J Msea 2020 139905

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 35

Journal Pre-proof

Grain refinement in hot working of 2219 aluminium alloy: On the effect of deformation
mode and loading path

X. Zeng, X.G. Fan, H.W. Li, M. Zhan, S.H. Li, T.W. Ren, K.Q. Wu

PII: S0921-5093(20)30977-1
DOI: https://doi.org/10.1016/j.msea.2020.139905
Reference: MSA 139905

To appear in: Materials Science & Engineering A

Received Date: 15 March 2020


Revised Date: 3 July 2020
Accepted Date: 7 July 2020

Please cite this article as: X. Zeng, X.G. Fan, H.W. Li, M. Zhan, S.H. Li, T.W. Ren, K.Q. Wu, Grain
refinement in hot working of 2219 aluminium alloy: On the effect of deformation mode and loading path,
Materials Science & Engineering A (2020), doi: https://doi.org/10.1016/j.msea.2020.139905.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier B.V.


CRediT authorship contribution statement

X. Zeng: Conceptualization, Methodology, Investigation, Writing-original draft.


X. G. Fan: Writing-review&editing, Supervision, Project administration, Funding
acquisition.
H.W. Li: Supervision, Project administration.
M. Zhan: Supervision, Project administration.
S. H. Li: Supervision, Project administration.
T.W. Ren: Investigation.
K. Q. Wu: Investigation.

1
Grain refinement in hot working of 2219 aluminium alloy: on the

effect of deformation mode and loading path

X. Zenga,b, X.G. Fana,b,*, H.W. Lia,b, M. Zhana,b, S.H. Lic,d, T.W. Rena,b, K. Q. Wua,b

a. State Key Laboratory of Solidification Processing, School of Materials Science and

Engineering, Northwestern Polytechnical University, Xi’an, 710072, P.R. China

b. Shaanxi Key Laboratory of High-Performance Precision Forming Technology and Equipment,

Northwestern Polytechnical University, Xi’an, 710072, P.R. China

c. State Key Laboratory of Mechanical System and Vibration, Shanghai Jiao Tong University,

Shanghai 200240, China

d. Shanghai Key Laboratory of Digital Manufacture for Thin-walled Structures, Shanghai Jiao

Tong University, Shanghai 200240, China

*Corresponding author: Tel./fax: +86-029-8849-5632; E-mail: fxg3200@nwpu.edu.cn;

fxg3200@126.com

Abstract
To study the effect of deformation mode together with loading path on the
microstructure evolution of 2219 aluminium alloy at elevated temperature, plane
strain compression (PSC) and uniaxial compression (UC) along different loading
paths (rolling direction (RD), transverse direction (TD) and normal direction (ND)),
as well as equal-channel angular pressing (ECAP) along RD were performed on the
cubic and cylindrical samples cut from thick plate. Detailed microstructure evolution
was characterized by optical microscope and electron backscattered diffraction
(EBSD). The results indicate that the PSC deformation shows a higher stress level
than UC and an increasing discrepancy in flow softening occurs with the increasing
strain during PSC along different loading paths. The volume fraction of dynamic
recrystallization (DRX) after RD-UC is higher than that of RD-PSC, while the grain
refinement level is the lowest in the one pass ECAP. DRX is the main grain
refinement mechanism for RD-UC, while deformation bands (DBs) are dominant in
RD-PSC and ECAP. Due to the grain morphology anisotropy, a significant

1
microstructural difference was exhibited in the PSCed samples along different loading
paths. The grain refinement in hot working of 2219 aluminium alloy plate is greatly
related to the deformation modes and loading paths in view of the brick-like grains.
Key words: 2219 aluminium alloy; Deformation mode; Loading path; Microstructure
evolution; Grain refinement mechanism.
1. Introduction
Due to the merits of high strength, light weight and good weldability, 2219 aluminium
alloy is widely used in the critical structural components in the aviation and aerospace
industry [1-4]. Its deformation behavior that will provide guidance for the
optimization of processing parameters is of great interest to researchers [2-5]. For
instance, Kaibyshev et al. [5] have made a systematic study on the deformation
behavior of homogenized (solution treatment) 2219 aluminium alloy from 1.5×10−6 to
3×10−2s−1 and in the temperature range of 250-500°C. The stress exponent, activation
energy, etc. were calculated, and the deformation mechanism related to temperature
and strain rate was deduced. However, to the authors’ knowledge, there is little open
literature about 2219 aluminium alloy deformed at the strain rate larger than 1, though
the cracking often occurs at very high strain rate, e.g. the failure of propellant tank in
Ref. [1]. As the microstructure evolution mechanism will affect the deformation
behavior of material, thus deciding the property of the components, a further study of
the microstructure evolution mechanism of 2219 aluminium alloy at elevated
temperature and high strain rate is necessary.
At elevated temperature, although dynamic recovery (DRV) is the main restoration
mechanism for aluminium alloy with high stacking fault energy (SFE), dynamic
recrystallization (DRX) [6-8] plays an important role in the hot working, including
discontinuous dynamic recrystallization (DDRX), continuous dynamic
recrystallization (CDRX) and geometrical dynamic recrystallization (GDRX).
Amongst, CDRX is often induced by the transformation of sub-grains into grains,
while nucleation by grain boundaries bulging mechanism is dominant in most DDRX
process. Besides, with respect to the annealed microstructure that exists large numbers
of coarse second-phase particles, the effect of particle stimulate nucleation (PSN)
2
cannot be ignored during the subsequent hot working [9].
Although the volume fraction of DRX is usually the function of temperature, strain
and strain rate, etc., it has been proved that the kinetics of DRX is affected by stress
state. Thus, some researches about deformation modes involving different kinds of
stress states have been carried out in the past several years. Comparing with the
microstructure after hot plane strain compression (PSC) and uniaxial compression
(UC), Mandal et al. [10] studied the effect of stress state on DRX behavior of
austenitic stainless steel. They found that the volume fraction of DRX is lower in the
PSCed microstructure, since the deformation mode of UC may be more beneficial for
the generation of dislocations. Eventually, they concluded the stress state will affect
the DRX kinetics, but not alter the DRX nucleation mechanism of UC and PSC.
However, Athreya et al. [11] found that the DRX kinetics in PSC is much faster than
UC via detailed deformed microstructure study of Ni, during which the initiation of
DRX needs a lower critical stress and strain in PSC. As a result, the fraction of DRX
in PSCed sample is higher than that in UCed sample. In addition, Cheng et al. [12]
concluded the additional stress in PSC of IN718 superalloy contributes to the
propagation of shear bands, while those cannot be promoted by UC. Al-Samman et al.
[13] stated that different from UC, the plastic deformation of PSC is enforced to
proceed in the compression and extension directions, resulting in the discrepancy in
flow stress and texture evolution of AZ31 magnesium alloy. As for 2219 aluminium
alloy, Murty et al. [14] performed hot isothermal PSC along normal direction (ND)
and concluded that the uniaxial deformation cannot refine the grains effectively even
at large strain. Zeng et al. [15] carried out UC experiment in different loading
directions at 400ºC and found that the grains in the rolled sheet with a brick-like
morphology will aggravate the plastic flow anisotropy and result in diverse DRX
types.
By its nature, the introduction of DRX as well as deformation bands (DBs) aims at
refining the microstructure, since it can greatly increase the strength without lowering
the ductility of material. Owning to the excellent grain refinement capacity, equal
channel angular pressing (ECAP) is one of the most typical severe plastic deformation
3
(SPD) methods, which imposes a shear deformation at the intersection of the channel
[16]. It can even refine the grains to a nanometer level through enough passes and
different routes. During the last few decades, numerous studies about ECAP have
been carried out on copper [17, 18], aluminium [19] and it’s alloy [20-23] and
magnesium alloy [24, 25], etc. With respect to 2219 aluminium alloy, Mazurina et al.
[26, 27] conducted ECAP in a wide range of strain and temperature to make
systematic studies on the grain refinement mechanism. They found DBs that incubate
at low strain and propagate at moderate strain contribute to the grain fragmentation.
Meanwhile, the structural changes are classified into three stages according to the
magnitude of strain. Generally, shear deformation is thought to be significant for grain
refinement, which will facilitate the development of high angle grain boundaries
(HAGBs) [28]. However, according to the above results of Murty et al. [14] and
Mazurina et al. [26, 27], although a shear deformation is imposed during PSC and
ECAP, the grain refinement level is extremely limited at a given loading path and
strain for 2219 aluminium alloy. Thus, besides deformation modes, the effect of
loading paths on microstructure evolution should also be taken into consideration.
Davenport and Higginson [29] have reviewed the effect of loading path (strain path)
during hot working, which has a significant impact on the deformation mechanics,
microstructure evolution and texture development. Dislocation sub-structures are
susceptible to the strain path changes, which in turn affect the DRX kinetics. To refine
grains and homogenize microstructure, different loading paths are often applied to
ECAP [16, 24, 30], amongst route-B has the severest grain refinement ability [31]. As
the rolled thick plate with brick-like grains was used in this study, the subsequent
deformation along different loading paths would have a different influence on the
microstructure evolution and thus on the mechanical property.
In order to study the effect of deformation modes together with loading paths on the
microstructure evolution of 2219 aluminium alloy at elevated temperature, PSC and
UC experiments along different loading paths as well as ECAP along RD were carried
out on the thick plate candidate. The related grain refinement mechanisms are
carefully discussed, which will provide guidance for tailoring the mechanical property
4
of 2219 aluminium alloy.
2. Experiment
The annealed 2219 aluminium alloy thick plate with the thickness of 14.5mm (ND)
was employed in this study, whose mainly chemical composition was
Al-6.5Cu-0.36Mn-0.21Fe-0.18Zr (Wt%). The optical microstructure in the three
planes is shown in Fig. 1(a-c). On the RD×TD plane, the grains show an irregular
shape. The average spacing of grain boundaries is about 69.3µm along TD, while the
long axis of some grains is larger than 570µm along RD. On the TD×ND and RD×ND
planes, the grains present a more obviously ribbon-like morphology, where the
adjacent grain boundaries aligned with RD are almost parallel to each other. The
average spacing of grain boundaries is about 64.7µm along ND. It can be concluded
that the grains in the thick plate also present a brick-like shape. In addition, a
considerable volume fraction of coarse second-phase particles (Al2Cu) occur in the
original microstructure, whose volume fraction is about 8%. EBSD orientation map
(OM) of RD×ND plane is shown in Fig. 1(d).
(a) (b)

