A TEM Kikuchi pattern study of ECAP AA1200 via routes A, C, BC

https://doi.org/10.1016/j.matchar.2010.03.007Get rights and content

Abstract

Shear plastic deformation in ECAP occurs through the formation, movement and storage of dislocations. The differences which exist between shearing characteristics lead to important implications concerning the optimum processing route. It is known that the effectiveness of cell evolution into an array of high-angle boundaries (HABs) is in the order BC > C > A, or A > BC > C, depending on the two-channel-angle intersection of the ECAP die. In route A, large portions of HABs are continuously and progressively generated, while routes C and BC cause fully redundant deformation at each 2n passes, and each 4n passes, respectively. This study is focused on dislocation generation, storage and recombination during ECAP for routes A, C and BC. Kikuchi bands identified with TEM were used to quantitatively measure the cell and grain boundary misorientation. ECAP was performed on an AA1200 commercially pure aluminium alloy up to ε = 8.64. A different hierarchy for HAB generation efficiency was found. Thermal stability was studied by annealing the alloy at 0.5, 0.6, 0.7 TM (where TM is the alloy melting point) for 2 h after the severe plastic deformation. It appeared that, even if route BC involves the fastest microstructure grain refining, route C is likely to be the most stable upon reheating.

Introduction

Due to their superior mechanical and physical properties, ultra-fine-grained materials are currently attracting a great deal of attention. These properties include increased strength, low temperature superplasticity, and the possibility to produce new classes of materials, like high-strength aluminium alloys, that do not require age hardening [1], [2], [3]. Ultra-fine-grained materials may be prepared using various techniques including gas condensation and subsequent in situ consolidation under high vacuum or high energy ball milling and they may be produced in a wide range of materials (e.g. pure metals, metallic alloys including superalloys, intermetallics, and semiconductors) by subjecting these materials to a very high plastic strain [4], [5], [6], [7], [8], [9], [10], [11], [12], [13], [14], [15], [16], [17], [18], [19]. Cost-effective methods that can be used to produce bulk materials are of particular importance in this context. Severe plastic deformation (SPD) techniques are able to produce bulk materials with fine-grained structure. Among the various SPD techniques, equal-channel angular pressing (ECAP) [18] is able, in certain circumstances, to give grain sizes in the submicrometre range of 400–800 nm [20], [21], [22]. In particular, ECAP is an especially attractive processing method because it allows large bulk samples to be produced which are free from any residual porosity, inducing small shape changes (notably almost no cross-sectional change).

The evolution of the microstructure during ECAP, is closely driven by the specific pressing conditions, which means by the shearing deformation induced in the material at each pass through the die [23]. Plastic deformation of metals occurs as a result of the formation, movement and storage of dislocations. In fact, microstructure evolution during ECAP is directly linked to a complex evolution of dislocation networks and their recombination and annihilation phenomena. Imbalance of dislocations of opposite signs moving in opposite directions (or having some Burgers vector components of opposite sign and sliding direction) leads to accumulation, generally at the grain boundary, or at high-angle boundaries (HABs), of a surplus of dislocations of the same sign. The groups of excess dislocations in fact stimulate crystallographic slip. Therefore, the evolution and accumulation of misorientation across both low-angle boundaries (LABs) and HABs is closely linked to the crystallographic accommodation of each cell, grain and the neighbouring crystallites during the shearing deformation. In deformation route A the billet is extruded without rotation between passes (unidirectional straining path); route C is defined as a rotation of 180° between passes and route BC as a rotation of 90° in the same direction between passes [4], [13], [24], [25], [26], [27], [28], [29].

In a study of pure aluminium, deformed at room temperature by a 90° equal-channel die, at a strain ε = 4, Iwahashi et al. [11], [20] reported the effectiveness of cell evolution into an array of HABs as BC > C > A. In accordance with this observation, they found route BC to be the most effective route for producing equiaxed fine grain structure with HABs. Segal [30] arrived at the same conclusion showing that each pass of ECAP is likely to develop thin shear bands oriented along the shear plane. The above mentioned differences in shearing characteristics lead to important implications concerning the optimum processing route. In fact, grain refining is expected to be less effective in route A compared to routes BC and C. Route BC appears to be preferable to route C, which lacks any deformation in the plane normal to the pressing direction [4], [24]. In route A, the shear bands that form during the second pass intersect the billet axis at an angle of 45°, while the shear bands that form in the first pass change orientation to an angle of about 20° to the billet axis. The same procedure is repeated at the following passes and this continuously transforms the previous shear bands into an elongated shape along the flow direction. Route BC is effective in refining the microstructure, because four sets of shear bands are created after a fully redundant cycle (4n passes). Gholinia et al. [13] offered an alternative to Segal's conclusions. They suggested that shear strain is accumulated in the same direction by route A, making this route the most effective method for generating HABs. Both routes BC and C are redundant strain processes, and consequently are less efficient in creating HABs. However, Zhu and Li [4], [31] suggested that combinations of shear plane with texture and crystal structure play a primary role in grain refinement, while the accumulative strain plays a secondary role. These authors also claimed that their theory could explain why BC is the most effective deformation route in grain refinement for the 90° die while route A is the most effective for the 120° die.

