Next Article in Journal
Macrosegregation Evolution in Eutectic Al-Si Alloy under the Influence of a Rotational Magnetic Field
Next Article in Special Issue
Dynamic Response Analysis of Projectile Target Penetration Based on an FE-SPH Adaptive Coupling Method
Previous Article in Journal
Research Progress of Magnetic Field Regulated Mechanical Property of Solid Metal Materials
Previous Article in Special Issue
Effect of Heat Treatment and Test Temperature on the Strength Properties of Cast Heat-Resistant Nickel Base Inconel 718 Superalloy under Shock-Wave Loading
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Atomistic Investigation of Titanium Carbide Ti8C5 under Impact Loading

1
College of Mechanical & Electrical Engineering, HoHai University, Nanjing 210098, China
2
School of Mechanical, Medical and Process Engineering, Queensland University of Technology (QUT), Brisbane, QLD 4001, Australia
3
College of Civil Engineering and Architecture, Zhejiang University, Hangzhou 310058, China
4
State Key Laboratory of Explosion Science and Technology, Beijing Institute of Technology, Beijing 100081, China
5
State Key Laboratory of Mechanics and Control of Mechanical Structures, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China
*
Author to whom correspondence should be addressed.
Metals 2022, 12(11), 1989; https://doi.org/10.3390/met12111989
Submission received: 19 October 2022 / Revised: 10 November 2022 / Accepted: 17 November 2022 / Published: 20 November 2022
(This article belongs to the Special Issue Deformation and Fracture of Condensed Materials in Extreme Conditions)

Abstract

:
Titanium carbides attract attention from both academic and industry fields because of their intriguing mechanical properties and proven potential as appealing candidates in the variety of fields such as nanomechanics, nanoelectronics, energy storage and oil/water separation devices. A recent study revealed that the presence of Ti8C5 not only improves the impact strength of composites as coatings, but also possesses significant strengthening performance as an interlayer material in composites by forming strong bonding between different matrices, which sheds light on the design of impact protection composite materials. To further investigate the impact resistance and strengthening mechanism of Ti8C5, a pilot Molecular Dynamics (MD) study utilizing comb3 potential is carried out on a Ti8C5 nanosheet by subjecting it to hypervelocity impacts. The deformation behaviour of Ti8C5 and the related impact resist mechanisms are assessed in this research. At a low impact velocity ~0.5 km/s, the main resonance frequency of Ti8C5 is 11.9 GHz and its low Q factor (111.9) indicates a decent energy damping capability, which would eliminate the received energy in an interfacial reflection process and weaken the shock waves for Ti8C5 strengthened composites. As the impact velocity increases above the threshold of 1.8 km/s, Ti8C5 demonstrates brittle behaviour, which is signified by its insignificant out-of-plane deformation prior to crack initiation. When tracking atomic Von Mises stress distribution, the elastic wave propagation velocity of Ti8C5 is calculated to be 5.34 and 5.90 km/s for X and Y directions, respectively. These figures are inferior compared with graphene and copper, which indicate slower energy delocalization rates and thus less energy dissipation via deformation is expected prior to bond break. However, because of its relatively small mass density comparing with copper, Ti8C5 presents superior specific penetration. This study provides a fundamental understanding of the deformation and penetration mechanisms of titanium carbide nanosheets under impact, which is crucial in order to facilitate emerging impact protection applications for titanium carbide-related composites.

1. Introduction

In recent decades, tremendous efforts on the exploration of titanium matrix composites have been conducted because of their outstanding mechanical properties [1,2] and tuneable potential [2,3,4,5,6]. Titanium carbide is a widely adopted strengthening material in composites and has received immense attention from civil automobile, aerospace as well as military industries, owing to its superior bending rigidity [7], Young’s modulus [8], strength to weight ratio, wear and corrosion resistance [9], Oil/water separation capability [10], etc. To this end, an in situ synthesis of titanium carbide is realized via multiple approaches such as supersonic flame spraying process [11], high temperature and high-pressure environment synthesis [12], low temperature salt bath carburizing [13] and state-of-art high energy beam setups, such as laser cladding [14]. Micro-meter titanium carbide can be achieved via the dissolution-precipitation mechanism utilizing a high-energy laser.
Varieties of titanium carbides have been successfully realized so far; however, constrained by the lateral size of sheets and complex oxygen-containing function groups observed in the state-of-art synthesis process, most studies on monolithic titanium carbides have been carried out via in silico approaches such as Molecular Dynamics (MD) and Density Function Theory (DFT)-based simulation [15]. Deploying an empirical potential energy function and embedded atom method, a MD study suggests the bending rigidity of 2D titanium carbides can be as high as ~49.55 eV, which is much more significant than graphene and MoS2 (2.3 and 9.61 eV, respectively) [7]. A DFT study on functionalized titanium carbides suggests that the oxygen group not only enhances ideal strength, but also induces a strong anisotropy for titanium carbides as the material subjected to tensile loading [16]. Recently, a follow-up MD study utilizing COMB3 potential confirmed the above findings; moreover, the Young’s modulus of the assessed titanium-based carbides ranged between 133 and 517 GPa, which shows its potential in nanomechanics, nanoelectronics and energy storage applications [8].
In addition to having outstanding mechanical properties, pristine titanium carbides are also known for their strengthening material/matrix in varieties of composites. Titanium carbide (Ti8C5) thin film on a Ti6Al4V alloy surface realized via low temperature carburizing salt bath possesses significantly improved friction and wear properties and its surface hardness reaches ~953.28HV, which is 1.57 times higher than Ti6Al4V [13]. Moreover, Ti8C5 nanolayers/nanoparticles in situ formed at an interface of RGO/CuTi composite function as ‘rivets’ show sizes of up to ~100 nm. The presence of these Ti8C5 ‘rivets’ raises the interfacial strength, increases load transfer efficiency between layers and clearly presents a 53.9% improvement in tensile strength compared with composites without Ti8C5 in their interface [12]. The strengthening behaviour is also confirmed by a study on CNTs/Cu-Ti composites with strong Ti8C5 interfacial bonding [17]. Moreover, the quasi-static loading experiments’ shock loading result states that Ti8C5 formed between diamond and titanium under a sintering temperature improves the impact strength of diamonds as coatings, which clearly suggests their impact protection potential [18].
Considering the mechanical strengthening behaviour of Ti8C5 in composites (such as graphene-copper composite) under quasi-static uniaxial tensile loading, its performance under various impact loading condition is rarely systematically discussed in the literature; however, it is critical to facilitate its application in bullet-proof fields as well as promote the interface design of advanced composites for impact resist applications [12,19,20]. Because of the experimental manipulation complex and small lateral dimensions of available Ti8C5 sheets [12], a pilot study on impact resistance of Ti8C5 under various impact velocities is carried out utilizing the MD method. In this study, resonance pattern, stress distribution capability as well as the specific penetration energy of Ti8C5 are accessed. The performance and phenomenon of a Ti8C5 nanosheet under impact loading are mainly compared with a thin copper layer and single-layer graphene in this pilot study.