TD ND

RD TD
(c) (d)

ND ND

RD RD

Fig. 1. Optical photographs (OPs) (a-c) and orientation map (OM) (d) of the initial thick sheet: (a)

RD×TD plane, (b) TD×ND plane and (c, d) RD×ND plane.

PSC and UC experiments were carried out on Gleeble-3500 simulator along RD, TD

5
and ND. The Gleeble-3500 simulator with a direct resistance heating system is often
used to perform the thermo-mechanical testing such as tension, compression and
plane strain compression, etc. It can accurately control and record the temperature and
loading by the fully integrated digital closed loop control thermal and mechanical
testing system. More details about Gleeble simulator can be seen in Ref. [32].
Schematic diagrams of the deformed samples are presented in Fig. 2(a) and (b).
Before PSC and UC, the samples were heated to 400°C with the heating rate of
10°C/s and held for 5min. After PSC and UC at the strain rate of 10s-1, the samples
were water quenched. Cubic and cylindrical samples were respectively used in the
PSC and UC experiments, whose corresponding dimensions were
20(length)×14.5(width)×10(height)mm and Ф8×12mm. After deformation, the
samples were sectioned by wire-electrode cutting machine and the microstructure in
RD×ND plane as well as extra TD×ND plane for TD-PSCed sample was detected by
Leica DMI3000 optical microscope and Merlin Compact thermal field emission
scanning electron microscope. Details of the preparation methods of metallographic
and EBSD specimens can be seen in Ref. [15].
For the sake of maintaining the same equivalent strain in different hot deformation
modes, the reduction ratio of PSC and UC were 50% and 55%, respectively, which
can be calculated by the formula (1) and (2). Accordingly, the calculated equivalent
strain of the deformed samples was about 0.8. The formulas of equivalent strain ε̅P
of PSC [33] and ε̅ U of UC are given as follows:

̅ = (2/√3)ln[1/(1 − )] (1)

Where ΨP is the reduction ratio in PSC, and ΨP =50% in this study.


̅ = ln[1/(1 − )] (2)
Where ΨU is the reduction ratio in UC, and ΨU =55% in this study.
In addition, ECAP was carried out at 400°C with the pressing speed of 100mm/min.
Before ECAP, the sample was heated to 400°C with the heating rate of 10°C/min and
held for 30min for the purpose of temperature homogenization. The dimension of
ECAP sample was Φ10×50mm. To make a precise evaluation on the relationship

6
between the shear deformation and grain morphology, the RD×ND plane of brick-like
grains was parallel to the channel section of ECAP, as shown by Fig. 2(c). In other
words, the RD×ND plane of brick-like grains was vertical to the main shear plane that
can be referred to Refs. [17, 21]. The red lines on the cross-section of the billet in Fig.
2(c) denote the long axes of the initial grains. The equivalent strain ε̅E can be
calculated by the following formula [30]:
N ∅ φ ∅ φ
ε̅E = [2 cot + +φcosec + ] (3)
√3 2 2 2 2

Where N is the number of passes, ϕ is the channel angle and φ is the curvature angle.
In this study, N=1. ϕ and φ are 105° and 30°, respectively. Thus, the calculated
equivalent strain of one pass ECAP is also 0.8.
The strain rate at one-pass ECAP can be calculated by the following formula [34]:
1 ∅ φ ∅ φ √
E = √3 [2 cot 2 + 2 +φcosec 2
+2 ] (4)

Where v is the loading speed of plunger, D is the diameter of die channel.


