Mechanical properties of carbon nanotubes

Last updated

The mechanical properties of carbon nanotubes reveal them as one of the strongest materials in nature. Carbon nanotubes (CNTs) are long hollow cylinders of graphene. Although graphene sheets have 2D symmetry, carbon nanotubes by geometry have different properties in axial and radial directions. It has been shown that CNTs are very strong in the axial direction. [1] Young's modulus on the order of 270 - 950 GPa and tensile strength of 11 - 63 GPa were obtained. [1]

Contents

Strength

Carbon nanotubes are the strongest and stiffest materials yet discovered in terms of tensile strength and elastic modulus respectively. This strength results from the covalent sp2 bonds formed between the individual carbon atoms. In 2000, a multi-walled carbon nanotube was tested to have a tensile strength of 63 gigapascals (9,100,000 psi). (For illustration, this translates into the ability to endure tension of a weight equivalent to 6,422 kilograms-force (62,980 N; 14,160 lbf) on a cable with cross-section of 1 square millimetre (0.0016 sq in).) Further studies, such as one conducted in 2008, revealed that individual CNT shells have strengths of up to ≈100 gigapascals (15,000,000 psi), which is in agreement with quantum/atomistic models. Since carbon nanotubes have a low density for a solid of 1.3 to 1.4 g/cm3, its specific strength of up to 48,000 kN·m·kg−1 is the best of known materials, compared to high-carbon steel's 154 kN·m·kg−1.

Under excessive tensile strain, the tubes will undergo plastic deformation, which means the deformation is permanent. This deformation begins at strains of approximately 5% and can increase the maximum strain the tubes undergo before fracture by releasing strain energy.[ citation needed ]

Although the strength of individual CNT shells is extremely high, weak shear interactions between adjacent shells and tubes lead to significant reduction in the effective strength of multi-walled carbon nanotubes and carbon nanotube bundles down to only a few GPa. This limitation has been recently addressed by applying high-energy electron irradiation, which crosslinks inner shells and tubes, and effectively increases the strength of these materials to ≈60 GPa for multi-walled carbon nanotubes and ≈17 GPa for double-walled carbon nanotube bundles.

CNTs are not nearly as strong under compression. Because of their hollow structure and high aspect ratio, they tend to undergo buckling when placed under compressive, torsional, or bending stress.

Comparison of mechanical properties
Material Young's modulus (TPa) Tensile strength (GPa)Elongation at break (%)
Single-Walled Carbon Nanotubes (SWNT)E≈1 (from 1 to 5)13–5316
Armchair SWNTT0.94126.223.1
Zigzag SWNTT0.9494.515.6–17.5
Chiral SWNT0.92
MWNTE0.2–0.8–0.9511–63–150
Stainless steel E0.186–0.2140.38–1.5515–50
Kevlar–29&149E0.06–0.183.6–3.8≈2

EExperimental observation; TTheoretical prediction

Radial elasticity

On the other hand, there was evidence that in the radial direction they are rather soft. The first transmission electron microscope observation of radial elasticity suggested that even the van der Waals forces can deform two adjacent nanotubes. [2] Later, nanoindentations with atomic force microscope were performed by several groups to quantitatively measure radial elasticity of multiwalled carbon nanotubes [3] [4] and tapping/contact mode atomic force microscopy was also performed on single-walled carbon nanotubes. [5] Young's modulus of on the order of several GPa showed that CNTs are in fact very soft in the radial direction. A complete phase diagram giving the transition to the radially collapsed geometry as function of diameter, pressure and number of tube-walls has been produced from semiempirical grounds. [6]

Radial direction elasticity of CNTs is important especially for carbon nanotube composites where the embedded tubes are subjected to large deformation in the transverse direction under the applied load on the composite structure.

One of the main problems in characterizing the radial elasticity of CNTs is the knowledge about the internal radius of the CNT; carbon nanotubes with identical outer diameter may have different internal diameter (or the number of walls). In 2008, a method using an atomic force microscope was introduced to determine the exact number of layers and hence the internal diameter of the CNT. In this way, mechanical characterization is more accurate. [7]

Hardness

Standards single-walled carbon nanotubes can withstand a pressure up to 25 GPa without [plastic/permanent] deformation. They then undergo a transformation to superhard phase nanotubes. Maximum pressures measured using current experimental techniques are around 55 GPa. However, these new superhard phase nanotubes collapse at an even higher, albeit unknown, pressure.[ citation needed ]

The bulk modulus of superhard phase nanotubes is 462 to 546 GPa, even higher than that of diamond (420 GPa for single diamond crystal).