2. Materials and Methods

The mechanical behaviour of Ti8C5 under hypervelocity impact was assessed through a series of large-scale MD simulations performed with the aid of the open-source package LAMMPS [21]. A spherical diamond projectile with a radius of 25 Å and a square 50 × 50 nm2 Ti8C5 nanosheet [22] were considered. The (010) plane of the Ti8C5 nanosheet faced the projectile. Initial velocities ranging from 5 to 70 Å/ps (i.e., 0.5–7 km/s) were assigned to the projectiles. The boundaries of the Ti8C5 nanosheets were fixed during the impact process (Figure 1a), whereas the projectile was located above the centre of the Ti8C5 sample with an initial distance of ~20 Å. Considering the high thermal fluctuations generated during the impact process, the simulation temperature was set to be 10 K. Simulations with higher environment temperatures up to 700 K were also performed for validation, whereby the obtained penetration energy for the same impact velocities was maintained at the same level (Supplementary Materials, Figure S1). However, noting the negative correlation between temperature and material strength [23,24], a further increase to the environmental temperature may lead to lower penetration energy. A time step of 0.1 fs was selected for the simulations.
The third generation Charge-Optimized Many-Body potential (COMB3) was employed to describe the C-Ti atomic interactions in the Ti8C5 nanosheet, as well as the Cu–Cu interaction in copper, as it was fully optimized in order to determine the binding energy of carbon- and metal-based systems [25,26,27,28,29]. The COMB3 potential is defined as follows,
E C O M B 3 = V E S + V S H O R T + V v d W + V C O R R
where V E S is an electrostatic term, which is the sum of atomic ionization energies and energy related to the Coulomb interaction. V S H O R T , V v d W and V C O R R represent the short range, van der Waals and correction energy term, respectively. The Tersoff potential was utilized to simulate the atomic interactions in diamonds, which has been shown to describe the binding energy for carbonaceous system well [30]. The interactions between the diamond projectile and nanosheets were described by Morse potential [31]. Prior to the impact process, a conjugate gradient algorithm was applied to conduct energy minimization, and then the Nose–Hoover thermostat [32] was employed to equilibrate the whole system at 10 K for 2000 fs under the canonical ensemble. The equations of motion were integrated with time using the velocity Verlet algorithm [33] and no periodic boundary conditions were adopted. To reproduce the energy conversion from kinetic energy to potential energy, no thermostat was applied during the impact process. During the simulation, the atomic stress was calculated based on the virial stress Π α β [34],
Π α β = 1 Ω i ϖ i π i α β ,   π i α β = 1 ϖ i ( m i v i α v i β + 1 2 j i F i j α r i j β )
Here, π i α β is the atomic stress associated with atom i. ϖ i is the effective volume of the ith atom and Ω is the volume of the whole system. m i and v i are the mass and velocity of the ith atom, respectively. F i j and r i j are the force and distance between atoms i and j, respectively, and the indices α and β denote the Cartesian components. The volume of the Ti8C5 nanosheet is estimated by assuming the Ti8C5 as a continuum media with a thickness of h (13.6 Å). Considering that adopting different atomic volumes will alter the magnitude of the stress, it will not however change the trends of the results as focused on in this paper. Considering the complex stress states during impact, we tracked the Von Mises stress σ V M in the nanosheet based on the atomic virial stress (Equation (2)), as calculated by
σ V M = ( ( σ x σ y ) 2 + ( σ y σ z ) 2 + ( σ y σ z ) 2 + 6 ( σ x y 2 + σ x z 2 + σ y z 2 ) ) / 2
Initially, we focused on the impact performance of the Ti8C5 nanosheet under the impact velocity of 3 km/s. In the ideal vacuum condition, the total energy change in the projectile equals the change in the Ti8C5 nanosheet. It is notable that during the impact process, although the projectile exhibited an ignorable deformation, a notable amount of potential energy change in the projectile ( Δ E b a l l , p e ) was observed during the impact with Ti8C5 (Figure 1b). After perforation, the total energy loss in the projectile ( Δ E b a l l , t o t which is the sum of Δ E b a l l , p e and Δ E b a l l , k e ) remains constant (Figure 1c); thus, Δ E b a l l , t o t is taken as the penetration energy ( E p ) [35,36]. It was noticed that the NVE ensemble was adopted in the impact process and total energy of the system was constant during the impact process, and the energy loss in the projectile was equal to the energy gain in the nanosheet Δ E N S , t o t (where Δ E N S , t o t equals the sum of nanosheet potential energy gain Δ P e N S and kinetic energy gain of nanosheet Δ K e N S ). As such, E p = Δ E b a l l , t o t = Δ E b a l l , p e + Δ E b a l l , k e = Δ E N S , t o t = Δ P e N S + Δ K e N S .

