Materials and Design 88 (2015) 1082–1087
Contents lists available at ScienceDirect
Materials and Design journal homepage: www.elsevier.com/locate/jmad
Explosively welded multilayer Ni–Al composites I.A. Bataev a,⁎, T.S. Ogneva a, A.A. Bataev a, V.I. Mali a,b, M.A. Esikov a,b, D.V. Lazurenko a, Y. Guo c, A.M. Jorge Junior c,d a
Novosibirsk State Technical University, K. Marks 20, 630092 Novosibirsk, Russia Lavrentyev Institute of Hydrodynamics SB RAS, 15 Lavrentyev pr., 630090 Novosibirsk, Russia Department of Materials Engineering, Federal University of São Carlos, Via Washington Luiz, km 235, São Carlos 13565-905, SP, Brazil d LEPMI&SIMAP-CNRS, Institut Polytechnique de Grenoble, BP 75, St Martin d'Hères, 38402, France b c
a r t i c l e
i n f o
Article history: Received 22 July 2015 Received in revised form 15 September 2015 Accepted 16 September 2015 Available online 21 September 2015 Keywords: Explosion welding Composite Intermeltallics Quasicrystals
a b s t r a c t In this study, the structure of explosively welded Ni–Al multilayer composites was investigated. In particular, the interface between Ni and Al plates was studied using scanning electron microscopy (SEM), energy dispersive X-ray spectroscopy (EDX), and transmission electron microscopy (TEM). Continuous interlayers of mixed Al and Ni were found at the interface. The Al and Ni in these zones were heated above their melting temperatures, non-uniformly mixed, and rapidly solidified. Several intermetallic phases, including the decagonal phase and metastable Al9Ni2, were observed in these zones using electron diffraction. A cellular dislocation structure formed in the Ni plates and a polygonized dislocation structure formed in the Al plates due to the extremely high strain rate deformation and heating. Subsequent heat treatment at 620 °C led to the rapid formation of stable intermetallic layers at the interfaces. The growth of the intermetallic layers was considerably faster in the explosively welded composite than in the reference sample. © 2015 Elsevier Ltd. All rights reserved.
1. Introduction Explosion welding is one of the most common techniques for joining plates, sheets, bars and tubes of dissimilar metallic alloys [1–6]. This technique is typically used to produce bimetallic plates that resist corrosion and wear [5]. In explosive welding, the materials join because of the extremely high strain rate deformation of the thin surface layers when the materials collide at high speeds. It was initially assumed and has recently been proven using experiments and numerical simulations [7] that a cumulative jet appears between the colliding materials. This jet removes surface oxides and impurities so that the surfaces coming into contact are clean. Explosive welding is frequently regarded as solid-state joining [8]. However, numerous studies have reported the appearance of local molten areas at the interface between the welded materials (see, for instance, [9–12]). The mixing of dissimilar alloys in the molten areas may lead to the formation of intermetallic layers that influence the properties of the welded joints significantly. Although explosive welding has been known for a long time and is widely used in industry, relatively few papers accurately describe the structural changes that occur at the interface between the welded plates. In particular, the structure of the molten area has been poorly investigated. At the same time, the deformation at a high strain rate, elevated pressure at the contact point, and short duration of the
⁎ Corresponding author at: K. Marks 20, 630092 Novosibirsk, Russia. E-mail address:
[email protected] (I.A. Bataev).
http://dx.doi.org/10.1016/j.matdes.2015.09.103 0264-1275/© 2015 Elsevier Ltd. All rights reserved.
