Elsevier

Acta Materialia

Volume 57, Issue 15, September 2009, Pages 4508-4518
Acta Materialia

Dislocation–twin interaction mechanisms for ultrahigh strength and ductility in nanotwinned metals

https://doi.org/10.1016/j.actamat.2009.06.015Get rights and content

Abstract

Ultrafine polycrystalline metals containing nanotwins exhibit simultaneous ultrahigh strength and ductility. We study the plastic deformation of such materials through molecular dynamics simulations. Based upon these simulations, we trace the sequence of dislocation events associated with the initiation of plastic deformation, dislocation interaction with twin boundaries, dislocation multiplication and deformation debris formation. We report two new dislocation mechanisms that explain the observation of both ultrahigh strength and ductility found in this class of microstructures. First, we observe the interaction of a 60° dislocation with a twin boundary that leads to the formation of a {001}110 Lomer dislocation which, in turn, dissociates into Shockley, stair-rod and Frank partial dislocations. Second, the interaction of a 30° Shockley partial dislocation with a twin boundary generates three new Shockley partials during twin-mediated slip transfer. The generation of a high-density of Shockley partial dislocations on several different slip systems contributes to the observed ultrahigh ductility, while the formation of sessile stair-rod and Frank partial dislocations (together with the presence of the twin boundaries themselves) explain observations of ultrahigh strength. Our simulation highlights the importance of interplay between the carriers of and barriers to plastic deformation in achieving simultaneous ultrahigh strength and ductility.

Introduction

Refining the scale of microstructure in metallic systems commonly leads to increasing strength, but at the expense of decreased toughness. Recently, Lu et al. [1] synthesized ultrafine pure crystalline copper (Cu) containing a high-density of growth twins via a pulsed electrodeposition technique. The resulting material is unusual in that it simultaneously exhibits a high yield strength, high ductility, high strain-rate sensitivity and high electric conductivity [2], [3], [4]. Transmission electron microscopy (TEM) studies of these nanotwinned metals revealed dislocation pile-ups at the twin boundaries [2], suggesting that the enhanced mechanical strength is associated with the effectiveness of twin boundaries as barriers to dislocation motion. Jin et al. [5], [6] studied the mechanisms of interaction between dislocations and twins in different face-centered cubic (fcc) metals. While they found that these interactions can generate dislocation locks, the detailed interaction mechanisms are both material- and loading condition-dependent. Zhu et al. [7] showed that twin boundaries are deep traps for screw dislocations, suggesting that twin boundary-mediated slip transfer is the rate-controlling mechanism for the observed increased strain-rate sensitivity with increasing twin density. All of these studies indicate that the ultrahigh strength of nanotwinned crystalline metals is related to nanotwin-induced strain hardening. However, several important issues remain open. What are the mechanisms of twin formation in pulsed electrodeposition synthesis? What is the origin of the observed ultrahigh ductility? What are the atomistic mechanisms by which twin boundaries lead to strain hardening? In this paper, we employ molecular dynamics (MD) simulations to study the mechanisms that contribute to the observed ultrahigh strength and ductility of ultrafine, nanotwinned Cu.

In the MD simulations presented herein, we employ embedded atom method (EAM)-type interatomic potentials that are suitable for describing Cu in order to examine a microstructure containing an ultrafine array of growth twins under uniaxial, fixed true strain-rate conditions. We examine the deformation mechanism by monitoring the evolution of the dislocation–twin microstructure during the initiation and propagation of plastic deformation with atomic resolution. Slip transfer across twin boundaries was frequently observed, which indicates twin boundaries serve as barriers to dislocation motion and those barriers are not impenetrable. We observe the generation of a {001}110 Lomer dislocation from the interaction of a 60° full dislocation and a twin boundary. The subsequent dissociation of the Lomer dislocation results in the formation of Shockley, stair-rod and Frank partial dislocations. Generations of those Shockley partial dislocations activate a series of additional slip systems in the crystal, while the sessile (stair-rod and Frank) dislocations provide dislocation locks preventing further dislocation migration. The Lomer dislocations are important in the plastic deformation of the nanotwinned crystal as the source of the experimentally observed strain-hardening effect. We also observed a new mechanism in which a 30° partial dislocation interacts with the twin boundary to generate three new Shockley partial dislocations during the twin-mediated slip transfer. The gliding of the high-density Shockley partial dislocations contributes to the experimentally observed ultrahigh ductility, while the sessile dislocations formed through the dissociation of the Lomer dislocation contribute to the experimentally observed ultrahigh strength.

Section snippets

Methods

The {111} stacking fault energy in fcc metals, γSF, and the equilibrium separation between pairs of Shockley partial dislocations, are known to be sensitive to the choice of interatomic potential employed in atomistic simulations. Therefore, the first step in simulations of nanotwinned Cu is the identification of an interatomic potential function that produces realistic values of the stacking fault energy. While many interatomic potentials are available for pure Cu, we chose the EAM potential

Dislocation nucleation and evolution

The initial simulation cell contains a regular array of coherent twin boundaries, a pair of “low-angle tilt” grain boundaries, as shown in Fig. 1, and no dislocations. This microstructure is similar to the as-sputtered microstructures reported in Ref. [13]; in particular, both the simulations and experiments contain low-angle (∼9°) boundaries and a set of parallel coherent twin boundaries (∼15 nm apart). In the present study, we examine the dislocation nucleation mechanisms and the subsequent

Dislocation–twin interaction mechanisms

Dislocations can be blocked where they first make contact with twin boundaries. As more of the dislocation encounters the twin boundary, a long, straight dislocation segment forms at the intersection of the slip plane and the twin boundary. Because the dislocations intersecting the twin boundary are straight in fcc systems, we need only focus on three types of lattice dislocation: (i) a screw dislocation (AB on c plane for example); (ii) a 60° dislocation with a 30° leading partial and a 90°

Discussion

Many different dislocation generation mechanisms operate in coarse-grained materials. For example, a classic Frank–Read source can generate dislocation after dislocation, thereby providing a continuous supply of dislocations to sustain plastic deformation. However, when the grain size is on the nanometer scale, most of the dislocation generation mechanisms that operate in coarse-grained materials shut down, leaving a dearth of operative dislocation sources and resulting in what has become known

Conclusions

We present MD simulation results for the uniaxial tensile loading of Cu with a microstructure similar to that produced by a pulsed electrodeposition or a magnetron-sputtering technique (i.e. an ultrafine-grained microstructure containing a high-density of growth nanotwins). The plastic deformation of the initially dislocation-free sample is initiated by dislocation nucleation from the grain boundaries. The applied load couples to multiple slip systems producing both Schmid and non-Schmid

Acknowledgements

The authors grateful acknowledge useful discussions with Prof. Peter Gumbsch. All simulations were performed using Lammps Molecular Dynamics Simulator [19], [20] and visualized in Atomeye [21].

References (21)

  • M. Dao et al.

    Acta Mater

    (2006)
  • L. Lu et al.

    Acta Mater

    (2005)
  • Z. Jin et al.

    Scripta Mater

    (2006)
  • Z.H. Jin et al.

    Acta Mater

    (2008)
  • E. Ma

    Scripta Mater

    (2003)
  • L. Murr et al.

    Scripta Metall

    (1978)
  • S.J. Plimpton

    J Comput Phys

    (1995)
  • J. Li

    Modell Simulat Mater Sci Eng

    (2003)
  • L. Lu et al.

    Science

    (2004)
  • L. Lu et al.

    Science

    (2009)
There are more references available in the full text version of this article.

Cited by (0)

View full text