3. Results and Discussion

3.1. Deformation Characteristics

To acquire the deformation process, Figure 2a–c (top panel) illustrates the atomic configurations of Ti8C5 at different deformation stages, namely the crack initiation, crack expansion and fully developed cracks, and the results are compared with the copper nanosheet that has a similar number of atom layers under the same experiment setup (Figure 2d–f). As expected, the crack initiates at the geometry centre, where accumulation of stress is observed. The nanosheet experiences insignificant out-of-plane deformation ~11.49 Å prior to the propagation of cracks and stress (Figure 2a) compared with copper (Figure 2d) and carbonaceous nanosheets [36,37,38], which signifies brittle behaviour. During the crack propagation phase, the stress that accumulates around the impact region starts to re-distribute and the stress distribution in the deformed region of Ti8C5 exhibits a circular pattern (Figure 2b,c), whereas copper (lattice orientation is [100], [010], [001] for x, y and z direction respectively) demonstrates an X-shaped pattern in the highlighted region (Figure 2d–f). In addition to the stress accumulation, dislocation is also often observed in this highlighted region. This phenomenon is considered as a result of the lattice structure and lattice orientation. Unlike the armchair and zigzag kicking fracture mechanisms observed in carbonaceous nanosheets [36,37,38], the cracks in Ti8C5 and copper propagate along all directions and the same crack shape is roughly maintained during the whole process. Full perforation of Ti8C5 takes place at ~4.0 ps, which is significantly earlier than copper and the fully developed cracks for Ti8C5 are ~16.8% bigger than copper in terms of radius (Figure 2c).
Higher impact velocity induces more severe local deformation. As such, extensive elastic deformation of the nanosheet is not expected. When tracking the number of discrete debris in nanosheets after the crack propagation stage (Figure 3), it presents a positive correlation with impact velocity. For an ultra-high impact velocity ~6 km/s, the contact region melts immediately as the projectile reaches the nanosheet and creates more discrete debris for all the cases presented in Figure 3. Among all tested samples, Ti8C5 generally presents a higher number of discrete debris in all examined impact velocity amplitudes. Moreover, for nanosheets subject to the high impact velocity of 6 km/s, most debris of Ti8C5 tends to be comprised of smaller atom groups (compared with copper), whereas the debris of graphene is more likely to be comprised of discrete atoms. Additionally, we conducted additional simulations by varying the time step from 0.1 fs to 0.5 fs, from which the same results were obtained.

3.2. Resonance and Damping

Energy dissipation associated with resonance and damping are rarely discussed in the past shock loading or hyper velocity impact studies; however, these characteristics are essential for strengthening the matrix in laminar composites, as materials with different resonance frequencies and outstanding quality factors ( Q ) could eliminate the energy in the interfacial reflection process and weaken shock waves [20,39]. Utilizing the same experiment set up for all cases, oscillations in a relatively short period of 10 ps generated via various velocity impacts are recorded for amplitude and frequency comparisons. Two close observation points labelled in Figure 1a are selected for all the cases, ensuring that the captured oscillation is not influenced by crack propagation. Generally, increasing the impact velocity, oscillation amplitude and frequency does not cause significant variations in all cases. In addition, compared with graphene (Figure 4c,d) and copper (Figure 4e,f) nanosheets, Ti8C5 (Figure 4a,b) has significantly lower vibrational frequencies and higher amplitudes.
For a more quantitative vibrational evaluation and to calculate the energy dissipation capability of the Ti8C5 nanosheet due to the damping process, nanosheets are subjected to a low impact velocity of 0.5 km/s, which do not penetrate the nanosheets or damage the lattice structure significantly. In all simulation scenarios, projectiles bounce back after contact. In such assessments, the external energy of nanosheets right after the departure of the projectile is recorded (Figure 5) for the calculation of the Q factor based on estimation schemes derived by previous researchers [40,41]. Here, external energy is the difference of Ti8C5 potential energy before and after the separation of the two objects. Q is defined as the ratio between the total system energy and the average energy loss in one radian at resonant frequency, i.e., Q = 2 π E v / Δ E v , where E v is the total energy of the vibrating system and Δ E v is donated as the dissipated energy in one cycle of vibration. Q is assumed to be constant during vibration; therefore, the correlation between maximum energy ( E v , n ) after the n vibration cycles and initial energy ( E v , 0 ) in the first vibration cycle can be denoted as E v , n = E v , 0 ( 1 2 π / Q ) n . According to the external energy change of Ti8C5 presented in Figure 5a, a clear decay pattern is determined. It was noticed that the NVE ensemble adopted in this study and total energy in the vibration phase is constant, and the loss in potential energy contributes to both kinetic energy and the temperature rise of the system. Because of the remarkable decay of the oscillation amplitude, the Q factor of Ti8C5 possesses a small value of~111.9, and the corresponding main resonance frequency for Ti8C5 is calculated via discrete Fourier transform, being ~11.9 GHz (Figure 5b). For the graphene and copper nanosheet impact simulation, their main resonance frequencies were ~34.01 GHz and ~43.03 GHz, respectively, whereas the Q factors were 413.37 and 137.17, respectively. Considering the external energy variation and Q Factor in the 600 ps time domain, Ti8C5 clearly demonstrates a better damping capability over single layer graphene. Although the Q factor calculation results can be affected by the presence of a fixed boundary, all the simulations were conducted with the same fixed boundary setup, which guarantees a fair comparison.

