This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Brought to you by:
Paper

Interfacial effect on deformation and failure of Al/Cu nanolaminates under shear loading

, , , , , and

Published 26 July 2018 © 2018 IOP Publishing Ltd
, , Citation Chao Lv et al 2018 J. Phys. D: Appl. Phys. 51 335301 DOI 10.1088/1361-6463/aad2a8

0022-3727/51/33/335301

Abstract

The interface effects on deformation and failure of Al/Cu nanolaminates under shear loading are investigated by molecular dynamics and analytical methods, including interface orientations and repeat layer spacing. Interface orientations play a dominant role in dislocation evolution over repeat layer spacing. Interfacial stress affects nucleation and propagation of dislocations in nanolaminates, and is modeled with an analytical form. The yield and failure of Al/Cu nanolaminates are mainly controlled by dislocation evolutions in Al and Cu layers, respectively. In particular, when repeat layer spacing is less than 12 nm, the shear strength of Al(1 0 0)/Cu(1 0 0) nanolaminates decreases with decreasing repeat layer spacing due to the interaction between stacking faults and dislocations. For better shear performance, the minimum repeat layer spacing of Al/Cu nanolaminates is about 12 nm, and the interface orientations can be tailored.

Export citation and abstract BibTeX RIS

1. Introduction

In recent years, metal nanolaminates have been of sustained interest due to their ultrahigh hardness and excellent resistance to radiation damage [1, 2]. Experimental and theoretical results demonstrate that interlayer interfaces play an important role in their physical properties, such as magnetic properties [3], strength [4], work hardening and deformation evolution [5] under various loading conditions.

Semi- and incoherent interlayer interfaces serve as nucleation sites of dislocations under in-plane and out-plane loading [610], while it is difficult for dislocation nucleation at perfect coherent interlayer interfaces [11, 12]. The interlayer interface also acts as a strong barrier to slip transmission due to core spreading of glide dislocations within interfaces [1315]. In addition, the mechanical properties of metal nanolaminates also depend on the repeat layer spacing (λ) [1620], and can be described with the dislocation pile-up-based Hall–Petch model for $\lambda> 100$ nm, single-dislocation-based confined layer slip model (CSL) for $2 < \lambda <100$ nm, and the interface crossing of single dislocations when $1 < \lambda <2$ nm, respectively. For instance, Li et al studied Al(1 1 1)/Cu(1 1 1) nanolaminates under out-of-plane tension, and found that there was an optimal λ range for achieving high tensile strength, which was associated with the density of stacking fault tetrahedra [21]. Gang et al examined the coherent and semi-coherent interfaces of Cu(1 1 1)/Ni(1 1 1) nanolaminates under in-plane tension, and found that the ability for semi-coherent interface to hinder dislocation slide was dependent on λ [22]. Moreover, it is found that the interfaces highly influence the mechanical behavior of Al and Al alloy [23]. Kuksin et al found that the binary Al–Cu system affected the dislocation evolution under shock loading [24, 25], which is also demonstrated by Meng et al [26]. Tiwary et al studied the effect of lamellar thickness on the mechanical of Al–Cu eutectic structure [27]. Verestk et al found that the critical resolved shear stress is proportional to the copper concentration [28]. The work of Ma et al showed that the interface played an important role in the nanoindentation hardness of Cu–Zr–Al/Cu nanolaminates [29].

However, deformation of metal nanolaminates under shear has rarely been investigated, although the shear response of nanomaterials is interest for their engineering applications [30, 31] as a primary deformation mode of structural materials. In this work, the mechanical response of Al/Cu nanolaminates at room temperature under shear loading is investigated through molecular dynamics (MD) and analytical methods. The effects of repeat layer spacing and orientations of constituent crystals at interfaces (or simply, interfacial orientations) on deformation of Al/Cu nanolaminates are explored, and the underlying mechanisms of the dislocation-interface and dislocation–dislocation interactions are revealed. We find that interfacial orientations play a key role in dislocation evolution, which affects yield and failure of Al/Cu nanolaminates. Nonetheless, the effect of repeat layer spacing can be neglected. This indicates that shear response of Al/Cu nanolaminates can be mainly controlled by tailoring the interfacial orientations. The mechanisms of shear deformation are also useful for the design and application of composites with similar interfacial structures.

