You are on page 1of 29

A

PROJECT REPORT
ON

COLD FORGING OF Al-Mg ALLOY


In fulfilment of the requirements for the degree of
BACHELOR IN TECHNOLOGY
In
Metallurgy & Materials Engineering
NIFFT, RANCHI
Under the supervision of
Mr. KOUSHIK SIKDAR
Assistant professor,

Dept. of Materials & Metallurgical Engineering


NIFFT, RANCHI
Submitted by:
SATYAM KUMAR (203/16)
SAURABH SHASHANK (219/16)

NATIONAL INSTITUTE OF FOUNDRY & FORGE


TECHNOLOGY
HATIA, RANCHI-834003

1
CERTIFICATE

This is to certify that the seminar report entitled “Cold Forging of Al-Mg alloy” submitted
by

Mr. SATYAM KUMAR (203/16) &

Mr. SAURABH SHASHANK (219/16)

In partial fulfillment of the degree of BACHELOR OF TECHNOLOGY in Metallurgy &


Materials Engineering is a Bonafide seminar work carried out by him under my guidance.

In my opinion, the work fulfills the requirement for which it is being submitted and I wish
him best of luck for his bright future.

Date: Mr. Koushik SIkdar

Place: NIFFT, Ranchi Assistant Professor,

Dept. of Materials and Metallurgical


Engg.

NIFFT, Ranchi

2
LIST OF CONTENT

ABSTRACT ………………………………………………………. 4

ACKNOWLEDGEMENT…………………………………………. 5

LIST OF FIGURES……………………………………………….6-7
LIST OF TABLES …………………………………..….................8

CHAPTER 1: LITERATURE REVIEW…………………………9-12

1.1 INTRODUCTION
1.2 PRINCIPAL USES OF MAGNESIUM
1.3 PHYSICAL AND CHEMICAL PROPERTIES OF PURE Mg
1.4 DEFORMATION CHARACTERISTICS OF MAGNESIUM

CHAPTER 2: PROJECT ANALYSIS…………………………13-18

2.1 PRODUCTION OF WORK PIECE


2.2 SAMPLE PREPARATION
2.3 HARDNESS MEASUREMENTS
2.4 EFFECT OF SOLUTE ELEMENTS ON C/A RATIO

CHAPTER 3: CONCLUSION…………………………………….19

CHAPTER 4: REFERENCES ……………………………….20-21

3
ABSTRACT

The microstructural development in an AZ31 magnesium alloy


during cold multi-directional forging followed by annealing is
investigated in a wide range of cumulative strains up to 5. The
kinetics of recrystallization is accelerated accompanying with
increase in the Johnson–Mehl–Avrami–Kolmogorov slope from
2.6 to 4.7 as well as an appearance of a linear Johnson–Mehl–
Avrami–Kolmogorov relationship. The grain size evolved after
annealing decreases with repeated multi-directional forging and
approaches around 1 lm in high strain. A strong initial texture is
broken gradually with repeated multi-directional forging and
almost disappears at severe high strain of 5. These can be
because various variants of deformation twins and their
intersections introduced by multi-directional forging are
homogeneously developed in high density by repeated changing
of the loading directions during multi-directional forging. The
annealing process and the mechanisms occurring after cold multi-
directional forging are discussed comparing with those after hot
multi-directional forging.

4
ACKNOWLEGEMENT

I have Great pleasure in expressing my profound sense of


Gratitude to Asst. Professor DR.KAUSHIK SIKDAR,
Department of METALLURGY & MATERIALS, NIFFT, RANCHI
for his unforgettable cooperation, inspiration for our project work
and also we are grateful to him for his precious suggestions,
encouragement and assistance throughout the course of our
project work and also our career and character building in mean
time. I sincerely thank to all dept. staffs for their assistance during
our project work.
I am expressing my sincere thanks to all library staff of NIFFT,
who directly or indirectly helped us during our project practical
work.

I also thank all the well-wishers whose cooperation and valuable


suggestions have helped us in completing the project successfully

SATYAM KUMAR (203/16) &


SAURABH SHASHANK (219/16)
B.TECH, MME

5
LIST OF FIGURES
Figure 1: The Mg-rich end of the Mg–Al phase diagram.

Figure 2: Scheme of forging set for: (a) h/d = 0.8 and (b) h/d = 2.5
where 1 is the hollow punch; 2 container; 3 thermocouple; 4
mandrel; 5 work piece; 6 heating and 7 is the ejector.

Figure 3: Plots of load for forging as-cast AZ31 Mg alloy using


billet h/d = 0.8; forgings obtained at: (a) 300, (b) 230, (c) 200 ◦C.

Figure 4: Plot of load for forging as-cast AZ31 Mg alloy using billet
h/d = 2.5; forgings obtained at: (a) 200 and (b) 280◦C.

Figure 5: Forging pressures required for the upsetting of


magnesium alloy billets between flat dies. (a) Alloy AZ80A; strain
rate: 0.11 s-1. (b) Alloy AZ61A; strain rate: 0.11 s-1. (c) Alloy
AZ31B; strain rate: 0.7 s-1.