In this study, v=100mm/min and D=10mm. Thus, the calculated average strain rate of
one pass ECAP is about 0.36s-1, which is in good agreement with the result of finite
element (FE) simulation (at the range of 0.1-0.7s-1 from the outer corner to inner
corner) introduced as follows.
As a widely used FE simulation software for forged products, Deform software is
often applied to the simulation of metal bulk forming process. Usually, the forming of
axisymmetric parts can be simulated by Deform-2D software, while the parts with
complex shape or complex forming operations can be simulated by Deform-3D
software. More details about Deform software can be seen in Refs. [35, 36]. Thus, in
order to indicate the equivalent strain distribution, Deform-2D software was
employed to simulate the UC deformation and Deform-3D software was used to
simulate PSC and ECAP deformation. The stress-strain curves obtained by ND-UC at
400ºC in the range of 10-3s-1-10s-1 and the shear friction of 0.3 were used in the
isothermal FE models. To maintain the strain rate at 10s-1, the speed of punch (Top die)
was defined as a function of stroke, which was the same as that in the actual
experiment. I.e., at high strain rate, the displacement-time curve was used in the

7
Gleeble-3500 simulator to control the displacement of punch. The formulas of speed
and displacement of punch are shown in the following:
lave=H0Ψt/N (5)
vn= { H0-(n-1) lave} (6)
Sn=nlave (7)
Where H0 is the initial height of sample, Ψt is the total reduction ratio, N is the stroke
number of the punch and lave is the average displacement of punch at each stroke. vn is
the speed of punch at n-stroke (1≤n≤N) and is the strain rate. Sn is the total
displacement of punch at n-step. In this study, we chose N=20.
The shape of ND-UCed sample after experiment and FE simulation was compared to
verify the accuracy of FE simulation. Due to the impact of friction condition, the
deformation is uneven. As shown in Fig. 2(d-f), the equivalent strain at the center of
PSCed sample is 1.06, while it is 1.1 and 0.86 for the UCed and ECAPed sample,
respectively.

Plunge
(a) (b) (c)
D
Billet
Die

RD
Inner

ND Outer
(d) (e) (f)

Fig. 2. Schematic diagrams and equivalent strain distribution of the deformed samples: (a, d) PSC,

(b, e) UC and (c, f) ECAP.

To get an overview of PSCed microstructure, the optical microstructure of the PSCed


sample along three different loading directions is shown in Fig. 3(a-c), which are
stitched by many optical photographs at the same magnification. Notably, the material

8
flowing lines can be clearly seen after PSC. In Fig. 3(a), the material flowing lines
retain straight after ND-PSC at the center, but the spacing of grain boundaries
decreases, which will be confirmed by EBSD maps. At the shear zone, the material
flowing lines are inclined, but it seems that the flow lines are not cut off. However,
RD-PSCed and TD-PSCed samples show definitely different flow lines from those of
ND-PSCed sample, in which the straight HAGBs turn to bend towards the external
side, as shown in Fig. 3(b) and (c). Although the material flowing lines in TD-PSCed
sample are almost similar to those in RD-PSCed sample, they are not as apparent as
those in the RD-PSCed sample. This is ascribed to the more distorted grain
morphology on TD×ND plane after TD-PSC. As different microstructure
characteristic presents at different zones due to the heterogeneous deformation, the
typical zones of the PSCed sample are marked in Fig. 3(d), i.e., the hard deformation
zone, shear zone, small deformation zone and large deformation zone, which are
marked by I, II, III and IV, respectively. The corresponding equivalent strains are 0.55,
0.91, 0.80 and 1.06 for I zone, II zone, III zone and IV zone. Besides, the
microstructure evolution at the typical zones of the PSCed sample was characterized
by EBSD, which would be presented in Section 3.

(a) (b)

ND RD

RD ND

(c) (d)

I
II
III IV

TD

ND Compression direction

Fig.3. OPs (a-c) and the marked position (d) of the PSCed sample: (a) ND-PSCed sample, (b)
RD-PSCed sample, (c) TD-PSCed sample and (d) marked position.

9
3. Results
3.1 True stress-strain curves
Fig. 4 shows the true stress-strain curves of UC and PSC along RD, TD and ND, in
which they are named RD-UC, TD-UC, ND-UC, RD-PSC, TD-PSC and ND-PSC,
respectively. The friction effect during UC has been modified by using the formulas in
Ref. [15]. The samples subjected to PSC show a higher stress level. The same trend
has been found by Al-Samman et al. [13], who ascribed this phenomenon to the
broadening suppression in PSC. All of the curves show an obvious restoration
characteristic. Just like the UC deformation of thin sheet in Ref. [15], the annealed
2219 aluminium alloy shows a non-prominent strength anisotropy, as proved by the
almost superposed curves in the three directions of UC. However, it is interesting to
find that the difference among the three curves increases with the increasing strain in
PSC. Thereinto, the flow stress of TD-PSC is the lowest and that of ND-PSC is the
highest at high strain level, implying that the flow softening is related to the loading
paths. The related mechanisms in charge of the deformation behavior will be
discussed later.

120

100
True stress (MPa)

80
RD-UC RD-PSC
60 TD-UC TD-PSC
ND-UC ND-PSC
40

20

0
0.0 0.2 0.4 0.6 0.8
True strain

Fig. 4. True stress-true strain curves of the samples deformed by PSC and UC with different

loading paths.

3.2 Microstructure evolution of PSC


Fig. 5 shows the microstructure at the different zones of ND-PSCed sample. At the
hard deformation zone in Fig. 5(a), the spacing of HAGBs is comparable to that of the
initial microstructure in Fig. 1(d). However, there exist intragranular orientation
10
gradients in the ribbon-like grains, which indicate the emergence of dislocations and
the resultant stored energy [37]. In Fig. 5(b), the grains mainly produce an inclined
geometric orientation at the shear zone, of which the intersection angle is near 60°
between the long axis of ribbon-like grains and compression direction (ND). The
average HAGB spacing of ribbon-like grains is about 32µm, not including the narrow
bands in the grain interior. Some microbands occur in the interior of grains with large
HAGB spacing, marked by red arrows. Although the grains still keep initial shape at
the small deformation zone in Fig. 5(c), the spacing of HAGBs decreases to about
27µm. The coarse second-phase particles aggravate the heterogeneous deformation in
grain interior and the distortion of grain boundaries. At the large deformation zone
(Fig. 5d), lots of fine grains are developed near the original grain boundaries and the
coarse second-phase particles, which are certified to be DRXed grains at a larger
magnification. Meanwhile, some microbands with the orientation <001>//TD occur in
the interior of ribbon-like grains with large HAGB spacing and almost unchanged
crystal orientation.
To illuminate the characteristics of substructure and grain boundaries, EBSD
detection at a larger magnification of 2000× was conducted on the typical zones of
ND-PSCed sample, as shown in Fig. 6. The grain interior mainly fills with sub-grain
cells accompanying with low angle grain boundaries (LAGBs). To avoid introducing
artifacts, only the LAGBs with misorientation of 2°-15° are marked in white. Few
DRXed grains occur near the coarse second-phase particles and microbands in Fig.
6(a), while the banded structures with sub-grain cells are dominant in Fig. 6(b). More
DRXed grains and serrations are shown in Fig. 6(c).

11
(a) (b)

(c) (d)

ND

RD

Fig.5. OMs of the ND-PSCed sample at different zones: (a) the hard deformation zone (I), (b) the

shear zone (II), (c) small deformation zone (III), and (d) large deformation zone (IV).

(a) (b) (c)

ND

RD
TD

Fig.6. Band contrast (BC)+Grain boundaries (GBs)+OM maps of the ND-PSCed sample at the

large magnification of 2000×: (a) the shear zone (II), (b) small deformation zone (III) and (c) large

deformation zone (IV).