Wettability

The surface wettability of CNT is of importance for its applications in various settings. Although the intrinsic contact angle of graphite is around 90°, the contact angles of most as-synthesized CNT arrays are over 160°, exhibiting a superhydrophobic property. By applying a voltage as low as 1.3V, the extreme water repellant surface can be switched to a superhydrophilic one.[ citation needed ]


Kinetic properties

Multi-walled nanotubes are multiple concentric nanotubes precisely nested within one another. These exhibit a striking telescoping property whereby an inner nanotube core may slide, almost without friction, within its outer nanotube shell, thus creating an atomically perfect linear or rotational bearing. This is one of the first true examples of molecular nanotechnology, the precise positioning of atoms to create useful machines. Already, this property has been utilized to create the world's smallest rotational motor. Future applications such as a gigahertz mechanical oscillator are also envisioned.

Defects

As with any material, the existence of a crystallographic defect affects the material properties. Defects can occur in the form of atomic vacancies. High levels of such defects can lower the tensile strength by up to 85%. An important example is the Stone Wales defect, otherwise known as the 5-7-7-5 defect because it creates a pentagon and heptagon pair by rearrangement of the bonds. Because of the very small structure of CNTs, the tensile strength of the tube is dependent on its weakest segment in a similar manner to a chain, where the strength of the weakest link becomes the maximum strength of the chain.

Plastic deformation

A typical 3D material undergoes plastic deformation, meaning that the deformation is permanent, by the movement of 1D dislocations through the material. During this process, these dislocations can interact with each other and multiply. Because CNTs are themselves 1D materials, the well-known generation and multiplication mechanisms (such as a Frank-Read source) for 1D dislocations do not apply. [8]

Instead, CNTs undergo plastic deformation through the formation and movement of defects, primarily topological defects such as the Stone Wales defect or 5-7-7-5 defect. The 5-7-7-5 defect can also be thought of as a pair of 5-7 defects, in which each defect is adjacent to one 5-membered and two 7-membered rings. [9] This defect structure is metastable, so it takes an energy of several eV to nucleate, or form. In addition, the defect moves by the separate migration of the 5-7 defect pairs. This motion is also associated with an energy barrier. The exact energy depends on the configuration and chirality of the particular CNT. The activation energy for the formation of these defects in a CNT of diameter and chiral angle can be estimated as eV, where is the external strain. [10] [11] This activation energy barrier partially explains the low ductility of CNTs (~6-15%) at room temperature. However, it can be overcome at high temperatures or with the application of suitable strains. [12] For example, the defect is nucleated at positions experiencing high tensile stress in armchair-type CNTs, and at positions experiencing high compressive stress in zigag-type CNTs. [13]

Applied stresses can overcome the energy barrier needed to move 5-7 defect pairs. Another way of understanding this is that when strained, a CNT releases strain by forming these defects spontaneously. For example, in (5,5) tubes, a critical tensile strain of ~5% results in defect generation. The defect structure reduces strain because the heptagon geometry is able to stretch more than the original hexagonal rings, while the C-C bond remains about the same length. [14] Bending the tubes beyond a critical curvature has the same effect. This behavior can be approximated by a simple, semi-quantitative analysis. Applying a stress over a tube of length and diameter does work approximately equal to on the tube, where is the Burgers vectors for the defect, is the bending curvature, and relates the Young's modulus of the CNT to that of graphene. The energy increase resulting from the defect creation and the separation of the 5-7 pairs is approximately given by . Here, is the dislocation core energy and gives the interaction energy between the defect pairs. Defect motion occurs when the work done by an applied stress overcomes it, such that the required bending curvature is inversely proportional to the diameter of the CNT: . [13] Similarly, thermal vibrations can provide the energy required for defect nucleation and motion. In fact, a combination of stress and high temperature is required to induce observable plastic deformation in CNTs. This has been achieved in the literature via the application of a current, which causes resistive heating in the material. [15] For CNTs subjected to temperatures above 1500K, elongations up to 280% have been reported. This kind of behavior is called superplasticity. [16] At these high temperatures, kinks may form and move by climb as well as glide. Climb of kinks is evidenced by the fact that they do not always move along the close-packed planes in CNTs, but rather along the length of a tube. When kinks do glide along close-packed planes in CNTs, they follow a helical path. It is proposed that elevated temperatures allow for the diffusion of vacancies, so that defects climb through a process similar to that observed in 3D crystalline materials. [17]

