Sei sulla pagina 1di 9

Available online at www.sciencedirect.

com
Available online at www.sciencedirect.com

ScienceDirect
ScienceDirect
Procedia Manufacturing 00 (2018) 000–000
Available
Availableonline atatwww.sciencedirect.com
online www.sciencedirect.com
Procedia Manufacturing 00 (2018) 000–000 www.elsevier.com/locate/procedia
www.elsevier.com/locate/procedia
ScienceDirect
ScienceDirect 
Procedia Manufacturing 15 (2018) 1626–1634
Procedia Manufacturing 00 (2017) 000–000
www.elsevier.com/locate/procedia
17th International Conference on Metal Forming, Metal Forming 2018, 16-19 September 2018,
17th International Conference on MetalToyohashi,
Forming, Metal
JapanForming 2018, 16-19 September 2018,
Toyohashi, Japan
Microstructure characterisation in alloy 825
Manufacturing Microstructure
Engineering characterisation
Society International Conferencein2017,
alloy 8252017, 28-30 June
MESIC
Munir Al-Saadia,b,*, Fredrik Sandberg
a,b,
2017, Vigo
a
, Andrey Kasaravb, Stefan Jonssonbb, Pär Jönssonbb
b
a (Pontevedra), Spain
Munir Al-Saadi *, Fredrik Sandberg , Andrey Kasarav , Stefan Jonsson , Pär Jönsson
a
R&D, AB Sandvik Materials Technology, SE-811 81Sandviken, Sweden
Costing models for capacity optimization in Industry 4.0: Trade-off
b
b
Department of Materials
a
Science
R&D, and Engineering,
AB Sandvik MaterialsKTH Royal Institute
Technology, SE-811of81Sandviken,
Technology,Sweden
SE-100 44 Stockholm, Sweden
Department of Materials Science and Engineering, KTH Royal Institute of Technology, SE-100 44 Stockholm, Sweden

between used capacity and operational efficiency


Abstract
Abstract A. Santanaa, P. Afonsoa,*, A. Zaninb, R. Wernkeb
Hot-compression tests down to 0.7 in nominal strain were conducted at a strain rate of 1s-1 between 800 and 1200 °C.
Hot-compression
The tests down
material was taken from to
the0.7 ina University
nominal
columnar strain
and of were
equiaxed
Minho, conducted
zones
4800-058 of at a Portugal
straincast
a continuous
Guimarães, rate strand
of 1s-1of
between 800 Furthermore,
Alloy 825. and 1200 °C.
The material was taken from the columnar b
and equiaxed
Unochapecó, zones
89809-000 of a
Chapecó,continuous
SC, Brazil
recrystallization was determined using LOM and FEG-SEM equipped with an EBSD-detector. cast strand of Alloy
The825.
trueFurthermore,
stress of the
recrystallization
columnar structurewaswas
determined usingthan
a 0.2% lower LOM forand
the FEG-SEM equipped withfor
equiaxed microstructure an the
EBSD-detector. The truetemperatures.
studied deformation stress of the
columnar
The opticalstructure was images
and EBSD a 0.2% lower
in boththan for the equiaxed
structures showed thatmicrostructure for the studied deformation
a dynamic recrystallization had almost temperatures.
not occurred.
The optical
Instead, and EBSDformation
a substructure images inwith
bothmostly
structures showed
low-angle thatboundaries
grain a dynamic dominated
recrystallization had almost not occurred.
was found.
Abstract
Instead, a substructure formation with mostly low-angle grain boundaries dominated was found.
© 2018 The Authors. Published by Elsevier B.V.
© 2018 the
Under The Authors.
concept Published
of "Industryby Elsevier
4.0", B.V.
production processes will be pushed to be on increasingly interconnected,
© 2018 The Authors.
Peer-review Published
under responsibility
responsibility by
of Elsevier B.V.
the scientific
scientific committeeof ofthe
the17th
17thInternational
InternationalConference
Conferenceon MetalForming.
Forming.
Peer-review under
information of the committee Metal
Peer-review under responsibility of the scientific committee of the 17th International Conference on Metal Forming.optimization
based on a real time basis and, necessarily, much more efficient. In this context, capacity
goes beyond
Keywords: Alloythe traditional
825; Columnar andaimequiaxed
of capacity maximization,
structure; Hot compressioncontributing also for
test; Microstructural organization’s profitability and value.
evaluation
Indeed,
Keywords:lean management
Alloy 825; Columnar and and continuous
equiaxed improvement
structure; Hot approaches
compression test; suggest
Microstructural capacity optimization instead of
evaluation
maximization. The study of capacity optimization and costing models is an important research topic that deserves
1. Introduction
contributions from both the practical and theoretical perspectives. This paper presents and discusses a mathematical
1. Introduction
model for capacity management based on different costing models (ABC and TDABC). A generic model has been
The wrought
developed and itforms of Alloy
was used 825 products,
to analyze e.g. solid
idle capacity bar
and to [1], forgings
design [2],towards
strategies or pipethe
andmaximization
tube materialsof[3] are normally
organization’s
air The
value. wrought
melted
The forms
Electric
trade-off Arc ofFurnace
Alloymaximization
capacity 825
(EAF),products,
Argone.g. solid bar
vsOxygen [1],efficiency
forgings [2],
Decarburization
operational or pipe and tube
is(AOD)-refined,
highlighted materials
andit
and then [3]that
either
is shown are normally
continuously
capacity
air melted Electric
cast or ingot cast.
optimization mightTheArc
hide Furnace
continuously (EAF),
operational cast Argon Oxygen Decarburization (AOD)-refined, and then either
products of Alloy 825 materials often contain large columnar grains adjacent
inefficiency. continuously
cast ormould
to 2017
© the ingot cast.
wallThe
The Authors. and continuously
fine equiaxed
Published cast
by Elsevier products
grains
B.V. in theofcenter.
Alloy 825The materials often contain
casting structure large columnar
breakdown is processed grains
by adjacent
thermo-
to the mould
Peer-review
mechanicalunder wall and fine
responsibility
processing equiaxed grains
of therolling,
(forging, scientific in the
committee
etc.) center. The casting
of thea Manufacturing
to obtain structure
uniform chemistry breakdown
Engineering is processed
Society International
and microstructure. by thermo-
Conference
Thereafter, the
mechanical
2017.
material processing
is subjected to (forging,
an appropriaterolling, etc.) to process
annealing obtain ato uniform
develop chemistry
the optimum andcombination
microstructure. Thereafter,
of a good the
corrosion
material is subjected to an appropriate annealing process to develop the optimum combination
resistance and mechanical properties [4]. Thus, the dynamic recrystallization can have a huge influence on the control of a good corrosion
Keywords:
resistanceCost
andModels; ABC; TDABC;
mechanical Capacity
properties [4]. Management; Idle Capacity;
Thus, the dynamic Operational Efficiency
recrystallization can have a huge influence on the control