Fig. 7 shows the microstructure of RD-PSCed sample, where the compression


direction is along the long axes of ribbon-like grains. At the hard deformation zone in
Fig. 7(a), the grains’ morphologies almost keep unchanged and their long axes are
aligned with RD. However, the spacing of HAGBs reduces to 33µm regardless of
some narrow microbands occurring near the initial grain boundaries. At the shear zone
in Fig. 7(b), some of the bent grains with small HAGBs spacing are broken up, while
those with large HAGB spacing show a good deformation accommodation except for
12
some microbands with new orientations occurring within them. Moreover, the grain
boundaries turn to be serrated, as shown by the red arrow, which will be confirmed at
large magnification. In Fig. 7(c), the grains suffer the most severely buckling
deformation. Even the grains with a large HAGB spacing were subdivided by the
inserted grains. Many newly formed HAGBs occur in the grain interior, resulting in a
disordered microstructure. In Fig. 7(d), the grains show an abnormal morphology at
the large deformation zone, which is similar to the central bulged marker pins in the
thickness direction during rolling in Ref. [38]. Microbands with diverse orientations
from the matrix occur in the interior of the so-called bulged grains. It seems that DBs
with geometrical orientation perpendicular to the compression direction occur in the
inhomogeneous microstructure. Few fine DRXed grains can be seen near the coarse
second-phase particles. Fig. 8 shows the microstructure at a larger magnification of
2000× at the shear and small deformation zones of RD-PSCed sample. In Fig. 8(a),
the serrations at shear zone are attributed to the sub-grains which spread in the interior
of banded grains. Moreover, DBs can be easily verified by the A-C-A type of
orientation change [6, 39]. In Fig. 8(b), the tip of new-formed tiny grains penetrated
into the coarse grains, resulting in the distorted grain boundaries.
(a) (b)

(c) (d)

ND

RD
Fig.7. OMs of the RD-PSCed sample at different zones: (a) the hard deformation zone (I), (b) the
shear zone (II), (c) small deformation zone (III), and (d) large deformation zone (IV).

13
(a) (b)

ND

RD

Fig.8. BC+GBs+OM maps of the RD-PSCed sample at the large magnification of 2000×: (a) the

shear zone (II) and (b) small deformation zone (III).

The microstructure of TD-PSCed sample is shown in Fig. 9, where both the OMs in
TD×ND plane and ND×RD plane were exhibited. At the shear zone (Fig. 9a), many
grains in an irregular shape occur in the coarse grains with twisty boundaries, owning
different crystal orientations from the matrix. At the large deformation zone (Fig. 9b),
the microstructure is extremely heterogeneous, but it seems that the long axes of the
grains are perpendicular to the compression direction (TD). Some microbands with
the crystal orientation of <113>//TD occur in the parent grains.
To investigate the in-plane deformation, the microstructure of TD-PSCed sample in
ND×RD plane is shown in Fig. 9(c) and (d). At the small deformation zone (Fig. 9c),
the grain boundaries still keep straight and the spacing of HAGBs is comparable to
that of the RD-PSCed sample. However, some sub-structures with gradually varying
crystal orientations (orientation gradients) occur in the ribbon-like grains, which are
inclined from the long axes of the grains, as shown by the red arrows. At the large
deformation zone in Fig. 9(d), the grains show twisted grain morphologies and
definitely different crystal orientations from that in the RD-PSCed sample on account
of the central bulging deformation. In addition, some fine equiaxed grains occur near
the bulged grain boundaries and second-phase particles due to DDRX.

14
(a) (b)

TD

ND
(c) (d)

ND

RD

Fig.9. OMs of TD-PSCed sample at different zones in TD×ND plane (a, b) and ND×RD plane (c,

d): (a) the shear zone (II), (b) large deformation zone (IV), (c) small deformation zone (III) and

(d) large deformation zone (IV).

3. 3 Microstructure evolution of UC
Figs. 10 show OMs after UC along ND and RD, whose observation positions are
marked in the OPs in Fig. 10(a) and (d). The grains in RD×ND plane still keep a
ribbon-like morphology and the grain boundaries in the long axis direction are aligned
with RD after ND-UC. However, the spacing of HAGBs decreases and the grain
boundaries turn to be serrated. Lots of LAGBs distribute in the grain interior,
contributing to the formation of cellular structures, as shown by the sub-grains in Fig.
10(c). Some transverse HAGBs transformed from LAGBs in grain interior divide the
ribbon-like grains, as marked by the white arrows in Fig. 10(b) and (c). Meanwhile,
some fine grains occur near the serrated grain boundaries and second-phase particles,
of which the former is evolved by rotation of sub-grains and the latter is nucleated by
PSN. Thus, both CDRX and DDRX occur during the ND-UC.
OMs of the RD-UCed sample are shown in Fig. 10(e) and (f). Obviously, the large
ribbon-like grains in RD×ND plane turn to be twisted, presenting an extremely
irregular shape. Lots of fine grains with different crystal orientations from the

15
surroundings occur around the bulged grain boundaries and the coarse second-phase
particles. It seems that many DRXed grains are in the orientation <001>//TD during
RD-UC, as shown in Fig. 10(f) at large magnification of 2000×. Besides, the
sub-grain cells in the deformed grains present an elongated shape along ND. The
results of UC in thick plate is similar to that of thin sheet in Ref. [15], other than the
discrepancy in the volume fraction of DRX, which may be related to the larger HAGB
spacing in the thick plate of annealed 2219 aluminium alloy.
Compression direction
(b) (c)
(a)

ND ND

RD RD
(e) (f)
(d)

RD ND

ND RD

Fig. 10. OPs and OMs at the large deformation zone of samples experienced UC along ND (a-c)

and RD (d-f): (b) at magnification of 200×, (e) at magnification of 500× and (c, f) at the

magnification of 2000×.

3.4 Microstructure evolution of ECAP


Fig. 11 shows the OMs at the sectioned central plane of ECAPed cylindrical sample,
where P1-P5 are along the diameter of the circular section. P1 and P5 are 0.5mm
away from the surface. P2 and P4 are 1.2 mm away from the surface. P3 is at the
center, i.e., 5mm away from the surface. Surprisingly, the grain refinement extent is
limited and DRX is sluggish via ECAP at 400°C. However, many DBs occur in the
interior of ribbon-like grains at P1 (Fig. 11b) and P2 (Fig. 11c) of the ECAPed
cylinder, but they are rare at P4 (Fig. 11e) and P5 (Fig. 11f). Notably, some of the DBs
also can be named shear bands that extend to several adjacent grains [6, 39]. As
shown in Fig. 11(c), the shear band marked by white arrows seems crossing the grain
16
boundary marked by white triangle. In this study, we don’t make a rigorous distinction
between DBs and shear bands, since most of bands are within one grain. The reduced
number of DBs from the inner to the outer may be attributed to the decreasing shear
strain. From P1 to P5, the equivalent plastic strains are 0.92, 0.89, 0.86, 0.66 and 0.63,
respectively.
(b) (c)
(a)
Sectioned position
Inner
ND

RD Outer
(d) (e) (f)

Fig. 11. EBSD observation position (a) and OMs after ECAP at the positions of (b) P1, (c) P2, (d)

P3, (e) P4 and (f) P5. P1 and P5 are 0.5mm away from the surface; P2 and P4 are 1.2 mm away

from the surface; P3 is at the center, i.e., 5mm away from the surface.