Under severe plastic deformation (SPD), and for every possible route in ECAP, dislocation boundaries evolve into a regular pattern of grain subdivisions belonging to two scales [32], [33], [34], [35], [36], [37], [38]: large-scale long and continuous dislocation boundaries, called geometrically necessary boundaries (GNBs), and small-scale incidental dislocation boundaries (IDBs) [33]. The former include micro-bands (MBs) and dense dislocation walls (DDWs) surrounding groups of equiaxed cells. The group of cells enclosed by GNBs are called cell blocks (CBs), while IDBs include ordinary cell boundaries. The boundary spacing and misorientation angle distribution of GNBs and IDBs evolve differently as a function of strain; in particular, they exhibit different morphologies at small to medium strains, but similar at high strains [32], [33], [34], [35], [36], [37], [38]. Generally, misorientation axes for IDBs are randomly distributed, whereas GNB orientation distribution clusters on preferred axes and some GNBs show a pattern of rotation around the transverse direction [39], [40], [41]. GNBs play an important part in the deformation process. In fact, to accommodate the strain imposed during deformation, additional dislocations arrange to form geometrically necessary high-dislocation regions associated with dislocation interactions and slip patterns that develop within a deforming grain [40]. Thus, the distance between neighbouring GNBs decreases with strain, approaching, as a limit value, the cell spacing, which decreases much less with strain. Therefore, whatever the route used, grains subdivide under the effect of newly introduced HABs on finer and finer length-scales, until ultimately a limit is reached where the HAB spacing converges with the cell size [42], [43]. The number of cells between GNBs therefore decreases with increasing strain, with a consequent evolution from cells that have only other cells as nearest neighbours towards cells that have both GNBs and cells as nearest neighbours [39]. The IDB mechanism of formation and evolution is the statistical trapping of glide dislocations, whereas the formation of GNBs is associated with differences in the slip systems operating in neighbouring crystallites or, alternatively, involving similar slip systems but having different shear amplitude [44]. In this study, the Kikuchi pattern approach was followed to determine the boundary character. Thus, it was possible to directly identify cells as IDBs, and HABs as GNBs.

The average spacing between dislocation boundaries (LABs and HABs) is a common microstructure parameter in deformed materials due to its inverse relationship with flow stress, as amply documented in literature [40], [45], [46]. Studies of misorientation angles and of the early cell formation process are less investigated (e.g. [2], [7], [12], [30], [47], [48]). Experimental studies on the correlations between the deformation pattern and the resulting microstructure have been carried out, especially by transmission electron microscopy (TEM) inspections [21], [33], [48], [49], [50], [51]. These observations have shown that a structural description, based on individual dislocations or dislocations arranged in smaller groups (tangles and cell walls), is valid at low strains. However with increasing strain the characteristic structural features are dislocations with relatively large misorientation across the boundaries.

Thermal stability is an important feature for materials subjected to severe plastic deformation. Refined microstructures must maintain a reasonably refined structure upon heating at least up to 0.5 TM (where TM is the alloy melting temperature). However, their high defect content in the form of crystal boundaries makes them susceptible to structural coarsening.

Several studies have reported the superplasticity behaviour of ECA pressed aluminium and other alloys, which ultimately implies a rather good thermal stability of the severely deformed microstructure [52], [53], [54], [55]. Langdon et al., in several studies [56], [57], [58], [59], [60], [61], have reported that a refined microstructure has a meaningful mechanical application only if this refined structure is not completely destroyed (i.e. recrystallized) upon heating. Of course, there will be temperature threshold-like limits that will depend on the specific aluminium alloy (in terms of secondary phases and grain refining dispersoids), the thermo-mechanical status of the alloy and the level of SPD.