2. Models and methods

An atomic configuration of Al/Cu nanolaminate is illustrated in figure 1(a) with alternating Al and Cu layers along the z-axis. The repeat layer spacings (λ) of Al and Cu are identical, as in accumulative roll bonding [32]. Since the nanotwins have an important influence on the plastic deformation of FCC metal [3335], we build the models without nanotwins to obtain clear mechanism of Al/Cu interfacial effects. Shear strain is calculated as $\gamma =\delta_x/L_y$ , where $\delta_x$ is the displacement along the x-axis and Ly is the length of simulation box along the y-axis (figure 1(b)). The failure shear stress (FSS), which is the critical shear stress where structural failure occurs under shear loading, is highly influenced by crystal orientation for pure face-centered cubic metal. For Al at 300 K, the maximum FSS is 3.4 GPa on {1 0 0}, while the minimum FSS is 1.8 GPa on {1 1 1} [36, 37]. For Cu at 300 K, the maximum and minimum FSS are 9.0 GPa on {1 0 0} and 4.9 GPa on {1 1 1}, respectively [37, 38]. To reveal the interfacial orientation effect on shear response of Al/Cu nanolaminates, four typical interfacial orientations are investigated in this work (figure 1(c)), denoted as Al(1 0 0)/Cu(1 0 0), Al(1 1 1)/Cu(1 1 1), Al(1 0 0)/Cu(1 1 1), and Al(1 1 1)/Cu(1 0 0) nanolaminates. There is a lattice mismatch of 12% between Al (lattice parameter a  =  0.405 nm) and Cu (a  =  0.361 nm). Thus, the layer interfaces of Al(1 0 0)/Cu(1 0 0) and Al(1 1 1)/Cu(1 1 1) nanolaminates are semi-coherent, and those of Al(1 0 0)/Cu(1 1 1) and Al(1 1 1)/Cu(1 0 0) nanolaminates are incoherent [39]. In addition, the MD results of Cao et al [39] show that the interfacial coherence is not sensitive to temperature if the temperature is below the melting point of Cu and Al. To eliminate initial dislocations or stress in nanolaminates, the in-plane dimensions of a simulation box (along the x- and y-axes) satisfy the following relations [40]:

Equation (1)

where Mi=x,y and Ni=x,y are integers, Li=x,y is the length of a simulation box along the i-axis, $a_{\rm Al}^{i=x, y}$ and $a_{\rm Cu}^{i=x, y}$ are the length of Al and Cu unit cell along the i-axis, respectively. In this work, the in-plane dimensions of Al(1 0 0)/Cu(1 0 0), Al(1 1 1)/Cu(1 1 1), Al(1 0 0)/Cu(1 1 1) and Al(1 1 1)/Cu(1 0 0) nanolaminates are 202.50 Å  ×  101.25 Å, 214.78 Å  ×  124.00 Å, 255.15 Å  ×  141.75 Å, and 220.51 Å  ×  133.92 Å, respectively. The data with error bars in the following discussion are statistically averaged results of five simulations per case.

Figure 1.

Figure 1. (a) Schematic of Al/Cu nanolaminates; (b) schematic of shear loading, applied in the xy-plane; (c) four typical interface types with different crystallographic orientations for Al and Cu crystals.

