribbons on Carbon Nanotubes. Materials 2022, 15, 8220. [Google Scholar] [CrossRef]
  43. Jensen, B.D.; Wise, K.E.; Odegard, G.M. The effect of time step, thermostat, and strain rate on ReaxFF simulations of mechanical failure in diamond, graphene, and carbon nanotube. J. Comput. Chem. 2015, 36, 1587–1596. [Google Scholar] [CrossRef]
  44. Liu, X.; Pan, D.; Hong, Y.; Guo, W. Bending Poisson Effect in Two-Dimensional Crystals. Phys. Rev. Lett. 2014, 112, 205502. [Google Scholar] [CrossRef] [Green Version]
  45. Feliciano, J.; Tang, C.; Zhang, Y.; Chen, C. Aspect ratio dependent buckling mode transition in single-walled carbon nanotubes under compression. J. Appl. Phys. 2011, 109, 084323. [Google Scholar] [CrossRef] [Green Version]
  46. Zhan, H.; Zhang, G.; Bell, J.M.; Tan, V.B.C.; Gu, Y. High density mechanical energy storage with carbon nanothread bundle. Nat. Commun. 2020, 11, 1905. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Rolling up of porous graphene (left) to phenine nanotubes (right). On the left, a six-membered ring of carbon in the unit cell of a graphene sheet is replaced by 6 hydrogen atoms, forming a periodically defected structure; the size of the unit cell is 12.96 Å × 7.48 Å. Once seamlessly rolled up along armchair or zigzag edges, armchair PNTs or zigzag PNTs are formed, and the diameter of the tubular structure can be simply estimated.
Figure 1. Rolling up of porous graphene (left) to phenine nanotubes (right). On the left, a six-membered ring of carbon in the unit cell of a graphene sheet is replaced by 6 hydrogen atoms, forming a periodically defected structure; the size of the unit cell is 12.96 Å × 7.48 Å. Once seamlessly rolled up along armchair or zigzag edges, armchair PNTs or zigzag PNTs are formed, and the diameter of the tubular structure can be simply estimated.
Materials 16 04706 g001
Figure 2. Evolution of cross-sectional geometry of armchair PNTs with diameter (ranging from 8.25 Å to 70.13 Å). For each case, the initially constructed atomic model, the structure after energy minimization, and those after structural relaxation at 1 K are provided.
Figure 2. Evolution of cross-sectional geometry of armchair PNTs with diameter (ranging from 8.25 Å to 70.13 Å). For each case, the initially constructed atomic model, the structure after energy minimization, and those after structural relaxation at 1 K are provided.
Materials 16 04706 g002
Figure 3. Twisting of zigzag carbon chains in armchair PNTs (a). The potential energy of the system decreases as a result of twisting (d). (b,c,e,f) present selected twisting angle evolution as a function of time for PNTs with different diameters.
Figure 3. Twisting of zigzag carbon chains in armchair PNTs (a). The potential energy of the system decreases as a result of twisting (d). (b,c,e,f) present selected twisting angle evolution as a function of time for PNTs with different diameters.
Materials 16 04706 g003
Figure 4. Evolution of cross-sectional geometry of zigzag PNTs with varying diameter (ranging from 8.25 Å to 157.19 Å). For each case, the initially constructed atomic model, the structure after energy minimization, and those after structural relaxation at 1 K are provided. The structures of a finite sized zigzag CNT are also provided on the lower right panel for comparison.
Figure 4. Evolution of cross-sectional geometry of zigzag PNTs with varying diameter (ranging from 8.25 Å to 157.19 Å). For each case, the initially constructed atomic model, the structure after energy minimization, and those after structural relaxation at 1 K are provided. The structures of a finite sized zigzag CNT are also provided on the lower right panel for comparison.
Materials 16 04706 g004
Figure 5. Mechanical properties of PNTs under tensile loadings. (a,b) show stress–strain relationship for armchair and zigzag PNTs with aspect ratio of 7/1, the ultimate stress and ultimate strain are plotted in (c,d) as a function of diameter.