Microstructure inspections of the deformation structures in ECAP materials have been studied by transmission electron microscopy (TEM) and by electron backscattered diffraction (EBSD) using FEGSEM. For example, Iwahashi et al. [10], [11], Liu et al. [32] studied microstructure using TEM, and Gholinia et al. [13], El-Danaf [62], Reihanian et al. [63] used high-resolution EBSD to measure boundary misorientation. However, the latter procedure involves a systematic overestimation of the HAB fraction. In fact, to date, the lower limit of boundary misorientation accuracy of the FEGSEM is typically of ϕ  2°. In practice this limitation requires the use of more reliable techniques, such as analysis of Kikuchi patterns in TEM, in order to achieve meaningful quantitative measurements [32], [48], [64], [65], [66]. For this reason, in this study, the Kikuchi pattern method was used for all the misorientation and size measurements of both cells and grains (LABs and HABs). The microstructure evolution and refining process, in the three routes (A, C, and BC), was studied up to the strain ε = 8.64. Thermal stability studies were carried out for three different temperatures (0.5 TM, 0.6 TM, and 0.7 TM) for all three routes: A, C and BC.

Section snippets

Experimental Details and Method

The chemical composition of commercially pure aluminium (AA1200) is (wt.%): 0.7 Si, 0.3 Fe, 0.1 Zn, 0.05 (Cu + Mn). The material was cast and machined by Hydro Aluminium (Norway) to rod-shaped bars, 10 mm in diameter and 100 mm in length, which were homogenized for 4 h at 540 ° C. The bars were subjected to severe plastic deformation by ECAP at room temperature in a dedicated pressing machine with pressing forces ranging between 20 and 70 kN. The pressing speed was 4 mm/min. ECAP was carried out using a

Results and Discussion

The initial homogenized microstructure consists of grains whose mean size is d = 330 μm [48]. After the first ECAP pass, shear bands developed in the grains containing mostly low-angle boundaries, but also a significant fraction of high-angle boundaries [48].

Fig. 3 gives the distribution of misorientation ϕ for routes A, C and BC. After the first ECAP pass 0.75 of the cell boundaries have low-angle character misorientation (ϕ < 15°). Steps in the curves indicate the presence of a pronounced texture

Conclusions

This study dealt with cell and grain evolution with strain in a severe plastic deformed AA1200 by ECAP via three different routes. A, C, and BC. A Φ = 90° ECAP channel-angle die were used. Boundary misorientation was quantitatively measured using Kikuchi patterns identified with transmission electron microscopy.

Results showed that:

  • (i)

    the distribution of misorientation, ϕ, is far from being randomly distributed after the maximum strain of ε = 8.64 (8 passes); a progressive and monotonic downward shift

Acknowledgment

The author wishes to thank Professor Wolfgang Blum for the fruitful discussions.

References (73)

  • M. Furukawa et al.

    Microhardness measurements and the Hall–Petch relationship in an Al–Mg alloy with submicrometer grain size

    Acta Mater

    (1996)
  • A. Shan et al.

    Estimation of friction during equal channel angular (ECA) pressing of aluminum alloys

    J Mater Proc Technol

    (2002)
  • Y. Iwahashi et al.

    Principle of equal-channel angular pressing for the processing of ultra-fine grained materials

    Scripta Mater

    (1996)
  • D.A. Hughes

    Microstructure evolution, slip patterns and flow stress

    Mater Sci Eng

    (2001)
  • D.A. Hughes et al.

    High angle boundaries formed by grain subdivision mechanisms

    Acta Mater

    (1997)
  • R.Z. Valiev et al.

    Principles of equal-channel angular pressing as a processing tool for grain refinement

    Program Mater Sci

    (2006)
  • M. Kamachi et al.

    A model investigation of the shearing characteristics in equal-channel angular pressing

    Mater Sci Eng

    (2003)
  • M. Furukawa et al.

    Factors influencing the shearing patterns in equal-channel angular pressing

    Mater Sci Eng

    (2002)
  • T.G. Langdon

    The principles of grain refinement in equal-channel angular pressing

    Mater Sci Eng

    (2007)
  • J.M. García-Infanta et al.

    Microstructural development during equal channel angular pressing of hypo-eutectic Al–Si casting alloy by different processing routes

    Mater Sci Eng

    (2008)
  • J.M. García-Infanta et al.

    Effect of the deformation path on the ductility of a hypoeutectic Al–Si casting alloy subjected to equal-channel angular pressing by routes A, BA, BC and C

    Scripta Mater

    (2008)
  • K. Oh-ishi et al.

