Dictra
Dictra
Dictra
School of Engineering
Department of Mechanical Engineering
by
Panagiota I. Sarafoglou
Dipl. Chemical Engineering, NTUA
M.Sc. Materials, NTUA
Submitted
in partial fulfilment of the requirements of the degree of
DOCTOR OF PHILOSOPHY
Supervisor
Professor Gregory N. Haidemenopoulos
Volos 2016
University of Thessaly
1. Prof. G. N. Haidemenopoulos Department of Mechanical Engineering
(Thesis advisor)
University of Thessaly
2. Prof. N. Aravas
Department of Mechanical Engineering
University of Patras
4. Prof. S. Pantelakis Department of Mechanical Engineering
and Aeronautics
National Technical University of Athens
5. Prof. D. I. Pantelis School of Naval Architecture and Marine
Engineering
National Technical University of Athens
6. Prof. E. Pavlatou
School of Chemical Engineering
Drexel University
7. Prof. A. Zavaliangos Department of Materials Science and
Engineering
ii
iii
vi
vii
The completion of this dissertation would not have been possible without the help of
several people whom I would like to thank;
I am grateful to my advisor Prof. G.N. Haidemenopoulos for trusting me and providing me
this opportunity to work on his alloyneering concept and for his encouragement and guidance
both personally and professionally. I owe him many thanks for his friendship and teaching during
these years.
I am very grateful to Prof. N. Aravas for serving in my thesis advisory committee and Prof.
D. Manolakos for serving in my thesis advisory committee and for providing guidance during my
years at NTUA.
I would like to acknowledge Professors D. Pantelis, E. Pavlatou, S. Pantelakis and
A.Zavaliangos for serving in my thesis examination committee.
I would like to express my gratitude to the Aluminium of Greece for providing the materials
for my thesis and especially El. Grigoriou, A. Triantafyllou and K. Liapis for the fruitful discussions
regarding the thesis.
I would like to thank the staff of the Laboratory of Materials of the University of Thessaly
but especially Eleni Kamoutsi and Eleana Pappa. During my thesis I had a great support from
them.
I am also glad to have worked with undergraduate students but also friends Spyros –
Alexandros Tousias, Margianna Tzini, John Aristeidakis and John Fanikos. They have been great
partners for me and I wish the best for them.
I would like to thank Lin Shang and Andreas Stieben (RWTH- Aachen) as well as Helen Rosolymou
(Chemical Engineering, NTUA) for their assistance with XRD measurements and Dr. Ch.Sarafoglou
(Naval Architecture & Marine Engineering) for her assistance with SEM/EDX.
Living in Volos has been exciting and I would like to thank my special friends:
The very first person I met here and best friend since then, Dimitra for just being who she is.
My amazing road-trip friends, Latharo, Vagner and Apostolis for their long coffee hours. Their
unconditional support has been essential all these years in many ways. My dear lawyer and friend
Helen and Dimitris for his music. I am not sure I would have done my thesis without them.
My friends from my NTUA years, Antonis Stefanidis, Kostas Ntaskas, Giorgos Chortis, for
being by my side.
I am grateful to my unique in many ways big perfect family. They have cherished with me
every great moment and supported me whenever I needed it. And I need to thank them in Greek
now,
ix
Last but not least my beloved Apostolis for all laughing, crying, travelling, fishing and finally
living…I would like to dedicate this thesis to him.
Gioula Sarafoglou
Volos 2016
Extrudable aluminum alloys of the 6xxx series are heat-treatable alloys based on the
Al-Mg-Si system. These alloys find significant applications in the building, construction and
transportation sectors due to the excellent strength/ductility properties combined with
reduced weight and oxidation resistance. Large scale ingots or billets, suitable for hot
extrusion, are produced with the direct-chill casting method. The as-cast material exhibits
low formability due to microstructural inhomogeneities, arising from the solidification
process and the alloy chemistry. Major inhomogeneities in the as-cast material include:
elemental microsegregation in the level of the secondary dendrite arms, grain boundary
segregation and the formation of several low-melting eutectics and intermetallic
compounds such as the iron-containing α-AlFeSi and β-AlFeSi, which degrade the
extrudablity of the as-cast billet.
These effects can be partially or completely eliminated with the application of a
homogenization treatment of the as-cast billets. The benefits of this treatment include the
following: removal of elemental microsegregation, dissolution of low melting eutectics,
transformation of the iron-containing intermetallic compounds, shape control (round-off)
of hard particles with sharp edges, dissolution of the grain-boundary Mg2Si phase and re-
precipitation (during cooling) with a more homogeneous in-grain distribution. All these
effects improve the extrudability and increase the response of the material to natural or
artificial aging.
New demanding light-weight applications in the automotive and aircraft sectors
require the development of high-strength alloy extrusions. Such applications include pillar
and chassis structural parts in vehicles and wing or fuselage reinforcement parts in
aircrafts. The increase of strength is possible with higher alloying with Mg and Si, in order
to form higher amounts of the strengthening phase Mg2Si (e.g. 6061 and 6082 alloys).
However increased alloying deteriorates the extrudability, leading to extremely low
extrusion speeds for the high-strength alloy systems. In the other end of the spectrum, i.e.
low-alloy low-strength applications (e.g. 6060 or 6063 alloys), the main industrial
By this it is meant that not only processing steps prior to extrusion are important but
steps following the extrusion process should be considered as well, since they set important
microstructural requirements to be achieved in the preceding process steps. For example,
the homogenization process is carried out in order to eliminate the microsegregation
10
1.2. State-of-the-art
As it will be shown by the literature review in the next sections, the problem of the
design of alloy compositions and homogenization treatments for enhanced extrudability
has been considered by experimental methods either in the laboratory or by industrial
trials. In most cases, parts of the process chain have been considered individually. Most
works consider the effect of composition or process parameters such as the
homogenization time and temperature or the effect of the cooling rate individually on the
outcome of the homogenization process. In this way, ‘best-practice’ homogenization recipes
have been established for certain alloys. For the design of a new alloy composition or the
design of a new homogenization treatment for enhanced extrudability, the traditional trial-
and-error methodology is still followed.
11
1.3. Methodology
The approach followed in this thesis has both computational and experimental
aspects. A flow chart of the methodology is depicted in Table 1.1.
Computational
Computational alloy thermodynamics, based on the CALPHAD approach, has been
applied for the calculation of equilibrium phase constitution as a function of alloy
composition and temperature. Isopleth phase diagrams in multicomponent alloys have
been calculated permitting the determination of the stability limits of the various phases in
the system. The relevant thermochemical database COST-507 for light alloys has been
modified to incorporate Manganese (Mn) in the α-AlFeSi iron intermetallic. The relevant
software used was Thermo-Calc. The program was also applied to carry out Scheil-Gulliver
solidification simulations and calculate the resulting microsegregation of elements and
phases in the as-cast microstructure. Segregation profiles for elements and phases across
the grain were determined as a function of alloy composition. The results were confirmed
experimentally.
Computational kinetics, based on the solution of the diffusion equations in
multicomponent systems (Onsager principle) was applied in order to calculate the rate of
diffusional phase transformations during homogenization. In this way the dissolution of the
Mg2Si phase, the transformation of iron intermetallics (β-AlFeSi to α-AlFeSi) as well as the
reprecipitation during cooling was simulated. This included the development of a dual-
grain model (DGM) to take into account the different grain sizes between neighboring
grains. The relevant software was DICTRA, which incorporates the so-called sharp interface
model for treating diffusion-controlled transformations. The interdiffusion coefficients
12
Experimental
13
14
15
The thesis provides results leading to progress beyond the state-of-the-art in the area
of the computational-based design of the homogenization process of extrudable aluminum
alloys. Such innovative aspects are:
The application of a computational thermodynamics and kinetics approach, for the
integrated simulation of the process chain including, solidification-induced
microsegregation and subsequent homogenization in multicomponent and
multiphase aluminum alloys.