Although the ribbon-like grain seems un-refined in the outer layers in Fig. 11(e) and
11(f), the evident intragranular orientation gradients are induced by shear deformation.
The microstructure at large magnification at P4 is investigated in Fig. 12. It can be
seen the large grains are divided by the fine grains in a banded distribution in Fig.
12(a). Meanwhile, DRX occurs near the grain boundaries in Fig. 12(c). In Fig. 12(d)
and (e), banded structures with the approximately same crystal orientations with the
matrix occur in the large ribbon-like grains, which present an intersection angle of
30°-40° with RD. It may be at the incubation stage of DBs.

17
(a) (b)(b) (c)

(d) (e)

ND

RD

Fig. 12. OMs after ECAP at large magnification of 2000× at P4.

4. Discussion
4.1 The effect of deformation mode
In order to study the effect of deformation mode, the microstructure of RD-UCed,
RD-PSCed and ECAPed samples is compared and the corresponding deformation
mechanism is discussed in this section.
In most cases, a uniform tensile deformation occurs in radial direction during UC
along axial direction, and UCed sample often presents a drum shape on account of
friction between the anvil and sample. However, it should be noted that the drum
RD-UCed sample is non-axisymmetric with an elliptic shape in TD×ND section due
to the plastic flow anisotropy, which is mainly caused by texture and aggravated by
the special grain morphology [15]. Thus, the RD-UCed sample will suffer a more
non-uniform deformation than the ND-UCed sample, and some DDRXed grains occur
at elevated temperature. Besides, large numbers of cell blocks (sub-grains) formed by
the ordered two-dimensional sub-structures (dislocation walls) distribute in the grain
interior uniformly at high strain rate, which are elongated vertically to the
compression direction.
Unlike the deformation mode of RD-UC, the deformation along TD is suppressed
during RD-PSC, so the sample is thinned at RD and stretched at ND under RD
compression. In other words, the PSC deformation is restricted in the RD×ND plane.
As a result, the brick-like grains after RD-PSC tend to be buckled, and

18
two-dimensional HAGBs are formed in the PSCed microstructure. At the center of
RD-PSCed sample, it appears that a bulging deformation caused by plastic
localization and instability occurs, resulting in the abnormal morphology of grains.
Moreover, adiabatic temperature may rise at high strain rate even for aluminium alloy
[40], which will lead to the generation of adiabatic shear bands [1].
To gain a deeper insight into the grain refinement of RD-PSC, a detailed study was
conducted on the microstructure at the center of PSCed sample. It seems that there is a
relationship among DRX, DB and grain crystal orientations. As a matter of fact, it is
widely accepted that DBs including transition bands, kink bands and shear bands are
preferential nucleation sites for DRX [6, 29]. In Fig. 13(b), the DRXed grains mainly
occur around the coarse second-phase particles, while some DRXed grains in banded
distribution occur in the large original grain with a <101>//TD orientation (green
color). The grains with <113>//TD orientation (light purple color) are not beneficial
for the generation of DRX, whose matrix are occupied by LAGBs in Fig. 13(c). For
the grains with the <111>//TD orientation (blue color) in Fig. 13(d), large numbers of
microbands occur in the initial large grains. Simultaneously, DBs can be obviously
seen in Fig. 13(e) marked by the red arrows, but they present a geometric orientation
vertical to RD.

19
(b)

(c) (a) (d)

(e)
ND
TD

RD

Fig. 13. OMs at the center (IV) of RD-PSC sample: (a) at the center and (b-e) at large

magnification.

As reviewed in the introduction, the kinetics of DRX is affected by the deformation


modes [10, 11]. The distribution of DRXed grains of RD-UCed and RD-PSCed
samples is shown in Fig. 14(a) and 14(b), where the DRX volume fractions of
RD-UCed sample and RD-PSCed sample are near 21.3% and 12.7%, respectively. It
is in the same trend with the result of Mandal et al. [10], but inversed to the
conclusion of Athreya et al. [11]. In their studies, Athreya et al. [11] found that the
difference in DRX fraction between the two deformation modes increases first with
the increasing strain rate and then decreases at high strain rate from 0.01s-1 to 1s-1,
while the UC and PSC experiments were carried out at 10s-1 by Mandal et al. [10].
Thus, the varying DRX kinetics at the two deformation modes may be largely
attributed to the impact of strain rate, which will be discussed later. On the other hand,
DDRX is dominant during RD-UC and RD-PSC due to the brick-like grain
morphology, since the compression direction is along the grains’ long axes and the
annealed microstructure shows a prominent PSN effect introduced by the coarse

20
second-phase particles. The corresponding misorientation angle distribution is shown
in Fig. 14(c) and (d), and the average misorientation angles of RD-UCed sample and
RD-PSCed sample are 14.74° and 11.14°, respectively. Although LAGBs play major
roles in the deformed microstructure in both of the two deformation modes, the
HAGBs with misorientation angle between 50° and 60° are in a large proportion after
RD-PSC, which may be attributed to the generation of DBs.
(a) (b)

0.20 0.20
(c) RD-UC (d) RD-PSC
0.16 fLAGBs=0.671 0.16 fLAGBs=0.804
θAve=14.74o θAve=11.14ο
Fraction
Fraction

0.12 0.12

0.08 0.08

0.04 0.04

0.00 0.00
10 20 30 40 50 60 10 20 30 40 50 60
Misorientation angle (°) Misorientation angle (°)

Fig. 14. The distribution of DRXed grains (in red) and misorientation angle distribution maps at

the center (IV) of RD-UCed sample (a, c) and RD-PSCed sample (b, d).

Theoretically, ECAP is considered to be a typical simple shear method [41, 42], which
will impose a shear strain uniformly on the RD×ND plane. However, there is a
microstructural difference at the different layers of the ECAPed sample due to the
friction condition and die geometry. As the RD×ND plane of brick-like grains was
vertical to the main shear plane, a given loading path was applied to ECAP. The
microstructure at the center of ECAPed sample in Fig. 11(d) is somewhat similar to
that at the shear zone of RD-PSC (Fig. 7b), though PSC is in a pure shear mode that is
similar to the strain state at the center of rolling [38, 41]. Besides, the ECAPed
microstructure in the outside layer in Fig. 12(e) is the same as that at the shear zone of
ND-PSC (Fig. 5b) to some extent. Thus the grain morphology as well as crystal