    Effect of strain path on evolution of deformation bands during ECAP of pure aluminum

    Mater Sci Eng

    (2005)
  • V.M. Segal

    Severe plastic deformation: simple shear versus pure shear

    Mater Sci Eng

    (2002)
  • S. Li

    Comments on “Texture evolution by shear on two planes during ECAP of a high-strength aluminum alloy”

    Scripta Mater

    (2009)
  • Q. Liu et al.

    Microstructure and strength of commercial purity aluminium (AA 1200) cold-rolled to large strains

    Acta Mater

    (2002)
  • D. Kuhlmann-Wilsdorf et al.

    Geometrically necessary, incidental and subgrain boundaries

    Scr Metall

    (1991)
  • P.L. Sun et al.

    Characteristics of submicron grained structure formed in aluminum by equal channel angular extrusion

    Mater Sci Eng

    (2000)
  • N. Hansen et al.

    Flow stress anisotropy caused by geometrically necessary boundaries

    Acta Metall Mater

    (1992)
  • B. Bay et al.

    Microstructural evolution in rolled aluminium

    Mater Sci Eng

    (1992)
  • R. Ørsund et al.

    Local lattice curvature and deformation heterogeneities in heavily deformed aluminium

    Scr Metall

    (1989)
  • G.J. Baxter et al.

    The influence of transient strain-rate deformation conditions on the deformed microstructure of aluminium alloy Al–1% Mg

    Acta Mater

    (1999)
  • D.A. Hughes et al.

    Scaling of microstructural parameters: misorientations of deformation induced boundaries

    Acta Mater

    (1997)
  • A. Godfrey et al.

    Scaling of the spacing of deformation induced dislocation boundaries

    Acta Mater

    (2000)
  • H. Weiland

    The determination of long-range misorientations in the microstructure of materials

    Acta Metall

    (1992)
  • M. Wagenhofer et al.

    Influences of strain rate and grain size on yield and serrated flow in commercial Al–Mg alloy 5086

    Scripta Mater

    (1999)
  • D. Kuhlmann-Wilsdorf

    Theory of plastic deformation: properties of low energy dislocation structures

    Mater Sci Eng

    (1989)
  • Cited by (42)

    • Enhancement of strength and ductility in 6061Al alloy processed by cross accumulative roll bonding

      2023, Transactions of Nonferrous Metals Society of China (English Edition)
    • Skin effect in 125 to 10 µm-thick 5N-Al foils

      2021, Materials Letters
      Citation Excerpt :

      Thus, ρdisl = 2ndisl/(lmeshtTEM), were, lmesh is the total mesh length, and tTEM is the local measured thickness determined by converged electron beam diffraction patterns. Cell (low-angle boundary, LAB) and GB (high-angle boundary, HAB) misorientation were measured by Kikuchi band patterns as fully described elsewhere in previous published works by this author [14,15]. The thin 5N-Al foils were inspected both at the central (core) zone and next to the surface.

    • The lamella structure of Al-Mg-Si matrix nanocomposites with isotropically high strength

      2020, Materialia
      Citation Excerpt :

      Besides, the differences in glide activity during strain path change can promote the formation of new slip systems and increase the dislocation density, further resulting in subgrain boundaries. The excess dislocations caused by more strain path change can enhance the strength [25]. As shown in Fig. 10, LBs formed in the CARBi sample, and many subgrains with sizes of ~ 9 μm were found inside lamellar grains.

    • A review on processing of aluminium and its alloys through Equal Channel Angular Pressing die

      2020, Materials Today: Proceedings
      Citation Excerpt :

      However, the addition of dispersoids can ensure stability at super-plastic temperatures. Also, route C has substantial grain size homogeneity and size stability as compared to routes A and BC [34]. Additionally, it is found using tests on a low purity aluminum alloy (99.1% Al) that ECAP makes the material less corrosion resistant [35].

    • Fatigue behavior of AA7075 aluminium alloy severely deformed by equal channel angular pressing

      2018, Journal of Alloys and Compounds
      Citation Excerpt :

      It is clear that after 1 pass of ECAP, deformation bands develop within the grains, width of grains decreases and elongated grains oriented approximately at 45° to the transverse direction (TD direction) appear in the microstructure. The mentioned changes in the microstructure caused by shear bands that formed due to the plastic strain localization during ECAP [22,29–31]. Fig. 2 shows TEM images of initial AA7075 aluminium alloy and ECAPed specimen in annealed condition.

    View all citing articles on Scopus
    View full text