3.3. Stress Distribution and Propagation

The above discussion illustrates the morphology of Ti8C5 under impact. It is of great interest to compare how the stress distribution and propagation pattern in Ti8C5 would be altered with the variation velocity amplitudes from a more statistical point of view. In this regard, the Von Mises atomic stress distribution and elastic stress wave propagation under different impact velocities are studied as they largely determine the impact resist performance. For a lower impact velocity of 3 km/s, stress throughout the Ti8C5 nanosheet prior to bond break presents a multimodal distribution pattern (Figure 6a). The main body of the data centred around 19 GPa is mainly due to impact whereas the peak on the left side is created by the fixed boundaries. As the impact increases, the peak on the left maintains the same level, whereas the mode of the main body decreases with its tail shifts towards the right (Figure 6b–d). Higher impact velocities lead to more severe stress concentrations around the impact region; moreover, stress distribution prior to bond breakage is dominated by the elastic stress propagation velocity, as the impact velocity rises much higher than the stress propagation rate, and ‘less connection’ between the main body and the tail is observed. Cumulative density function (CDF) presented in Figure 6e summaries the finding: the slope of each curve is inversely correlated with the impact velocity.
A nanosheet with a more significant elastic wave propagation velocity transfers momentum at a superior rate; thus, a more even stress distribution/energy delocalization is expected during ballistic impact. Theoretically, the elastic stress wave velocity v s in a solid material can be calculated based on its Young’s modulus γ and density ρ according to v s = Y / ρ [42,43]. In this study, v s is estimated via tracking the location of the highest Von Mises atomic stress, as studies suggests decent agreement between the two approaches [38]. The elastic wave propagation velocity for Ti8C5 is 5.34 and 5.90 km/s for X and Y directions, respectively, which presents a significant gap considering copper is 7.91 (along [100] direction) and 7.52 km/s (along [010] direction). As a result, a more evenly distributed pattern throughout the nanosheet is captured at the same time instance (Figure 7a) compared with Ti8C5 (Figure 6a). Graphene possesses the highest elastic wave propagation velocity, which is 20.11 and 19.84 km/s along the armchair and zigzag directions, respectively; as such, its stress distribution shows a unimodal shape (Figure 7b). Slopes of the three curves in the cumulative probability diagram indicate graphene and copper are likely to bear more loading prior to bond break, and it also suggests that Ti8C5 owns the least significant breaking stress among all the three materials.

3.4. Impact Resistance Evaluation

To quantitatively evaluate the impact resistance of Ti8C5, its penetration (Ep) and specific penetration energy,   E p * , under various velocity amplitudes are calculated and compared with graphene and copper. According to Figure 8a, the penetration energy of both Ti8C5 and copper show a consistent rate of growth in the whole velocity domain, whereas graphene experiences a notable slump in Ep at a relatively low impact velocity, which is due to the reduction in overall defamation [36]. After increasing the impact above 3 km/s, the same overall ranking is maintained. For all impact velocities, copper presents the most significant Ep and the gap between each category enlarges with impact velocity. Projectiles are not rigid bodies in this study; as a result, cracks in projectiles are found during the process if they were assigned ultra-high velocities (i.e., copper is impacted at 7 km/s).
Theoretically, for nanosheets with thickness h, which is far less than projectile diameter D, i.e., D/h >> 1 [44], it is regarded as a thin film and its penetration energy can be estimated by E p = Δ P e N S + Δ K e N S = ( ρ A s h ) v 2 / 2 + E d . The term ( ρ A s h ) v 2 / 2 refers to the minimum inelastic energy or kinetic energy transferred to the target nanosheet (where A s represents the strike face area and v stands for the impact velocity) and the second term E d represents other energy dissipation mechanisms such as elastic deformation or breakage of bonds. Considering the different morphology, mass as well as density differences of various nanosheets, specific or gravimetric penetration energy is adopted to evaluate their impact resistance capability, which is calculated as E p * = E p / ( ρ A s h ) and it can be further expressed as E P * = v 2 / 2 + E d * . Apparently, E d * is a figure of merit to evaluate the impact energy delocalization ability, which is dominated by elastic wave propagation velocity. Better energy delocalization ability prevents the development of local severe stress and helps the material to deform more locally prior to the arrival of bond break stress.
E P * in each group follows the material-independent energy dissipation baseline (i.e., v 2 / 2 ), and the overall ranking differs from the penetration energy (Figure 8a,b). At low impact velocities, ~2 km/s graphene shows superiority over the other two materials in terms of specific penetration energy, because of its most significant local deformation prior to bond break and, hence, a larger E d * term is expected, which is confirmed by the energy breakdown presented in Figure 9. For low impact velocities <3 km/s, graphene shows a more significant Δ P e N S   / Δ E N S , t o t ratio compared with the other two nanosheets. As impact velocity increases above 3 km/s, the elastic wave propagation velocity or energy delocalization ability becomes less dominant in determining the impact resistance, as all samples experience less overall deformation prior to bond break with a less significant Δ P e N S   / Δ E N S , t o t ratio observed. As a result, an increase in impact velocity closes the specific penetration energy gap between graphene and Ti8C5.