explosive welding process may result in unique and uncommon structures at the interface between welded materials. Nishida et al. [13,14] and Chiba et al. [15] observed the formation of amorphous and quasicrystalline phases in explosively welded Ni–Ti bimetals. In addition, they found amorphous structures in explosively welded Ti–steel bimetals. Amorphous phases were also observed at the interface between explosively welded Nb-foil and stainless steel plate by Bataev et al. [16]. The formation of amorphous and quasicrystalline structures in explosively welded materials is due to the very high cooling rate and the successful combination of the elements into an alloy. Because of their technological properties (good corrosion resistance, high strength, superior wear resistance, etc.), quasicrystals have attracted a great deal of attention from engineers. In comparison to crystalline and amorphous materials, one can say that quasicrystals belong to a third state of matter that exhibits quasiperiodic and orientational long-range order with the point group symmetry disallowed by crystallography [17]. Most of the quasicrystalline alloys are metastable phases formed by the rapid solidification (~ 106 K/s) of a melt and, after heat treatment, they become crystalline. Decagonal (D-phase) quasicrystals are two-dimensional and have a tenfold symmetry that combines a planar quasiperiodic atomic arrangement with a periodic order (ordinary translational periodicity) in the direction perpendicular to it. In contrast, icosahedral (I-phase) quasicrystals are threedimensional and have icosahedral point group symmetry (five-fold rotation axes). They are aperiodic in three dimensions [17]. The explosion-welding characteristics and the possibility of forming quasicrystals, which have interesting properties, in a certain range of
I.A. Bataev et al. / Materials and Design 88 (2015) 1082–1087
compositions, open the possibility of using explosion welding to fabricate materials with special properties. In this study, the formation of a quasicrystalline decagonal phase was observed at the interface of an explosively welded Al–Ni composite. The evolution of the dislocation structure of the metallic plates and the formation of metastable crystalline intermetallic structures during explosive welding were also reported. In addition, the layer-growth behaviors of explosively welded and cast samples after heat treatment were compared and discussed in light of the observed microstructures. 2. Materials and methods An Al–Ni composite was fabricated using four plates of commercially pure (cp-) Ni (50 × 25 × 1 mm) and three plates of cp-Al (50 × 25 × 0.5 mm). Ammonite 6GV with a detonation velocity of 4.2 km/s and a density of 0.9 g/cm3 was used as the explosive material. The welding scheme is shown in Fig. 1. After welding, the composite was studied by optical microscopy (OM) using a Carl Zeiss AxioObserver Z1m microscope, scanning electron microscopy (SEM) using a Carl Zeiss EVO 50 microscope coupled with an Oxford Instruments X-Act energy dispersive X-ray (EDX) spectroscope, and transmission electron microscopy (TEM) using a Tecnai G2 20 instrument coupled with an EDAX EDX system. EDX spectroscopy was used in the TEM and SEM investigations. The samples used in the OM and SEM investigations were crosssections of the composites that had been ground using SiC paper and polished using colloidal SiO2. The samples used in the TEM investigation were standard disks, 3 mm in diameter, that had been prepared using wire-cut electrical discharge machining. The disks were thinned using the following processes, in order: grinding using SiC paper, dimple grinding using Al2O3 powder in a Gatan dimple grinder (model 656), and ion-milling using a Gatan precision ion polishing system (model 691) at an energy of 5 KeV.
Fig. 1. A schematic representation of the welding experiments. Each of the gaps between two plates was 1 mm.