Figure 6: Effect of forging temperature on forging pressure


required for upsetting to a 10% reduction at hydraulic press
speeds for a magnesium alloy and an aluminum alloy

Figure 7: True stress–true strain (σ-ε) curves for AZ31 Mg alloy


obtained during MDF at a strain rate of 3X10 -3 s-1 and at 300 K. A
pass strain was Δε = 0.1. (a) Each σ-ε curve in 1 pass to 49
passes. (b) The envelopes of σL (or σT) – σ-ε curves, in which a
peak flow stress in each σ-ε curve is plotted against cumulative
strain (ΣΔε). The peak stresses (σL and σT) represented by a solid
and broken lines are the flow curves measured for the longitudinal
(extruded) and transverse directions of the Mg rod, respectively

Figure 8: Microstructures developed at cumulative strains of (a)


ΣΔε = 0.5 and (b) ΣΔε = 1.5 with Δε = 0.05, (c) ΣΔε = 3 and (d)
ΣΔε = 5 with Δε = 0.1 for AZ31 Mg alloy MDFed at 300 K.
6
Figure 9: Effect of severe high strain on the Vickers hardness vs.
annealing time curve for AZ31 Mg alloy processed by MDF at 300
K followed by annealing at 473 K.

Figure 10: Relationships between room temperature hardness


measured just after deformation (i.e. Xrex = 0) or full
recrystallization (i.e. Xrex = 1) for Mg alloy MDFed at 300 K
followed by annealing at 473 K

Figure 11: Fractional recrystallization vs. time (Xrex–t) curves at


473 K for AZ31 Mg alloy MDFed to cumulative strains of 1.5, 3
and 5 with a pass strain Δε of 0.1 at 300 K.

Figure 12: Fractional recrystallization vs. annealing time (Xrex–t)


curves (solid lines) and fractional softening vs. annealing time
(XH–t) curves (broken lines) for AZ31 Mg alloy MDFed to severe
high strains with a pass strains Δε of 0.1. The X H obtained by the
Vickers hardness test at 300 K is calculated by using the Eq. (1).
The typical results for the Mg alloy MDFed to ΣΔε = 0.2 and 1.5
with Δε = 0.05 are shown for comparison (10).

Figure 13: The JMAK plots for the recrystallization curves in Figs.
5 and 6, which show the effect of prior cumulative strain ΣΔε
ranging from 0.2 to 5 for Mg AZ31 MDFed at 300 K followed by
annealing at 473K.

Figure 14: Fully recrystallized (i.e. Xrex = 1) microstructures


developed at 473 K for AZ31 Mg alloy cold MDFed to cumulative
strains of (a) ΣΔε = 0.2, (b) ΣΔε = 0.8, (c) ΣΔε = 1.5 with Δε = 0.05
and (d) ΣΔε = 1.5, (e) ΣΔε = 3 and (f) ΣΔε = 5 with Δε = 0.1.

7
Figure 15: Cumulative strain dependence of the recrystallized
grain size evolved after full recrystallization (X rex = 1) for AZ31 Mg
alloy MDFed at 300 K with Δε = 0.05 and Δε = 0.1. The results for
the Mg alloy obtained during hot MDFed with Δε = 0.8 are shown
by a broken line for comparison (14).

Figure 16: Inverse pole figures of the initial texture (a) and a
deformation texture developed from ΣΔε = 1.5(b) to ΣΔε = 5 (c)
and the changes in the distribution of wire texture with the basal
planes lying parallel to the extruded direction (ε = 0) with cyclic
MDF of AZ31 Mg alloy (d). The result for the AZ31 Mg alloy
compressed in single-pass to ε = 1.2 at 623 K (broken line) is
shown for comparison.

LIST OF TABLES
Table 1: Classification of alloys in order of increasing forging
difficulty
Table 2: Forgeability of various work metals by high-energy rate
forging
Table 3: Recommended forging temperature ranges for
magnesium alloys
Table 4: Flash design and forging load

LITERATURE REVIEW

INTRODUCTION:
8
Cold Forging. High speeds in cold forging improve lubrication. Improved
lubrication lowers the frictional forces, and this in turn results in improved
metal flow and surface finish of the components. The process is practically
adiabatic; therefore, with soft materials, such as aluminum and copper,
some softening occurs. Forging forces and pressures are generally higher
than those obtained at conventional speeds, although energy requirements
may well be lower.

Table 1 Classification of alloys in order of increasing forging difficulty

Work metal Forgeability

Aluminium alloys Excellent for alloys such as 2014, 2024, and others hardened with copper. For
7075 and 7079, forging temperatures and reductions must be adjusted downward
to prevent rupturing. No improvements in shape detail obtained
Magnesium alloys Poor for such alloys as AZ31B and AZ80A. Alloys HM21A and HM31A are as
forgeable as in presses if temperatures are adjusted downward.