The development of a homogenization dual-grain model (DGM), for the evaluation of
the effect of different grain size between neighboring grains.
The description of temporal and spatial correspondence, of the β-AlFeSi to α-AlFeSi
transformation during homogenization.
The evaluation of the effect of excess-Si on the precipitation of metastable β’-Mg2Si
during cooling from the homogenization temperature by the application of the KWN
precipitation model.
The application of microstructural indices for the characterization of
homogenization state of extrudable aluminum alloys by quantitative metallography.
The development of design rules for the as-cast microstructure in terms of mapping
phase fractions of desirable and undesirable phases with respect to alloy
composition.
The development of design rules for selection of alloy compositions and processing
conditions for homogenization.
16
The structure of the thesis follows the process chain, of the extrudable Al-alloys.
Chapter 2 deals with the evolution of the as-cast microstructure and more specifically
with the microsegregation of elements and phases as a result of solidification. The mapping
of phase fraction in the Mg-Si composition space is also presented in this chapter.
The simulation of homogenization with the DGM is described in Chapter 3. The DGM
deals with the temporal and spatial evolution of the β to α – AlFeSi transformation. A
process map based on the time for the completion of the β to α-AlFeSi transformation is
presented.
The index–based quantification of the homogenization state is presented in Chapter 4.
Precipitation of the Mg2Si during homogenization cooling whith the KWN model is
presented in Chapter 5. The development of design rules for high extrudability is also
presented in Chapter 5.
The conclusions of this thesis and proposal for further research are presented in Chapter
6 and 7 respectively.
17
Extrudable aluminum alloys of the 6xxx series, based on the Al-Mg-Si system, are
subjected to a rather complicated process chain involving melt treatment, casting,
homogenization, preheating, extrusion, and ageing. The process chain begins with the direct
casting (DC casting) of large cylindrical billets. The as-cast microstructure, which depends
on composition and casting conditions, influences in a large degree the properties of
extrudable 6xxx Al–alloys. The as-cast material exhibits low formability due to
microstructural inhomogeneities, arising from the solidification process and the alloy
chemistry. Major inhomogeneities in the as-cast material, which degrade the extrudability
of the as cast billet include:
elemental microsegregation in the level of the secondary dendrite arms
grain boundary segregation and formation of low-melting eutectics
formation of intermetallic compounds
The research work described in this section focuses on the composition and
microstructural evolution of the as–cast billets since these key factors influence not only the
extrudability but the product properties of Al-alloys after extrusion has been completed.
The 6xxx series alloys, which are based on the ternary Al-Mg-Si alloy system, contain
magnesium (Mg) and silicon (Si) as major alloying elements for the formation of Mg2Si,
which is the major strengthening phase, providing precipitation hardening after aging.
Besides the intentional additions Mg and Si, transition metals and impurities such as
Fe, Mn and Cr are present. Even a small amount of these impurities causes the formation of
new intermetallic compounds (IMCs). The composition of the alloy as well as casting
conditions will influence the exact type and volume fraction of intermetallic phases that will
form during casting [1].
Iron (Fe), is the main impurity in these alloys forming intermetallic compounds in
combination with Si such as β-Al5FeSi and α–Al12Fe3Si. In addition, Mn can substitute for
Fe, and promote the formation of α-Al12(FeMn)3Si. These phases, from now on called β-
19
20
The present work is an effort to present in detail the effect of the major alloying
elements Mg and Si on microsegregation and the formation of Mg2Si and Fe intermetallics
during solidification of extrudable 6xxx Al-alloys. Mn is added in these alloys in order to
speed up the β→α transformation during homogenization. However the effect of Mn on IMC
formation in the as-cast microstructure has not been assessed. Therefore in this work the
effect of Mn is also quantified. In the first part, a computational thermodynamics-based
analysis of equilibrium phase constitution is presented, in order to identify the major
phases and their stability limits in the alloys. Solidification of the Al-billets is then modelled
by applying Scheil simulations in order to compute the mole fractions of all phases as well
as their segregation in the as-cast structure. Computed fractions are validated against
experimented measurements. The mole fraction of Mg2Si and the β-AlFeSi phase is then
mapped over the useful (0-1.2 mass %) Mg-Si composition space for 6xxx extrudable alloys.
All necessary thermodynamics calculations for the description of the Gibbs free
energy of the various phases in the system were performed with the Thermo-Calc software
[12], which is based on the CALPHAD approach [13]. This is a method that overcomes the
problem of the multi-dimensionality posed by a system with many components. A
thermodynamic description of a system requires assignment of thermodynamic functions
for each phase. The CALPHAD method employs a variety of models to describe the
temperature, pressure and concentration dependence of the free energy functions of the
various phases. The description refers to pure metals, stoichiometric and non-
stoichiometric phases and intermetallic compounds as well as substitutional and interstitial
solid solutions. It is based on the minimization of the total free energy of the system,
comprising of all phases, which take part in equilibrium of a multicomponent system.
The total free energy is the sum of the free energies of individual phases
21
where,
ni is the number of moles of phase i
For multicomponent systems the Gibbs free energy of a phase is the sum of four
contributions:
where
G mag is the magnetic contribution to the Gibbs free energy due to magnetic ordering.
The major phases and stability limits for the case of Al-Mg-Si-Fe system are described
in the COST507 light alloys database [14]. This is a public domain thermodynamic database
created in the course of the European concerted action program “Development of a
thermochemical and thermos-physical database for light alloys”. The Laboratory of
Materials at the University of Thessaly was a member of this consortium and contributed to
the development of the database. The description of the relevant phases of interest in the
COST507 database is given in the following sections.
FCC and Liquid
The free energy of the FCC substitutional solid solution phase (α-Al) as well as the
liquid phase are expressed according to equation (2) as:
22
o , LIQ
GMg H Mg (298.15K ) 8202.24 8.83T 8.01x1020 T 7 GMg
SER
(T )
(5)
o , LIQ
GMg H Mg (298.15K ) 8690.31 9.39T 1.03x1028 T 9 GMg
SER
(T )
(6)
SER
GMg (T )
is the free energy with respect to the reference state (Stable Element Reference,
SER) defined as the enthalpy of the pure metal at 298.15K.
For the Liquid phase, the ideal term Gid can be described as
Gid ,LIQ RT ( X Al ln X Al X Mg ln X Mg X Si ln X Si X Fe ln X Fe X Mn ln X Mn )
(7)
As can be seen from equation (7) the ideal term corresponds to the entropic contribution to
free energy.
23
24
X i io X i X j [
( p)
( X i X j )( p ) ] (10)
i i j p
where
is a normalized temperature,
with
T /Tc and Tc being the Curie temperature
which is composition-dependent and is given by
Tc X iTcio X i X j [Tc
( p)
( X i X j )( p ) ] (11)
i i j p
Tc
The values of and are stored for the FCC phase in the COST 507 database.