4. Conclusions

In summary, the fracture behaviour of monolayer Ti8C5 subject to various hypervelocity impacts is explored. For Ti8C5 under relatively a low impact velocity ~0.5 km/s, its main resonance frequency is 11.9 GHz, which is different from other matrices in composites such as copper and graphene. A resonance study on Ti8C5 also shows a low Q factor of 111.9 indicates decent energy damping capability. The combination of resonance frequency and low Q factor would help to eliminate energy in the interfacial reflection process and weaken shock waves for Ti8C5-strengthened composites. When increasing the impact velocity above the threshold of 1.8 km/s, a crack initiates at the geometry centre and the nanosheet experiences significant out-of-plane deformation before the propagation of the crack, demonstrating brittle behaviour. Elastic wave propagation velocity is an important figure in impact resistance assessments, as it positively correlates with the energy delocalization rate, which helps materials to deform more prior to bond break and thus absorb more energy. Tracking atomic Von Mises stress distribution, the elastic wave propagation velocity of Ti8C5 is calculated to be 5.34 and 5.90 km/s for X and Y directions, respectively, but these figures are not significant compared with graphene and copper, suggesting inferior energy delocalization rates and thus less energy dissipation via deformation prior to bond break. However, because of its relatively small mass density compared with copper, Ti8C5 presents superior specific penetration for all tested velocity amplitudes. This study provides a fundamental understanding of the deformation and penetration mechanisms of titanium carbide nanosheets under impact, which should shed lights on the design of titanium carbide- related composites for bullet-proof applications or shielding structures for aerospace systems’ protection applications for titanium carbide-related composites. We note that although these results are obtained from a Ti8C5 nanosheet with a small size, similar failure mechanisms are also expected for monolithic Ti8C5 nanosheets with a large scale.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/met12111989/s1, Figure S1: The penetration profile for Ti8C5 nanosheet subjected to impact velocity of 3 km/s. (a) Penetration energy of Ti8C5 nanosheet with environment temperature up to 700 K. (b) Ti8C5 nanosheet deformation with environment temperature of 10 K at simulation time of 2.8 ps. (c) Ti8C5 nanosheet deformation with environment temperature of 700 K at simulation time of 2.8 ps.