Standard image High-resolution image

MD simulations are performed with the LAMMPS package [41]. The Cai and Ye embedded atom method (EAM) potentials [42] are used to describe the atomic interactions among Al and Cu atoms. This potential has been verified in previous research, and the simulations of Al–Cu interfaces and Al–Cu alloy are in agreement with experimental results [27, 43, 44]. The accuracy of the potentials was also demonstrated by previous simulations [39, 45, 46]. Prior to shear loading, the Al/Cu nanolaminate configurations are relaxed with the isothermal-isobaric ensemble via a Nosé–Hoover thermostat at 300 K for 100 ps [47, 48]. The time step for integrating the equation of motion is 1 fs. Actually, it has been demonstrated that the process of dislocation propagation is independent of the shear strain rate at time scales sufficiently short to neglect creep and yet sufficiently long with respect to the sound speed of sound for strain rates below $1.5 \times10^9$ s−1 [38, 49]. Thus, shear loading is applied with the constant volume-temperature ensemble [50] at a strain rate of 109 s−1 and 300 K. Periodic boundaries conditions are applied along the x-, y- and z-axes in all simulations. Visualization and structural analysis are performed with OVITO [51].

3. Results and discussions

3.1. Typical shear response of Al/Cu nanolaminates with different interface orientations

To investigate the response of Al/Cu nanolaminates to shear loading, four typical stress–strain curves with different interface orientations are analyzed here, and the repeat layer spacing λ is 5 nm. On the basis of the inflection points on the shear stress–strain ($\tau-\gamma$ ) curves as well as dislocation evolutions, different deformation stages can be identified for each case of in-plane shear loading (figure 2).

Figure 2.

Figure 2. Stress–strain curves of (a) Al(1 0 0)/Cu(1 0 0) and (b) Al(1 1 1)/Cu(1 0 0) nanolaminates with $\lambda = 5$ nm under shear loading. Insets: atomic configurations at representative strains. ($\gamma_{\rm Y}, \ \tau_{\rm Y}$ ) and ($\gamma_{\rm F}, \ \tau_{\rm F}$ ) are the yield and failure point, respectively.

Standard image High-resolution image

For the Al(1 0 0)/Cu(1 0 0) and Al(1 1 1)/Cu(1 0 0) nanolaminates (figure 2), these stages are (I) elastic, (II) yield and (III) failure stages. In stage I (γ is below the yield strain $\gamma_{\rm Y}$ ), τ increases with γ. Since no dislocations are generated in Al or Cu layers as well as at the Al/Cu interfaces, $\tau= G^{\rm e}\gamma$ . Here $G^{\rm e}$ is the effective elastic shear modulus of the nanolaminates. The yield strain $\gamma_{\rm Y}$ for the Al(1 0 0)/Cu(1 0 0) and Al(1 1 1)/Cu(1 0 0) nanolaminates are 0.038 and 0.075, respectively. Since the shear strength of Cu(1 0 0) is higher than that of Al(1 0 0) and Al(1 1 1) [37], in-layer dislocations are generated in the Al layers, leading to the first inflection point on the $\gamma-\tau$ curves between stages I and II. This phenomenon is also consistent with the results of metal nanolaminates under tensile loading [22]. $\gamma_{\rm Y}$ and failure stress $\tau_{\rm F}$ at stage II are different for the Al(1 0 0)/Cu(1 0 0) and Al(1 1 1)/Cu(1 0 0) nanolaminates; this will be discussed in the following dislocation analysis. With the increase in γ, τ increases to $\tau_{\rm F}$ , which corresponds to the second inflection point on the $\gamma-\tau$ curves between stages II and III. The failure strains for Al(1 0 0)/Cu(1 0 0) and Al(1 1 1)/Cu(1 0 0) nanolaminates are $\gamma_{\rm F}=0.085$ and 0.095, respectively. Meanwhile, there is a remarkable difference in $\tau_{\rm F}$ between these two types of nanolaminates (see more discussion in dislocation analysis). At stage III, the nanolaminates undergo plastic flow with $\tau\sim1.0$ GPa, and the slip bands appear in the Cu layers. The τ-values of these nanolaminates in plastic flow are close to that of pure Cu and higher than that of pure Al [37].

Figure 3.