Related Research Articles

<span class="mw-page-title-main">Carbon nanotube</span> Allotropes of carbon with a cylindrical nanostructure

A carbon nanotube (CNT) is a tube made of carbon with a diameter in the nanometre range (nanoscale). They are one of the allotropes of carbon.

<span class="mw-page-title-main">Young's modulus</span> Mechanical property that measures stiffness of a solid material

Young's modulus is a mechanical property of solid materials that measures the tensile or compressive stiffness when the force is applied lengthwise. It is the modulus of elasticity for tension or axial compression. Young's modulus is defined as the ratio of the stress applied to the object and the resulting axial strain in the linear elastic region of the material.

<span class="mw-page-title-main">Ultimate tensile strength</span> Maximum stress withstood by stretched material before breaking

Ultimate tensile strength is the maximum stress that a material can withstand while being stretched or pulled before breaking. In brittle materials, the ultimate tensile strength is close to the yield point, whereas in ductile materials, the ultimate tensile strength can be higher.

In engineering, deformation refers to the change in size or shape of an object. Displacements are the absolute change in position of a point on the object. Deflection is the relative change in external displacements on an object. Strain is the relative internal change in shape of an infinitesimal cube of material and can be expressed as a non-dimensional change in length or angle of distortion of the cube. Strains are related to the forces acting on the cube, which are known as stress, by a stress-strain curve. The relationship between stress and strain is generally linear and reversible up until the yield point and the deformation is elastic. The linear relationship for a material is known as Young's modulus. Above the yield point, some degree of permanent distortion remains after unloading and is termed plastic deformation. The determination of the stress and strain throughout a solid object is given by the field of strength of materials and for a structure by structural analysis.

<span class="mw-page-title-main">Stress–strain curve</span> Curve representing a materials response to applied forces

In engineering and materials science, a stress–strain curve for a material gives the relationship between stress and strain. It is obtained by gradually applying load to a test coupon and measuring the deformation, from which the stress and strain can be determined. These curves reveal many of the properties of a material, such as the Young's modulus, the yield strength and the ultimate tensile strength.

In physics and materials science, elasticity is the ability of a body to resist a distorting influence and to return to its original size and shape when that influence or force is removed. Solid objects will deform when adequate loads are applied to them; if the material is elastic, the object will return to its initial shape and size after removal. This is in contrast to plasticity, in which the object fails to do so and instead remains in its deformed state.

The field of strength of materials typically refers to various methods of calculating the stresses and strains in structural members, such as beams, columns, and shafts. The methods employed to predict the response of a structure under loading and its susceptibility to various failure modes takes into account the properties of the materials such as its yield strength, ultimate strength, Young's modulus, and Poisson's ratio. In addition, the mechanical element's macroscopic properties such as its length, width, thickness, boundary constraints and abrupt changes in geometry such as holes are considered.

<span class="mw-page-title-main">Toughness</span> Material ability to absorb energy and plastically deform without fracturing

In materials science and metallurgy, toughness is the ability of a material to absorb energy and plastically deform without fracturing. Toughness is the strength with which the material opposes rupture. One definition of material toughness is the amount of energy per unit volume that a material can absorb before rupturing. This measure of toughness is different from that used for fracture toughness, which describes the capacity of materials to resist fracture. Toughness requires a balance of strength and ductility.

<span class="mw-page-title-main">Work hardening</span> Strengthening a material through plastic deformation

In materials science, work hardening, also known as strain hardening, is the strengthening of a metal or polymer by plastic deformation. Work hardening may be desirable, undesirable, or inconsequential, depending on the context.