1. Introduction
* Corresponding author: E-mail addresses: munir.al-saadi@sandvik.com, muniras@kth.se
* The cost of idle
Corresponding capacity
author: is a fundamental
E-mail addresses: information for companies
munir.al-saadi@sandvik.com, and their management of extreme importance
muniras@kth.se
in modern©production
2351-9789 systems.
2018 The Authors. In general,
Published it isB.V.
by Elsevier defined as unused capacity or production potential and can be measured
Peer-review
2351-9789
in underThe
several©ways:
2018 responsibility
Authors.
tons of theby
Published
of production, scientific
Elsevier
availablecommittee of the
B.V.hours of 17th International
manufacturing, etc.Conference on Metal Forming.
The management of the idle capacity
Peer-review under responsibility of the scientific committee of the 17th International Conference on Metal Forming.
* Paulo Afonso. Tel.: +351 253 510 761; fax: +351 253 604 741
E-mail address: psafonso@dps.uminho.pt

2351-9789 © 2017 The Authors. Published by Elsevier B.V.


Peer-review under responsibility of the scientific committee of the Manufacturing Engineering Society International Conference 2017.
2351-9789 © 2018 The Authors. Published by Elsevier B.V.
Peer-review under responsibility of the scientific committee of the 17th International Conference on Metal Forming.
10.1016/j.promfg.2018.07.294
Munir Al-Saadi et al. / Procedia Manufacturing 15 (2018) 1626–1634 1627
2 Munir Al Saadi / Procedia Manufacturing 00 (2018) 000–000