Table 2 Forgeability of various work metals by high-energy rate forging

The forgeability of magnesium alloys depends on three factors: the solidus


temperature of the alloy, the deformation rate, and the grain size. Only
forging-grade billet or bar stock should be used in order to ensure good
workability. This type of product has been conditioned and inspected to
eliminate surface defects that could open during forging, and it has been
homogenized by the supplier to ensure good forgeability. Table 3 lists the
compositions of magnesium alloys that are commonly forged, along with
their forging temperatures.
9
Table 3 Recommended forging temperature ranges for magnesium alloys

A high resistance to corrosion characterizes aluminium–magnesium alloys,


with Mg contents of 4–10%. Only Al–10% Mg castings respond to heat-
treatment. Aluminium–zinc–magnesium alloys have a relatively high
eutectic melting point, which make them suitable for castings that are
assembled by brazing. The as-cast alloys respond to both natural and
artificial ageing.
The global manufacturing using light metals and alloys is on the edge of
substantial growth and opportunity. Magnesium (Mg) being the lightest
metal, in the form of alloys, especially intrigues researchers at academic
and industrial level. There are, however, still some challenges which limit
10
practical application of Mg alloys. One of them is the poor formability at
ambient temperature due to hexagonal close packed structure with limited
number of operative slip systems. Microstructural refinement is an effective
way for improving the ductility and strength of Mg wrought products [1,2]
and several methods have been adopted to achieve it, such as equal
channel angular extrusion [3], accumulative roll bonding [4], multi-
directional forging [5] and accumulative back extrusion [6]. The authors [7]
found that, during hot deformation of an AZ31 Mg alloy, ultrafine-grained
(UFG) structures are dynamically formed through grain fragmentation by
kink bands in relatively
low strain, and concluded that the hot deformation can be controlled by in
situ or continuous dynamic recrystallization (cDRX). They also studied the
annealing behaviours that appear in the UFGed Mg alloys processed by
hot deformation, and concluded that the annealing process can be
controlled mainly by grain coarsening assisted by recovery accompanying
with few changes in texture, that is continuous static recrystallization
(cSRX) [8]. As far as the authors know, however, there are few systematic
reports available on the annealing behaviours of cold-deformed Mg alloys
and especially on those after large strain deformation. The authors
overcame the latter by using multi-directional forging (MDF) [5] and have
studied systematically the annealing behaviour of cold-deformed Mg alloy,
and examined the effect of prior strain, pass strain of MDF and temperature
on the annealing behaviour for the Mg alloy [9,10]. The aims of the present
work are to investigate the structural development taking place in the same
AZ31 Mg alloy during severe
Cold deformation and subsequent isothermal annealing. The MDF with a
pass strain of 0.1 was carried out at room temperature in a wide range of
severe strains up to 5. The behaviours of recrystallization occurring after
severe cold deformation were studied comparing with those appearing in
low strain [9, 10]. The effect of severe plastic deformation on grain
refinement and texture changes as well as of the subsequent annealing on
them is analysed in detail, and the mechanisms operating during the
processes are discussed.

11
Fig 1 The Mg-rich end of the Mg–Al phase diagram.

Experimental procedure

A commercially produced AZ31 Mg alloy was provided as a hot extruded


rod with a diameter of 19 mm. The chemical composition was as follows: Al
2.68, Zn 0.75, Mn 0.68, Cu 0.001, Si 0.003, Fe 0.003, and balance Mg (all
in mass %). A rectangular sample with a dimension of 8.9 mm, 8.4 mm and
8.0 mm (the axis ratio, 1.1:1.05:1.0) were machined from the rod parallel to
the extrusion direction and used for MDF with a pass strain of Δε = 0.1. A
cubic sample with 8 mm in each length was partly used for MDF with a
pass strain of Δε = 0.05. These samples were annealed at 733 K for 7.2 ks
and then furnace cooled, leading to the evolution of equiaxed grains with
an average size of about 22.5 μm. MDF was carried out in repeated
compression at a constant true strain rate of 3 X10 -3 s-1 and at around 300
K. The samples were deformed up to cumulative strain ΣΔε of 5 by
repeated MDF with a pass strain Δε of 0.1 with changing the loading axis at
angle of 90 through three mutually perpendicular axes (i.e. X to Y to Z to X
to . . ..) from pass to pass. MDF was applied because it has been
demonstrated to be the most effective methods for homogeneous
development of such high-density twins and their intersections [10]. The
compression axis in the 1st pass was always parallel to the extruded
direction of the Mg rod. Deformed samples were cut to plates with 2–3 mm
thickness parallel to the last compression axis, and then isothermally
annealed for a period between 0.1 ks and 100 ks at 473 K. Each plate was
mechanically and electrolytically polished and then etched in a solution of
12
6% picric acid and 94% methanol. The microstructures evolved during
deformation and subsequent annealing was examined by using optical
microscopy (OM) and orientation imaging microscopy (OIM). The grain size
and the fraction of recrystallization were measured using the linear line
interception and point-counting methods applied to those microstructures.
Crystallographic orientation and texture changes were examined by
scanning electron microscope (SEM) or OIM. The Vickers hardness tests
were carried out on each plate at room temperature under a condition of 3
N.

Fig.2. Scheme of forging set for: (a) h/d = 0.8 and (b) h/d = 2.5 where 1 is the hollow punch; 2 container;
3 thermocouple; 4 mandrel; 5 work piece; 6 heating and 7 is the ejector.