β-AlFeSi phase
GAl14 Fe3Si3 (T ) 14H Al (298.15) 3H Fe (298.15) 3H Si (298.15) GAlf 14 Fe3Si3 (T ) 14GAlSER (T ) 3GFeSER (T ) 3GSiSER (T )
(12)
As mentioned above, the left hand side in (12) is the free energy relative to the
reference state, called stable element reference (SER), with the enthalpies of the elements
GAlf 14 Fe3Si3
evaluated at room temperature (298.15K). The term is the temperature-
dependent energy of formation of the intermetallic compound and is stored together with
α-AlFeSi phase
25
The site occupation in the fourth sublattice is expressed by the site fraction
yi(4) where
Ga AlFeSi y Al
(4)
G1 ySi(4)G2 0.0992RT ( y (4)
Al ln y Al ySi ln ySi )
(4) (4) (4)
(13)
where the first two terms correspond to the free energies of the terminal stoichiometric
compounds and the third is the entropic term.
The free energies of the terminal compounds are
and
26
The Mg2Si phase is also treated as a stoichiometric phase. The Gibbs free energy is
then
In the present study the thermodynamic database for light non-ferrous alloys COST
507 was used [14]. COST-507 database contains thermodynamic data for the 19 elements
most commonly encountered in Al alloys (Al, B, C, Ce, Cr, Cu, Fe, Li, Mg, Mn, N, Nd, Ni, Si, Sn,
V, Y, Zn, Zr). According to the thermodynamic description in the COST507 database Mn does
not enter in the description of the, α-AlFeSi. Therefore this database was amended by
recent thermodynamic descriptions, where a higher number of ternary and quaternary
systems have been assessed [17]. In the present work the description of the α-AlFeSi phase
was amended in order to introduce Mn in the iron sublattice. This was accomplished by the
introduction of the relevant thermodynamic quantities in the GES module of Thermo-Calc,
where the composition and temperature dependence of the various thermodynamic
parameters are stored.
27
Due to the limited diffusion in the solid the average composition is lower than the
composition it would have, if solidification proceeded under global thermodynamic
equilibrium. It is possible to calculate the composition of the solid at the interface Cs as a
function of the fraction solid fs that has formed, by considering a mass balance in Figure 2.2.
28
When a small quantity of solid dfs forms, solute is rejected into the liquid and the
liquid composition increases by dCL. The mass balance gives
fs C L
df L dC
0 (1 f S ) C CL (1 L k )
o (18)
As mentioned above, equation (19) predicts a lower composition in the solid than the
equilibrium prediction, at the same time it predicts a higher fraction of remaining liquid to
be solidified. Towards the end of solidification a rapid increase in the solute concentration
of the liquid as well as an increase in the concentration of the solid to the value Csa results in
the formation of a non-equilibrium eutectic [20]. Therefore the Scheil-Gulliver equation
29
30
31
Table 2.2: Metallographic procedure for 6060, 6082 and 6063 alloys
Stages Alloy
6060 6082 6063
Grinding
SiC Papers (grit)
320 x x x
500 x x
1000 x x x
2400 x x x
Polishing
Diamond Paste
3μm x x x
1μm x x x
Coloidal Silica x x x
Electro Polishing
stainless steel for
cathode;specimen is anode.
Anodize 40-80 s at approximately
x x
0.2 A/cm2 (1.3 A/in.2, or about 20
V dc). Check results on microscope
with crossed polarizers
Chemical etching
Keller's x
Modified Poultons x
The volume fractions of the Mg2Si and α+β AlFeSi phases have been measured,
through quantitative methods based on point counting, using 1000 grit on 10 sections
32
Equilibrium calculations are presented in this section, which aim at identifying the
major phases and their stability limits in the alloys. For the 6060 alloy, the Si isopleth
section for composition 0.48Mg, 0.03Mn and 0.2Fe (in mass %), is depicted in Figure 2.4a.
Similarly the Mg isopleth for 0.4Si, 0.03Mn and 0.2Fe (in mass %) is depicted in Figure 4b.
The Mg2Si phase is stable bellow 500οC. The α-AlFeSi phase is the first phase stabilized
bellow the liquidus temperature. The range of stability of α-AlFeSi extends to lower
temperatures depending on Mg content. On the contrary Si stabilizes the β-AlFeSi phase.
Additionally the isopleths of Figure 2.4 indicate that transformation of α to β-AlFeSi takes
place as the temperature drops below the solidus. For this reason as mentioned above, a
kinetic calculation, described in section 4.4.4, has been performed in order to determine the
amount of β-AlFeSi phase, which forms after solidification.
33
34
Typical Scheil solidification paths, for the studied alloys are shown in Figure 2.6a-d as
temperature vs mass fraction solid. The phase sequence during solidification is FCC→α-
AlFeSi→β-AlFeSi→Mg2Si→Si (diamond). The arrows depict the temperature and the mass
fraction of solid where each phase starts to form. Solidus temperatures for these alloys
were calculated and results are depicted in Figure 2.7.
(a) (b)
(c) (d)
Figure 2.6: Typical solidification paths for a)6082, b)6063, c)6005, d)6060
35
The mole fraction of phases, formed during solidification, was calculated as a function
of temperature and is shown in Figure 2.8 for 6082 alloy. Formation of α-AlFeSi starts at
622οC while further formation of α-AlFeSi stops, when the β-AlFeSi starts to form at 588οC.
The Mg2Si phase starts to form at 585οC and continuous to grow simultaneously with the β-
AlFeSi phase until the quaternary eutectic temperature (554oC) is reached and the
remaining liquid transforms to a mixture of FCC, β-AlFeSi, Mg2Si and Si.
The phase fractions in the as-cast microstructure are 4.14x10-3 α-AlFeSi, 7.16 x10-4 β-
AlFeSi, 5.32x10-3 Mg2Si and 9.18 x10-4 Si-diamond.
Figure 2.8: Mole fraction of phases as a function of temperature during solidification of Al-
0.38Mg-0.4Si-0.2Fe-0.03Mn (mass%)
36
(a) (b)
Figure 2.9: Microsegregation of phases (a) and microsegregation of alloying elements (b)
of Al-0.38Mg-0.4Si-0.2Fe-0.03Mn (mass%) alloy during solidification
A parametric study was performed for the four variants of alloy 6060, with the
compositions given in Table 1. Both variants, 6060_1 and 6060_3, contain 0.38%Mg but
with 0.4 and 0.47% Si respectively, while both variants 6060_2 and 6060_4 contain
0.48%Mg with 0.4 and 0.47%Si respectively.
37
(a) (b)
(c) (d)
Figure 2.10: Segregation of phases for 6060_1 and 6060_3 a) α-AlFeSi, b)β-AlFeSi, c) Mg2Si,
d) Si (Diamond)
38
(c) (d)
Figure 2.11: Segregation of phases in 6060-2 and 6060_4 alloys, a) α-AlFeSi, b)β-AlFeSi, c)
Mg2Si, d) Si (Diamond)
39
(c) (d)
Figure 2.12: Segregation of elements in 6060_1 and 6060_3 alloys a) Si, b)Mg, c)Fe, d)Mn
40
(c) (d)
Figure 2.13: Segregation of elements in 6060_2 and 6060_4 alloys a) Si, b)Mg, c)Fe, d)Mn
Segregation of elements for 6060_1 and 6060_3 is depicted in Figure 2.12, while for
the 6060_2 and 6060_4 in Figure 2.13. The elements depicted are Si in (a), Mg in (b), Fe in
(c) and Mn in (d). Elemental microsegregation is evident in all cases; however the elements
do not exhibit the same degree of microsegregation. A microsegragation index δi is defined
as
CM Co
Co
(20)
where CM is the maximum composition in the segregation profile and Co is the composition
of the first solid to form (at zero weight fraction solid). The calculated values of the
segregation index appear in Table 2.3. The element possessing the highest
41
The Scheil simulation described above defines the phase fractions present in the
microstructure at the end of solidification. According to the isopleth sections in Figure 2.4
transformation of α to β-AlFeSi takes place below the solidus temperature during
equilibrium cooling. However, taking into account the high cooling rates of the solidified
ingots, following the direct-chill casting operation, it should be evaluated whether this
transformation produces any appreciable amount of β-AlFesi phase. A kinetic simulation
using the DICTRA computational kinetics software was performed for the 6082 alloy. The
geometric model used in a previous work [25] was employed and is shown in Figure 2.14.