Figure 3. Stress–strain curves of (a) Al(1 0 0)/Cu(1 1 1) and (b) Al(1 1 1)/Cu(1 1 1) nanolaminates with $\lambda = 5$ nm under shear loading. Insets: atomic configurations at representative strains. ($\gamma_{\rm Y}, \ \tau_{\rm Y}$ ) is the failure point.

Standard image High-resolution image

Compared with the Cu(1 0 0)-stacking nanolaminates (figure 2), there are only two stages (the yield stage is absent) for Al(1 0 0)/Cu(1 1 1) and Al(1 1 1)/Cu(1 1 1) nanolaminates (figure 3). Similar to the Cu(1 0 0)-stacking nanolaminates, stage I also represents elastic deformation. The difference in failure strain $\gamma_{\rm F}$ is minor between Al(1 0 0)/Cu(1 1 1) and Al(1 1 1)/Cu(1 1 1) nanolaminates, but drastic in failure stress $\tau_{\rm F}$ : the $\gamma_{\rm F}$ -values for Al(1 0 0)/Cu(1 1 1) and Al(1 1 1)/Cu(1 1 1) are 0.09 and 0.085, and the $\tau_F$ -values are 2.5 GPa and 1.9 GPa, respectively. Especially, there is approximately a stage of periodic evolution for the Al(1 1 1)/Cu(1 1 1) nanolaminates, a common phenomenon for FCC metal under shear loading in the (1 1 1) plane [38].

3.2. Effect of repeat layer spacing

In general, the repeat layer spacing λ, as well as interface orientations, affects the mechanical properties of metallic nanolaminates [14, 22]. Thus, it is necessary to study the effect of λ on yield and failure of Al/Cu nanolaminates. Here, λ is varied from 4 to 100 nm, a common range for metallic nanolaminates [1620].

According to the stress superposition principle for composites with interfaces [38], the effective shear stress in nanolaminates $\tau^{\rm e}$ is given as:

Equation (2)

where $\tau^{\rm Al}$ and $\tau^{\rm Cu}$ denote shear stresses in Al and Cu layers, $ \newcommand{\e}{{\rm e}} \eta^{\rm Al} = \eta^{\rm Cu} = 0.5$ are the volume fractions of Al and Cu layers, and $\tau^{\rm i}$ is the shear stress at Al/Cu interfaces. Equation (2) can be rewritten in terms of λ as

Equation (3)

Here, the $\overline{\tau}^{\rm Al}$ and $\overline{\tau}^{\rm Cu}$ are the shear stresses when Al and Cu layers are in the plastic flow stage, and $\gamma^{\rm Al}_{\rm F}$ and $\gamma^{\rm Cu}_{\rm F}$ are the failure strains of Al and Cu layers, respectively. In general, there are three stages of deformation for Al/Cu nanolaminates. When $\gamma < \gamma^{\rm Al}_{\rm F}$ , the $\tau-\gamma$ curve is elastic for both Al and Cu layers. When $\gamma^{\rm Al}_{\rm F} < \gamma < \gamma^{\rm Cu}_{\rm F}$ , plastic deformation occurs in Al layers but not in Cu layers. In this semi-plastic situation, the stress in Al layers in plastic flow becomes $\overline{\tau}^{\rm Al}$ , while Cu layers are still in the elastic stage. When $\gamma > \gamma^{\rm Cu}_{\rm F}$ , both Al and Cu layers undergo plastic flow. The critical stresses and strains for pure Al and Cu [37, 38] are listed in table 1 for comparison. For shearing of FCC metals in (1 1 1) plane, the flow stress periodically increases and decreases due to recovery and generation of slip bands [3638]. Thus, the $\tau_{\rm F}$ and $\overline{\tau}$ for (1 1 1) shearing are identical in table 1. The $\gamma_{\rm F}$ -values of Al and Cu layers are reduced due to the Al/Cu interfaces compared to their counterparts for pure Al and Cu.

Table 1. Critical stress and strain for pure Al and Cu under shear loading.