Machines and Dies

Machines. Hydraulic presses or slow-action mechanical presses are the


most commonly used machines for the open-die and closed-die forging of
magnesium alloys. In these machines, magnesium alloys can be forged
with small corners and fillets and with thin web or panel sections. Corner
radii of 1.6 mm ( in.), fillet radii of 4.8 mm ( in.), and panels or webs 3.2 mm
( in.) thick are not uncommon. The draft angles required for extraction of
the forgings from the dies can be held to 3° or less. Magnesium alloys are
seldom hammer forged or forged in a rapid-action press, because they will
crack unless exacting procedures are used. Alloys ZK60A, AZ31B, and
HM21A are more easily forged by these methods than AZ80A, which is
13
extremely difficult to forge. Cracking can occur also in moderately severe,
unsupported bending. Magnesium alloys generally flow laterally rather than
longitudinally. This characteristic must be considered in the design of tools.
Dies. Because forging temperatures for magnesium alloys are relatively low
(Table 1), conventional low-alloy hot-work tool steels are satisfactory
materials for forging dies. Dies are finished to a smooth, highly polished
surface to prevent surface roughness, scratches, or imperfections on the
forging. The high polish also promotes metal flow during forging. Wet
abrasive blasting and extremely fine abrasive finishing papers are used to
produce a smooth finish on die-impression surfaces.

Fig.3. Plots of load for forging as-cast AZ31 Mg alloy using billet h/d = 0.8; forgings obtained at: (a) 300,
(b) 230, (c) 200 ◦C.

Fig. 4. Plot of load for forging as-cast AZ31 Mg alloy using billet h/d = 2.5; forgings obtained at: (a) 200
and (b) 280◦C.

Forging Practice
14
Forging pressures for the upsetting of magnesium alloy billets between flat
dies are shown in Fig. 5. At normal press forging speeds, the forging
pressure increases and then decreases slightly with increased upset
reduction, probably because work metal temperature increases during
forging.

Fig. 5 Forging pressures required for the upsetting of magnesium alloy billets between flat dies. (a) Alloy
AZ80A; strain rate: 0.11 s-1. (b) Alloy AZ61A; strain rate: 0.11 s-1. (c) Alloy AZ31B; strain rate: 0.7 s-1.

Forging load and pressure in closed-die forging vary greatly with the shape
being forged. Relatively small changes in flash dimensions, for example,
can result in appreciable changes in the forging load:

15
Table 4: Flash design and forging load

Forging temperature has a marked effect on forging pressure requirements.


Figure 6 shows the magnitude of this effect for magnesium alloy AZ31B in
comparison with aluminium alloy 6061. As Table 2 shows, at normal forging
temperatures, AZ31B requires greater forging pressure than carbon steel,
alloy steel or aluminium and requires less than stainless steel. Magnesium
alloys flow less readily than aluminium into deep vertical die cavities. If two
dies are needed for a typical aluminium structural forging, the same part in
a magnesium alloy may require three dies for successful forging.

Fig. 6 Effect of forging temperature on forging pressure required for upsetting to a 10% reduction at
hydraulic press speeds for a magnesium alloy and an aluminum alloy.

16
Results and discussion

3.1. Deformation behaviour in high cumulative strain


The σ ε curves during MDF at room temperature are shown in Fig. 1. Each
σ-ε curve during MDF of 1 pass to 49 passes and the σ-ε curves in
successive forging up to 50 passes are plotted against strain and
accumulative strain in Fig. 1a and b, respectively. It is clearly seen that the
Mg alloy can be stably deformed up to severe large strain of ΣΔε = 5 during
cold MDF, although it fractured in a brittle fashion at ε≈0.15 during single
pass compression [9, 10]. The σ L and σT in Fig. 1b are the flow stress
peaks appearing in each σ-ε curve when the Mg samples were deformed
parallel or perpendicular to the extrusion direction, respectively. The
enveloped flow curves for the σL and σT are shown by solid and dashed
lines in Fig. 1b. These flow curves show a rapid strain hardening in low
strain below 2–3, followed by its gradual decrease appearing in high strain.
It is clearly seen in Fig. 1b that the σ L is always higher than the σ T in low
strain and approaches a similar value in high strain. The former may result
in a strong extrusion texture of the present Mg alloy, i.e. the basal plane
roughly parallel to the extrusion (longitudinal) direction [7]. The latter result,
i.e. σL ≈ σT, in high strain, therefore, may suggest some disappearance of
such a strong texture accompanied by repeated MDF to severe large
strain. This will be discussed in detail in later. A typical microstructure
evolved at ΣΔε = 3 and ΣΔε = 5 with Δε = 0.1 is represented in Fig. 2
together with that developed at ΣΔε = 0.5 and ΣΔε = 1.5 with Δε = 0.05.
During cold MDF to low strains, deformation twins with rather broad width
were evolved and crossing the original grains, as seen in Fig. 2a and b.
With further MDF, several kinds of twins evolved become much narrow and
finer accompanying with their frequent intersection with one another.
Microstructures composed of high density and finer twins are uniformly
developed in original grain interiors at ΣΔε = 3 (Fig. 2c and d), where the
initial grain boundaries are hardly visible in the microstructures developed.
It is pointed out in elsewhere [9,10] that more than two types of twins
developing and intersecting with one another are classified roughly into two
typical types, i.e. the so-called tension twins with {10-12} and the
compression twins with {10-11}. These twin intersections may become the
nucleation sites for recrystallization, because high strain concentrations are
evolved around them [11].