of the grain size [5-8]. However, differences in the grain size within the material may occur, which is often observed.
These differences are caused by the differences in the grain structure, columnar and equiaxed, and/or by hot-working
conditions such as the deformation temperature, strain and strain rate. In practice, these microstructural changes can
include recovery and dynamic recrystallization effects. In addition, the understanding of overall microstructure
textures is very important [9]. This texture is to a large extent responsible for the anisotropy in the mechanical
properties of the material and will determine the properties of the bar products.
The effect of a recrystallization process on the crystallographic texture was studied by using the inverse pole
figures. There is a lack of research around the behavior of hot deformation. A few studies exist, but they have only
focused on the dynamic recrystallization (DRX) microstructure evolution of an alloy 825 during hot deformation at a
very high reduction degree (true strain 0.7-2.5) or at high strain rates (> 1 s-1) [10-12]. Therefore, the primary
objectives of the present work are to employ optical microscopy and electron backscatter diffraction (EBSD)
techniques to study the DRX microstructure evolution of an alloy 825. Specifically, the focus is on deformation
temperatures between 800 to 1200 °C at a true strain of 0.7 and a nominal strain rate of 1 s-1.

2. Experimental work

Cylindrical compression samples of 15 mm in height and 10 mm in diameter were cut and machined with the
compression axis parallel to the long axis of the columnar and equiaxed grains. Thereafter, the columnar and equiaxed
samples were etched with a mixture of 85 ml H2O, 10 ml HCl, and 5 ml CH3COOH. The chemical composition of
both structures used in the study emanated from the same heat, which is shown in Table 1. The micro-samples
exhibited a random orientation of columnar grains which contain small TiN inclusions and equiaxed grains containing
low levels of microporosities. For this study, a monotonic-hit compression test was performed. All samples were
initially heated to the homogenization temperature of 1200 °C for 100 seconds. Then, the samples were cooled down
to the deformation temperature at a rate of 5 °C/s. The samples were held at each deformation temperature for 30
seconds. Thereafter, the sample was compressed with a load of 100 kN in vacuum before being air-pressure quenched
after each test to a temperature of 100 to 200 °C.

Table 1. Chemical composition for investigated specimens, wt. %.

C Si Mn Cr Ni Mo Cu Ti Fe N
0.007 0.2 0.7 22.0 Bal. 2.5 1.6 0.8 32 0.009

The deformed columnar and equiaxed specimens were sectioned through the longitudinal axis and
metallographically prepared for the investigation using Light Optical Microscopy (LOM). The sectioned surfaces were
electrolytically etched in a solution of 10 g Oxalic acid and 100 ml H2O, for 3-60s and using a voltage of 6V to prepare
the samples for optical metallographic examinations. The photomicrographs presented were taken from the center of
the longitudinal sections. Subsequently, the LOM samples were used for hardness tests, so the elongated non-
recrystallized grain structure could be observed together with any recrystallized grains. The samples were tested for
hardness on a Vickers (Diamond Pyramid) hardness tester, using a 1.0 kg load and each sample was tested 10 times.
The reported measurement values were average values of those 10 measurements. In addition, standard deviations
were also estimated for each investigated sample. Subsequently, the second half compressed samples were taken and
electrolytically polished to prepare for the for SEM-EBSD analysis. Specifically, the samples were jet polished at
temperatures between 8 and 18 °C in a 3 M H2SO4 ethanol solution (630 ml ethanol, 123 ml H2SO4). The electrolytic
polishing voltage, current and time were 30-40 V, 1-2A, and approximately 30 s, respectively. The EBSD data
collection and subsequent indexing was performed using the AZtec 3.3 software and the used EBSD measurements
in a ‘FEG-SEM’ equipped with a ‘Nordlys F’ EBSD detector [13]. In this study, the EBSD data acquisition was
performed by indexing Ni-based superalloy phase FCC (gamma phase). The gamma phase can be selected
automatically from the AZtec software using the Twist program supplied by Oxford Instruments. For each condition,
double EBSD scans were acquired covering an area of approximately 1 mm2 and 0.11 mm2 employing a step size of
1.5 and 0.5 µm respectively. From the acquired EBSD data, several microstructural parameters were evaluated using
the Channel5 software (TANGO) [14, 15]. In doing this, Σ3 (60°<111>) twin boundaries were ignored before
1628 Munir Al-Saadi et al. / Procedia Manufacturing 15 (2018) 1626–1634
Munir Al Saadi/ Procedia Manufacturing 00 (2018) 000–000 3

estimating the recrystallized grain size. Furthermore, the minimum misorientation angle for recrystallized grains was
taken 5° to separate them from sub-grains, having an angle of 2° before estimating the DRX fraction. Grain boundaries
with misorientation values below 2° are discarded, because of the angular resolution of the instrument. Moreover,
HAGBs corresponds to grain boundaries having misorientation angles � >10° and LAGBs boundaries having
misorientation angles between 2° < � < 10°. In this study, the sub-grains were not taken in to account. If recrystallized
grain size boundaries with misorientation angles were lower than 5°, they were not considered as separate grains. Also,
inverse pole figures were used to identify fiber texture from EBSD data. The Channel5 software (HKL Mambo) was
employed to import and analyze the EBSD scan data. Also, an EBSD IPF colour key was used for the crystal
orientation mapping. A fibre texture of <100> represents the crystal direction that is parallel to the samples
longitudinal, Compression Direction (CD).