Metal Shear plane Failure strain ($\gamma_{\rm F}$ ) Failure stress ($\tau_{\rm F}$ )/GPa Stress in plasticity flow ($\overline{\tau}$ )/GPa
Al (1 0 0) 0.128 3.25 0.50
Al (1 1 1) 0.095 1.79 1.79
Cu (1 0 0) 0.142 8.91 1.00
Cu (1 1 1) 0.106 2.54 2.54

However, as shown in section 3.1, the semi-plastic stage ($\gamma^{\rm Al}_{\rm F} <\gamma < \gamma^{\rm Cu}_{\rm F}$ ) only exists in the Cu(1 0 0)-stacking nanolaminates. Although the $\tau_{\rm F}$ -value of Cu(1 1 1) is higher than those of Al(1 0 0) and Al(1 1 1), the semi-plastic stage is absent for the Cu(1 1 1)-stacking nanolaminates. Thus, the interfacial stress plays an important role in the shear response of nanolaminates, which depends on interface orientations and repeat layer spacing λ.

The results for λ ranging from 4 to 20 nm are illustrated in figure 4. $\lambda= 50$ and 100 nm are also explored, and the results are consistent. If we assume $\tau^{\rm i}(\lambda)=0$ , the yield strain of these nanolaminates should be equal to that of pure Al. However, as shown in figure 4(a), $\gamma_{\rm Y}$ is less than that of pure Al, especially for the Al(1 0 0)/Cu(1 0 0) nanolaminates. In addition, as shown in figure 4(b), $\tau_{\rm F}$ of Al(1 1 1)/Cu(1 0 0) nanolaminates is higher than $\tau^{\rm Al(1\, 1\, 1)}_{\rm F}$ , while $\tau_{\rm F}$ of Al(1 0 0)/Cu(1 0 0) nanolaminates is less than $\tau^{\rm Al(1\, 0\, 0)}_{\rm F}$ . Following equation (3), when $\gamma < \gamma_{\rm Y}$ ,

Equation (4)
Figure 4.

Figure 4. (a) $\gamma_{\rm Y}$ λ and (b) $\tau_{\rm Y}$ λ curves of nanolaminates consisting of Cu(1 0 0) crystals. $\gamma_{\rm Y}$ and $\tau_{\rm Y}$ represent the shear yield strain and stress, respectively. $\gamma^{\rm Al(1\, 0\, 0)/(1\, 1\, 1)}_{\rm F}$ and $\tau^{\rm Al(1\, 0\, 0)/(1\, 1\, 1)}_{\rm F}$ represent the shear yield strain and stress of Al in (1 0 0)/(1 1 1) plane, respectively. λ is the repeat layer spacing.