The calculation domain consists of two regions. The first region corresponds to the fcc
matrix phase while the α-AlFeSi is taken as a dispersed spheroidal phase. The second region
was attached to the right of the first region and corresponds to the β-AlFeSi phase. The
width of the fcc region was taken to be one half of the mean secondary dendrite arm
spacing (10 μm) and the width of the β-AlFeSi corresponded to the mole fraction of the β-
AlFeSi as calculated by the Scheil simulation.
42
For the composition of the 6082 alloy, the mole fraction of β-AlFeSi corresponds to a
width of the β-AlFeSi region equal to 0.0111 μm. The temperature during the simulation
was considered to drop from 600 to 200 οC in 1 h corresponding to a cooling rate of 0.11Ks-
1. In order to take into account the impeding effect of the dispersed α-AlFeSi phase on the
diffusivities of the elements in the fcc matrix phase, a labyrinth factor [26], taken equal to
the volume fraction of the α-AlFeSi phase [27] was employed. The result of the simulation is
given in Figure 2.15 which shows the position of fcc/β-AlFeSi interface as a function of
simulation (cooling) time.
Figure 2.15: The position of fcc/β-AlFeSi interface as a function of simulation (cooling) time
The final position of the interface at 9.99977 μm, to the left of the initial position
(10μm), corresponds to an increase of the volume fraction of β-AlFeSi phase by 2.07 % of
the amount of β-AlFeSi formed during solidification. This value is considered negligible and
43
The purpose of this section is to provide the necessary experimental data, regarding
phase identification and phase fractions, for the validation of the computational
methodology used in this work. As mentioned above, the validation was performed on alloy
6082, since it exhibited the largest fraction of intermetallics due to its higher alloy content.
The as-cast microstructure of 6082 is depicted in the SEM micrograph Figure 2.16a-c.
The EDS signal from areas 1 and 2 consisted of Mg and Si indicating that this phase is Mg2Si.
The signal from area 3 consisted of Al, Fe, Si and Mn indicating that it is the α-AlFeSi phase,
which contains Mn. The signal from area 4 consisted of Al, Fe and Si (without Mn) and was
identified as β-AlFeSi. In order to further verify the phase identification, the composition of
the α-AlFeSi was determined quantitatively.
(a) (b)
(c)
Figure 2.16: The as-cast microstructure of 6082 alloy as observed in the SEM
44
45
Besides the identification of phases in the SEM, it is important to validate the matrix
composition, adjacent to the intermetallics, close to the boundary. The results are shown in
Figure 2.18, with a SEM microstructure in (a), the Scheil simulation for the
microsegregation of elements in (b) and an EDS spectrum from the matrix region adjacent
to the boundary in (c). The measured compositions are compared against the calculated
data from the Scheil simulation in Table 2.5. The agreement between measured and
calculated compositions is considered satisfactory.
46
47
Using the procedure outlined in the experimental procedures section, the SEM phase
identification was correlated with optical micrographs. Typical microstructures are
depicted in Figure 2.19a-c for 6082, 6063 and 6060 respectively. The microstructure
consists of Al-rich solid solution dendrites while the intermetallic phases are segregated at
the secondary dendrite arm boundaries. These phases are identified as Mg2Si (black), α-
AlFeSi (grey plate-like with rounded edges) and β-AlFeSi (grey plate-like). These
observations are also in very good qualitative agreement with available metallographic
observations in the literature [2, 31].
In Figure 2.20 6082 alloy is depicted and the morphology of the intermetallic
compound α-AlFeSi appears with a typical lamellar morphology known as “Chinese-script”
structure.
48
(b)
(c)
Figure 2.19: Typical microstructure of alloys (a) 6082, (b)6063 and
(c)6060
49
Further metallographic examination has been performed in the 6082 alloy. The Scheil
simulation on 6082 (Figure 2.6a) indicated that the last liquid solidifies by a quaternary
eutectic consisting of FCC + β-AlFeSi + Mg2Si + Si at 544οC. The amount of this quaternary
eutectic is small compared with the primary phases and forms when less than 10% liquid
remains, in agreement with ref. [31]. The quaternary eutectic constituent is depicted in
Figures 2.21a-b.
The presence of spherical Mg2Si and eutectic phases are also shown in Figure 2.21b.
(a) (b)
Figure 2.21: The (a) quaternary eutectic structure and spherical Mg2Si and eutectics of
6082 alloy in as-cast condition
50
51
52
The results of the Scheil simulations have shown that increasing the Mn level from a
low 0.03% to a high 0.45mass% has two major effects, for the high-strength alloys with
more than 0.4mass% Si:
reduction of β-AlFeSi mole fraction
reduction of Mg2Si mole fraction
These reductions are expressed for β-AlFeSi as
f 0.2 Mn f 0.03Mn
f
f 0.03Mn
(21)
and for Mg2Si as
2 Si f Mg 2 Si
0.2 Mn 0.03 Mn
f Mg
f Mg 2 Si 0.03 Mn
f Mg 2 Si
(22)
where f, denotes the mole fractions of these phases at the two levels of Mn content 0.03 and
0.2 mass%. These reductions are plotted in Figure 2.23 for the β-AlFeSi phase and Fig.24a, b
for the Mg2Si phase as a function of Si content (above 0.4mass%) for various levels of Mg
ranging from 0.5 to 1.2 mass%. In general the reduction in the β-AlFeSi is up to 15% while
the respective reduction in the Mg2Si can reach 18%. Regarding the β-AlFeSi phase, with
increasing the Si content (for fixed Mg content) the reduction first increases and then
decreases. The maximum reduction of β-AlFeSi is located between 0.5-0.6%Si depending on
53
Figure 2.23: Mn-induced reduction of β-AlFeSi phase as a function of Si alloy content for
various levels of Mg in the alloy
Regarding the reduction in the Mg2Si phase, two behaviours are sheen. The first is
shown in Figure 2.24a for the 0.4-0.6 Mg level (mass%) where the reductions in Mg2Si are
higher and can reach values up to 27% depending on the Si level. In this range the lowest
reduction in Mg2Si (2%) appears for the 0.8Si-0.6Mg composition. The second is shown in
Figure 2.24b for the higher level of Mg between 0.7 and 1.2 mass% where the reductions in
Mg2Si are lower and limited to values up to 13%. Here the minimum reduction (0%) in
Mg2Si appears for the 0.5Si-1.2Mg composition. It is clear that the Mg-Si compositions for
the highest reduction in β-AlFeSi do not coincide with those for lowest reduction in Mg2Si.
