Next Article in Journal
Preparation of Battery-Grade Lithium Carbonate with Lithium-Containing Desorption Solution
Next Article in Special Issue
Strain Amplitude Dependence of High Damping Grp/Mg97Zn1Y2 Composites Ranging from Anelastic to Microplastic
Previous Article in Journal
Effect of the Thermomechanical Treatment on the Corrosion of UNSM Processed Inconel 718: An Electrochemical Study
Previous Article in Special Issue
Study on the Relationship between High Temperature Mechanical Properties and Precipitates Evolution of 7085 Al Alloy after Long Time Thermal Exposures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dynamic Mechanical Properties of Ti–Al3Ti–Al Laminated Composites: Experimental and Numerical Investigation

College of Mechatronic Engineering, North University of China, Taiyuan 030051, China
*
Author to whom correspondence should be addressed.
Submission received: 28 August 2021 / Revised: 15 September 2021 / Accepted: 16 September 2021 / Published: 19 September 2021

Abstract

:
The Ti–Al3Ti–Al laminated composites with different Al contents were prepared by vacuum hot pressing sintering technology. The effects of Al content on the dynamic mechanical properties of the composites were studied using the combination of Split Hopkinson Pressure Bar experiment and finite element analysis. The results showed that different Al content changes the fracture mode of the composites. The laminated composites without Al have higher brittleness and lower fracture strain. The Ti–Al3Ti–Al laminated composites containing 10–15%Al have better dynamic mechanical properties than those without Al, but the subsequent increase of Al content is not conducive to the improvement of strength. However, when the Al content in the specimen reaches 30%, the dynamic mechanical properties of the composites decrease, multi-crack phenomenon and relatively large strain occur, and the Al extruded from the layers fills the crack.

Graphical Abstract

1. Introduction

Ti–Al3Ti laminated composites, as a kind of lightweight high performance structure material, are composed of ductile Ti layers and intermetallic compound Al3Ti layers. Their unique laminated structure and special failure modes make these composites possess excellent dynamic impact resistance. Therefore, the laminated composites are usually used in armor protection, aviation aerospace engineering, and other fields, and have a promising application prospect [1,2,3,4,5,6].
At present, scholars have researched into the mechanical properties of Ti–Al3Ti laminated composites, including: three-point bending [7,8], quasi-static compression [9,10], and anti-penetration performance [11,12,13,14,15,16,17], etc. Adharapurapu [18] et al., reported that the toughness and fatigue properties of Ti–Al3Ti laminated composites were improved obviously due to the bridging effect of ductile titanium. Cao [19] et al., through the combination of experiment and simulation, found that the damage modes of Ti–Al3Ti laminated composites include traverse cracks, crack deflection, and delamination, which enabled the composites to absorb more energy. Zhou [20] et al., found that the fracture modes of the Ti–Al3Ti interface in the composites varied as a function of strain rate.
It is noted that during the vacuum hot-pressing sintering, different preparation processes would lead to a certain degree of Al still remaining, and form Ti–Al3Ti–Al laminated composites. Yuan [21] et al., reported that compared to Ti–Al3Ti laminated composites, Ti–Al3Ti–Al laminated composites have better static and dynamic compression strength. Price R.D. [22] prepared Ti–Al3Ti–Al laminated composites by controlling the reactive foils sintering process. The dynamic compression study showed that the composites with 3–15% Al have the best mechanical properties. Du [23] et al., carried out multiple hot rolling of Ti–Al laminated plates at 500 °C and found that the tensile strength of this composite increase with the increasing of the volume fraction of Al. Patselov [24] et al., prepared Ti–Al3Ti–Al laminated composite with 15% volume fraction Al. Through the three-point bending experiment, they found that the fatigue signs in the form of grooves were clearly visible in the aluminum layer at the end of the fatigue fracture zone.
The current research mainly focuses on their influence on the overall mechanical properties of laminated composites [25,26,27,28,29,30], but the systematic studies on the strengthening mechanism of Al are rare. In this study, Ti–Al3Ti–Al laminated composites were prepared by endothermic semi-solid reaction [31]. The compression behavior and crack propagation of Ti–Al3Ti–Al laminated composites under high strain rate were studied. The influences of Al layer on the fracture properties were analyzed emphatically. Finally, through finite element simulation, the influences of different Al contents on the dynamic mechanical properties of Ti–Al3Ti–Al laminated composites were analyzed.