Figure 5. Mechanical properties of PNTs under tensile loadings. (a,b) show stress–strain relationship for armchair and zigzag PNTs with aspect ratio of 7/1, the ultimate stress and ultimate strain are plotted in (c,d) as a function of diameter.
Materials 16 04706 g005
Figure 6. Evolution of twisting angle for a small diametered PNT under tensile loadings. Three chains are selected for the calculation; the locations are indicated in the inset.
Figure 6. Evolution of twisting angle for a small diametered PNT under tensile loadings. Three chains are selected for the calculation; the locations are indicated in the inset.
Materials 16 04706 g006
Figure 7. Failure process of PNTs under tensile loadings. The stress distribution for each PNT is provided and the stress level can be referred to the color bars. (a) armchair PNTs; (b) zigzag PNTs.
Figure 7. Failure process of PNTs under tensile loadings. The stress distribution for each PNT is provided and the stress level can be referred to the color bars. (a) armchair PNTs; (b) zigzag PNTs.
Materials 16 04706 g007
Figure 8. PNTs under compressive loadings. (a,b) show the stress–strain relationship for the armchair and zigzag PNTs. The critical stress and critical strain for buckling are shown in (c,d).
Figure 8. PNTs under compressive loadings. (a,b) show the stress–strain relationship for the armchair and zigzag PNTs. The critical stress and critical strain for buckling are shown in (c,d).
Materials 16 04706 g008
Figure 9. Buckling process of PNTs under compression. The stress distribution for each PNT is provided and the stress level can be seen from the color bars. (a) buckling process of an armchair PNT with D = 16.5 Å, (b) buckling process of a zigzag PNT with D = 11.91 Å.
Figure 9. Buckling process of PNTs under compression. The stress distribution for each PNT is provided and the stress level can be seen from the color bars. (a) buckling process of an armchair PNT with D = 16.5 Å, (b) buckling process of a zigzag PNT with D = 11.91 Å.
Materials 16 04706 g009
Figure 10. Energy storage density of PNTs under tensile loadings.
Figure 10. Energy storage density of PNTs under tensile loadings.
Materials 16 04706 g010
Table 1. Summarized mechanical properties of armchair PNTs.
Table 1. Summarized mechanical properties of armchair PNTs.
Diameter16.50 Å24.75 Å33.00 Å41.25 Å61.88 ÅUnit
Ultimate stress under tension36.3634.8034.1734.4334.72GPa
Ultimate strain under tension0.2090.1860.1790.1860.191-
Young’s modulus239219210199197GPa
Energy storage density34,48828,24824,98426,78427,792KJ/Kg
Critical stress under compression9.302.381.691.781.48GPa
Critical strain under compression0.040.030.0360.0390.023-
Table 2. Summarized mechanical properties of zigzag PNTs.
Table 2. Summarized mechanical properties of zigzag PNTs.
Diameter11.91 Å23.82 Å33.34 Å42.87 Å61.92 ÅUnit
Ultimate stress under tension30.7829.3429.3029.8729.79GPa
Ultimate strain under tension0.2050.160.1450.1550.155-
Young’s modulus114160173166159GPa
Energy storage density22,29617,56817,54418,13217,196KJ/Kg
Critical stress under compression4.742.522.132.070.95GPa
Critical strain under compression0.270.050.030.040.03-
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

Liu, Y.; Wang, R.; Wang, L.; Xia, J.; Wang, C.; Tang, C. Size- and Chirality-Dependent Structural and Mechanical Properties of Single-Walled Phenine Nanotubes. Materials 2023, 16, 4706. https://doi.org/10.3390/ma16134706

AMA Style

Liu Y, Wang R, Wang L, Xia J, Wang C, Tang C. Size- and Chirality-Dependent Structural and Mechanical Properties of Single-Walled Phenine Nanotubes. Materials. 2023; 16(13):4706. https://doi.org/10.3390/ma16134706

Chicago/Turabian Style

Liu, Yanjun, Ruijie Wang, Liya Wang, Jun Xia, Chengyuan Wang, and Chun Tang. 2023. "Size- and Chirality-Dependent Structural and Mechanical Properties of Single-Walled Phenine Nanotubes" Materials 16, no. 13: 4706. https://doi.org/10.3390/ma16134706

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