17
Fig. 7. True stress–true strain (σ-ε) curves for AZ31 Mg alloy obtained during MDF at a strain rate of
3X10-3 s-1 and at 300 K. A pass strain was Δε = 0.1. (a) Each σ-ε curve in 1 pass to 49 passes. (b) The
envelopes of σL (or σT) – σ-ε curves, in which a peak flow stress in each σ-ε curve is plotted against
cumulative strain (ΣΔε). The peak stresses (σL and σT) represented by a solid and broken lines are the
flow curves measured for the longitudinal (extruded) and transverse directions of the Mg rod, respectively

a) b)

c) d)
Fig. 8. Microstructures developed at cumulative strains of (a) ΣΔε = 0.5 and (b) ΣΔε = 1.5 with Δε = 0.05,
(c) ΣΔε = 3 and (d) ΣΔε = 5 with Δε = 0.1 for AZ31 Mg alloy MDFed at 300 K.

18
Fig. 9. Effect of severe high strain on the Vickers hardness vs. annealing time curve for AZ31 Mg alloy
processed by MDF at 300 K followed by annealing at 473 K.

Fig. 10. Relationships between room temperature hardness measured just after deformation (i.e. X rex = 0)
or full recrystallization (i.e. Xrex = 1) for Mg alloy MDFed at 300 K followed by annealing at 473 K

3.2. Annealing behaviour after MDF in severe cumulative strain

The Vickers hardness versus annealing time (Hv–t) curves for the AZ31 Mg
alloy, cold MDFed to ΣΔε 1.5, 3 and 5 followed by isothermal annealing at
473 K, are represented in Fig. 3. The Hv–t curves show characteristic
changes occurring in three steps with reheating time, i.e. a roughly
constant or slightly decrease in Hv in the first stage, a following rapid drop
in the second one and finally again a constant value of Hv in the third one
at longer times. The rapid drop in Hv occurs in shorter reheating time with
increase in ΣΔε and is resulted from discontinuous static recrystallization
(dSRX) in the stage 2 [9, 10]. It is seen also in Fig. 3 that increasing prior

19
strain ΣΔε brings about an increases in Hv for not only an as deformed
state, but also a fully annealed one at long reheating times. This is
confirmed in Fig. 4, where both the harnesses for the fractional
recrystallization Xrex of 0 and 1 are plotted against prior cumulative strain.
The Hv for Xrex = 0, i.e. as-deformed state, increases rapidly with strain in
low strain below ΣΔε = 2–3 and then gradually in high strain. Such a
behaviour of Hv is roughly similar to those for the envelope curves for σ L
and σT in Fig. 1b. The Hv for X rex = 1 steadily increases with ΣΔε,
suggesting that the grain size evolved after full annealing may decrease
with prior strain. This will be discussed in detail in later. The fractional
recrystallization versus annealing time (X rex–t) curves for three cumulative
strains of 1.5, 3 and 5 are depicted in Fig 5. Each X rex–t curve shows a
gradual increase at early times and then a sharp and rapid increase in a
short period of time followed by a full recrystallization (i.e. X rex = 1). Such a
behaviour of the Xrex–t curves may roughly correspond to that for Hv–t
curves in Fig. 3.

3.3. Kinetics of recrystallization

The fractional softening XH was obtained by using the data of Fig. 3 and the
following equation [12, 13],
XH = (Hε – Ht) / (Hε – H0)
Where Hε, Ht and H0 are the hardness for as-deformed structures evolved
at various strains (e), an intermediate structure at annealing time (t) and a
fully annealed one at long times, respectively. Fig. 6 shows the
relationships between XH and t and between Xrex and t indicated by dashed
and solid lines, respectively. To clarify the effect of prior strain on these
curves in a wide range of prior strain, the results for a pass strain of 0.05
described in [10] are also given in Fig. 6 for comparison. It is clearly seen in
Fig. 6 that both the curves for XH and Xrex are shifted to shorter reheating
time with increase in prior cumulative strain and this effect appears more
clearly in low strain below ΣΔε = 1.5. It is noted in the result for ΣΔε = 1.5
that, with increase in pass stain Δε, both the curves for X H and Xrex are
clearly shifted to shorten time. It is also noted in Fig. 6 that the XH is
always higher than the Xrex during early annealing, but approaches the
Xrex at long times. The difference between X H and Xrex, ΔXH–rex, appears to
be larger at early stages of annealing (and at lower cumulative strains), and
decreases with time and becomes negligible small or nearly zero at long
times. It is interesting to note that the ΔX H–rex for ΣΔε = 5 is nearly zero in
the periods of time investigated within the experimental scatters. It is
20
discussed in [10] that such behaviours of the X H and the ΔXH–rex may result
from any inhomogeneity of deformed microstructures developed during
MDF, as discussed in detail in later. The kinetics recrystallization are
discussed sometimes by using the following Johnson–Mehl–Avrami–
Kolmogorv (JMAK) equation for isothermal recrystallization [13],

X = 1- exp (-Atk)

Where the exponent (k) is called as the JMAK exponent. Fig. 7 shows the
JMAK plots obtained from the results in Figs. 5 and 6. To clarify the effect
of prior strain in a wide range of strain, the results for a pass strain of 0.05
[10] are represented by broken lines in Fig. 7. It is remarkable to note that
the JMAK plots for Δε = 0.1 indicated by solid line are approximated by
linear lines within the experimental frame investigated, while those for Δε =
0.05 shown by broken line break clearly at long times. The average slope,
i.e. the value of k in Eq. (2), for Δε = 0.1 is around 4.7 and that for Δε = 0.05
is 2.6 in an early stage of annealing, both of which hardly change with prior
strain in the strain range investigated. These behaviours were discussed in
[10] in association with decrease in the inhomogeneity or increase in the
homogeneity of deformed microstructures developed by repeated MDF, as
shown in Fig. 2. It is concluded that an increase of ΣΔε (and also Δε) under
MDF conditions can result in the formation of microstructures containing
higher densities of much finer twins, followed by a development of higher
flow stresses as well as higher harnesses in high cumulative strain (Figs. 1
and 4). These may lead to an acceleration of the kinetics of recrystallization
as well as an increase of the JMAK exponent from 2.6 to 4.7. These may
also bring about an increase in a critical X rex at which the JMAK plots for Δε
= 0.05 break at long times and an appearance of a linear JMAK plot for Δε
= 0.1 in Fig. 7.