2. Experimental Method

2.1. Materials Preparation

Ti–Al3Ti–Al laminated composites were prepared by the hot-pressing sintering method. Al alloy foils and TC4 alloy foils of different thickness were used as the initial components of the Ti–Al3Ti–Al laminated composites. The compositions of TC4 alloy foils and Al alloy foils are shown in Table 1. The processed and cleaned TC4 foils and Al foils were stacked alternately, then placed in the hot-pressing sintering furnace (ZT-40-21Y, Shanghai Chenhua Science Technology CO. Ltd, Shanghai, China). The temperature was then raised to 600–710 °C. The Ti–Al3Ti–Al laminated composites with residual Al were obtained by adjusting the thickness ratio of the laminates and controlling the heat preservation time. The prepared materials were cut into φ8 mm cylindrical specimens through the line cutting process.

2.2. Mechanical Experiment

The Split Hopkinson Pressure Bar (SHPB) (Harbin Institute of Technology, Harbin, China) experiment is one of the most important research methods to investigate the dynamic mechanical properties of the material at strain rates of 102−104 S−1. The diameter and the length of the bar were 12 mm and1200 mm, respectively. The specimen was 8 mm in diameter and 6 mm in thickness. The SHPB apparatus consists of a high pressure air gun, a striker bar, an incident bar, and a transmitter bar as shown in Figure 1.
The impact bar launched by the high-pressure air gun impacted the input bar axially and resulted in the formation of loading pulse, which propagated along the bar to the specimen. The specimen then deformed under the action of the pulse. Meanwhile, the reflection pulse was transmitted back to the incident bar, and the transmission pulse was transmitted to the transmission bar. The gauge at the center of the bar was used to record the strain. The incident, reflected, and transmitted signals were recorded on the oscilloscope in the form of voltage, and the voltage was converted to the actual incident ε i , reflected ε r and transmitted ε t .
Then the pattern strain rate ε ˙ s , strain σ s , and stress ε s were:
ε ˙ s = d ε s d t = C 0 l 0 ( ε i ε r = ε t )
σ s = E A 2 A s ( ε i + ε r + ε t )
ε s = C 0 l s 0 t 0 ( ε i ε r ε t ) d t
where: A S is the cross-sectional area of the test piece, and l S is the thickness of the test piece. C 0 is the elastic longitudinal wave velocity of the waveguide rod.
The main equipment parameters are: high-strength steel was selected for all the bars, the elastic modulus is 200 GPa, the density is 7800 cm/m3, and Poisson’s ratio is 0.29. The length of the impact bar is 200 mm, and the length of the incident bar and transmission bar is 1200 mm. The impact rate is 35 m/s.

2.3. Numerical Simulation Method

Ansys LS-DYNA (Ansys 17.0 LS-DYNA, Ansys Inc., Pittsburgh, PA, USA) was used to simulate the Split Hopkinson Pressure Bar experiment and study the dynamic mechanical properties of Ti–Al3Ti–Al laminated composites. Both the Split Hopkinson Pressure Bar and the impacted specimen are cylindrical, so the three-dimensional axisymmetric finite element model was established (Figure 2) using the APDL command stream. As shown in Figure 2b, the laminated composite model consists of Al layers, Ti layers, and Al3Ti layers. To make the calculation process smoother and the calculation results more accurate, SOLID 185 elements were selected to represent Al layers, Ti layers, and Al3Ti layers. A common node setting was adopted to simulate the interface combination. Ti layers, Al3Ti layers, and Al layers in laminated composites were meshed with different amounts of quadrilateral elements in terms of their thickness. The total number of elements was around 43,200, the time per step is 1μs.
The bar and specimen were simplified into coaxial cylinders. To reduce errors, mesh encrypted treatment was carried out on the bar near the specimen. The diameter of the bullet was 12 mm and its length was 20 mm, and the speed was set to 30 m/s. The diameter and the length of the bar were 12 mm and1200 mm, respectively. The specimen was 8 mm in diameter and 6 mm in thickness. The dimensions of each layer of the model are shown in Table 2. The model without Al contains 25 TC4 layers and 24 Al3Ti layers (Model 1), the other models contain 24 layers of Al, 25 layers of TC4, and 48 layers of Al3Ti.
The JOHNSON–COOK constitutive model and Mie–Gruneisen state equation were used to model TC4 and Al plates. Detailed parameters are shown in Table 3 [32,33].
The Johnson–Holmquist–Ceramics (JH-II) [34] constitutive model was used to describe the brittle Al3Ti material; the specific parameters are shown in Table 4 [32].