3. Results and discussion

3.1. Hot Deformation Flow Curves

The true stress-true strain of columnar grain and equiaxed Alloy 825 deformed at different deformation
temperatures at a constant strain of 0.7 and a strain rate of 1.0 s-1 is shown in Fig 1.

True strain: 0.7 E-800°C


Strain rate: 1s-1 C-800°C
E-900°C
C-900°C
E-1000°C

C-1000°C

E-1100°C
C-1100°C
E-1200°C
C-1200°C

Fig. 1. Stress-strain curves showing flow stress of the as columnar and equiaxed during hot compression at various temperatures and a constant
strain rate and strain. C stands for Columnar specimens and E stands for Equiaxed specimens. Data are given for temperatures between 800 and
1200 °C. The designation of each run is composed of the name of the initial structure (Columnar or Equiaxed) and deformation temperatures
between 800-1200 °C.

In general, the flow stress decreases with an increased temperature and for equiaxed curves showed slightly above
that of the columnar over all deformation temperatures. This is probably due to the fact that the grains in the equiaxed
zone in the as-cast sample have stronger orientations compared to those of the columnar grain zone. The shape of the
flow curves is similar in both structures and similar to what is reported for other nickel-based and Fe-Ni-Cr
superaustenitic stainless steel grades [16]. At a deformation temperature of 800 °C and for the as columnar and
equiaxed samples, the stress-strain curve shows a continuous work hardening feature that is commonly associated
with a dynamic recovery (DRV) being the dominant restoration mechanism. None of the columnar and equiaxed
samples deformed reached the steady-state flow stress as no stress peak was visible. At a higher deformation
temperatures of 900 and 1000 °C, the flow stress-strain curves show work hardening character prior to a steady state
flow at true strains from 0.2 to 0.35. This indicates that a dynamic recrystallization has not occurred during the
deformation. In the equiaxed case and when the deformation occurred at 1100 °C, the material exhibits a slight drop
in flow stress after a strain of 0.5. This possibly indicates that a thermal softening had occurred. Most of the curves
display a shape by recall for a dynamic recovery. A fully dynamic recovery behaviour was only observed at
temperatures of 1100 and 1200 °C. However, dynamic recrystallization conditions were apparently not achieved in
most of the tests performed for all deformation temperatures. Earlier studies [11, 12] showed that the maximum stress
Munir Al-Saadi et al. / Procedia Manufacturing 15 (2018) 1626–1634 1629
4 Munir Al Saadi / Procedia Manufacturing 00 (2018) 000–000

values obtained are higher than those found in this study when being compressed at a strain rate of 5s-1. The strain rate
in steady state during deformation conditions is often described as a function of peak true stress and of deformation
temperature as obeys an Arrhenius-type equation (Sellers and Tegart) [17], Eq. (1).

�� � ������
= ������(��� )� ��� �− ���, (1)
��

where A, n, and α are material dimensionless constants. R stands for the universal gas constant, 8.314 Jmol-1K-1. T
stands for the absolute temperature, K. ����� is deformation activation energy for a hyperbolic-sine equation, kJ/mole.
��
stands for a strain rate, s-1. �� stands for the peak true stress, MPa.
��

3.2. Effect of deformation temperatures on microstructure

Fig. 2 exhibit the compressed columnar and equiaxed microstructures of the Alloy 825 samples after a hot
compression for a nominal true strain of 0.7 and a strain rate of 1 s�� . At low deformation temperatures, such as
800 °C, 900 °C, and 1000 °C, no recrystallization occurred in overall microstructure of the compressed samples in
both structures, as seen in Fig. 2(a)-(f). This is in a good agreement with the results in the ref. [12]. However, a small
fraction of recrystallization was observed on some grain boundaries of the equiaxed structure, as shown by the bold
arrows in the micrograph Fig. 2(d) and (f). It has been reported that recrystallization occurs inside the deformation
bands (DBs) and that it has been is observed at a strain rate of 5 � �� when using a high strain of 1.2 [12]. However,
fully recrystallized microstructure is observed at temperatures above 1050 °C when deformed using a strain rate of 10
s-1 and a strain below 1.5 [10]. At higher deformation temperatures of 1100 °C - 1200 °C, Fig. 2(g)-(j), a partial
recrystallization occurred. At higher deformation temperatures of 1100 °C - 1200 °C, Fig. 2(g)-(j), a partial
recrystallization occurred. This occurs by the necklace structure in the micrographs in both structures and by of
recrystallized grains formed on grain boundaries. However, an almost complete to fully recrystallization has been
observed at a temperature of 1100 °C when deformed with a strain rate of 5 s-1[12]. In addition, a fraction of
recrystallization in equiaxed structure was 0.2% larger than in the columnar structure. It was noted that no intermetallic
phases or internal crack formations were observed at low or high deformation temperatures. Hence, It was concluded
that dynamic recrystallization conditions in both deformed samples were apparently not achieved in most of the tests
performed and not for any deformation temperatures.

3.3. Micro-hardness of deformed columnar/equiaxed Alloy 825

Table 2 shows a relationship between the hardness and deformation temperature for both the columnar and
equiaxed grain structures in the temperature range between 800 °C to 1200 °C. The investigated columnar or equiaxed
material showed a decreasing hardness with an increasing deformation temperature at a nominal strain of 0.7 and a
strain rate of 1 � �� . However, the equiaxed grain structure showed a 0.01 to 0.2% higher hardness than the columnar
grain structure over all deformation temperatures.

Table 2. Micro-hardness (HV) of deformed columnar/equiaxed Alloy 825, with ± standard deviation.

Structure Deformation temperature, [°C]


800 900 1000 1100 1200
Columnar 248 ±4 186 ±5 173 ±4 166 ±5 157 ±4
Equiaxed 250 ±5 229 ±15 181 ±5 167 ±4 159 ±4

3.4. Grain misorientation distribution of deformed samples

Fig. 3(a), (b) shows the recrystallized grain distribution of misorientation angles for columnar and equiaxed
deformed microstructures for alloy 825 samples, when using a 5o grain definition. The low angle grain-boundary
(LAGB) represents a recrystallized grain boundary with a misorientation angle of 5o (composed of arrays of
dislocations [6]) and the high angle grain boundary (HAGB) represents a recrystallized grain boundary with a
misorientation angle of 60o (composed of grain growth). Therefore, the increasing amount of LAGB for columnar
1630 Munir Al-Saadi et al. / Procedia Manufacturing 15 (2018) 1626–1634
Munir Al Saadi/ Procedia Manufacturing 00 (2018) 000–000 5

and/or equiaxed deformed samples indicate that the grains are refined over all studied deformation temperatures.
Moreover, the amount of LAGB decreases with increasing deformation temperatures.

(b)
(a)

100µm 100µm

(c)
(d)

100µm 100µm

(f)
(e)

100µm 100µm

(g) (h)

100µm 100µm

(i) (j)

100µm 100µm

Compression direction

Fig. 2. In left column: Microstructures of deformed columnar structure. In right column: Microstructures of deformed equiaxed structure
at deformation temperatures of (a), (b) 800 °C, (c), (d) 900 °C, (e-f) 1000 °C, (g), (h) 1100 °C and (i), (j) 1200 °C and at a nominal strain
0.7 and strain rate of 1 ) '( . Microstructures were revealed using an Oxalic acid etchant and using 200x magnifications in the LOM before
reproduction. In (d) and (f), the bold arrows showing some small recrystallized grains on the grain boundary.
Munir Al-Saadi et al. / Procedia Manufacturing 15 (2018) 1626–1634 1631
6 Munir Al Saadi / Procedia Manufacturing 00 (2018) 000–000

As seen in Fig. 4, a substructure is dominating for both columnar and equiaxed structures. Also, a big amount of
recovery and annihilating dislocations occurred with increasing deformation temperatures, but no full recovery was
obtained. Furthermore, the increasing amount of HAGB in the equiaxed specimen indicates that the grain increases
in size and that grain growth has occurred as well as that the grain has become coarser as in the case of the results at
the 1100 and 1200 °C temperatures, Fig. 4.