21
Fig. 11. Fractional recrystallization vs. time (X rex–t) curves at 473 K for AZ31 Mg alloy MDFed to
cumulative strains of 1.5, 3 and 5 with a pass strain Δε of 0.1 at 300 K.
3.4. Microstructures developed during severe cold MDF and
annealing

Fig. 8 shows the typical grain structures that evolved after full annealing of
AZ31 subjected to cold MDF from ΣΔε = 0.2 to ΣΔε = 5 at 473K with Δε =
0.05 (Fig. 8a–c) and Δε = 0.1 (Fig. 8d–f). A finer-grained structure
developed with increasing strain. The average grain size evolved after full
recrystallization (drex) was measured in the AZ31 alloy MDFed with two
pass strains of 0.05 and 0.1. The d rex versus prior strain ΣΔε curves are
represented by solid lines in Fig. 9. It is clearly seen that the d rex decreases
with increase in prior strain and more rapidly with increasing of pass strain,
while the drex for Δε = 0.1 looks like approaching a saturation value of
around 1 μm in severe cumulative strain. These results may be caused by
the deformed microstructures containing higher density of finer twins
developed by changing of forging axis during cold MDF (Fig. 2) [9, 10]. The
density of deformation twins and the intersections, being nucleation sites
for dSRX, increase with increasing ΣΔε and De, and so are distributed
more homogeneously in severe high strain. These may lead to the
development of finer recrystallized grain size in high cumulative strain and
the changes in the some annealing characteristics, e.g. ΔX H–rex -0 (Fig. 6)
and a linear JMAK plots with a higher exponent of 4.7 (Fig. 7) in severe
cumulative strain. By the way, the dynamic grain size versus cumulative
strain curve for the same AZ31 Mg alloy under hot MDF conditions [14] is
represented by dashed line in Fig. 8 for comparison. Here MDF with a pass
strain of 0.8 was carried out with decreasing temperature from 623 K in first
pass to 403 K in 7th pass. The drex evolved during hot MDF decreases
drastically with repeated MDF and may approach even below 0.2 μm at
large cumulative strains. It is concluded, therefore, that grain refinement
takes place more effectively during hot MDF under dropping temperature
conditions in comparison with that occurring during cold MDF followed by
annealing. Next the mechanisms of grain refinement operating during hot
deformation and during cold deformation followed by annealing will be
discussed in comparison. During hot deformation of Mg alloys, grain
refinement taking place results from grain fragmentation due to the
formation of high angle boundaries of kink bands, finally followed by the
formation of new fine grains in high strain, which is continuous DRX [7]. In
contrast, new grain formation after cold MDF followed by annealing are
resulted from the operation of conventional dSRX even after severe

22
cumulative deformation, as confirmed in [9,10] and the present study. It has
been recently reported in [15–17] that severe plastic deformation (SPD) of
conventional cubic materials results in the formation of ultrafine grains
(UFGs) with below submicron sizes. The latters can be associated with the
development of various kinds of deformation bands in grain interiors, such
as micro shear band (MSB) and kink band with persistent nature. It is
discussed in elsewhere [7,16,17], that the formation of UFGs based on
MSBs and kink bands can be controlled by strain-induced continuous
reactions assisted by dynamic recovery, that is called as continuous DRX
(cDRX). In Mg alloys with a few slip systems, kink bands in replacement of
MSBs are frequently evolved at relatively low strains below 1 during hot
deformation [7]. High density dislocations evolved can accumulate and
rearrange inside the individual boundaries of MSB or KB under dynamic
recovery operating conditions, and then these individual boundaries can
transform into continuous ones connected with each other during
subsequent severe plastic deformation. Finally, the micro crystallites
fragmented and surrounded by MSBs or KBs may transform into fine grains
with high angle boundaries during further deformation. During cold
deformation of Mg alloys, in contrast, high density twins and dislocations
are introduced in grain interiors (Fig. 2). It is well known that deformation
twins have generally coherent and incoherent boundaries in the
neighbouring regions and coherent boundaries contain quite a few
dislocations [11]. Such coherent twin boundaries are considered to be
hardly transformed into continuous and so incoherent ones, and so to be
hard to surround the micro crystallites fragmented by twins at ambient
temperature. This may be why UFGs are hardly evolved even at severe
high strains in Mg alloys during cold MDF. During hot deformation of Mg
alloys having deformation twins, however, UFGs may be evolved based on
such twin boundaries [18, 19]. Coherent twin boundaries can transform into
random angle ones through frequent cross slip and rearrangement of
dislocations by assistance of dynamic recovery during hot deformation.