3. Results and Discussion

3.1. Dynamic Mechanical Properities

Figure 3 shows the dynamic compressive stress–strain curves of Ti–Al3Ti–Al laminated composites with different Al contents. The maximum peak stresses of 0%, 15%, and 30%Al content are 1156 MPa, 1175 MPa, and 927 MPa, respectively, and the fracture strain is 0.094, 0.103, and 0.122 respectively. It can be seen that Ti–Al3Ti–Al laminated composite with 15% Al content shows a higher compressive strength. When Al contents increase from 0% to 15%, the fracture strain increases by 9.6%. While when Al contents increase from 15% to 30%, the fracture strain increases by 18.4%.
In the strain range of 0–0.01, the stress growth rate gradually decreases with the increase of Al contents, indicating that the presence of Al makes the specimen need more strain to achieve compactness. In the strain range of 0.02–0.08, the stress–strain curves of the specimens are relatively flat due to the significant thermal softening effect of Al that counteracted the effect of strain hardening [35,36]. Some studies [37] have shown that the plastic deformation of aluminum alloys at high strain rates is mainly characterized by adiabatic heating and shear strain localization.
The fracture characteristics of the specimens with different Al contents impacted by SHPB are shown in Figure 4. For the specimen without Al (Figure 4a), the obvious brittleness fracture is occurred, and three cracks appear on the surface.
For the specimen containing about 15% Al (Figure 4b), it can be seen that there is only one main crack on the surface, about 45° to the axis, which means the shear fracture occurred in the specimen. A small amount of Al extrudes from the layer.
For the specimen containing about 30% Al (Figure 4c), the multi-cracks phenomenon appeared in the specimen and caused the whole specimen to fracture. Large amounts of Al are squeezed out of the cracks. The deformation degree of the specimen is large.
By contrast, the brittleness of the specimen without Al is relatively large. The specimen with 15% Al has the best impact resistance. Huang’s research [30] found that local plastic deformation of the Al layer slows down excessive interlayer residual stress. Price’s research [22] also showed that residual Al can slow down Al3Ti damage and prevent separation.

3.2. Simulation Calculation

3.2.1. The Simulated Dynamic Stress-Strain Curves

Figure 5 shows the simulated dynamic stress–strain curves of Ti–Al3Ti–Al laminated composites with different Al contents. The compressive strength of the material with 0%, 10%, 20%, and 30% Al content is 1166.0 MPa, 1181.4 MPa, 1025.2 MPa, and 913.3 MPa respectively, and the fracture strain is 0.090, 0.095, 0.113, and 0.145 respectively. For the Ti–Al3Ti–Al laminated composites with 10% Al content (Ti:Al3Ti:Al = 2:7:1), its fracture strain is greater than that of the materials without Al, and its compressive strength is the highest. It has a good balance between brittleness and toughness. When the Al content is 20%, the compressive strength of the material decreases by 13.2% compared with that of 10% Al, but the fracture strain increases by 18.9%. When the Al content increases to 30%, the compressive strength continues to decrease and the fracture strain continues to increase. Konieczny [38] also pointed out that the amount of residual aluminum on the intermetallic compound centerline increased the ductility of the composites.