1083
Heat treatment was performed to produce metal-intermetallic laminated (MIL) composites and to evaluate the transformation of the structure caused by heat. The explosively welded samples were heated to 620 °C in an air atmosphere. The Al–Ni multilayer composites produced by casting molten Al between Ni plates as described in [18] were heat treated at the same time as the reference sample. The effect of explosive welding on the growth rate of the intermetallic layer was evaluated. The microhardness of each polished cross-section was measured using a Wolpert Group 402 MVD Vickers hardness tester. A four-sided diamond pyramid was used as an indenter. The load on the indenter was 10 g. 3. Results Cross-sections of the welded samples are shown in Figs. 2 and 3. It is noticeable that, due to plastic deformation, the final thicknesses of the Al and Ni plates were 0.43 and 0.89 mm, respectively, which caused increases in the length and width of the composite. Although in most of research on explosive welding no significant difference in the thicknesses of the plates before and after welding was observed, in this case, due to the low strength and high ductility of pure Ni and Al, the reduction in thickness was quite significant (approximately 14% for the Al and 11% for the Ni). Images captured at higher magnifications show continuous interlayers between the Al and Ni plates. The interlayers consist of both Ni and Al and are hereinafter referred to as mixing zones. SEM images show non-uniform distributions of Al and Ni in different mixing zones, which may have been caused by the high solidification rate (Fig. 3). Most of the regions in the mixing zones have dendritic morphologies with an average dendrite size of less than 2 μm (Fig. 3b). The EDX measurements of the mixing zones confirm the variations in the elemental composition. The elemental composition, measured in several locations at each interface (Fig. 3a) using an energy of 20 KeV, averages 36 at.% Ni and 64 at.% Al. This may be a result of the interaction volume of the beam, which includes penetration depths of approximately 10 μm for Al and 1 μm for Ni, and may occur because the melting point of Al is lower than that of Ni. However, in Fig. 3b and Table 1, in which the numbers correspond to points at which EDX analyses were performed at an energy of 3 KeV, the penetration depth was reduced to approximately 40 and 11 nm, for Al and Ni, respectively. As one can observe, the composition begins at the Al-rich side, passes through a region with a NiAl composition, and finishes on the Ni-rich side of the binary phase diagram, indicating that several compositions are possible and suggesting that the temperature was much higher than 1638 °C, which is the melting temperature of the NiAl phase. The results of the TEM investigation are shown in Fig. 4. Interfaces #2, 3, 4 and 5 (Fig. 4) were analyzed; however, no significant differences
Fig. 2. A multilayer Ni–Al composite produced by explosive welding.
1084
I.A. Bataev et al. / Materials and Design 88 (2015) 1082–1087
Fig. 3. A SEM image of interface #4 (shown in Fig. 2) showing non-uniform distribution of Al and Ni (a) and the dendritic morphology of the phases (b) in the mixing zones.
in their structures or phase compositions were observed. The Al and Ni areas located near the mixing zones have polygonal and cellular dislocation structures, respectively (Fig. 4 a,b). The mixing zone contains different types of intermetallic compound. Al3Ni, AlNi, Al9Ni2 and the decagonal phase (D-phase) were found using the electron diffraction technique (Fig. 5). Note that the last two phases are metastable and had previously been observed in samples produced using twopiston splat cooling [19,20]. The results of the EDX investigations show that the average composition of the decagonal phase is 39 at.% Ni and 61 at.% Al. This suggests that the range of compositions that allows the D-phase to form is even larger than that observed elsewhere [21] and may be justified at very high cooling rates. The microhardness profile across the section of the composite after the explosive welding is shown in Fig. 6. It was noted that the mixing zones had much higher hardness (from 450 to 700 HV) in comparison with the initial materials (100 and 30 HV for Ni and Al respectively), which confirmed the formation of hard intermetallic phases at the interface. The hardness of Ni and Al plates also increased up to 190 and 50 HV, respectively, after the welding. In the subsequent sections, the formation of the structures shown in Figs. 3–5 is discussed. Because the structures of the mixing zones result from the solidification process and those of the adjacent areas result from a solid-state process, the formation processes of these structures are treated separately. One of the possible applications of explosively welded metallic multilayer composites is their subsequent transformation into metalintermetallic laminated (MIL) composites [22]. This transformation typically occurs as a result of a reactive sintering process at elevated temperatures [23]. Fig. 7a and b compare the microstructures of growth layers after the annealing of explosively welded and cast composites, and Fig. 7c compares their growth over time; it clearly shows the linear growth of the cast samples and the logarithmic growth of the explosively welded samples. Heating the composite to 620 °C results in the formation of two intermetallic layers with different compositions between the Ni and Al layers (Fig. 7). An EDX analysis revealed that the intermetallic phase
Table 1 The chemical composition at each point marked in Fig. 3b. Point #