<span class="mw-page-title-main">Yield (engineering)</span> Phenomenon of deformation due to structural stress

In materials science and engineering, the yield point is the point on a stress-strain curve that indicates the limit of elastic behavior and the beginning of plastic behavior. Below the yield point, a material will deform elastically and will return to its original shape when the applied stress is removed. Once the yield point is passed, some fraction of the deformation will be permanent and non-reversible and is known as plastic deformation.

In materials science, hardness is a measure of the resistance to localized plastic deformation, such as an indentation or a scratch (linear), induced mechanically either by pressing or abrasion. In general, different materials differ in their hardness; for example hard metals such as titanium and beryllium are harder than soft metals such as sodium and metallic tin, or wood and common plastics. Macroscopic hardness is generally characterized by strong intermolecular bonds, but the behavior of solid materials under force is complex; therefore, hardness can be measured in different ways, such as scratch hardness, indentation hardness, and rebound hardness. Hardness is dependent on ductility, elastic stiffness, plasticity, strain, strength, toughness, viscoelasticity, and viscosity. Common examples of hard matter are ceramics, concrete, certain metals, and superhard materials, which can be contrasted with soft matter.

The specific strength is a material's strength divided by its density. It is also known as the strength-to-weight ratio or strength/weight ratio or strength-to-mass ratio. In fiber or textile applications, tenacity is the usual measure of specific strength. The SI unit for specific strength is Pa⋅m3/kg, or N⋅m/kg, which is dimensionally equivalent to m2/s2, though the latter form is rarely used. Specific strength has the same units as specific energy, and is related to the maximum specific energy of rotation that an object can have without flying apart due to centrifugal force.

<span class="mw-page-title-main">Potential applications of carbon nanotubes</span>

Carbon nanotubes (CNTs) are cylinders of one or more layers of graphene (lattice). Diameters of single-walled carbon nanotubes (SWNTs) and multi-walled carbon nanotubes (MWNTs) are typically 0.8 to 2 nm and 5 to 20 nm, respectively, although MWNT diameters can exceed 100 nm. CNT lengths range from less than 100 nm to 0.5 m.

Methods have been devised to modify the yield strength, ductility, and toughness of both crystalline and amorphous materials. These strengthening mechanisms give engineers the ability to tailor the mechanical properties of materials to suit a variety of different applications. For example, the favorable properties of steel result from interstitial incorporation of carbon into the iron lattice. Brass, a binary alloy of copper and zinc, has superior mechanical properties compared to its constituent metals due to solution strengthening. Work hardening has also been used for centuries by blacksmiths to introduce dislocations into materials, increasing their yield strengths.

Colossal carbon tubes (CCTs) are a tubular form of carbon. In contrast to the carbon nanotubes (CNTs), colossal carbon tubes have much larger diameters ranging between 40 and 100 μm. Their walls have a corrugated structure with abundant pores, as in corrugated fiberboard, where the solid membranes have a graphite-like layered structure.

<span class="mw-page-title-main">Optical properties of carbon nanotubes</span> Optical properties of the material

The optical properties of carbon nanotubes are highly relevant for materials science. The way those materials interact with electromagnetic radiation is unique in many respects, as evidenced by their peculiar absorption, photoluminescence (fluorescence), and Raman spectra.

Carbon nanotube springs are springs made of carbon nanotubes (CNTs). They are an alternate form of high-density, lightweight, reversible energy storage based on the elastic deformations of CNTs. Many previous studies on the mechanical properties of CNTs have revealed that they possess high stiffness, strength and flexibility. The Young's modulus of CNTs is 1 TPa and they have the ability to sustain reversible tensile strains of 6% and the mechanical springs based on these structures are likely to surpass the current energy storage capabilities of existing steel springs and provide a viable alternative to electrochemical batteries. The obtainable energy density is predicted to be highest under tensile loading, with an energy density in the springs themselves about 2500 times greater than the energy density that can be reached in steel springs, and 10 times greater than the energy density of lithium-ion batteries.

Tube-based nanostructures are nanolattices made of connected tubes and exhibit nanoscale organization above the molecular level.

<span class="mw-page-title-main">Gallium nitride nanotube</span>

Gallium nitride nanotubes (GaNNTs) are nanotubes of gallium nitride. They can be grown by chemical vapour deposition.