3.2.2. Energy Analysis

Figure 6a shows the energy variation curves of the specimens with different Al contents during the SHPB experiment. The time of 0 μs is the time when the incident wave just contacts the specimen. The total energy absorbed by the specimens with 0%, 10%, 20%, and 30% Al content is 29.03 J, 30.29 J, 35.49 J, and 39.88 J, respectively. It can be seen that with the increase of Al content, the energy absorbed by the specimen increases gradually. This indicates that the higher the Al content, the stronger the energy absorption capacity of the laminated composites. After the energy absorption reaches the peak, there is a small decline. This is due to the recovery of elastic deformation and energy release of the specimen.
For the Ti layers of the composites (Figure 6b), with the increase of Al content, the energy absorption of the Ti layer is 18.38 J, 16.88 J, 17.02 J, and 19.30 J, respectively. It can be seen that the energy absorbed by the Ti layers in the Ti–Al3Ti–Al laminated material containing 10% Al is the lowest. However, the energy absorption of Ti layers in the laminated composites without Al is higher than that of 10% Al and 20% Al. In addition, although the volume fraction of the Ti layer only accounts for 20%, its energy absorption is about half of the total energy absorption, which indicates that the ductile layer Ti is the main energy absorption layer, increasing the toughness of laminated composites.
Figure 7 shows the deformation degree of the Ti layer of Ti–Al3Ti–Al laminated composites with different Al contents. It can be seen that with the increase of Al content, the maximum strain of the Ti layer in each sample is 0.498, 0.434, 0.445, and 0.537, respectively. The strain of Ti is the lowest at 10% Al content, indicating that the risk of fracture of the composites is low. This corresponds to the result of Figure 7b above.

3.2.3. Comparison of Failure Morphology

Figure 8 shows the stress nephogram of the longitudinal section of the specimen after impact. Different Al contents lead to different deformation degrees and failure mode of the specimen. It can be seen that Ti–Al3Ti–Al laminated composites containing 10% Al have the best protective performance under dynamic impact.
For the specimen without Al (Figure 8a), the stress distribution is not uniform and a large number of microcracks are generated. However, the Ti layer can effectively prevent the propagation of microcracks, thus the specimen has no overall failure.
For the specimen containing 10% Al (Figure 8b), the whole specimen is relatively intact with few failure units, indicating that Al plays a positive role in it. The overall stress distribution of the specimen is uniform, and there is no high-stress concentration phenomenon.
When the Al content reaches 20% (Figure 8c), the interlayer stress increases, and the phenomenon of stratification occurs.
When the Al content reaches 30% (Figure 8d), more failure elements appear in the specimen, indicating that large cracks have formed, because with the increase of Al content, the strength and stiffness of the material decrease, and the ability of resisting impact load weakens.

4. Conclusions

The dynamic mechanical properties of Ti–Al3Ti–Al laminated composites were studied by SHPB experiment and numerical simulation. The conclusions can be summarized as follows:
(1)
The Ti–Al3Ti–Al laminated composites with Al content of 10–15% had excellent compressive strength and failure strain.
(2)
The laminated composites without Al showed brittle fracture; there were three main cracks on the surface of the specimen. The laminated composites containing 15% Al had only one main crack, and ruptured in a direction approximately 45° from the vertical axis. When the Al content reached 30%, the specimen revealed multiple cracking properties, and the deformation was large.
(2)
When the Al content was 10%, the overall stress distribution of the specimen was uniform and had the best impact resistance.