Author Contributions

Conceptualization, K.X. and H.Z.; methodology, K.X.; software, K.X.; validation, K.X. and H.Z.; formal analysis, K.X.; data curation, K.X.; writing—original draft preparation, K.X.; writing—review and editing, K.X., H.Z., J.S., J.W., Z.Z. and X.Z.; visualization, K.X.; supervision, Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by the National Natural Science Foundation of China (12102127), Natural Science Foundation of Jiangsu Province (BK20190164), Fundamental Research Funds for the Central Universities (B210202125), the China Postdoctoral Science Foundation (2021M690872) and Changzhou Health Commission Technology Projects (ZD202103).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fang, M.; Han, Y.; Shi, Z.; Huang, G.; Song, J.; Lu, W. Embedding boron into Ti powder for direct laser deposited titanium matrix composite: Microstructure evolution and the role of nano-TiB network structure. Compos. B 2021, 211, 108683. [Google Scholar] [CrossRef]
  2. Wen, Y.; Wu, Y.; Hua, L.; Xie, L.; Wang, L.; Zhang, L.-C.; Lu, W. Effects of shot peening on microstructure evolution and mechanical properties of surface nanocrystal layer on titanium matrix composite. Mater. Des. 2021, 206, 109760. [Google Scholar] [CrossRef]
  3. Xie, L.; Wu, Y.; Yao, Y.; Hua, L.; Wang, L.; Zhang, L.-C.; Lu, W. Refinement of TiB reinforcements in TiB/Ti-2Al-6Sn titanium matrix composite via electroshock treatment. Mater. Charact. 2021, 180, 111395. [Google Scholar] [CrossRef]
  4. Xie, H.; Jin, Y.; Niu, M.; Liu, N.; Wang, J. Effect of Tribofilm Induced by Nanoparticle Addition on Wear Behavior of Titanium-Matrix Composite. Tribol. Lett. 2021, 69, 18. [Google Scholar] [CrossRef]
  5. Niu, X.; Liu, Y.; Chen, X.; Liu, J.; Sun, Z.; Song, Y. Growth behavior of short fatigue cracks in a unidirectional SiC fiber-reinforced titanium matrix composite under spectrum loading. Theor. Appl. Fract. Mech. 2021, 114, 102980. [Google Scholar] [CrossRef]
  6. Niu, H.Z.; Yin, B.G.; Zhang, H.R.; Zang, M.C.; Tan, H.; Zhang, D.L. Multiphase polymorphic nanoparticles reinforced titanium matrix composite produced by selective electron beam melting of a prealloyed composite powder. Scr. Mater. 2021, 200, 113916. [Google Scholar] [CrossRef]
  7. Borysiuk, V.N.; Mochalin, V.N.; Gogotsi, Y. Bending rigidity of two-dimensional titanium carbide (MXene) nanoribbons: A molecular dynamics study. Comput. Mater. Sci. 2018, 143, 418–424. [Google Scholar] [CrossRef]
  8. Hatam-Lee, S.M.; Esfandiar, A.; Rajabpour, A. Mechanical behaviors of titanium nitride and carbide MXenes: A molecular dynamics study. Appl. Surf. Sci. 2021, 566, 150633. [Google Scholar] [CrossRef]
  9. Rao, V.R.; Ramanaiah, N.; Sarcar, M.M.M. Tribological properties of Aluminium Metal Matrix Composites (AA7075 Reinforced with Titanium Carbide (TiC) Particles). Int. J. Adv. Sci. Technol. 2016, 88, 13–26. [Google Scholar] [CrossRef]
  10. Li, Z.-K.; Liu, Y.; Li, L.; Wei, Y.; Caro, J.; Wang, H. Ultra-thin titanium carbide (MXene) sheet membranes for high-efficient oil/water emulsions separation. J. Membr. Sci. 2019, 592, 117361. [Google Scholar] [CrossRef]
  11. Guo, J.; Xia, C. Wear Behavior and Mechanism of Thermal Sprayed TiC-NiCr Coating. Nonferrous Met. Eng. 2021, 11, 48–53. [Google Scholar]
  12. Chu, K.; Wang, F.; Wang, X.-H.; Li, Y.-B.; Geng, Z.-R.; Huang, D.-J.; Zhang, H. Interface design of graphene/copper composites by matrix alloying with titanium. Mater. Des. 2018, 144, 290–303. [Google Scholar] [CrossRef]
  13. Liu, X.; Liu, L.; Sun, B.; Li, B.; Ke, X. The preparation of titanium carbide thin film of Ti6Al4V at low temperature and study of friction and wear properties. J. Meas. Eng. 2016, 4, 167–172. [Google Scholar]
  14. Zhou, L.; Liu, Y.; Dong, S.; Zhu, L.; Li, Y. Synthesis of Titanium Carbide Particle in Laser Melted-pool in situ. J. Wuhan Univ. Technol. Sci. Ed. 2020, 34, 1315–1320. [Google Scholar] [CrossRef]
  15. Borysiuk, V.N.; Mochalin, V.N.; Gogotsi, Y. Molecular dynamic study of the mechanical properties of two-dimensional titanium carbides Tin + 1Cn (MXenes). Nanotechnology 2015, 26, 265705. [Google Scholar] [CrossRef] [Green Version]
  16. Fu, Z.H.; Zhang, Q.F.; Legut, D.; Si, C.; Germann, T.C.; Lookman, T.; Du, S.Y.; Francisco, J.S.; Zhang, R.F. Stabilization and strengthening effects of functional groups in two-dimensional titanium carbide. Phys. Rev. B 2016, 94, 104103. [Google Scholar] [CrossRef]
  17. Xiong, N.; Bao, R.; Yi, J.; Tao, J.; Liu, Y.; Fang, D. Interface evolution and its influence on mechanical properties of CNTs/Cu-Ti composite. Mater. Sci. Eng. A 2019, 755, 75–84. [Google Scholar]
  18. Mark, J.J.; Zhang, X. Effect of using titanium interlayers on the properties of vitrified diamond tools. Int. J. Comput. Mater. Sci. Surf. Eng. 2010, 3, 86–94. [Google Scholar] [CrossRef]
  19. Liu, X.; Wang, F.; Wang, W.; Wu, H. Interfacial strengthening and self-healing effect in graphene-copper nanolayered composites under shear deformation. Carbon 2016, 107, 680–688. [Google Scholar] [CrossRef]
  20. Liu, X.; Wang, F.; Wu, H.; Wang, W. Strengthening metal nanolaminates under shock compression through dual effect of strong and weak graphene interface. Appl. Phys. Lett. 2014, 104, 231901. [Google Scholar] [CrossRef]
  21. Steve, P. Fast parallel algorithms for short-range molecular dynamics. J. Comput. Phys. 1995, 117, 1–19. [Google Scholar]
  22. Jain, A.; Ong, S.P.; Hautier, G.; Chen, W.; Richards, W.D.; Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.; Ceder, G.; et al. The Materials Project: A materials genome approach to accelerating materials innovation. APL Mater. 2013, 1, 011002. [Google Scholar] [CrossRef] [Green Version]
  23. Zhao, H.; Aluru, N.R. Temperature and strain-rate dependent fracture strength of graphene. J. Appl. Phys. 2010, 108, 064321. [Google Scholar] [CrossRef] [Green Version]
  24. Yi, L.; Yin, Z.; Zhang, Y.; Chang, T. A theoretical evaluation of the temperature and strain-rate dependent fracture strength of tilt grain boundaries in graphene. Carbon 2013, 51, 373–380. [Google Scholar] [CrossRef]