Ni (at.%)
Al (at.%)
1 2 3 4 5 6 7 8
26 69 68 44 45 44 47 40
74 31 32 56 55 56 53 60
located closer to the Ni plate is composed of Ni2Al3, and the intermetallic phase located closer to the Al plate is composed of NiAl3, which is in agreement with previously published results [24,25]. The microhardnesses of the Ni and Al layers of the explosively welded multilayer composites decreased significantly after annealing (Fig. 8), which suggests that the dislocation structure was rearranged. At the same time, the Ni2Al3 interlayer was harder than the intermetallic layers that appeared in the mixing zones after explosive welding. 4. Discussion 4.1. The formation of structures in the mixing zones Based on the results of the TEM and SEM investigations, it is obvious that the formation of structures in the mixing zones is associated with rapid solidification from the liquid phase. The dendrites in Fig. 3b are the first symptom of this. These micron-sized dendrites are mostly equiaxed, which indicates that very high constitutional supercooling occurred. The smallest dendrites result from reduced branching, which, in turn, is caused by increased undercooling. The prerequisite for the formation of such dendrites is that the molten material (liquid) be supercooled, or undercooled, below the freezing point ahead of the solid–liquid interface. The process begins with the growth of a spherical solid nucleus in the undercooled melt (Fig. 4c). As the sphere grows, its morphology is destabilized, and its spherical shape becomes distorted. Therefore, the solid shape begins to show preferred growth directions within the crystal. In addition, and more importantly, because the primary arms of each dendrite are approximately 300 nm and the secondary arms are approximately 79 nm, the spacing between the secondary arms, if there is any, may be extremely small. Using an empirical relationship to describe the spacing of the secondary arms, Anantharaman and Suryanarayana [26] created graphs for Al alloys that correlate the spacing with the cooling rate. For the minimum value that they found (50 nm), the cooling rate was greater than 109 K/s. In addition, the structures shown in Fig. 5 are, in many respects, similar to those produced by the two-piston splat cooling of an Ni–Al melt found by Pohla and Ryder [19]. However, two-piston splat cooling produces samples with an apparent layered structure with layers in a certain location containing Al9Ni2 and D-phases; however, an irregular distribution of these phases throughout the mixing zones was observed in the explosively welded samples. This has several possible causes. First, the material in the mixing zones is turbulently mixed during the welding process [16]. Typically, the welding time for small samples is measured in microseconds. Therefore, homogenization of the turbulently mixed areas cannot be completed. The structure is also influenced by the high pressure and the shock waves. According to numerical simulations [7,27,28,29], the pressure at the contact point exceeds 10 GPa. However, in actual experiments, this value may change from point to
I.A. Bataev et al. / Materials and Design 88 (2015) 1082–1087
1085
Fig. 4. The structures of the welded materials near the interface: a — the cellular dislocation structure of Ni; b — the polygonized dislocation structure of Al; c — the typical solidification structure in the mixing zones. These images were captured at interface #4 (shown in Fig. 2).
Fig. 5. Images and the corresponding electron diffraction patterns of the intermetallic phases formed during solidification in the mixing zones: a — AlNi; b — Al9Ni2; c — Al3Ni; d — decagonal phase. These images were captured at interface # 4 (shown in Fig. 2).
1086
I.A. Bataev et al. / Materials and Design 88 (2015) 1082–1087
Fig. 6. The distribution of the microhardness of the welded composite.
point due to variations in the explosives properties and complex interactions between shock waves and expansion waves. Therefore, neighboring areas can have different solidification histories, leading to a complex arrangement of phases after solidification. The formation of metastable phases such as Al9Ni2 and, remarkably, the decagonal phase, is very important and of scientific and technological interest; these phases represent markers that describe the solidification history. Note that quasicrystals typically appear in alloys with specific e/a, and their formation is typically observed in experiments involving rapid solidification of those alloys. Athanasiou proposed the following semi-empirical criteria for the formation of quasicrystals in binary Al-TM (TM = transition metal) alloys [30]: (i) The cooling rate must be 106 K/s, (ii) the ratio of the atomic radii, R, of the constituents must satisfy 1.04 b R b 1.16, and (iii) the concentration of the transition metal must be 14–25 at.%. Specifically, Shuleshova et al. [21] have reported a variety of compositions for the D-phase in Al100–xNix alloys. From the above discussion regarding dendrites and the precipitation of special second phases, one can say that these features show that the cooling rate during solidification was extremely high and could be comparable with that of two-piston splat cooling or melt spinning. This conclusion is in agreement with the results of numerical simulations, which show that the cooling rate during explosive welding is on the order of 108 K/s [31]. Therefore, the structures that arise in the mixing zones of explosively welded composites, such as the ones described above, can be predicted by analyzing the large amount of data obtained from experiments on rapidly solidified melts.