21
orientation has a significant impact on the microstructure evolution regardless of the
microstructural difference induced by shear modes.
DBs/shear bands are generally thought to propagate in the coarse grains deformed at
low temperature and high strain rate [12, 43, 44]. However, DBs can be evidently
seen in the RD-PSCed and ECAPed microstructure as shown in Fig.13 and Fig. 11.
According to studies of Mazurina et al. [26, 27], the generation of DBs is
temperature-independent at the early stage under low to moderate strain, though
grain/sub-grain structures occur in the interior of DBs with the increase of
deformation temperature. Furthermore, DBs will result in the grain fragmentation at
moderate strain, thereby leading to the uniform grain refinement at large strain, which
is belong to the CDRX process. Different from their studies, in which the initial billet
has been homogenized by solution treatment, a complete annealing is conducted on
the initial rolling thick plate in this study. Consequently, lots of coarse second-phase
particles precipitate in the brick-like microstructure, which will aggravate the
heterogeneity of deformation. Moreover, as larger numbers of slip systems can be
activated by tensile or compressive loading comparing with torsion or shear loading
[29], the deformation of ECAP as well as RD-PSC is more uneven. Thus, the shear
deformation imposed on the long axes of grains accelerates the generation of DBs
even at elevated temperature. However, as Segal [41, 42] has also summarized that
pure shear is beneficial for the generation of cell structures rather than grain
refinement, while simple shear is the ideal deformation mode for grain refinement and
structure modification though it does not contribute considerably development of
HAGBs at minor strain, the grain refinement level is limited at the one-pass ECAP. So,
in order to refine the microstructure greatly, not only is the changing deformation
mode needed, but also a lager strain should be imposed.
In fact, the effect of strain rate also cannot be neglected during the hot working of
2219 aluminium alloy that is sensitive to strain rate at elevated temperature [15]. Due
to limitation of equipment, the strain rate during ECAP is about 0.36s-1, which is
lower than that during PSC and UC (10s-1). In the opened literature, DRX is inclined
to occur at low strain rate and high temperature, while a considerable volume fraction
22
of LAGBs are developed at high strain rate [15, 40]. However, the DRX volume
fractions of ECAPed sample is about 8.2%, lower than that in the RD-UCed and
RD-PSCed samples, as shown in Fig. 15. On the contrary, the fraction of LAGBs in
the ECAPed sample is the highest. These suggest that the influence of simple shear
deformation mode and brick-like grain morphology on microstructure development
prevails over that of strain rate during ECAP. In the limited studies, Berbon et al. [45]
have investigated the influence of pressing speed at the range of 10-2-10mm/s on the
microstructure evolution of Al and Al-1%Mg alloy during ECAP, and found that the
equilibrium grain size is not susceptible to the pressing speed. Similarly,
Muñoz-Morris and Morris [20] stated that the elongated cellular substructures are
formed and grain size reduces little at high strain rate level by performing ECAP at
standard pressing speed (20mm/min), faster speed (60mm/min) and slower speed
(6mm/min). Thus, DBs as well as large numbers of LAGBs are generated during
ECAP along RD, as shown in Fig. 15(c). Likewise, Jia et al. [21] also found DBs
occurred in the Al-8Zn alloy during the one pass room temperature ECAP at strain
rate of 2×10-2s-1, though a higher strain rate deformation is more beneficial for the
generation of DBs. Moreover, the cellular structures/sub-grains are developed in the
interior of DBs, as shown by the band contrast (BC) map in Fig. 15(d). They are
similar to the bamboo-like substructure, but like no other the microbands that are
double-dislocation wall structure in study of Xue et al. [17].

23
(a) 0.20
(b) ECAP
0.16 fLAGBs=0.826
θAve=9.19o

Fraction
0.12

0.08

0.04

0.00
10 20 30 40 50 60
Misorientation angle (°)

(c) (d)

ND

RD

Fig. 15. The distribution of DRXed grains (in red) (a) and misorientation angle distribution map (b)

at the center (P3) of ECAPed sample, OM+GBs map (c) and BC map (d) at 2000×.

4.2 The effect of loading path


As mentioned previously, the initial thick plate with brick-like grains was produced
by rolling followed by subsequent complete annealing. Thus, the anisotropic grain
morphology will have an influence on the microstructure evolution as well as
deformation behavior of 2219 aluminium alloy under PSC and UC with different
loading paths. Amongst, the effect of loading paths during UC has been studied in Ref.
[15] on a thin rolled 2219 aluminium alloy sheet. Although the thickness of the plate
used in this study is larger than the thin sheet, the microstructure deformation
mechanism is almost not changed. I.e., both CDRX and DDRX occur in ND-UC,
while DDRX is dominated in RD-UC and TD-UC. However, due to the larger HAGB
spacing in the thick plate, the volume fraction of DRX is lower than that of thin sheet,
which further proves that the microstructure evolution is also related to geometrical
characteristic of grains under different loading paths. Therefore, in this section, we
mainly focus on the effect of loading paths on the microstructure evolution of PSC.
In ND-PSC, although the spacing of HAGBs decreases after deformation, the volume
fraction of DRX is limited, and the fine equiaxed grains mainly occur near the grain

24
boundaries and second-phase particles. At 50% reduction, the HAGB spacing of some
large grains cannot decrease to the spacing of 1-2 sub-grains, which is the prerequisite
for the occurrence of GDRX [8]. Thus GDRX is not enabled to initiate. Meanwhile,
as Murty et al. [14] have found, the microstructure after PSC along ND was hardly to
be refined, who concluded the unidirectional deformation was unable to refine the
microstructure and thus can be substituted by multidirectional deformation. So, the
limited DRX primarily originates from the dislocation nucleation mechanism such as
PSN of the coarse second-phase particles.
For RD-PSC and TD-PSC, the compression is along the long axes of the brick-like
grains, resulting in a bulging deformation on the microstructure. Consequently, quite
diverse microstructure characteristic is showed in the RD-PSCed and TD-PSCed
sample from that of ND-PSCed sample. Namely, unlike the approximately straight
grain boundaries in the ND-PSCed sample, the bulged grain boundaries are presented
at the large deformation zone of RD-PSCed and TD-PSCed sample, making up an
extremely heterogeneous microstructure. Although RD-PSC and TD-PSC are carried
out at elevated temperature, the transient instability at 10s-1 in the in-plane
deformation results in the occurrence of DBs. In view of the limited volume fraction
of DDRX, DBs are thought to be the main grain subdivision mechanism. In addition,
since the initial long axis of grains in RD is larger than that of TD, there is also a
microstructure difference between the two loading paths. At the large deformation
zone, different from the bulged grain boundaries of RD-PSCed sample in Fig. 7(d)
and Fig. 13(a), the long axes of the grains turn to be perpendicular to the compression
direction (TD) with a 90° rotation after TD-PSC in Fig. 9(b). Meanwhile, the grains’
crystal orientations also change greatly.
The pole figures (PFs) of initial billet and PSCed sample are shown in Fig. 16.
Compared with the initial billet, the texture intensity decreases after PSC both along
RD and TD, but it is almost unchanged after ND-PSC. The main grain crystal
orientations consisting of a texture {100}<-512> in initial billet deviates 20° after
ND-PSC as shown by the white arrows in Fig. 16(a) and 16(b), while the texture types
change greatly after RD-PSC and TD-PSC in Fig. 16(c) and 16(d). I.e., the texture
25
types are {111}<01-1> and {110}<1-13> in the RD-PSCed sample, but those change
to {110}<114> and {111}<315> after TD-PSC. Therefore, the loading paths also have
an impact on the texture development.

(a) (b)

(c) (d)

Fig. 16. PFs at the center (IV) of (a) the initial billet, (b) ND-PSCed sample, (c) RD-PSCed

sample and (d) TD-PSCed sample.