3.5. Percentage/Number/Size of recrystallized grains

Fig. 4(a) shows a plot of the percentage of recrystallized DRX grains and recrystallized grain sizes as a function of
the deformation temperatures for the compressed columnar and equiaxed samples. A fine step size of 1.5 µm was used
for scans of about 1 x 1 mm2 in this area. In columnar compressed samples, recrystallization did not occur. However,
some recrystallized grains were sparsely observed. They had nucleated around TiN-particles at temperatures of 800,
900 and 1000 °C. At the same temperatures, recrystallized grain sizes of about 5 µm were observed. These did not
change with increasing temperatures. A degree of roughly a 1% recrystallization and up to a 20 µm recrystallized
grain size was observed at deformation temperatures between 1100 and 1200 °C. As mentioned by Mitra [12],
recrystallized DRX grains sizes lies in the range 8 to 10 µm when deformed at strain rates of 1 to 10 s-1. Also, Niikura
et al [10] showed that the recrystallized grain size is not only dependent on the temperature and strain rate, but also
on the initial grain size and strain for significant levels of recrystallization degrees. Thus, the recrystallized grain size
was estimated to have values of up to 15 µm when deformed at a temperature of 1050 °C and at a high reduction of a
strain above 2 when the initial grain size of investigated material used was 370 µm [10].
Analysed data of the number of recrystallized grains in the columnar and equiaxed structures are plotted in Fig.4b.
The number of recrystallized grains increases with an increased deformation temperature. This is most likely due to
that the old recrystallized grains grow, which resulted in that no new grains were created in most of the recrystallized
grains. In equiaxed deformed samples, the percentage of recrystallized grains increased slightly with an increased
temperature. Specifically, they were not more than two per cent at a temperature of 1000 °C. Furthermore, less than a
7% recrystallization degree was observed at a deformation temperature of 1100°C. In addition, the recrystallized grain
sizes also became slightly larger with increasing temperatures up to 1100 °C. At higher deformation temperatures than
1200 °C, the recrystallized grain size and recrystallization degree were approximately 35 and 26%, respectively. This
is shown by the necklace structure in the micrographs for the equiaxed samples in Fig. 5 (d) and (f). The number of
recrystallized grains increased with an increased temperature up to a temperature of 1000 °C and did almost not rise
at higher deformation temperatures, as seen in Fig. 4(b).

30 x104 30 x104
(a) 800°C
(b) 800°C

25 900°C 25 900°C
Frequency

Frequency

1000°C 1000°C
20 20
LAGB 1100°C 1100°C

15 1200°C 15 1200°C

10 10
HAGB
5 5

0 0
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
Misorientation angle [°] Misorientation angle [°]

Fig. 3. Grain misorientation distribution of (a) columnar and (b) equiaxed deformed samples over a range of 800 °C to 1200 °C and at nominal
strain of 0.7.

The recrystallized grains which formed beyond 1100 °C may have attributed to the deformation heating and /or the
start of the static recrystallization after the end of the deformation test. Furthermore, a higher deformation temperature
increased the rate of recovery and recrystallization, as seen in Fig. 3(b).
1632 Munir Al-Saadi et al. / Procedia Manufacturing 15 (2018) 1626–1634
Munir Al Saadi/ Procedia Manufacturing 00 (2018) 000–000 7

35 40 10000

Recrystallized Grain Size, [µm]


Columnar grain structure
Fully Recrystallized DRX Grains-Columnar
(a) (b)
Percentage Recrystallized DRX Grains

Fully Recrystallized DRX Grains-Equiaxed 35 Equiaxed grain structure

Number of Recrystallized Grains


30
Recrystallized Grain Size (µm)-Columnar
30 1000
25 Recrystallized Grain Size (µm)-Equiaxed
25
20
20 100
15
15
10 10
10

5 5

0 0 1
700 800 900 1000 1100 1200 1300 700 800 900 1000 1100 1200 1300
Deformation Temperature [°C] Deformation Temperature, [°C]

Fig. 4. (a) Variation of percentage of recrystallized DRX grains and recrystallized grain sizes in columnar and equiaxed grains structures and (b)
plot of several recrystallized grains in columnar and equiaxed grain structures. Data are from deformation temperature for Alloy 825 after hot
compression using strain rate of 1) '( and strain of 0.7.

It was concluded that the percentage of recrystallized grains in the as-columnar state showed a somewhat lower
value, 0.01-0.2%, than in the as equiaxed state. It is well known that solutes or impurities prevent dislocation
movements, so that they retard a further recovery. Since recovery in this material is dominant, the effects of solute on
the recovery may be seen explicitly.