References

  1. 1 2 M.-F. Yu; et al. (2000). "Strength and Breaking Mechanism of Multiwalled Carbon Nanotubes Under Tensile Load". Science. 287 (5453): 637–40. Bibcode:2000Sci...287..637Y. doi:10.1126/science.287.5453.637. PMID   10649994.
  2. R. S. Ruoff; et al. (1993). "Radial deformation of carbon nanotubes by van der Waals forces". Nature. 364 (6437): 514. Bibcode:1993Natur.364..514R. doi:10.1038/364514a0. S2CID   4264362.
  3. I. Palaci; et al. (2005). "Radial Elasticity of Multiwalled Carbon Nanotubes". Physical Review Letters. 94 (17): 175502. arXiv: 1201.5501 . Bibcode:2005PhRvL..94q5502P. doi:10.1103/PhysRevLett.94.175502. PMID   15904310. S2CID   8090975.
  4. M.-F. Yu; et al. (2000). "Investigation of the Radial Deformability of Individual Carbon Nanotubes under Controlled Indentation Force". Physical Review Letters. 85 (7): 1456–9. Bibcode:2000PhRvL..85.1456Y. doi:10.1103/PhysRevLett.85.1456. PMID   10970528.
  5. Y.H.Yang; et al. (2011). "Radial elasticity of single-walled carbon nanotube measured by atomic force microscopy". Applied Physics Letters. 98 (4): 041901. doi:10.1063/1.3546170.
  6. Y.Magnin; et al. (2021). "Collapse phase diagram of carbon nanotubes with arbitrary number of walls. Collapse modes and macroscopic analog". Carbon. 178: 552.
  7. M. Minary-Jolandan, M.-F. Yu (2008). "Reversible radial deformation up to the complete flattening of carbon nanotubes in nanoindentation". Journal of Applied Physics. 103 (7): 073516–073516–5. Bibcode:2008JAP...103g3516M. doi:10.1063/1.2903438.
  8. Shima, Hiroyuki; Sato, Motohiro, eds. (2013). "Chapter 6: Topological Defects". Elastic and Plastic Deformation of Carbon Nanotubes. CRC Press. pp. 81–110. ISBN   978-9814364157.
  9. P. Zhang; et al. (1998). "Plastic Deformations of Carbon Nanotubes". Physical Review Letters. 81 (24): 5346–5349. doi:10.1103/PhysRevLett.81.5346.
  10. T. Dumitrica; et al. (2004). "SStrain-rate and temperature dependent plastic yield in carbon nanotubes from ab initio calculations". Applied Physics Letters. 84 (15): 2775. doi:10.1063/1.1695630.
  11. L.G. Zhou; et al. (2003). "Formation energy of Stone-Wales defects in carbon nanotubes". Applied Physics Letters. 83 (6): 1222–1224. doi:10.1063/1.1599961. hdl: 10397/4230 .
  12. M. Mori (2011). "Elastic and Plastic Deformation of Carbon Nanotubes". Procedia Engineering. 14: 2366–2372. doi: 10.1016/j.proeng.2011.07.298 . hdl: 2115/54292 .
  13. 1 2 H. Mori; et al. (2006). "Energetics of plastic bending of carbon nanotubes". Physical Review B. 74 (16): 165418. doi:10.1103/PhysRevB.74.165418.
  14. M. B. Nardelli; et al. (1998). "Mechanism of strain release in carbon nanotubes". Physical Review B. 57 (8): R4277. doi:10.1103/PhysRevB.57.R4277.
  15. Y. Nakayama; et al. (2005). "Current-Induced Plastic Deformation of Double-Walled Carbon Nanotubes". Japanese Journal of Applied Physics. 44: L720. doi:10.1143/JJAP.44.L720.
  16. J.Y. Huang; et al. (2006). "Superplastic carbon nanotubes". Nature. 439 (7074): 281. doi: 10.1038/439281a . PMID   16421560. S2CID   4407587.
  17. J.Y. Huang; et al. (2006). "Kink Formation and Motion in Carbon Nanotubes at High Temperatures". Physical Review Letters. 97 (7): 075501. doi:10.1103/PhysRevLett.97.075501. PMID   17026242.