Fig. 8. The microhardness distribution in an explosively welded composite that was heat treated for 3 h at 620 °C.
From a technological perspective, metastable phases affect the mechanical properties of welded bimetals and multilayer composites as well as features of the structural transformations that occur during subsequent heat treatment, which will be discussed in Section 4.3. 4.2. The structures of the Al and Ni near the interface The formation of a structure near the interface (Fig. 4a and b) occurs with hot severe plastic deformation and can be described fairly well using classical approaches to plastic deformation with a subsequent rearrangement of the dislocation structure. Usually, significant plastic deformation results in the multiplication of dislocations and the formation of a cellular structure similar to the one shown in Fig. 4a for Ni. In addition, Meyers observed the formation of a cellular dislocation structure in shock-loaded samples of different materials and proposed a mechanism to explain this phenomenon [32]. Therefore, the formation of this structure near the interface of two explosively welded plates seems to be reasonable. The subsequent flow of heat through this type of structure leads to the rearrangement of dislocations and the formation of dislocation walls, as shown in Fig. 4b for Al. The evolution of polygonization depends significantly on the melting temperature and the stacking fault energy. Dislocation climbing and cross-slipping, which are necessary for polygonization, occur relatively easily in aluminum because of its low melting temperature (660 °C) and high stacking fault energy (166 mJ/m2 [33]). In contrast, Ni has a relatively high melting temperature (1455 °C) and a moderate
Fig. 7. A comparison of the intermetallic layers thicknesses in explosively welded (a) and cast (b) multilayer composites after annealing at 620 °C for 3 h and (c) a graph of the intermetallic layers thickness versus time.
I.A. Bataev et al. / Materials and Design 88 (2015) 1082–1087
stacking fault energy (125 mJ/m2). Therefore, in this case, polygonization is decelerated and the cellular structure is retained. Therefore, the impacts of time and temperature in explosive welding enable polygonization in Al but not in Ni. The formation of dislocations explains the hardening of Ni and Al after the welding process. 4.3. The influence of heat treatment on structural transformations of explosively welded composites Fig. 7c shows how the thickness of the intermetallic layer varies with the heating time in explosively welded and cast Ni–Al composites. For shorter heating times, the intermetallic growth rate is significantly higher for the explosively welded samples than it is for the cast samples. One possible explanation for this is that the growth of the intermetallic layer during the sintering reaction typically begins relatively slowly due to the presence of oxide films and impurities at the interface. Once the films have been broken, the growth process accelerates significantly [23]. Because explosive welding removes oxides and impurities from the surfaces [34], the growth of intermetallic layers is significantly faster. In addition, the mixing zones at the interfaces can be considered sites at which the nucleation of intermetallic compounds is accelerated because the preliminary mixing of materials at these sites significantly reduces the amount of diffusion necessary for this type of phase transition. As the temperature increases, metastable Al9Ni2 and the D-phase are easily transformed into more stable intermetallic phases, which have a positive effect on the reaction rate. Because dislocation lines are typically considered high-diffusivity paths, the increased dislocation density near the Al–Ni interface after welding may also contribute positively to the growth of the intermetallic layer. 