Author Contributions

Conceptualization, Methodology, J.M.; Project administration, Writing—review & editing, M.Y.; Investigation, L.Z.; Visualization, Z.W.; Supervision, Validation, Funding acquisition, K.W. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the financial supports provided by the National Natural Science Foundation of China (51201155), the Natural Science of Shanxi Province (2012011019-1, 2012011007-1), the Chinese Education Ministry Foundation for Doctors (20101420120006), and the National Key Research and Development Program of China (2018YFC1901704).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw/processed data required to reproduce these findings cannot be shared at this time due to technical or time limitations.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, D.K.; Cizek, P.; Hodgson, P.; Wen, C.E. Ultrafine equiaxed-grain Ti/Al composite produced by accumulative roll bonding. Scr. Mater. 2010, 62, 321–324. [Google Scholar] [CrossRef]
  2. Milman, Y.V.; Miracle, D.B.; Chugunova, S.I.; Voskoboinik, I.V.; Podrezov, Y.N. Mechanical behaviour of Al3Ti intermetallic and l12 phases on its basis. Intermetallics 2001, 9, 839–845. [Google Scholar] [CrossRef]
  3. Foadian, F.; Soltanieh, M.; Adeli, M.; Etminanbakhsh, M.A. Study on the Formation of Intermetallics During the Heat Treatment of Explosively Welded Al-Ti Multilayers. Metall. Mater. Trans. A 2014, 45, 1823–1832. [Google Scholar] [CrossRef]
  4. Rohatgi, A.; Harach, D.J.; Vecchio, K.S.; Harvey, K.P. Resistance-curve and fracture behavior of Ti–Al3Ti metallic–intermetallic laminate (MIL) composites. Acta Mater. 2003, 51, 2933–2957. [Google Scholar] [CrossRef]
  5. Suzuki, A.; Kosugi, N.; Takata, N.; Kobashi, M. Microstructure and compressive properties of porous hybrid materials consisting of ductile Al/Ti and brittle Al_3Ti phases fabricated by reaction sintering with space holder. Mater. Sci. Eng. 2020, 776, 139000. [Google Scholar] [CrossRef]
  6. Yang, R.; Zhao, F.; Wang, L.Z.; Zhang, Z.Y.; Zhang, B.B. Fast fabrication of Al3Ti/Al composites using combustion synthesis and melt dilution. Mater. Lett. 2021, 298, 130036. [Google Scholar] [CrossRef]
  7. Peng, L.; Li, H.; Wang, J. Processing and mechanical behavior of laminated titanium–titanium tri-aluminide (Ti–Al3Ti) composites. Mat. Sci. Eng. A 2005, 406, 309–318. [Google Scholar] [CrossRef]
  8. Adharapurapu, R.R.; Vecchio, K.S.; Jiang, F.C.; Rohatgi, A. Fracture of Ti-Al3Ti metal-intermetallic laminate composites: Effects of lamination on resistance-curve behavior. Metall. Mater. Trans. A 2005, 36, 3217–3236. [Google Scholar] [CrossRef]
  9. Li, T.Z.; Grignon, F.; Benson, D.J.; Vecchio, K.S.; Olevsky, E.A.; Jiang, F.C.; Rohatgi, A.; Schwarz, R.B.; Meyers, M.A. Modeling the elastic properties and damage evolution in Ti–Al3Ti metal–intermetallic laminate (MIL) composites. Mat. Sci. Eng. A 2004, 374, 10–26. [Google Scholar] [CrossRef]
  10. Li, T.Z.; Jiang, F.C.; Olevsky, E.A.; Vecchio, K.S.; Meyers, M.A. Damage evolution in Ti6Al4V–Al3Ti metal-intermetallic laminate composites. Mat. Sci. Eng. A 2007, 443, 1–15. [Google Scholar] [CrossRef]
  11. Fan, X.L.; Yuan, M.N.; Qin, Q. Failure mechanisms of Ti-Al3Ti metal-intermetallic laminate composites under high-speed impact. Rare. Metal. Mat. Eng. A 2018, 47, 2615–2620. [Google Scholar]
  12. Pereira, J.M.; Revilock, D.M.; Ruggeri, C.R.; Emmerling, W.C.; Altobelli, D.J. Ballistic Impact Testing of Aluminum 2024 and Titanium 6Al-4V for Material Model Development. J. Aerospace Eng. 2014, 27, 56–465. [Google Scholar] [CrossRef] [Green Version]