  25. Fonseca, A.F.; Liang, T.; Zhang, D.; Choudhary, K.; Phillpot, S.R.; Sinnott, S.B. Graphene-Titanium Interfaces from Molecular Dynamics Simulations. ACS Appl. Mater. Interfaces 2017, 9, 33288–33297. [Google Scholar] [CrossRef]
  26. Zhang, D.; Fonseca, A.F.; Liang, T.; Phillpot, S.R.; Sinnott, S.B. Dynamics of graphene/Al interfaces using COMB3 potentials. Phys. Rev. Mater 2019, 3, 114002. [Google Scholar] [CrossRef] [Green Version]
  27. Jiang, T.; Zhang, X.; Vishwanath, S.; Mu, X.; Kanzyuba, V.; Sokolov, D.A.; Ptasinska, S.; Go, D.B.; Xing, H.G.; Luo, T. Covalent bonding modulated graphene-metal interfacial thermal transport. Nanoscale 2016, 8, 10993–11001. [Google Scholar] [CrossRef]
  28. Cingolani, J.S.; Deimel, M.; Köcher, S.; Scheurer, C.; Reuter, K.; Andersen, M. Interace between graphene and liquid Cu from molecular dynamics simulations. J. Chem. Phys. 2020, 153, 074702. [Google Scholar] [CrossRef]
  29. Lin, F.; Xiang, Y.; Shen, H.S. Tunable Positive/Negative Young’s Modulus in Graphene—Based Metamaterials. Adv. Theory Simul. 2021, 4, 2000130. [Google Scholar] [CrossRef]
  30. Sha, Z.; Branicio, P.; Pei, Q.; Sorkin, V.; Zhang, Y. A modified Tersoff potential for pure and hydrogenated diamond-like carbon. Comput. Mater. Sci. 2013, 67, 146–150. [Google Scholar] [CrossRef]
  31. Mylvaganam, K.; Zhang, L. Ballistic resistance capacity of carbon nanotubes. Nanotechnology 2007, 18, 475701. [Google Scholar]
  32. Hoover, W.G. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A 1985, 31, 1695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Verlet, L. Computer “experiments” on classical fluids. I. Thermodynamical properties of Lennard-Jones molecules. Phys. Rev. 1967, 159, 98. [Google Scholar] [CrossRef] [Green Version]
  34. Diao, J.; Gall, K.; Dunn, M.L. Atomistic simulation of the structure and elastic properties of gold nanowires. J. Mech. Phys. Solids 2004, 52, 1935–1962. [Google Scholar] [CrossRef]
  35. Xia, K.; Zhan, H.; Ji, A.; Shao, J.; Gu, Y.; Li, Z. Graphynes: An alternative lightweight solution for shock protection. Beilstein J. Nanotechnol. 2019, 10, 1588–1595. [Google Scholar] [CrossRef] [PubMed]
  36. Xia, K.; Zhan, H.; Hu, D.; Gu, Y. Failure mechanism of monolayer graphene under hypervelocity impact of spherical projectile. Sci. Rep. 2016, 6, 33139. [Google Scholar] [CrossRef] [Green Version]
  37. Yin, H.; Qi, H.J.; Fan, F.; Zhu, T.; Wang, B.; Wei, Y. Griffith Criterion for Brittle Fracture in Graphene. Nano Lett. 2015, 15, 1918–1924. [Google Scholar] [CrossRef] [Green Version]
  38. Xia, K.; Zhan, H.; Zhang, X.; Li, Z. Graphdiyne family-tunable solution to shock resistance. Mater. Res. Express. 2020, 7, 115602. [Google Scholar] [CrossRef]
  39. Liao, C.-Y.; Ma, C.-C. Transient behavior of a cantilever plate subjected to impact loading: Theoretical analysis and experimental measurement. Int. J. Mech. Sci. 2020, 166, 105217. [Google Scholar]
  40. Wei, Y.; Zhan, H.; Xia, K.; Zhang, W.; Sang, S.; Gu, Y. Resonance of graphene nanoribbons doped with nitrogen and boron: A molecular dynamics study. Beilstein J. Nanotechnol. 2014, 5, 717–725. [Google Scholar]
  41. Zheng, Z.; Zhan, H.; Nie, Y.; Xu, X.; Qi, D.; Gu, Y. Single layer diamond—A new ultrathin 2D carbon nanostructure for mechanical resonator. Carbon 2020, 161, 809–815. [Google Scholar] [CrossRef]
  42. Hernandez, S.A.; Fonseca, A.F. Anisotropic elastic modulus, high Poisson’s ratio and negative thermal expansion of graphynes and graphdiynes. Diam. Relat. Mater. 2017, 77, 57–64. [Google Scholar] [CrossRef] [Green Version]
  43. Zhang, Y.; Pei, Q.; Wang, C. Mechanical properties of graphynes under tension: A molecular dynamics study. Appl. Phys. Lett. 2012, 101, 081909. [Google Scholar] [CrossRef] [Green Version]
  44. Lee, J.-H.; Loya, P.E.; Lou, J.; Thomas, E.L. Dynamic mechanical behavior of multilayer graphene via supersonic projectile penetration. Science 2014, 346, 1092–1096. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Impact simulation setup for the Ti8C5 model, (010) with the plane of the Ti8C5 facing the projectile. (a) Top view of the schematic setup. Insert shows the unit cell thickness of the (001) plane Ti8C5, and atoms labelled in silver and bule are carbon and titanium, respectively; (b) Side view of the schematic setup; the thickness of the model is equal to the thickness of the Ti8C5 unit cell; (c) Energy change in projectile as a function of time for impact velocity of 2 km/s. ∆Eball,ke and ∆Eball,pe represent the kinetic and potential energy change of the projectile, respectively. The green balls represent the diamond projectile at a simulation time of 0, 2 and 3.9 ps. Significant deformation is not observed in the projectile during the impact process.
Figure 1. Impact simulation setup for the Ti8C5 model, (010) with the plane of the Ti8C5 facing the projectile. (a) Top view of the schematic setup. Insert shows the unit cell thickness of the (001) plane Ti8C5, and atoms labelled in silver and bule are carbon and titanium, respectively; (b) Side view of the schematic setup; the thickness of the model is equal to the thickness of the Ti8C5 unit cell; (c) Energy change in projectile as a function of time for impact velocity of 2 km/s. ∆Eball,ke and ∆Eball,pe represent the kinetic and potential energy change of the projectile, respectively. The green balls represent the diamond projectile at a simulation time of 0, 2 and 3.9 ps. Significant deformation is not observed in the projectile during the impact process.
Metals 12 01989 g001
Figure 2. Von Mises stress distribution of nanosheets under impact velocity of 3km/s. (ac) Atomic configuration of Ti8C5: (a) Von Mises stress distribution pattern at a simulation time of 1.3 ps; the insert represents stress accumulation and formation of cracks in the impact region; (b) Stress redistribution/propagation at a simulation time of 2.3 ps; (c) Final atomic configuration at 4.0 ps. (df) Atomic configuration of copper: (d) Von Mises stress distribution pattern at a simulation time of 2.0 ps; (e) Stress propagation phase at 3.0 ps; (f) final atomic configuration at 6.0 ps.