5. Conclusion The interfaces between layers in explosively welded Ni–Al multilayer composites contain unique metastable phases such as Al9Ni2 and D-phase. The structures in the mixing zones have features in common with the structures produced by the rapid solidification of a melt. Therefore, the formation of quasicrystalline and other metastable phases in explosively welded materials is very possible. Note, however, that the process of explosive welding is much more complex due to the heterogeneous distribution of chemical elements over the volume of the melt and the effect of pressure variations. The structures of the unmelted zones adjacent to the interface are caused by the high strain rate deformation and the simultaneous influence of the temperature. The formation of quasicrystalline phases, which are typically exceptionally brittle, may explain the difficulties that arise during the explosive welding of some combinations of alloys. However, metastable structures at the interface significantly accelerate the formation of stable intermetallics, which can be used in the fabrication of MIL composites. Acknowledgments This work was financially supported by the Ministry of Education and Science of the Russian Federation, state task No 11.1892.2014/K Project Code 1892. References [1] B. Wang, X. Luo, B. Wang, S. Zhao, F. Xie, Microstructure and its formation mechanism in the interface of ti/nicr explosive cladding bar, J. Mater. Eng. Perform. 24 (2014) 1050–1058. [2] M. Ghosh, S. Chatterjee, Characterization of transition joints of commercially pure titanium to 304 stainless steel, Mater. Charact. 48 (2002) 393–399. [3] W. Deng, M. Lu, Q. Xu, Effect of Detonation Velocity on Interface and Properties of Al/Ti Composite Tube Under Explosive Welding, Hanjie Xuebao/Transactions of the China Welding Institution 35, 2014 39–42.
1087
[4] X. Guo, J. Tao, Z. Yuan, L. Zhang, X. Sun, Interface and properties of explosive welded ta1/al clad tube, Rare Met. Mater. Eng. 41 (2012) 139–142. [5] J. Banker, Industrial application of explosion clad (review), Paton Weld. J. (2009) 42–45. [6] S. Mroz, G. Stradomski, H. Dyja, A. Galka, Using the explosive cladding method for production of mg–al bimetallic bars, Arch. Civil Mech. Eng. 15 (2015) 317–323. [7] S.P. Kiselev, V.I. Mali, Numerical and experimental modeling of jet formation during a high-velocity oblique impact of metal plates, Combust. Explos. Shock Waves 48 (2012) 214–225. [8] K. Ferjutz, J.R. Davis, ASM Handbook: volume 6: welding, brazing, and soldering, Asm Handbook, ASM International, Materials Park, OH, 1993. [9] B. Crossland, Explosive welding of metals and its application, Oxf. Ser. Adv. Manuf. 2 (1982). [10] P. Manikandan, K. Hokamoto, A.A. Deribas, K. Raghukandan, R. Tomoshige, Explosive welding of titanium/stainless steel by controlling energetic conditions, Mater. Trans. 47 (2006) 2049–2055. [11] B.A. Greenberg, M.A. Ivanov, A.V. Inozemtsev, A.M. Patselov, M.S. Pushkin, A.M. Vlasova, Microheterogeneous structure of local melted zones in the process of explosive welding, Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 46 (2015) 3569–3580. [12] R. Mendes, J.B. Ribeiro, A. Loureiro, Effect of explosive characteristics on the explosive welding of stainless steel to carbon steel in cylindrical configuration, Mater. Des. 51 (2013) 182–192. [13] M. Nishida, A. Chiba, Y. Morizono, M. Matsumoto, T. Murakami, A. Inoue, Formation of nonequilibrium phases at collision interface in an explosively welded ti/ni clad, Mater. Trans. JIM 36 (1995) 1338–1343. [14] M. Nishida, A. Chiba, Y. Honda, J.