3.6. Dynamic recrystallized, substructured and deformed grain evolution

Fig. 5(a)-(f) shows recrystallized fraction maps in the compressed columnar and equiaxed grain structures at
deformation temperatures of 900, 1100 and 1200 °C and at a nominal strain of 0.7 and a strain rate of 1) '( . Generally,
the microstructure in Fig. 5 shows that almost no recrystallization has occurred at the studied deformation temperatures
900-1200 °C. Instead, a dynamic recovery (DRV) has been the dominant mechanism. Fig. 5(a-b) demonstrates that
no dynamic recrystallization has occurred at a deformation temperature of 900 °C for neither the columnar nor the
equiaxed microstructures. The overall microstructures after deformation of the columnar specimens at temperatures
of 1100 and 1200 °C are compared to the equiaxed specimens in Fig. 5(c)-(f). The features are obviously different.
Essentially, a larger deformed structure and more recrystallized grains are clearly seen in the equiaxed grain structure
compared to the columnar structure. Hence, the recrystallized fraction images in both structures reveal that a
substructure formation dominated. Furthermore, that it was higher in the case of a columnar structure compared to an
equiaxed structure, for the studied deformation temperatures of 1100 and 1200 °C.

3.7. Microstructural Evolution

The EBSD orientation mapping images of the deformed specimens at deformation temperatures of 900, 1100 and
1200 °C at a nominal strain of 0.7 and strain rate of 1) '( are outlined in Fig. 6(a)-(f). The colour of the EBSD grains
refers to the different orientations. Thus, a similar colour indicates that the grains have similar orientation, as can be
seen in Fig. 6g. At a deformation temperature of 900 °C, deformed grains were dominating in the entire columnar and
equiaxed structures, as shown in Fig. 6(a)-(b). At temperatures between 1100 and 1200 °C, the deformed columnar
structure showed a strong preference for a <100> pole direction oriented parallel to the compression axis. This
direction clearly dominates the overall microstructure in the studied samples. At temperatures of 1100 and 1200 °C,
the compressed equiaxed structure shows a relatively random texture within the recrystallized grains. It also tends to
have a <101> fibre texture, which is the stable compression texture for FCC materials. In the as-equiaxed structure,
partial DRX occurs through the so-called necklace mechanism. This leads to the formation of some individually
deformed grain boundaries, as shown by a bold black arrow in Fig. 6 (d and f).
Munir Al-Saadi et al. / Procedia Manufacturing 15 (2018) 1626–1634 1633
8 Munir Al Saadi / Procedia Manufacturing 00 (2018) 000–000

(a) (c) (e)

200µm

(b) (d) (f)

Fig.5. Recrystallized/substructured/deformed fraction map of columnar grain structure of Alloy 825 after a uniaxial compression to nominal strain
of 0.7 and a strain rate of 1 ) '( at following deformation temperatures. (a), (b) 900 °C, (c), (d) 1100 °C and (e), (f) 1200 °C. On the upper row:
Columnar compressed specimens. On the lower row: Equiaxed compressed specimens. Blue regions= fully recrystallized grains, Yellow regions=
substructured and recovered material, Red= Deformed material. Each overlaid with HAGB (black) and LAGB (white).

(a) (c) (e) (g) 111

(b) (d) (f) 001 101


Pole Figure
colouring

200µm

Fig. 6. Inverse pole figure maps showing distribution of crystallographic poles oriented parallel to compression axis for an Alloy 825 after uniaxial
compression to a nominal strain of 0.7and a strain rate of 1) '( . Data are given for deformation temperatures of (a), (b) 900 °C, (c), (d) 1100 °C,
and (e)-(f) 1200 °C. Also, in (g) key used for the colouring of the pole figure maps shown in a-f is given. On the upper row: columnar grain
structures. On the lower row: equiaxed grain structures. (d) and (f) the bold arrows showing recrystallized grains.