Standard image High-resolution image

Here, $G^{\rm Al}$ and $G^{\rm Cu}$ are the shear modulus of Al and Cu, respectively. At 300 K, the shear moduli of Al in the (1 0 0) and (1 1 1) planes are 26 GPa and 25 GPa, while the shear moduli of Cu in these two planes are 68 GPa and 28 GPa, respectively [3638]. We found that the shear modulus of nanolaminates was determined by interfacial combinations and independent of repeat layer spacing. According to our MD simulations, the shear moduli $G^{\rm Al(1\, 0\, 0)/Cu(1\, 0\, 0)}$ , $G^{\rm Al(1\, 0\, 0)/Cu(1\, 1\, 1)}$ , $G^{\rm Al(1\, 1\, 1)/Cu(1\, 0\, 0)}$ and $G^{\rm Al(1\, 1\, 1)/Cu(1\, 1\, 1)}$ are about 45 GPa, 26 GPa, 41 GPa and 24 GPa, respectively. Assuming $\tau^{\rm i}$ is negligible in equation (4), the effective shear modulus of A/B nanolaminates is approximatively given as $G^{\rm A/B}=(G^{\rm A}+G^{\rm B})/2$ , which is in agreement with our MD results. For the Cu(1 0 0)-stacking nanolaminates, $G^{\rm Cu(1\, 0\, 0)}$ is much larger than $G^{\rm Al(1\, 0\, 0)}$ and $G^{\rm Al(1\, 1\, 1)}$ . Thus, $\tau^{\rm e}$ is generally larger than $\tau^{\rm Al}$ at the same γ, and $\tau_{\rm Y}$ of the Al(1 1 1)/Cu(1 0 0) nanolaminates is higher than $\tau^{\rm Al(1\, 1\, 1)}_{\rm F}$ . However, the $\tau-\gamma$ curves (figure 2) show a remarkable premature yield for the Al(1 0 0)/Cu(1 0 0) nanolaminates; $\gamma_{\rm Y}$ of the Al(1 1 1)/Cu(1 0 0) nanolaminates is close to $\gamma_{\rm F}$ , resulting in the anomalous relationship of $\tau^{\rm Al(1\, 0\, 0)/Cu(1\, 0\, 0)}_{\rm Y} < \tau^{\rm Al(1\, 0\, 0)}_{\rm F}$ . This phenomenon is explained through dislocation analysis in the following section. Moreover, both $\gamma_{\rm Y}$ and $\tau_{\rm Y}$ are weakly dependent on λ. This indicates that the interfacial stress plays the dominant role in yield, which is in turn determined by the interface orientations instead of λ.

As mentioned in section 3.1, with increased shear strain, plastic deformation occurs in Cu layers, which leads to the structural failure of nanolaminates immediately. In this situation, both Al and Cu layers undergo plastic flow. Before structural failure, the effective stress of nanolaminates is

Equation (5)

$\gamma_{\rm F}$ of Al/Cu nanolaminates ranges from 0.075 to 0.1, and is still independent of λ (figure 5(a)). Since there is a slight difference in $\gamma_F$ for these nanolaminates, the interface orientations do not have a strong effect on $\gamma_{\rm F}$ . On the other hand, although $\tau_{\rm F}$ of the nanolaminates is also independent of λ (figure 5(b)), there is a remarkable difference in $\tau_{\rm F}$ among nanolaminates with different interface orientations: $\tau^{\rm Al(1\, 1\, 1)/Cu(1\, 0\, 0)}_{\rm F} \approx \tau^{\rm Al(1\, 0\, 0)/Cu(1\, 0\, 0)} > \tau^{\rm Al(1\, 0\, 0)/Cu(1\, 1\, 1)}_{\rm F} > $ $ \tau^{\rm Al(1\, 1\, 1)/Cu(1\, 1\, 1)}_{\rm F}$ . The results for $\lambda =50$ and 100 nm are also consistent with figure 5. However, for the Al(1 0 0)/Cu(1 0 0) nanolaminates with $\lambda <12$ nm, $\tau_{\rm F}$ increases with increasing λ; thus, the interfacial stress $\tau^{\rm i}$ is influenced by λ in such cases.

Figure 5.

Figure 5. (a) $\gamma_{\rm F}$ λ and (b) $\tau_{\rm F}$ λ curves of Al/Cu nanolaminates. $\gamma_{\rm F}$ and $\tau_{\rm F}$ represent the shear failure strain and stress, respectively. λ is the repeat layer spacing.

Standard image High-resolution image

From equation (5), we can obtain shear strength of nanolaminates, i.e. failure shear stress $\tau^{\rm e}_{\rm F}$ , as

Equation (6)

where $\tau^{\rm l}=\frac{1}{2}\overline{\tau}^{\rm Al}+\frac{1}{2}\tau^{\rm Cu}_{\rm F}$ is the average stress in metal layers. As shown in figure 5(b), the shear strength of these nanolaminates are less than $\tau^{\rm l}$ , but higher than $\overline{\tau}$ of pure Al and Cu. Moreover, $\tau_{\rm F}$ of Cu(1 0 0)-stacking nanolaminates is considerably higher than that of Cu(1 1 1)-stacking nanolaminates, and $\tau_{\rm F}$ of Al(1 1 1)/Cu(1 1 1) nanolaminates is the lowest. Along with the previous analysis of yield, we demonstrate that both yield and failure of Al/Cu nanolaminates are mainly controlled by interface orientations. The effect of repeat layer spacing λ can be neglected in most situations except for Al(1 0 0)/Cu(1 0 0) nanolaminates.