54
Alloys 6060 and 6063 contain 0.03 mass% Mn while alloy 6082 contains 0.45 mass%
Mn. Therefore, Scheil calculations for phase fraction mapping were conducted for these two
Mn levels. The mole fraction of β-AlFeSi and Mg2Si, is depicted in the contour plots of fixed
mole fractions in Figures 2.25 (a) and (b) respectively as a function of Mg and Si content for
the two Mn levels of 0.03 and 0.45 mass %. The mole fraction of the β-AlFeSi phase
decreases towards the lower right corner of the contour plot, indicating that low β-AlFeSi
can be obtained at lower Si and higher Mg contents. The 0.45 % Mn iso-β lines lie above the
respective 0.03 % Mn iso-β lines, indicating that for fixed Mg content, higher Mn allows the
use of higher Si, without further increase of β-AlFeSi (Si stabilizes β-AlFeSi). The spacing
between the iso-β lines for the two Mn levels increases towards the upper left corner of the
contour plot, indicating that the effect of Mn (in reducing β-AlFeSi) becomes stronger at
higher Si and lower Mg contents.
55
Regarding the Mg2Si phase (Figure 2.25b) it is clear that the mole fraction increases
towards the upper right corner of the contour plot, indicating that an increase of both Mg
and Si results in an increase of Mg2Si, with Mg having a stronger effect. The 0.45 % Mn iso-
Mg2Si lines lie above the 0.03 % Mn lines, indicating that a higher Mg content is required to
obtain the same Mg2Si mole fraction in the microstructure.
Mapping of both the Mg2Si phase and the β-AlFeSi phase across the Mg-Si space can be very
useful for alloy design purposes. Such maps are shown in Figure 2.26 (a) and (b) for the two
Mn levels of 0.03 and 0.45 % Mn respectively. The mole fractions of the β-AlFeSi phase
range from 0.001 to 0.005 while those for the Mg2Si range from 0.002 to 0.009.
56
The position of alloys 6060, 6063 and 6082 is also indicated on the maps. All three
alloys possess a low fraction of β-AlFeSi indicating an excellent extrudability potential. In
addition, the 6082 alloy possesses a higher fraction of Mg2Si indicating a higher strength
potential compared to 6060 and 6063. The Mg2Si is maximized towards the upper right
corner while the β-AlFeSi phase is minimized towards the lower right corner of the maps.
This allows selection of optimum compositions for favourable as-cast microstructures in
terms of the two phases. The Mn effect can be summarized as the potential to move to
higher Mg and Si compositions in order to obtain higher amount of Mg2Si without at the
same time increasing the β-AlFeSi phase in the as-cast microstructure.
57
From the results presented in this chapter the following conclusions can be drawn:
Computational alloy thermodynamics, based on the CALPHAD approach, has been applied
for the calculation of equilibrium phase constitution as a function of alloy composition and
temperature. Isopleth phase diagrams in multicomponent alloys have been calculated
permitting the determination of the stability limits of the various phases in the system.
This approach was also applied to carry out Scheil-Gulliver solidification simulations
and calculate the resulting microsegregation of elements and phases in the as-cast
microstructure.
Segregation profiles for elements and phases across the grain were determined as a
function of alloy composition.
Standard metallographic techniques including optical microscopy and scanning
electron microscopy (SEM) were applied to reveal the microstructure of the as-cast and
homogenized materials. Phase fractions were measured by applying quantitative image
analysis. The results of the simulations were confirmed experimentally.
The variation of the mole fractions of the extrudability-limiting β-AlFeSi phase and the
strengthening Mg2Si phase with alloying elements can be mapped over the useful range (0-
1.2 mass%) in the Mg-Si composition space.
The constructed maps indicate that low mole fractions of β-AlFeSi are associated with
lower Si and higher Mg compositions. On the other hand, high mole fractions of Mg2Si are
associated with both higher Si and Mg compositions, with Mg possessing a stronger effect.
Construction of maps for different levels of Mn has shown that addition of Mn could allow
for higher alloying with Mg and Si, in order to obtain higher amount of Mg2Si, without at the
same time increasing the β-AlFeSi intermetallic phase in the as-cast microstructure
58
[1] G. Mrowka-Nowotnik, J. Sieniawski, M. Wierzbinska, Arch. Mater. Sci. Eng. 28(2007) 69-
76.
[2] Y.L. Liu, S.B. Kang, J. Mater. Sci. 32 (1997) 1443-1447.
[3] P.Mukhopadhyay Inter. Sch. Res. Net. ISRN Metallurgy (2012)
[6] W.D. Fei, S.B. Kang, J. Mater. Sci. Lett. 14 (1995) 1795-1797
[7] H. Tanihata, T. Sugawara, K. Matsuda, S. Ikeno, J. Mater. Sci. 34 (1999) 1205-1210.
[8] A. Verma, S. Kumar, P.S. Grant, K.A.Q. O’Reilly, J. Alloys Compd. 55 (2013) 274-282.
[9] A. Wimmer, J. Lee, P. Schumacher, Berg-und Hutten. Monatshefte 157 (2012) 301-305.
[10] Y.L. Liu, S.B. Kang, J. Mater. Sci. 32 (1997) 1443-1447
[11] Y. Du, Y.A. Chang, S.D. Liu, B.Y. Huang, F.Y. Xie, Y. Yang, S.L. Chen, Z. Metallkunde 96
(2005) 1351-1362
[12] J.O. Andersson, T. Helander, L. Höglund, P. Shi, B. Sundman, Calphad 26 (2002) 273
312.
[13] H.L. Lukas, S.G. Fries, B. Sundman, Computational thermodynamics, the Calphad
method, Cambridge university press, 2007
[14] J. Ansara, A.T. Dinsdale, M.H. Rand: COST 507: Thermochemical database for light
metal alloys, European Community publication, 2, EUR 18449 (1991).
[15] G. Inden, Physica 103B (1981) 82-100
[16] B. Sundman, Agren, J., Journal of Physics and Chemistry of Solids 42 (1981) 297-301.
[17] J. Lacaze, L. Eleno, B. Sundman, Metall. Mater. Trans. A 41A (2010) 2208-2215.
[18] E. Scheil, E. Metallkde, 34 (1942) 70.
[19] C.H. Gulliver, J. Inst. Metals, 9 (1913) 120.
[20] R. W. Balluffi, S. M. Allen, W. G. Carter, “Kinetics of Materials” Wiley- Interscience
(2005)
[21] N. Belov , D. Eskin, A. Aksenov Elsevier (2005) 17-33
[22] M. Wu, A. Ludwig, Metall. Mater. Trans. A 38A (2007) 1465-1475.
[23] M. Wu, J. Li, A. Ludwig, A. Kharicha, Comput. Mater. Sci. 79 (2013) 830-840
[24] ASTM Specification, E-522 (1998)
[25] Haidemenopoulos, G.N., Kamoutsi, H., Zervaki, A.D., Jour. Mat. Proc. Tech. 212 (2012),
2255-2260
[26] Rayleigh, J.W, Ph. Mag. 34, (1892) 481–502
[27] Zhang, W., Du, Y., Peng, Y., Wen, G., Wang, S., In. Jour. R. Met. H. Mat. 104, (2013) 721-
735
[28] Mondolfo L.F., 1976. Aluminum Alloys: Structure and properties. Butterworths,
London-Boston
[29] Warmuzek, M., Rabczak, K., Sieniawski, J., Jour. Mat. Proc. Tech. 175, (2005) 421-426.
[30] Liu, Y.L., Kang, S.B., Kim, H.W., Mat. Let. 41 (1999), 267-272.
[31] Mrowka-Nowotnik, G., Sieniawski, J., Wierzbinska, M., Arch. Mat. Sc.Eng. 28 (2007), 69-
76.