4. Conclusions

Hot-compression tests, by a thermomechanical simulator in the temperature range of 800 °C to 1200 °C at a strain
of 0.7 and a strain rate of 1s-1, were performed on the columnar and equiaxed-grain specimens of a cast blooms Alloy
825 material. The purpose was to study the flow deformation behaviour and to characterize the deformed structure.
The microstructural characterizations in the deformed samples were carried out by using LOM and EBSD techniques.
Emphasis has been placed particularly on the dynamic recrystallized (DRX) microstructure evolution. Based on the
results, the made specific conclusions are the following:
1634 Munir Al-Saadi et al. / Procedia Manufacturing 15 (2018) 1626–1634
Munir Al Saadi/ Procedia Manufacturing 00 (2018) 000–000 9

1) In the columnar and equiaxed deformed structures, the flow stress and microhardness decreases with an
increased deformation temperature at a constant strain rate of 1s-1 and a nominal strain of 0.7. The true stress
and Vickers hardness of the columnar structure were a 0.8% and 0.2% respectively lower than for the
equiaxed microstructure for the studied deformation temperatures. However, the shape of the stress-strain
flow curves are similar in both structures.
2) The optical images of columnar specimens deformed at 800 °C-1000 °C show that elongated grain boundaries
exist but that no recrystallization has occurred at temperatures of 1100 °C and 1200 °C. However, less than
26% of the samples show indications of that recrystallization has occurred. In the equiaxed microstructure,
compressed at 800 °C-1000 °C, almost similar behaviours as found in the columnar structure with more
subgrains and grain boundaries. In addition, the optical images pictured for deformations in the temperature
range of 1100 °C-1200 °C showed a necklace structure and the formation of a limited number of refined
recrystallized grains at the grain boundaries.
3) EBSD grain maps and the microstructures in both structures showed that almost no recrystallization had
occurred at the studied deformation temperatures. Furthermore, fraction images to determine the degree of
recrystallization revealed that a formation of a substructure had taken place during the deformation.
4) IPF maps show that large non-recrystallized grains dominate over all the microstructures in both columnar
and equiaxed compressed samples. The results indicated a strong preference for their <100> and <101> poles
respectively to be oriented parallel to the compression axis.

Acknowledgements

MA would like to thank the colleagues at R&D Sandvik AB for helpful comments during this work. Furthermore,
MA wish to express his deep gratitude to Sandvik Materials Technology for their grant which made this study possible.

References
[1] ASTM B425-11, “Standard specification for Ni-Fe-Cr-Mo-Cu alloy (UNS N08825, UNS N08221, and UNS N06845) Rod and Bar”.
[2] ASTM B564-15, “Standard specification for nickel alloy forgings”.
[3] ASTM B423-11 “Standard specification for nickel-iron-chromium-molybdenum-copper alloy (UNS N08825, N08221, and N06845) seamless
pipe and tube”.
[4] E.L. Raymond, Mechanisms of sensitization and stabilization of incoloy nickel-lron-chromium alloy 825, CORROSION, 24-6 (1968) 180–188.
[5] E.O. Hall, The deformation and ageing of mild steel: II characteristics of the luders deformation, Proceedings of the physical society, 64 (1951)
742–746.
[6] N.J. Petch, Journal of the iron and steel institute, 197 (1953) 25.
[7] F.J. Humphreys, M. Hatherly, “Recrystallization and related annealing phenomena”, 2nd edition, (2004).
[8] G. Krauss, “Deformation, processing, and structure”, 1982 ASM Materials Science Seminar, St. Louis, Missouri, October, (1982) 23–24.
[9] U.F. Kocks, C.N. Tomé, H.R. Wenk, Texture and anisotropy preferred orientations in polycrystals and their effect on materials properties,
Cambridge University Press, (1998).
[10] M. Niikura, K. Takahashi, C. Ouchi, Microstructural change of austenite in hot working with a very high reduction, Transactions ISIJ, 27-6
(1987) 485–491.
[11] Y. Man, L. Jing-she, T. Hai-yan, B. Yao-zong, Hot working characteristics of corrosion-resistant alloy G3 and 825, JISI, 18-4 (2011) 68–72.
[12] M. Basirat, H. Fredriksson, “Plastic deformation and recrystallization in incoloy 825” KTH industrial engineering and management, Doctoral
thesis, Stockholm, Sweden, (2013).
[13] www.oxford-instruments.com/EBSD
[14] R.D. Doherty, D.A. Hughes, F.J. Humphreys, J.J. Jonas, D.J. Jensen, M.E. Kasnner, W.E. King, T.R. McNelley, H.J. McQueen, A.D. Rollet,
Current issues in recrystallization: a review, Material Science and Engineering: A, 238-2 (1997) 219–274.
[15] Help of Channel5 software.
[16] A. Momeni, K. Dehghani, Metallurgical Material Transition A, 42A (2010) 2011–1925.
[17] C.M. Sellars, W.J.M. Tegart, Hot workability, International Meterials Reviews, 17 (1972) 1–24.

Potrebbero piacerti anche