  13. Vecchio, K.S. Synthetic multifunctional metallic-intermetallic laminate composites. JOM 2005, 57, 25–31. [Google Scholar] [CrossRef]
  14. Snipes, J.S.; Ramaswami, S.; Grujicic, M. Penetration resistance and ballistic-impact behavior of Ti/TiAl3 metal/intermetallic laminated composites (MILCs): A computational investigation. Aims. Mater. Sci. 2016, 3, 686–721. [Google Scholar]
  15. Zelepugin, S.A.; Zelepugin, A.S.; Tolkachev, V.F. Numerical Simulation of Multilayer Composites Failure under Dynamic Loading. Appl. Mech. Mater. 2015, 756, 408–413. [Google Scholar] [CrossRef]
  16. Zelepugin, S.A.; Mali, V.I.; Zelepugin, A.S.; Ilina, E. Failure of Metallic-Intermetallic Laminate Composites under Dynamic Loading. AIP Conf. Proc. 2012, 1426, 1101–1104. [Google Scholar]
  17. Harach, D.J. Processing, properties, and ballistic performance of titanium-aluminum titanium metal-intermetallic laminate (MIL) composites. Diss. Abstr. Int. 2000, 61, 4922. [Google Scholar]
  18. Adharapurapu, R.R.; Vecchio, K.S.; Jiang, F.C.; Rohatgi, A. Effects of ductile laminate thickness, volume fraction, and orientation on fatigue-crack propagation in Ti-Al3Ti metal-intermetallic laminate composites. Metall. Mater. Trans. A 2005, 36, 95–1608. [Google Scholar] [CrossRef]
  19. Cao, Y.; Guo, C.H.; Zhu, S.F.; Wei, N.X.; Javed, R.A.; Jiang, F.C. Fracture behavior of Ti/Al3Ti metal-intermetallic laminate (MIL) composite under dynamic loading. Mat. Sci. Eng. A 2015, 637, 235–242. [Google Scholar] [CrossRef]
  20. Zhou, P.J.; Guo, C.H.; Wang, E.H.; Wang, Z.M.; Chen, Y.; Jiang, F.C. Interface tensile and fracture behavior of the Ti/Al3Ti Metal-Intermetallic Laminate (MIL) composite under quasi-static and high strain rates. Mat. Sci. Eng. A 2016, 665, 66–75. [Google Scholar] [CrossRef]
  21. Yuan, M.N.; Yao, Y.H.; Han, F.Z.; Wang, Z.J. Evaluation of the compressive and anti-penetration properties of Ti-Al3Ti-Al laminated composites. Adv. Compos. Lett. 2020, 29, 1–5. [Google Scholar] [CrossRef]
  22. Price, R.D.; Jiang, F.C.; Kulin, R.M.; Vecchio, K.S. Effects of ductile phase volume fraction on the mechanical properties of Ti-Al3Ti metal-intermetallic laminate (MIL) composites. Mater. Sci. Eng. A 2011, 528, 3134–3146. [Google Scholar] [CrossRef] [Green Version]
  23. Du, Y.; Fan, G.H.; Yu, T.B.; Hansen, N.; Geng, L.; Huang, X.X. Laminated Ti-Al composites: Processing, structure and strength. Mater. Sci. Eng. A 2016, 673, 572–580. [Google Scholar] [CrossRef] [Green Version]
  24. Patselov, A.M.; Gladkovskii, S.V.; Lavrikov, R.D.; Kamantsev, I.S. Fracture toughness of Ti–Al3Ti–Al–Al3Ti laminate composites under static and cyclic loading conditions. Russ. Metall. 2015, 2015, 811–815. [Google Scholar] [CrossRef]
  25. Chen, W.H.; He, W.J.; Chen, Z.J.; Jiang, B.; Liu, Q. Extraordinary room temperature tensile ductility of laminated Ti/Al composite: Roles of anisotropy and strain rate sensitivity. Int. J. Plast. 2020, 133, 102806. [Google Scholar] [CrossRef]
  26. Harach, D.J.; Vecchio, K.S. Microstructure evolution in metal-intermetallic laminate (MIL) composites synthesized by reactive foil sintering in air. Metall. Mater. Trans. A 2001, 32, 1493–1505. [Google Scholar] [CrossRef]
  27. Wu, H.; Cui, X.P.; Geng, L.; Fan, G.H.; Pang, J.C.; Wei, L.S. Fabrication and characterization of in-situ TiAl matrix composite with controlled microlaminated architecture based on SiC/Al and Ti system. Intermetallics 2013, 43, 8–15. [Google Scholar] [CrossRef]