3.3. Dislocation analysis

To reveal the mechanisms underlying macroscopic phenomena of plasticity and failure observed above, we perform atomic-level dislocation analysis below.

During the yield stage of Cu(1 0 0)-stacking nanolaminates, there is a remarkable premature yield for Al(1 0 0)/Cu(1 0 0) nanolaminates, which results in an anomalous relationship of $\tau^{\rm Al(1\, 0\, 0)/Cu(1\, 0\, 0)}_{\rm Y} \ll \tau^{\rm Al(1\, 0\, 0)}_{\rm Y}$ . For Al(1 0 0)/Cu(1 0 0) nanolaminates (figure 6(a)), semi-coherent interfaces are generated due to large mismatch (12%) of lattice parameters, which lead to the misfit network (perfect dislocations, blue lines) between Al and Cu layers [21, 52, 53]. The perfect dislocations may transform into new dislocations even in low shear strains [22] in two ways: (i) one interfacial misfit dislocation transforms to one stair-rod ($\frac16\langle 1\, 1\, 0 \rangle$ ) and two leading Shockley partials ($\frac16\langle 1\, 1\, 2 \rangle$ ) dislocations [22, 52]; and (ii) one interfacial misfit dislocation changes to one Frank partial ($\frac13\langle 1\, 1\, 1 \rangle$ ) and one Shockley partial ($\frac16\langle 1\, 1\, 2 \rangle$ ) dislocation [54]. These Shockley dislocations extend into Al layers, while the stair-rod dislocations remain at the interfaces. These dislocations merge under a small shear strain of about 0.038, and this eventually leads to the premature yield of Al(1 0 0)/Cu(1 0 0) nanolaminates. In contrast, the interfaces in Al(1 1 1)/Cu(1 0 0) nanolaminates are incoherent. As shown in figure 6(b), the Shockley dislocations can extend into Al layers only when shear strain reaches a value close to $\gamma^{\rm Al(1\, 1\, 1)}_{\rm Y}$ , and the plastic deformation of Al layers is induced by trailing Shockley dislocations emitted from neighboring interfaces. These processes are completely determined by the interface orientations, and yield of Cu(1 0 0)-stacking nanolaminates is independent of repeat layer spacing λ.

Figure 6.

Figure 6. Dislocation evolution of (a) Al(1 0 0)/Cu(1 0 0) (3D perspective) and (b) Al(1 1 1)/Cu(1 0 0) nanolaminates (in xz and yz planes). Only atoms with HCP local structure are shown for clarity. Burgers vectors and dislocation lines are also noted. The dash lines represent the positions of interfaces.

Standard image High-resolution image

The stress–strain curves show that the yield stage is absent for the Cu(1 1 1)-stacking nanolaminates. As shown in figure 7, there are no stacking faults before plastic deformation occurs in Al layers, since $\gamma^{\rm Cu(1\, 1\, 1)}_{\rm Y}$ is close to $\gamma^{\rm Al(1\, 1\, 1)}_{\rm Y}$ and less than $\gamma_{\rm Y}^{\rm Al(1\, 0\, 0)}$ . However, when shear strain reach the yield shear strain of Cu(1 1 1)-stacking nanolaminates, dislocations are generated from interfaces and extend into Al and Cu layers simultaneously, and both Al and Cu layers undergo plastic flow, leading to a structural failure. Thus, failure occurs in the nanolaminates immediately after the initial dislocation activities. This is consistent with the effect of yield strain difference on dislocation evolution of Cu/Ta nanolaminates [55]. Moreover, according to equation (3), the interfacial dislocation evolution results in strengthening for Al layers but softening for Cu layers. This effect is also in agreement with a previous study [56]. Since the interfacial stress is determined by interface orientations (rather than λ), the failure of nanolaminates is also independent of λ.