23
Fig. 12. Fractional recrystallization vs. annealing time (X rex–t) curves (solid lines) and fractional softening
vs. annealing time (XH–t) curves (broken lines) for AZ31 Mg alloy MDFed to severe high strains with a
pass strains Δε of 0.1. The X H obtained by the Vickers hardness test at 300 K is calculated by using the
Eq. (1). The typical results for the Mg alloy MDFed to ΣΔε = 0.2 and 1.5 with Δε = 0.05 are shown for
comparison (10).

Fig. 13. The JMAK plots for the recrystallization curves in Figs. 5 and 6, which show the effect of prior
cumulative strain ΣΔε ranging from 0.2 to 5 for Mg AZ31 MDFed at 300 K followed by annealing at 473K.

Fig. 14. Fully recrystallized (i.e. Xrex = 1) microstructures developed at 473 K for AZ31 Mg alloy cold
MDFed to cumulative strains of (a) ΣΔε = 0.2, (b) ΣΔε = 0.8, (c) ΣΔε = 1.5 with Δε = 0.05 and (d) ΣΔε =
1.5, (e) ΣΔε = 3 and (f) ΣΔε = 5 with Δε = 0.1.

3.5. Micro textures developed during severe cold MDF

24
The as-hot-extruded AZ31 Mg alloy has a strong initial texture as described
in [7–9,10] and is confirmed from the data for ε = 0 in Fig. 10. Here the
relative intensity of the texture orientations measured in the inverse pole
figures Fig. 10a–c is plotted against φ (Fig. 10d), the angle between the
compression axis and the basal plane [20, 21]. During single pass hot
compression at 623 K, the alignment of the basal plane initially parallel to
the compression axis is rotated gradually by compression and approaches
perpendicular to compression axis at ε = 1.2, as indicated by broken line in
Fig. 10. Hot deformation of Mg alloy can be controlled primarily by the
motion of basal dislocation and their accumulation and rearrangement-
taking place inside kink bands [11]. The latters can contribute to the
formation of the basal (stable) texture as well as new fine grains in high
strain [7, 11, and 21].
It is important to note in Fig. 10, in contrast, that the initial strong texture of
the Mg alloy is broken gradually with repeated cold MDF and almost
disappears in severe large strain, e.g. at ΣΔε = 5. This is also supported by
the results of Fig. 1 that the σ L being higher than the σT in low strain
approaches the σT during repeated MDF. Recently, Miura et al. reported
similar results for a Mg alloy AZ61 deformed by cold MDF [22], and
explained that the relatively low textural evolution may be derived by the
small pass strains to avoid large crystal rotations and, in combination with
the changes in the forging axis and twinning, the grain orientation
randomization effectively proceeds with repeated MDF. It is concluded in
the present study, therefore, that cold MDF followed by annealing will be an
important technique not only for randomization and/or controlling of the
stable texture (Fig. 10), but also for the formation and controlling of UFGs
in Mg alloy (Fig. 9). During cyclic cold MDF, various variants of deformation
twins as well as their intersections may be developed more frequently due
to repeated changing of the loading directions during MDF, resulting in
randomization and controlling of the stable texture in Mg alloy.

25
Fig. 15 Cumulative strain dependence of the recrystallized grain size evolved after full recrystallization
(Xrex = 1) for AZ31 Mg alloy MDFed at 300 K with Δε = 0.05 and Δε = 0.1. The results for the Mg alloy
obtained during hot MDFed with Δε = 0.8 are shown by a broken line for comparison (14).

Fig. 16. Inverse pole figures of the initial texture (a) and a deformation texture developed from ΣΔε =
1.5(b) to ΣΔε = 5 (c), and the changes in the distribution of wire texture with the basal planes lying parallel
to the extruded direction (ε = 0) with cyclic MDF of AZ31 Mg alloy (d). The result for the AZ31 Mg alloy
compressed in single-pass to ε = 1.2 at 623 K (broken line) is shown for comparison.

Conclusions

26
An as-extruded AZ31 Mg alloy with a strong texture was processed by
MDF at ambient temperature, followed by annealing at 473 K, and the
structural development during the processes was studied in a wide range of
cumulative strains up to 5. The main results obtained are summarized as
follows.

(1) The Mg alloy is successfully deformed to severe high strain up to 5 at


ambient temperature by using MDF with a pass strain of 0.1. The density of
much finer deformation twins evolved increases with repeated MDF and
then the deformed microstructure is developed more homogeneously with
increase in cumulative strain.

(2) Cold MDF followed by annealing results in the formation of new grain
structure due to the operation of discontinuous static recrystallization. With
increase in cumulative strain and pass strain, the kinetics of
recrystallization is accelerated accompanying with increase in the JMAK
slope as well as an appearance of a linear JMAK relationship. The grain
size evolved after annealing decreases with repeated MDF and
approaches around 1 μm in high cumulative strain.

(3) A strong texture evolved in the extruded rod is broken gradually with
repeated MDF and almost disappears at severe high strain of 5. This can
be because various variants of deformation twins and their intersections
are introduced by MDF and homogeneously developed in high density
through repeated changing of the loading directions during MDF.

(4) Cold MDF of Mg alloy does not result in dynamic formation of new
grains even at severe high strain of 5, because coherent twin boundaries
hardly transform into incoherent grain boundaries. Hot MDF, in contrast,
leads to grain refinement at low strains below 1 accompanied by
continuous dynamic recrystallization.