  28. Peng, L.M.; Wang, J.H.; Li, H.; Zhao, J.H.; He, L.H. Synthesis and microstructural characterization of Ti–Al3Ti metal–intermetallic laminate (MIL) composites. Scripta. Mater. 2005, 52, 243–248. [Google Scholar] [CrossRef]
  29. Fan, M.Y.; Domblesky, J.; Jin, K.; Qin, L.; Cui, S.Q.; Guo, X.Z.; Kim, N.; Tao, J. Effect of original layer thicknesses on the interface bonding and mechanical properties of TiAl laminate composites. Mater. Des. 2016, 99, 535–542. [Google Scholar] [CrossRef] [Green Version]
  30. Huang, M.; Xu, C.; Fan, G.H.; Maawad, E.; Gan, W.M.; Geng, L.; Lin, F.X.; Tang, G.Z.; Wu, H.; Du, Y.; et al. Role of layered structure in ductility improvement of layered Ti-Al metal composite. Acta Mater. 2018, 153, 235–249. [Google Scholar] [CrossRef]
  31. Yuan, M.N.; Li, L.L.; Wang, Z.J. Study of the microstructure modulation and phase formation of Ti-Al3Ti laminated composites. Vacuum. 2018, 157, 481–486. [Google Scholar] [CrossRef]
  32. Cao, Y.; Zhu, S.F.; Guo, C.H.; Vecchio, K.S.; Jiang, F.C. Numerical Investigation of the Ballistic Performance of Metal-Intermetallic Laminate Composites. Appl. Compos. Mater. 2015, 22, 437–456. [Google Scholar] [CrossRef]
  33. Senthil, K.; Arindam, B.; Iqbal, M.A.; Gupta, N.K. Ballistic Response of 2024 Aluminium Plates Against Blunt Nose Projectiles. Procedia Eng. 2017, 173, 363–368. [Google Scholar] [CrossRef]
  34. Holmquist, T.J.; Johnson, G.R. Modeling prestressed ceramic and its effect on ballistic performance. Int. J. Impact. Eng. 2005, 31, 113–127. [Google Scholar] [CrossRef]
  35. Tiamiyu, A.A.; Basu, R.; Odeshi, A.G.; Szpunar, J.A. Plastic deformation in relation to microstructure and texture evolution in AA 2017-T451 and AA 2624-T351 aluminum alloys under dynamic impact loadings. Mater. Sci. Eng. A 2015, 636, 379–388. [Google Scholar] [CrossRef]
  36. Xiong, Y.Y.; Li, N.; Jiang, H.W.; Li, Z.G.; Xu, Z.; Liu, L. Microstructural Evolutions of AA7055 Aluminum Alloy Under Dynamic and Quasi-static Compressions. Acta Metall. Sin. Engl. 2014, 27, 272–278. [Google Scholar] [CrossRef]
  37. Odeshi, A.G.; Adesola, A.O.; Badmos, A.Y. Failure of AA 6061 and 2099 aluminum alloys under dynamic shock loading. Eng. Fail. Anal. 2013, 35, 302–314. [Google Scholar] [CrossRef]
  38. Konieczny, M. Relations between Microstructure and Mechanical Properties in Laminated Ti-Intermetallic Composites Synthesized Using Ti and Al Foils. Key. Eng. Mater. 2013, 592–593, 728–731. [Google Scholar] [CrossRef]
Figure 1. Schematic of a Split Hopkinson pressure bar apparatus.
Figure 1. Schematic of a Split Hopkinson pressure bar apparatus.
Metals 11 01489 g001
Figure 2. (a) The finite element mesh SHPB experiment system (local), (b) Finite element model of Ti–Al3Ti–Al specimen, (c) Metallographic diagram of Ti–Al3Ti–Al specimen.
Figure 2. (a) The finite element mesh SHPB experiment system (local), (b) Finite element model of Ti–Al3Ti–Al specimen, (c) Metallographic diagram of Ti–Al3Ti–Al specimen.
Metals 11 01489 g002
Figure 3. Dynamic compressive stress–strain curves of Ti–Al3Ti–Al laminated composites with different Al contents.
Figure 3. Dynamic compressive stress–strain curves of Ti–Al3Ti–Al laminated composites with different Al contents.
Metals 11 01489 g003
Figure 4. The specimen after fracture and the fracture morphology observed under electron microscope: (a) Without Al, (b) containing about 15% Al, (c) containing about 30% Al.