59
Homogenization is a significant link of the process chain of the extrudable 6xxx Al-
alloys. It follows casting and takes place before extrusion. The main purpose of
homogenization is, therefore, to remove the detrimental effects of microsegregation left
over by the casting process and to “condition” the material for the extrusion process. In that
respect, homogenization can be considered as a “key process”, since it controls in a large
degree the extrudability of the alloy. The homogenization has received considerable
attention, mostly in experimental terms, both in laboratory and industrial scale. There is a
large amount of published work, briefly reviewed in the next section, on the effects of the
homogenization parameters on the degree of homogenization. However, despite its
importance, the work on simulation of the homogenization process is limited. This is due to
the complexity of the process, since there is a series of phase transformations, which take
place concurrently, in a multiphase and multicomponent system, such as the Al-Mg-Si-Fe-
Mn system under consideration. The most important phenomena, which take place during
homogenization of the as-cast billets, are the following:
1. Removal of the microsegregation of the alloying elements
2. Removal of non-equilibrium low melting point eutectics, particles and segregation
gradients which provide areas with low melting point in order to avoid cracking or tearing
during subsequent hot working
3. Shape control (spheroidization) of remaining intermetallics and dispersoids with sharp
edges such as Fe-based intermetallics, which give poor workability
4. Formation of secondary particles (dispersoids) for grain size control during extrusion or
rolling.
5. The dissolution of the Mg2Si phase and the reprecipitation during cooling
6. The β → α-AlFeSi transformation.
The homogenization process parameters are the homogenization temperature,
homogenization time and the cooling rate from the homogenization temperature. The
correct selection of the homogenization temperature ensures that all necessary
61
62
63
64
65
The key elements of the computational approach to simulate the homogenization process
are the following:
The fact that the grain size in the as-cast microstructure is not uniform is incorporated in
the model. A Dual-Grain Model (DGM) has been developed to serve as the computational
region. The region represents two neighboring grains, in contact, with significantly
different grain sizes. In this way the fact that the transformation rate is higher in the
smaller grains is taken into account. The DGM region serves as the representative volume
element (RVE) for the solution of the homogenization problem.
The initial conditions in terms of composition and phases present in the DGM correspond
to the actual microsegregation of the as-cast microstructure. The composition as well as
the phase fractions profiles that were introduced as initial conditions was the Scheil
microsegregation profiles, calculated from the relevant Scheil simulations for the specific
alloy. The Scheil simulations were presented in Chapter 2.
Two key phase transformations are treated concurrently in both grains of the DGM. These
are the Mg2Si dissolution and the β to α-AlFeSi transformation. Both are diffusional
transformations and their rate is controlled by the diffusion of alloying elements. Since
the diffusion of Mg and Si is faster than the diffusion of Fe and Mn, it is expected that the
dissolution of Mg2Si is faster than the β to α-AlFeSi transformation. In the DGM model,
only lattice diffusion through the matrix has been considered. Grain boundary diffusion or
dislocation assisted diffusion were not considered, since in the homogenization process
the diffusion takes place from the boundaries to the grain interiors in order to eliminate
the microsegregation gradients. In addition, the dislocation density of the material has
normal values, since the material has not undergone any cold deformation before
homogenization.
The intermetallic phases were treated as non-diffusion phases. This means that they act
entirely as sinks or sources of alloying elements. Diffusion is restricted exclusively to the
matrix.
66
Figure 3.1: The Dual Grain Model considering the different grain sizes between
neighboring grains
The total region size is 110μm considering the two grains in contact. Due to the
symmetry, only half of the grains were considered.
In order to input the concentration profile of alloying elements as well as the phase
fractions profiles resulting from the Scheil simulations as initial conditions, the fraction
solid axis fs was converted to a distance axis via the relation
d
xs f s (1)
2
67
This diffusion problem was solved with the computational kinetics package DICTRA
[25]. One basic model assumption is that the growth and dissolution rates of the precipitate
phases are very high compared with the diffusion rates in the matrix. Alternatively, the
matrix diffusion is the controlling mechanism of the overall kinetics. This assumption is
tolerable for the high homogenization temperatures, since the growth-dissolution rates are
very high and the particles reach the equilibrium state very fast. Consequently the local
conditions are rather concentration dependent than time-dependent. Locally, most of the
solute is trapped into the particles, reducing the concentration gradients and the diffusion
rate in the matrix. The “Dispersed Phase” module in DICTRA was employed as shown in
Figure 3.2. It treats problems involving diffusion through microstructures containing
dispersed precipitates of secondary phases. Diffusion is assumed to take place only in the
matrix phase. The dispersed phases are considered as “non-diffusion phases”. They act as
point sinks or sources of solute atoms in the simulation and their fraction is calculated from
the average composition in each node, assuming local equilibrium. The growth or
dissolution of phases leads to adjustments in the concentration profiles of the elements to
be used in the next time-step of the calculation. What changes during the calculations in
each time step is the volume fraction of phases and the local matrix compositions
(composition profiles) through the matrix diffusion.
The mathematical formulation of the multicomponent diffusion problem is based on
the Onsager principles and is described here briefly. A full description is presented in [26-
28].
68
stating, that the rate of entropy production (free energy dissipation) is the sum of the
products between forces and fluxes. In the diffusion case the “forces” are the chemical
potential gradients, which drive diffusion, and “fluxes” are the diffusive fluxes. The second
Onsager principle states that the forces are linearly related to the fluxes by the expression
N
J i Lik Fkj (2)
k 1
In the V-frame, where the volume is fixed, the fluxes of the N components are related
with the expression
N
J
i 1
i 0 (3)
meaning that there are N-1 independent fluxes. The i are the atomic volumes of the
69
since there are N-1 independent concentrations. Additionally the Gibbs-Duhem relation
imposes an additional constraint that only N-1 chemical potentials can vary independently
so that
i N 1 i c j
(9)
x j 1 c j x
where Lik are the Onsager coefficients related to the atomic mobilities and kj are the
relevant thermodynamic factors which express the dependence of the chemical potential on
composition. It is to be noted that in DICTRA, the mobilities are stored in the so-called
mobility database, while the thermodynamic factors are calculated via a link to Thermo-
Calc. In this way DICTRA takes into account the composition dependence of the diffusion
coefficient. The analysis of a N component system requires N-1 independent concentrations
and (N-1)2 independent interdiffusivities. In the Al-Mg-Si-Fe-Mn system considered in this
problem, the diffusion matrix contains 16 interdiffusivities.
Coming back to the DGM, the solution of the diffusion equation is performed under the
following boundary conditions. Considering a closed system, the boundary conditions are:
J i (0, t ) J i ( L, t ) 0 (13)
70
where cis ( x) and f kScheil are the compositions and phase fractions respectively resulting
described by a frequency factor M ko and activation energy Gi* and these are related in
equation:
M io G*
Mi exp( i ) (12)
RT RT
where R is the gas constant and T is the absolute temperature. Both M i and Gi* are
71
Regarding the time step used in the simulations, the maximum time step allowed, or
the highest time step actually taken in the calculation, would affect the final result of the
simulation. This effect will be most evident if one or more of the diffusing species have low
solubility in the matrix phase. When this is the case, a supersaturation is created in the
matrix phase during a diffusion step and if too large a time step is allowed, then too much
supersaturation is created. For this reason a time step sensitivity analysis has been
performed to assure that there is no time step dependence of the solution.
In order to validate the results of the computations of the Dual Grain Model,
laboratory homogenization treatments were performed in selected alloys listed in Table 2.1
in order to study the progress of homogenization. The homogenization treatments were
performed in the temperature range of 540oC-580oC for 30min, 4hr, 6hr and 32 hr. In order
to alter the cooling rate, three cooling methods were used: water Quenching (Q), Air cooling
(AC) and Forced air cooling (FC). The effect of the cooling rate is discussed in Chapter 5.