The microstructural difference is in charge of the varying true stress-true strain curves
in Fig. 3. Jha et al. [1] ascribed the local reduction of flow stress to the undissipated
heat generated by the transient deformation at high strain rate, accompanying with the
occurrence of adiabatic shear band formation. Nayan et al. [40] also found that flow
softening is closely related to shear banding and recrystallization, and shear banding
is interdependent with flow softening. Thus, considering the limited DDRX during
PSC along the grains’ long axes (RD and TD), the formation of DBs at high strain rate
are mainly in charge of the softening behavior of 2219 aluminium alloy. On the other
hand, the calculated Taylor factors are 3.255, 2.747 and 3.240 at RD, TD and ND
compression, respectively, implying that the deformation along TD is the easiest
owing to the favorable crystal orientations. Thus the crystal orientations/textures also
have a significant impact on the deformation behavior of 2219 aluminium alloy
during PSC. Except for the different grain refinement mechanism, the varying texture
types during PSC along different loading paths may largely account for the increasing
discrepancy in flow softening. Therefore, it can be concluded that the texture along
with the brick-like grain morphology is responsible for the prominent plastic
anisotropy during PSC. Considering the heterogeneity and complexity of actual
deformation, different loading paths are often utilized in the plastic forming to
improve the material property, such as multi-directional forging [7, 16, 25]. For ECAP,
26
multi-pass [18, 22, 23] and changing routes [24, 31] are often used to increase the
cumulative strain and refine the grains.
5. Conclusions
In this paper, the effect of deformation modes and loading paths on microstructure
evolution of 2219 aluminium alloy thick plate at elevated temperature was carefully
studied and the related grain refinement mechanisms were analyzed. It can be
summarized:
1) Compared with UC, the PSC deformation shows a higher stress level, but a
severer flow softening. The discrepancy in flow stress increases with increasing
strain during PSC along different loading paths. The changed texture types along
with the grain morphology related grain refinement mechanism are responsible for
the prominent plastic anisotropy during PSC.
2) The volume fraction of DRX after RD-UC is higher than that after RD-PSC, while
the grain refinement level is the lowest in the one pass ECAP. DDRX is
dominated in RD-UC, but DB is the main grain refinement mechanism for
RD-PSC and ECAP. Simultaneously, it seems that the incubation of DBs is crystal
orientation related, which mainly occur in the grains with the <111>//TD
orientation during RD-PSC.
3) Due to the morphology anisotropy of the brick-like grains, a significant
microstructural difference was exhibited under different loading paths of PSC. I.e.,
the RD-PSCed sample shows a so-called bulged microstructure at the large
deformation zone, while the grains seem experiencing a 90° rotation after TD-PSC.
For ND-PSC, the grain morphology is almost unchanged except for the decreasing
HAGB spacing.
4) The microstructure at the shear zone of RD-PSC and ND-PSC is similar to that of
the ECAPed sample to some degree, during which the shear deformation imposed
on the long axes of grains accelerates the generation of DBs even at elevated
temperature. The highest fraction of LAGBs as well as DBs suggests that the
influence of simple shear deformation mode and brick-like grain morphology on
microstructure development prevails over that of strain rate during ECAP.
27
CRediT authorship contribution statement
X. Zeng: Conceptualization, Methodology, Investigation, Writing-original draft. X. G.
Fan: Writing-review&editing, Supervision, Project administration, Funding
acquisition. H.W. Li: Supervision, Project administration. M. Zhan: Supervision,
Project administration. S. H. Li: Supervision, Project administration. T.W. Ren:
Investigation. K. Q. Wu: Investigation.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.

Acknowledgement
The authors gratefully acknowledge the support of the National Natural Science
Foundation of China (No. 51790175). The authors also appreciate Dr. Yan-Si Liang of
Hefei University of Technology for his selfless guidance and the reviewers for their
constructive suggestions.
References
[1] Abhay K. Jha, S.V.S.N. Murty, K. Sreekumar, P.P. Sinha, High strain rate
deformation and cracking of AA 2219 aluminium alloy welded propellant tank, Eng.
Fail. Anal. 16 (2009) 2209-2216.
[2] L. Liu, Y. Wu, H. Gong, F. Dong, A.S. Ahmad. Modified kinetic model for
describing continuous dynamic recrystallization behavior of Al 2219 alloy during
hot deformation process. J. Alloy. Compd. 817 (2020) 153301.
[3] H. He, Y. Yi, S. Huang, W. Guo, Y. Zhang. Effects of thermomechanical treatment
on grain refinement, second-phase particle dissolution, and mechanical properties
of 2219 Al alloy. J. Mater. Process. Tech. 278 (2020) 116506.
[4] Y.C. Lin, Q. Wu, D. G. He, X. H. Zhu, D.Y. Liu, X. H. Li. Effects of solution time
and cooling rate on microstructures and mechanical properties of 2219 Al alloy for
a larger spun thin-wall ellipsoidal head. J. Mater. Res. Technol. 9(2020) 3566-3577.
[5] R. Kaibyshev, O. Sitdikov, I. Mazurina, D.R. Lesuer, Deformation behavior of a
2219 Al alloy, Mater. Sci. Eng. A 334(2002) 104-113.
[6] F.J. Humphreys, M. Hatherly. Recrystallization and Related Annealing Phenomena