-i. Hirazumi, K. Horikiri, Electron microscopy studies of bonding interface in explosively welded ti/steel clads, ISIJ Int. 35 (1995) 217–219. [15] A. Chiba, M. Nishida, Y. Morizono, Microstructure of bonding interface in explosively-welded clads and bonding mechanism, Mater. Sci. Forum 465 (2004) 465–474. [16] I. Bataev, K. Hokamoto, H. Keno, A. Bataev, I.A. Balaganskii, A.V. Vinogradov, Metallic glass formation at the interface of explosively welded nb and stainless steel, Met. Mater. Int. 21 (2015) 713–718. [17] L. Bendersky, Quasicrystal with one-dimensional translational symmetry and a tenfold rotation axis, Phys. Rev. Lett. 55 (1985) 1461–1463. [18] T.S. Sameyshcheva, I.A. Bataev, A.A. Bataev, P.S. Yartsev, I.A. Polyakov, Metallic — intermetallic composites produced by vacuum casting and annealing of ni and al, Proceedings of the 7th International Forum on Strategic Technology 2012, pp. 275–278. [19] C. Pohla, P.L. Ryder, Crystalline and quasicrystalline phases in rapidly solidified Al–Ni alloys, Acta Mater. 45 (1997) 2155–2166. [20] X.Z. Li, K.H. Kuo, Decagonal quasicrystals with different periodicities along the tenfold axis in rapidly solidified Al–Ni alloys, Philos. Mag. Lett. 58 (1988) 167–171. [21] O. Shuleshova, D. Holland-Moritz, W. Lser, G. Reinhart, G.N. Iles, B. Bchner, Metastable Formation of Decagonal Quasicrystals During Solidification of Undercooled Al– Ni Melts: In Situ Observations by Synchrotron Radiation, EPL 86, 2009. [22] I.A. Bataev, A.A. Bataev, V.I. Mali, D.V. Pavliukova, Structural and mechanical properties of metallic–intermetallic laminate composites produced by explosive welding and annealing, Mater. Des. 35 (2012) 225–234. [23] D.J. Harach, K.S. Vecchio, Microstructure evolution in metal–intermetallic laminate (mil) composites synthesized by reactive foil sintering in air, Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 32 (2001) 1493–1505. [24] M. Konieczny, R. Mola, P. Thomas, M. Kopcial, Processing, microstructure and properties of laminated ni-intermetallic composites synthesised using ni sheets and al foils, Arch. Metall. Mater. 56 (2011) 693–702. [25] G.A. López, S. Sommadossi, W. Gust, E.J. Mittemeijer, P. Zieba, Phase characterization of diffusion soldered Ni/Al/Ni interconnections, Interface Sci. 10 (2002) 13–19. [26] T. Anantharaman, C. Suryanarayana, Rapidly Solidified Metals: A Technological Overview, Trans Tech Publications, 1987. [27] X. Wang, Z. Y., H. Liu, Z. Shen, Y. Hu, W. Li, Numerical study of the mechanism of explosive/impact welding using smoothed particle hydrodynamics method, Mater. Des. 35 (2012) 210–219. [28] X.J. Li, F. Mo, X.H. Wang, B. Wang, K.X. Liu, Numerical study on mechanism of explosive welding, Sci. Technol. Weld. Join. 17 (2012) 36–41. [29] A.A.A. Mousavi, S.T.S. Al-Hassani, Numerical and experimental studies of the mechanism of the wavy interface formations in explosive/impact welding, J. Mech. Phys. Solids 53 (2005) 2501–2528. [30] N.S. Athanasiou, Formation, characterization and magnetic properties of some ternary Al–Cu-m (m = transition metal) quasicrystals prepared by conventional solidification, Int. J. Mod. Phys. B 11 (1997) 2443–2464. [31] W.D. Liu, K.X. Liu, Q.Y. Chen, J.T. Wang, H.H. Yan, X.J. Li, Metallic glass coating on metals plate by adjusted explosive welding technique, Appl. Surf. Sci. 255 (2009) 9343–9347. [32] M. Meyers, L. Murr, Shock Waves and High-Strain-Rate Phenomena in Metals, Plenum Press, New York, 1981. [33] J. Hirth, J. Lothe, Theory of Dislocations, Wiley, New York, 1982. [34] A. Deribas, Physics of Explosive Hardening and Welding, Nauka, Novosibirsk, 1980.