REFERENCES

[1] Barnett MR, Keshavarz Z, Beer AG, Atwell D. Influence of grain size on
the
compressive deformation of wrought Mg–3Al–1Zn. Acta Mater
2004;52:5093–103.

27
[2] Ma AB, Jiang JH, Saito N, Shigematsu I, Yuan YC, Yang DH, et al.
Improving both
strength and ductility of a Mg alloy through a large number of ECAP
passes.
Mater Sci Eng A 2009;513–514:122–7.
[3] Kim WJ, Hong SI, Kim YS, Min SH, Jeong HT, Lee JD. Texture
development and
its effect on mechanical properties of an AZ61 Mg alloy fabricated by equal
channel angular pressing. Acta Mater 2003;51:3293–307.
[4] Pérez-Prado MT, Del Valle JA, Ruano OA. Grain refinement of Mg–Al–
Zn alloys
via accumulative roll bonding. Scripta Mater 2004;51:1093–7.
[5] Xing J, Soda H, Yang XY, Miura H, Sakai T. Ultra-fine grain
development in an
AZ31 magnesium alloy during multi-directional forging under decreasing
temperature conditions. Mater Trans 2005;46:1646–50.
[6] Fatemi-Varzaneh SM, Zarei-Hanzaki A. Processing of AZ31 magnesium
alloy by a new noble severe plastic deformation method. Mater Sci Eng A
2011;528:1334–9.
[7] Yang XY, Miura H, Sakai T. Dynamic evolution of new grains in AZ31
magnesium alloy during hot deformation. Mater Trans 2003;44:197–203.
[8] Yang XY, Miura H, Sakai T. Isochronal annealing behavior of AZ31
magnesium alloy after hot deformation. Mater Trans 2005;46:2981–7.
[9] Yang XY, Okabe Y, Miura H, Sakai T. Annealing of a AZ31 magnesium
alloy after interrupted cold deformation. Mater Des 2012;36:626–32.
[10] Yang XY, Okabe Y, Miura H, Sakai T. Effect of pass strain and
temperature on recrystallisation in AZ31 magnesium alloy after interrupted
cold deformation. J Mater Sci 2012;47:2823–30.
[11] Martin E, Jonas JJ. Evolution of microstructure and microtexture
during the hot deformation of Mg–3% Al. Acta Mater 2010;58:4253–66.
[12] Sakai T, Ohashi M, Chiba K, Jonas JJ. Recovery and recrystallization
of
polycrystalline nickel after hot working. Acta Metall 1988;36:1781–90.
[13] Humphreys FJ, Hatherly M. Recrystallization and related annealing
phenomena. 2nd ed. Oxford: Pergamon; 2005.
[14] Xing J, Yang XY, Miura H, Sakai T. Mechanical properties of AZ31
magnesium alloy after severe plastic deformation. Mater Trans
2008;49:69–75.
[15] Valiev RZ, Islamgaliev RK, Alexandrov IV. Bulk nanostructured
materials from severe plastic deformation. Prog Mater Sci 2000;45:103–89.
28
[16] Sitdikov O, Sakai T, Goloborodko A, Miura H, Kaibyshev R. Grain
refinement in coarse-grained 7475 Al alloy during severe hot forging.
Philos Mag 2005;85:1159–75.
[17] Sakai T, Belyakov A, Miura H. Ultrafine grain formation in ferritic
stainless steel during severe plastic deformation. Metall Mater Trans A
2008;39:2206–14.
[18] Sitdikov O, Kaibyshev K, Sakai T. Dynamic recrystallization based on
twinning in coarse-grained magnesium. Mater Sci Forum 2003;419–
422:521–6.
[19] Miura H, Nakao Y, Sakai T. Enhanced grain refinement by mechanical
twinning in a bulk Cu-30 mass% Zn during multi-directional forging. Mater
Trans 2007;48:2539–41.
[20] Ion SE, Humphreys FJ, White SH. Dynamic recrystallisation and the
development of microstructure during the high temperature deformation of
magnesium. Acta Metall 1982;30:1909–19.
[21] Yang XY, Ji Z, Miura H, Sakai T. Dynamic recrystallization and texture
development during hot deformation of AZ31 magnesium alloy. Trans
Nonferrous Met Soc China 2009;19:55–60.
[22] Miura H, Maruoka T, Yang XY, Jonas JJ. Microstructure and
mechanical
properties of multi-directionally forged Mg–Al–Zn alloy. Scripta Mater
2012;66:49–51.
[23] B.L. Mordike, T. Ebert, Magnesium: properties–applications–potential,
Mater. Sci. Eng. A 302 (2001) 37–45.
[24] M.M. Avedesian, ASM Specialty Handbook. Magnesium and
Magnesium
Alloys, The Materials Information Society, 1999.
[25] D. Lee, et al., Warm precision forging and upsetting for the extruded
billet of AZ61 magnesium alloy, in: Proceedings of the 10th International
Conference on Metal Forming, vol. 2, Krak´ow, Poland, 2004, pp.
621–626.
[26] B.P.P.A. Goveia, J.M.C. Rodrigues, P.A.F. Martins, Fracture predicting
in bulk metal forming, Int. J. Mech. Sci. Pergamon 38 (4) (1996) 367–372.
[27] ASM Handbook of Forging

29

You might also like