28
(Second Edition), Elsevier, 2004.
[7] T. Sakai, A. Belyakov, R. Kaibyshev, H. Miura, J.J. Jonas. Dynamic and
post-dynamic recrystallization under hot, cold and severe plastic deformation
conditions, Prog. Mater. Sci. 60 (2014) 130-207.
[8] K. Huang, R.E. Logé, A review of dynamic recrystallization phenomena in
metallic materials, Mater. Des. 111 (2016) 548-574.
[9] K. Huang, K Marthinsen, Q.L. Zhao, R. E. Logé, The double-edge effect of
second-phase particles on the recrystallization behaviour and associated mechanical
properties of metallic materials, Prog. Mater. Sci. 92 (2017) 284-359.
[10] S. Mandal, A. K. Bhaduri, V. S. Sarma. Influence of State of Stress on Dynamic
Recrystallization in a Titanium-Modified Austenitic Stainless Steel, Metall. Mater.
Trans. A 43 (2012) 410-414.
[11] C.N. Athreya, S. Suwas, V. Subramanya Sarma, Role of stress state on dynamic
recrystallization behaviour of Ni during hot deformation: Analysis of uniaxial
compression and plane strain compression, Mater. Sci. Eng. A 763 (2019) 138153.
[12] L. Cheng , X. Xue , B. Tang , D. Liu , J. Li , H. Kou, J. Li, Deformation behavior
of hot-rolled IN718 superalloy under plane strain compression at elevated
temperature, Mater. Sci. Eng. A 606 (2014) 24-30.
[13] T. Al-Samman, B. Ahmad, G. Gottstein, Uniaxial and Plane Strain Compression
Behaviour of Magnesium Alloy AZ31: A Comparative Study, Mater. Sci. Forum,
550 (2007) 229-234.
[14] S.V.S. Narayana Murty, Aditya Sarkar, P. Ramesh Narayanan, P.V.
Venkitakrishnan, J. Mukhopadhyay, Microstructure and micro-texture evolution
during large strain deformation of aluminium alloy AA 2219, Mater. Sci. Eng. A
677 (2016) 41-49.
[15] X. Zeng, X.G. Fan, H.W. Li, M. Zhan, S.H. Li. Grain morphology related
microstructural developments in bulk deformation of 2219 aluminum alloy sheet at
elevated temperature, Mater. Sci. Eng. A 760 (2019) 328-338.
[16] Ghader Faraji, Hyoung Seop Kim and Hessam Torabzadeh Kashi. Severe Plastic
Deformation: Methods, Processing and Properties (First Edition), Elsevier, 2018.
29
https://doi.org/10.1016/C2016-0-05256-7.
[17] Q. Xue, I. J. Beyerlein, D. J. Alexander, G. T. Gray III. Mechanisms for initial
grain refinement in OFHC copper during equal channel angular pressing. Acta
Mater. 55(2007) 655-668.
[18] M. Y. Alawadhi, S. Sabbaghianrad, Y. C. Wang, Y. Huang, T. G. Langdon.
Characteristics of grain refinement in oxygen-free copper processed by
equal-channel angular pressing and dynamic testing. Mater. Sci. Eng. A 775 (2020)
138985
[19] J. Fakhimi Derakhshan, M.H. Parsa, H.R. Jafarian. Microstructure and
mechanical properties variations of pure aluminum subjected to one pass of
ECAP-Conform process. Mater. Sci. Eng. A 747 (2019) 120-129
[20] M.A. Muñoz-Morris, D.G. Morris. Severe plastic deformation processing of
Al-Cu-Li alloy for enhancing strength while maintaining ductility. Scripta Mater.
63 (2010) 304-307.
[21] H. Jia, S. Jin, Y. Li. Formation of Σ3{110} incoherent twin boundaries through
geometrically necessary boundaries in an Al-8Zn alloy subjected to one pass of
equal channel angular pressing. J. Alloy. Compd. 762(2018)190-195.
[22] M. Suresh, A. Sharma, A.M. More, R. Kalsar, A. Bisht, N. Nayan, S. Suwas.
Effect of equal channel angular pressing (ECAP) on the evolution of texture,
microstructure and mechanical properties in the Al-Cu-Li alloy AA2195. J. Alloy.
Compd. 785 (2019) 972-983.
[23] L. Romero-Resendiz, I.A. Figueroa, C. Reyes-Ruiz, J.M. Cabrera, C. Braham, G.
Gonzalez. Residual stresses and microstructural evolution of ECAPed AA2017.
Mater. Charact. 152 (2019) 44-57.
[24] M.S.Arun, Uday Chakkingal. A constitutive model to describe high temperature
flow behavior of AZ31B magnesium alloy processed by equal-channel angular
pressing. Mater. Sci. Eng. A 754 (2019) 659-673.
[25] M.A. Salevati, F. Akbaripanah, R. Mahmudi, K.H. Fekete, A. Heczel, J. Gubicza.
Comparison of the effects of isothermal equal channel angular pressing and
multi-directional forging on mechanical properties of AM60 magnesium alloy.
30
Mater. Sci. Eng. A 776 (2020) 139002.
[26] I. Mazurina, T. Sakai, H. Miura, O. Sitdikov, R. Kaibyshev, Grain refinement in
aluminum alloy 2219 during ECAP at 250°C, Mater. Sci. Eng. A 473 (2008)
297-305.
[27] I. Mazurina, T. Sakai, H. Miura, O. Sitdikov, R. Kaibyshev, Effect of deformation
temperature on microstructure evolution in aluminum alloy 2219 during hot ECAP,
Mater. Sci. Eng. A 486(2008) 662-671.
[28] M. R. Barnett, X. Ma, A. Oudin, Generating Misorientations in Shear
Deformation Structures, Mater. Trans. 45 (2004), 2151-2156.
[29] [16] S. B, Davenport, R.L. Higginson, Strain path effects under hot working: an
introduction, J. Mater. Process. Tech. 98 (2000) 267-291.
[30] R. Z. Valiev, T. G. Langdon. Principles of equal-channel angular pressing as a
processing tool for grain refinement. Prog. Mater. Sci. 51 (2006) 881-981.
[31] H. H. Lee, W. Kim, K. C. Jung, S. Seo, J. K. Lee, H. L. Park, K.T. Park, H. S.
Kim. Circumferential twisting during route B equal-channel angular pressing. J.
Mater. Process. Tech. 259 (2018) 305-311.
[32] E. H. El-Shenawy. Physical simulation technology for thermo-mechanical
processing of metallic alloys using Gleeble system. Materials Today: Proceedings,
2020. https://doi.org/10.1016/j.matpr.2019.12.339.
[33] G. Xiao, N. Zhu, J. Long, Q. Xia, W. Chen, Research on precise control of
microstructure and mechanical properties of Ni-based superalloy cylindrical parts
during hot backward flow spinning, J. Manuf. Process. 34 (2018)140-147.
[34] A. Vinogradov, Y. Estrin. Analytical and numerical approaches to modelling
severe plastic deformation. Prog. Mater. Sci. 95 (2018) 172-242.
[35] G.D. Satish, N.K. Singh, R.K. Ohdar. Preform optimization of pad section of
front axle beam using DEFORM. J. Mater. Process. Tech. 203 (2008) 102-106.
[36] P. Vishwakarma, A. Sharma, 3D Finite Element Analysis of milling process for
non-ferrous metal using deform-3D. Materials Today: Proceedings 26 (2020)
525-528.
[37] M. Zouari, R. E. Logé, N. Bozzolo. In Situ Characterization of Inconel 718
31
Post-Dynamic Recrystallization within a Scanning Electron Microscope. Metals
7(2017) 476.
[38] M. P. Black, R. L. Higginson, C. M. Sellars, Effect of strain path on
recrystallisation kinetics during hot rolling of Al-Mn, Mater. Sci. Tech-lond 17
(2001) 1055-1060.
[39] X. Fan, X. Zeng, H. Yang, P. Gao, M. Meng, R. Zuo, P. Lei. Deformation
banding in β working of two-phase TA15 titanium alloy. Trans. Nonferrous Met.
Soc. China 27(2017) 2390-2399.
[40] N. Nayan, N. P. Gurao, S.V.S. Narayana Murty, A. K. Jha, B. Pant, S.C. Sharma,
K. M. George. Microstructure and micro-texture evolution during large strain
deformation of an aluminium-copper-lithium alloy AA 2195, Mater. Des. 65 (2015)
862-868.
[41] V. M. Segal, Severe plastic deformation: simple shear versus pure shear, Mater.
Sci. Eng. A 338 (2002) 331-344.
[42] V. M. Segal, Equivalent and effective strains during severe plastic deformation
(SPD), Phil. Mag. Lett. 2019, DOI: 10.1080/09500839.2019.1584411.
[43] Q. Liu, N. Hansen, Macroscopic and Microscopic Subdivision of a Cold-Rolled
Aluminium Single Crystal of Cubic Orientation, Proc. R. Soc. Lond. A 454 (1998)
2555-2592.
[44] H. W. Son, C. H. Cho, J. C. Lee, K. H. Yeon, J. W. Lee, H. S. Park, S. K. Hyun.
Deformation banding and static recrystallization in high-strain-rate-torsioned
Al-Mg alloy. J. Alloy. Comp. 814 (2020) 152311.
[45] P. B. Berbon, M. Furukawa, Z. Horita, M. Nemoto, T. G. Langdon. Influence of
Pressing Speed on Microstructural Development in Equal-Channel Angular
Pressing. Metall. Mater. Trans. A 30A (1999) 1989-1997.

32
Declaration of interests

The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

None

You might also like