Figure 7.

Figure 7. Dislocation evolution of (a) Al (1 0 0)/Cu(1 1 1) and (b) Al(1 1 1)/Cu(1 1 1) nanolaminates. Only atoms with HCP local structures are shown for clarity. Dislocation lines are also noted. The dashed lines represent the positions of interfaces.

Standard image High-resolution image

In particular, $\tau_{\rm F}$ of Al(1 0 0)/Cu(1 0 0) nanolaminates decreases with decreasing λ for $\lambda <12$ nm, while $\gamma_{\rm F}$ is not influenced by λ (figure 5). Due to stress concentration at interfaces, misfit dislocations are transformed and extend into metal layers [52, 57, 58]. Dislocations transformation leads to a premature yield for Al(1 0 0)/Cu(1 0 0) nanolaminates, and there are only leading Shockley dislocations generated without the emission of trailing Shockley dislocations. Thus, stacking faults do not disappear and even become longer [52, 59]. As shown in figure 8, the two types of stacking faults generated through Shockley partial dislocations, are at 54.7° and 125.3° with the interfaces, respectively. Thus, these stacking faults are not parallel, and there must be interactions between the stacking faults and dislocations. The stacking faults impede the extension of dislocation, which leads to the increase of strain and stress [60]. With increasing λ, the length of stacking faults increases as shown in figure 8. Thus, since the probability of dislocation interactions increases, $\tau_{\rm F}$ of Al(1 0 0)/Cu(1 0 0) nanolaminates increases with λ. However, with increasing λ, the probability of dislocation interactions reaches a limit, and the related shear stress is about 3.0 GPa [36, 37]. Thus, $\tau_{\rm Y}$ of Al(1 0 0)/Cu(1 0 0) nanolaminates increases with increasing λ for $\lambda <12$ nm, while it becomes a constant when $\lambda \geqslant 12$ nm. The low strength of Al/Cu nanolaminates with $\lambda <12$ nm should be considered in fabrication, since it is easier to tailor λ [32] than interface orientations [61, 62].

Figure 8.

Figure 8. Dislocation evolution of Al(1 0 0)/Cu(1 0 0) nanolaminates with different λ. Only atoms with HCP local structure are shown for clarity. The dashed lines represent the positions of interfaces.

Standard image High-resolution image

Conclusion

Shear deformation of Al/Cu nanolaminates is explored as regards interface orientations and repeat layer spacings. Dislocation evolution is mainly controlled by interface orientations rather than repeat layer spacing. The nucleation and propagation of dislocations in nanolaminates are affected by interfacial stress. There is a remarkable premature yield for Al(1 0 0)/Cu(1 0 0) nanolaminates due to the merge of dislocations generated from interfaces. For Cu(1 1 1)-stacking nanolaminates, dislocations generated from interfaces extend into Al and Cu layers simultaneously, which lead to the absence of a yield stage. The yield is mainly determined by the crystal orientation of Al layers, while failure, by that of Cu layers. Moreover, $\tau_{\rm Y}$ of Al(1 0 0)/Cu(1 0 0) nanolaminates decreases with decreasing λ for $\lambda < 12$ nm. This is caused by the unstable dislocation extension near interfaces. Repeat layer spacing of Al/Cu nanolaminates should be over 12 nm, and advanced fabrication should consider tailoring interface orientations for higher shear strength.

Acknowledgments

This work was partially supported by the National Natural Science Foundation of China (Grant No. 11327803), the Science Challenge Project (Grant No. TZ2018001) and the project of Youth Innovation of Science and Technology of Sichuan Province (Grant No. 2016TD0022). The numerical calculations were performed at the Supercomputing Center of the Peac Institute of Multiscale Sciences.

Please wait… references are loading.
10.1088/1361-6463/aad2a8