Figure 2. Von Mises stress distribution of nanosheets under impact velocity of 3km/s. (ac) Atomic configuration of Ti8C5: (a) Von Mises stress distribution pattern at a simulation time of 1.3 ps; the insert represents stress accumulation and formation of cracks in the impact region; (b) Stress redistribution/propagation at a simulation time of 2.3 ps; (c) Final atomic configuration at 4.0 ps. (df) Atomic configuration of copper: (d) Von Mises stress distribution pattern at a simulation time of 2.0 ps; (e) Stress propagation phase at 3.0 ps; (f) final atomic configuration at 6.0 ps.
Metals 12 01989 g002
Figure 3. Impact deformation and discrete debris of nanosheets under various impact velocities. Amount of debris after impact for (a,b) Ti8C5, (c,d) Copper and (e,f) Graphene are presented, respectively.
Figure 3. Impact deformation and discrete debris of nanosheets under various impact velocities. Amount of debris after impact for (a,b) Ti8C5, (c,d) Copper and (e,f) Graphene are presented, respectively.
Metals 12 01989 g003
Figure 4. Z-direction oscillation of adjacent carbon and titanium atoms located at the highlighted position in Figure 1a of the Ti8C5 nanosheet under impact velocities of (a) 3 km/s and (b) 6 km/s; z-direction oscillation of 2 carbon atoms in graphene nanosheets (similar position as the Ti8C5) under impact velocities of (c) 3 km/s and (d) 6 km/s, respectively; z-direction oscillation of 2 copper atoms (similar position as the Ti8C5) under impact velocities of (e) 3 km/s and (f) 6 km/s, respectively.
Figure 4. Z-direction oscillation of adjacent carbon and titanium atoms located at the highlighted position in Figure 1a of the Ti8C5 nanosheet under impact velocities of (a) 3 km/s and (b) 6 km/s; z-direction oscillation of 2 carbon atoms in graphene nanosheets (similar position as the Ti8C5) under impact velocities of (c) 3 km/s and (d) 6 km/s, respectively; z-direction oscillation of 2 copper atoms (similar position as the Ti8C5) under impact velocities of (e) 3 km/s and (f) 6 km/s, respectively.
Metals 12 01989 g004
Figure 5. Oscillation of external energy of nanosheet after the separation of projectile and nanosheet. (a) External energy variation of Ti8C5 over time period of 600 ps and (b) The corresponding frequency spectrum of Ti8C5 over time period of 600 ps.
Figure 5. Oscillation of external energy of nanosheet after the separation of projectile and nanosheet. (a) External energy variation of Ti8C5 over time period of 600 ps and (b) The corresponding frequency spectrum of Ti8C5 over time period of 600 ps.
Metals 12 01989 g005
Figure 6. Von Mises atomic stress distribution of Ti8C5 prior to bond break, when it subject to impact velocity: (a) 3 km/s; (b) 4 km/s; (c) 5 km/s; (d) 6 km/s. (e) Cumulative density function (CDF) of the Von Mises atomic stress distribution before the crack initiation for different nanosheets under diverse impact velocities.
Figure 6. Von Mises atomic stress distribution of Ti8C5 prior to bond break, when it subject to impact velocity: (a) 3 km/s; (b) 4 km/s; (c) 5 km/s; (d) 6 km/s. (e) Cumulative density function (CDF) of the Von Mises atomic stress distribution before the crack initiation for different nanosheets under diverse impact velocities.
Metals 12 01989 g006
Figure 7. Stress distribution diagram for (a) copper and (b) graphene subject to an impact velocity of 3 km/s; (c) Cumulative density function (CDF) of the Von Mises atomic stress distribution before the crack initiation for different nanosheets under an impact velocity of 3km/s (red, blue and brown lines represent copper, graphene and Ti8C5, respectively).
Figure 7. Stress distribution diagram for (a) copper and (b) graphene subject to an impact velocity of 3 km/s; (c) Cumulative density function (CDF) of the Von Mises atomic stress distribution before the crack initiation for different nanosheets under an impact velocity of 3km/s (red, blue and brown lines represent copper, graphene and Ti8C5, respectively).
Metals 12 01989 g007
Figure 8. Performance of nanosheets’ impact with projectiles of various velocity amplitudes. (a) Penetration energy and (b) Specific penetration energy as a function of impact velocity for Ti8C5, graphene and Copper nanosheets.
Figure 8. Performance of nanosheets’ impact with projectiles of various velocity amplitudes. (a) Penetration energy and (b) Specific penetration energy as a function of impact velocity for Ti8C5, graphene and Copper nanosheets.
Metals 12 01989 g008
Figure 9. Energy breakdown for various nanosheets at the moment of perforation. Figures (ac) represent the kinetic energy gain ( Δ K e N S ) and the potential energy gain ( Δ P e N S ) for Ti8C5, copper and the graphene nanosheet, respectively.
Figure 9. Energy breakdown for various nanosheets at the moment of perforation. Figures (ac) represent the kinetic energy gain ( Δ K e N S ) and the potential energy gain ( Δ P e N S ) for Ti8C5, copper and the graphene nanosheet, respectively.
Metals 12 01989 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xia, K.; Zhan, H.; Shao, J.; Wang, J.; Zheng, Z.; Zhang, X.; Li, Z. Atomistic Investigation of Titanium Carbide Ti8C5 under Impact Loading. Metals 2022, 12, 1989. https://doi.org/10.3390/met12111989

AMA Style

Xia K, Zhan H, Shao J, Wang J, Zheng Z, Zhang X, Li Z. Atomistic Investigation of Titanium Carbide Ti8C5 under Impact Loading. Metals. 2022; 12(11):1989. https://doi.org/10.3390/met12111989

Chicago/Turabian Style

Xia, Kang, Haifei Zhan, Jianli Shao, Jiaqiu Wang, Zhuoqun Zheng, Xinjie Zhang, and Zhiyong Li. 2022. "Atomistic Investigation of Titanium Carbide Ti8C5 under Impact Loading" Metals 12, no. 11: 1989. https://doi.org/10.3390/met12111989

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