Figure 4. The specimen after fracture and the fracture morphology observed under electron microscope: (a) Without Al, (b) containing about 15% Al, (c) containing about 30% Al.
Metals 11 01489 g004
Figure 5. The simulated dynamic stress-strain curves of Ti–Al3Ti–Al laminated composites with different Al contents.
Figure 5. The simulated dynamic stress-strain curves of Ti–Al3Ti–Al laminated composites with different Al contents.
Metals 11 01489 g005
Figure 6. Energy absorption–time curves of Ti–Al3Ti–Al laminated composites with different Al contents: (a) the whole; (b) Ti layers.
Figure 6. Energy absorption–time curves of Ti–Al3Ti–Al laminated composites with different Al contents: (a) the whole; (b) Ti layers.
Metals 11 01489 g006
Figure 7. The deformation degree of Ti layer in Ti–Al3Ti–Al laminated composites with different Al contents: (a) 0% Al, (b) 10% Al, (c) 20% Al, (d) 30%.
Figure 7. The deformation degree of Ti layer in Ti–Al3Ti–Al laminated composites with different Al contents: (a) 0% Al, (b) 10% Al, (c) 20% Al, (d) 30%.
Metals 11 01489 g007
Figure 8. Stress nephogram of longitudinal section of Ti–Al3Ti–Al laminated composites with different Al contents: (a) 0% Al, (b) 10% Al (c), 20% Al, (d) 30%.
Figure 8. Stress nephogram of longitudinal section of Ti–Al3Ti–Al laminated composites with different Al contents: (a) 0% Al, (b) 10% Al (c), 20% Al, (d) 30%.
Metals 11 01489 g008
Table 1. TC4 and Al alloy composition (mass fraction) %.
Table 1. TC4 and Al alloy composition (mass fraction) %.
AlloyTC4AlloyAl
Al5.50–6.75Si≤0.5
V3.5–4.5Fe≤0.5
Fe≤0.5Cu3.8–4.9
C≤0.1Mn0.30–0.9
O≤0.2Mg1.2–1.8
N≤0.05Ni≤0.1
Table 2. Size parameters of the target model containing different percentages of Al.
Table 2. Size parameters of the target model containing different percentages of Al.
-Thickness and Number of Al Alloy Layers (μm/n)Thickness and Number of TC4 Alloy Layers (μm/n)Thickness and Number of Al3Ti Alloy Layers (μm/n)Proportion of Al Content (%)
Model 10/048/25200/240
Model 225/2448/2587.5/4810
Model 350/2448/2575/4820
Model 475/2448/2562.5/4830
Table 3. Material parameters of TC4 and Al layers [32,33].
Table 3. Material parameters of TC4 and Al layers [32,33].
-Ρ (g/cm3)E (GPa)νA (GPa)B (GPa)CmnT melt (K)
Al2.71710.342654260.0151.10.93-
TC44.428113.80.342109810920.0141.10.931878
Table 4. Material parameters of Al3Ti layers [32].
Table 4. Material parameters of Al3Ti layers [32].
Ρ (g/cm3)E (GPa)νABCMNTPHEL (GPa)
3.352160.170.850.310.0130.210.290.21.842
D1D2K1 (GPa)K2 (GPa)K3 (GPa)-----
0.021.852.012.60-----
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ma, J.; Yuan, M.; Zheng, L.; Wei, Z.; Wang, K. Dynamic Mechanical Properties of Ti–Al3Ti–Al Laminated Composites: Experimental and Numerical Investigation. Metals 2021, 11, 1489. https://0-doi-org.brum.beds.ac.uk/10.3390/met11091489

AMA Style

Ma J, Yuan M, Zheng L, Wei Z, Wang K. Dynamic Mechanical Properties of Ti–Al3Ti–Al Laminated Composites: Experimental and Numerical Investigation. Metals. 2021; 11(9):1489. https://0-doi-org.brum.beds.ac.uk/10.3390/met11091489

Chicago/Turabian Style

Ma, Jian, Meini Yuan, Lirong Zheng, Zeyuan Wei, and Kai Wang. 2021. "Dynamic Mechanical Properties of Ti–Al3Ti–Al Laminated Composites: Experimental and Numerical Investigation" Metals 11, no. 9: 1489. https://0-doi-org.brum.beds.ac.uk/10.3390/met11091489

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