After the homogenization heat treatment, the specimens were prepared for standard
metallographic examination in order to reveal the microstructure and apply quantitative
image analysis. The evolution of the elemental profiles (removal of microsegregation) was
performed by SEM-EDX measurements. Phase characterization using EDS elemental
72
73
The results of the homogenization Dual Grain Model (DGM) are presented in this
section for the 6082 alloy. The evolution of the volume fraction of phases with time is
depicted in Figure 3.4a. Figure 3.4a depicts the isothermal holding at 540ο C up to 32h
(115200 sec). The Mg2Si dissolves quickly at the early stages of homogenization (below 500
sec). After an induction period of about 700 sec, the transformation β-AlFeSi→α-AlFeSi
starts. This is indicated by the reduction of the volume fraction of the β-AlFeSi phase and
the corresponding increase of the α-AlFeSi phase. The β-AlFeSi phase dissolves at a time
about 6000 sec (100 min), where the α-AlFeSi obtains its highest volume fraction and then
the volume fraction remains constant in the remaining holding time. In addition to the
holding period of 32 h, Figure 3.4b describes the cooling process with a cooling time of 2h
and the preheating prior to extrusion to 500οC which is obtained in 1h. Mg2Si re-
precipitation takes place during cooling from the homogenization temperature and the
volume fraction of Mg2Si is now much higher than the Mg2Si present in the as-cast
microstructure (at t=0). The respective values are 0.65% and 0.1% respectively. This is
attributed to the additional amount of Si coming from the dissolution of the Si-diamond
phase and the β-AlFeSi→α-AlFeSi transformation during homogenization, since the α-
AlFeSi phase is leaner in Si. Thus there is more Mg and Si available in the matrix to form
Mg2Si during cooling from the homogenization temperature. This Mg2Si dissolves again
during pre-heating prior to extrusion. Another interesting effect is depicted in Figure 3.4b.
The reverse transformation α-AlFeSi→β-AlFeSi seems to take place during cooling
and the direct transformation β-AlFeSi→α-AlFeSi takes place during billet preheating. This
is in line with phase stabilities since β-AlFeSi is stable at low temperatures. However in the
cooling rates encountered in industrial practice, the reverse transformation α-AlFeSi→β-
AlFeSi does not take place. Therefore after cooling from the homogenization temperature,
the iron intermetallic present in the microstructure is the α-AlFeSi phase, provided that the
direct transformation β-AlFeSi→α-AlFeSi during homogenization has been completed.
The spatial evolution of the Mg2Si dissolution and the β-AlFeSi→α-AlFeSi transformation
are presented below.
74
(b)
Figure 3.4 Temporal evolution of phase transformations during the homogenization
process a) Isothermal holding at 540οC for 32 h and b) holding at 540ο C for
32h cooling down (500οC/2h), preheating 500οC/1h.
75
While the temporal evolution of the integrated volume fractions was depicted in
Fig.3.4, with the DGM model is possible to monitor the spatial evolution of phase fractions
at specific times during homogenization. The dissolution of the Mg2Si phase from the grain
boundaries and its re-precipitation in the grain interiors during cooling from the
homogenization temperature has been investigated in this way. The reduction of the mole
fraction of Mg2Si during homogenization at 540oC is shown in Figure.3.5. The t=0 curve is
the Mg2Si profile at the beginning of homogenization (as-cast condition, Scheil profile). As
seen in Figure.3.5, the Mg2Si phase dissolves rapidly and the complete dissolution of Mg2Si
takes only about 6 min at 580oC for 6082 alloy.
Figure 3.5: The spatial evolution of the Mg2Si phase from the grain boundaries and its re-
precipitation in the grain interiors
The dissolution of the Mg2Si is faster in the small grain (20μm) due to the shorter
diffusion distances. In addition the Mg2Si profile spreads away from the boundary in the
large grain. This is due to Mg and Si diffusion towards the grain interior causing
precipitation of Mg2Si in new locations. Then this Mg2Si also dissolves and in this way the
Mg and Si diffuse to the grain interior. Following the complete dissolution of Mg2Si during
isothermal holding at 580oC, the Mg2Si phase is re-precipitated during cooling from the
76
(a)
(b)
The intermediate stages of the transformation are depicted in Fig.3.7 (a-d) for times
15, 45min, 1h and 1 ½ hours. The following observations can be made:
77
(a)
(b)
78
(d)
The effect of homogenization temperature and grain size on the dissolution of the
Mg2Si and the transformation of β→α-AlFeSi has been studied. As stated above, grain size
influences the transformation kinetics since it controls the distance over which diffusion of
alloying elements takes place. Homogenization temperature influences the diffusion rate.
For the 6082 alloy the weight fraction of Mg2Si phase is depicted in Figure 3.8a for 3
different homogenization temperatures, 540, 560 and 580oC for homogenization time
t=2sec. It is obvious that the Mg2Si dissolution is much faster at the high homogenization
temperature. Similarly the weight fraction of β-AlFeSi phase is depicted in Figure 3.8b for
the same temperatures and for homogenization time t=500sec. It is also evident that the
transformation rate increases with the homogenization temperature.
79
b)
Figure 3.8: The weight fraction of a)Mg2Si phase and b)β-AlFeSi phase for 3
different homogenization temperatures, 540, 560 and 580oC for
homogenization time t=2sec and 500sec respectively
Furthermore the dissolution of Mg2Si and β-AlFeSi has a different transformation rate
depending from the grain size. At 580οC and t=2sec the Mg2Si dissolution has been
completed only at the 20μm grain. The same stands the β-AlFeSi phase at t=500sec.
80
As stated in the introduction of this chapter, one of the aims of homogenization is the
removal of microsegregation, which is present in the as-cast microstructure.
The evolution of the concentration profiles of Mg, Si, Fe and Mn in the FCC phase during
homogenization at 540oC as predicted by the DGM model is given in Figure 3.9 a-d The
homogenization times considered are 0.5, 4, 8 and 32 hours. The t=0 curve corresponds to
the as-cast microsegregation profile (Scheil profile). The following remarks can be made:
The concentration profiles for Mg, Si and Mn become more uniform with
homogenization time. The profile for Fe does not change appreciably due to the very
low diffusivity of Fe in Al.
The average concentration of Mg and Si in the matrix interior (away from the
boundary) increase with homogenization time. This explains the larger amount of
Mg2Si precipitated during cooling discussed in the previous section.
The profiles homogenize faster in the smaller grain
The compositional fluctuations close to the boundary are due to the dissolution of
Mg2Si and the β-AlFeSi→α-AlFeSi transformation. These fluctuations decay with
homogenization time.
The spatial evolution of the concentration profiles is consistent with the spatial
evolution of the phase fractions.
81
c) d)
Figure 3.9: The evolution of the concentration profiles of Mg, Si, Fe and Mn in the FCC
phase during homogenization at 540oC
82
Figure 3.10: Homogenization process map of 6082 alloy based on the β-AlFeSi→α-AlFeSi
transformation
The purpose of this section is to provide the necessary experimental data, regarding
the validation of the phase transformations take place during homogenization e.g Mg2Si
dissolution, β→α-AlFeSi transformation and alloying element elimination.
As mention in previous chapter the validation was performed on alloys 6082 and on 2 6060
alloys. Homogenization heat treatments at 560οC for 30min, 4h, 8h and 32h have been
applied. The microstructure was revealed using the procedure outlined above i.e. light
83
84
Figure3.11: The effect of the homogenization time of the 6082 aluminum alloy for 30min 2,4 and 8h
.
85
(a)
(b) (c)
(d) (f)
Figure 3.12: SEM micrographs of 6082 alloy a) as- cast condition b) 0.5, c) 4h, d)8h, e)32h
86
87
The XRD spectrum for the 6060 alloy homogenized at 560oC for 0.5h is depicted in
Figure 3.16. The β-AlFeSi peaks have been replaced with α-AlFeSi peaks, indicating that the
transformation has been completed. The above results are in general agreement with the
DGM simulation results.
88
89
(b)
As indicated in Figure.3.18, the agreement between the simulation result and the
experimental data regarding the evolution of the Mg concentration with homogenization
time is good. The mapping for Mg indicates that Mg is absent from the boundary regions
90
(a)
(b)
91
(a)
(b)
As shown in Figure 3.20, the DGM model overestimates the Mn concentration in the
matrix especially at the short homogenization times, the difference decreases at longer
92
(a)
(b)
93
From the work presented above the following conclusions can be drawn:
The DGM model has been developed to treat the homogenization process in
multicomponent and multiphase Al-alloys exhibiting a large variability of the as-cast grain
size. With the model it is possible to simulate the temporal and spatial evolution of phase
fractions and element concentration during homogenization.
Regarding the evolution of phase fractions during homogenization, the predictions of
the DGM model have been validated experimentally with XRD analysis. The evolution of the
β-AlFeSi→α-AlFeSi transformation is predicted by the DGM and is confirmed by XRD in the
same time scale.
The DGM predictions regarding the evolution of Fe and Mg concentrations with
homogenization time are in excellent agreement. There are only some discrepancies in the
profiles of Si and Mn. The Si discrepancy is due to the Si-diamond phase not taken into
account in the DGM. The Mn discrepancy is due to the Mn retention in the α-AlFeSi phase.
The DGM predicts the fast dissolution of Mg2Si during homogenization and its re-
precipitation during cooling. It also predicts its dissolution during preheating prior to
extrusion.
The DGM can describe the temporal and spatial evolution of the β- -AlFeSi
transformation. The spatial evolution exhibits an exact spatial correspondence. In addition
the starts at a certain distance from the boundary and the transformation front moves
towards the boundary.
Both the Mg2Si dissolution and the β-AlFeSi→α-AlFeSi transformation are faster in the
smaller grain as predicted by the DGM. In addition the concentration profiles of the
elements homogenize faster in the smaller grain.
The DGM predicts correctly the effect of homogenization temperature, with the rate of
both the Mg2Si dissolution and the β-AlFeSi→α-AlFeSi transformation to increase with the
homogenization temperature.
A preliminary attempt to develop homogenization process maps has been performed
using the DGM. These maps are, at present, based on Mg2Si dissolution and the β-AlFeSi→α-
AlFeSi transformation and can be used for the design of the homogenization heat treatment.
94
95
4.3. Methodology
The chemical composition of the 6060_3 alloy investigated is shown in Table 2.2
(Chapter 2). Three homogenization heat treatments, in laboratory scale, were investigated,
following the simulation conditions from Chapter 3, holding at 540ο, 560ο, and 580oC for 4h
followed by air cooling (Table 4.1). These conditions ensured the complete transformation
of β→α-AlFeSi.
98
After the homogenization heat treatment, the specimens were prepared for standard
metallographic examination involving optical microscopy, SEM and image analysis as
discussed in previous Chapters, in order to reveal the microstructure. The as-cast as well as
the homogenized microstructures were characterized for intermetallic phases and the
particles were categorized in three morphological types as rounded particles, pinched
particles and particles exhibiting a neckless formation. It should be noted that pinched
particles are those that are in the initial stage of separation in smaller rounded particles
towards the formation of a neckless group. The number of images processed and the
number of particles measured for each alloy appears in Table 4.2.
Table 4.2: Number of images analyzed and number of particles measured per billet for the
quantification of the homogenization state
Number Number
Alloy
of Images of particles
6060_3 58 106
6060_3A 57 161
6060_3B 56 150
6060_3C 58 133
99
Circularity is a measure of
how closely a particle
resembles a circle. It varies 𝑝2
Circularity
from zero to one in magnitude 𝐶=
4𝜋𝐴
with a perfect circle having a
value of one
100
2
1
3
5
4
6
(c)
Figure 4.1: SEM image used for the measurement of indices: (a) Low magnification image,
(b) high magnification isolation of the group of particles, (c) image J display
used for the measurement of the indices
101
The as-cast microstructure depicted in Fig. 4.2a and consists of characteristic “Chinese
– script” morphology of intermetallic α-AlFeSi phase. After homogenization treatment at
540οC and 560οC this morphology still remains in the microstructure. In the contrary, in the
580οC temperature the α-phase seems to have reached a state of satisfactory
spheroidization.
103
(c) (d)
Figure 4.2: Metallographic images, (a) as-cast, (b) homogenized at 540oC, (c) 560oC and (d)
580oC
104
Figure 4.3: Images indicating the rounding of the edges of the particles during
homogenization
At the second stage an initial perturbation can be observed in the particles. Particles
take an ellipse shape with more rounded edges. This stage characterized by the trend of the
particles to be separated into smaller rounded particles. These discontinuities are depicted
in Figure.4.4. The driving force for this behavior is reduction of strain energy associated
with large particles.
105
The reduction of surface energy of the α-phase is the driving force for spheroidization.
With this process, the total interface area between the matrix and the α-AlFeSi phase is
reduced. The particles finally adopt a spherical shape at the third stage as depicted in
Figure. 4.5
Figure 4.5: Images revealing the spherical shape at the third stage
It was found that for the two homogenization temperatures 540ο and560ο C, although
the morphological changes towards spheroidization were activated; only the first stage was
completed. On the other hand for the homogenization at 580οC the α-particles reached the
third stage of spheroidization involving the neckless formation.
The mean values of microstructural indices, Aspect ratio, Feret, Circularity and
Circularity of Edges have been measured for all conditions investigated. The mean value
reported possesses a 95% confidence while the upper and lower points are the 97.5% and
2.5% cut-offs to exclude extreme values. From these data, which are shown in Figure 4.6 a-d
some general remarks can be made.
Regarding the Aspect Ratio (Figure 4.6a) the heat treatments of 6060_3B and 6060_3C
alloys (560ο and 580ο C respectively) have a lower value compared to 6060_3A alloy
(540οC). However 3C alloy, has a lower Feret value (Figure 4.6b) and almost equal with 3A
106
(a) (b)
(c) (d)
Figure 4.6: The values of indices for the as-cast and homogenized alloys: (a) aspect ratio, (b)
feret, (c) circularity and (d) circularity of edges
107
4.7. References
108
An important step in the process chain of extrudable Al-alloys is the cooling from the
homogenization temperature. As discussed in the previous chapters of this thesis,
homogenization involves holding the material at a high temperature for a certain time
period (holding time). During this time several transformations of the as-cast
microstructure take place, including the dissolution of Mg2Si, the transformation of β-AlFeSi
to α-AlFeSi, the spherodization of the α-AlFeSi and the reduction of elemental
microsegregation. During cooling from the homogenization temperature, new
transformations are activated, the most important being the re-precipitation of Mg2Si. The
assumption is made here that although the β-AlFeSi is the thermodynamically stable phase
at low temperatures, the transformation of α-AlFeSi to β-AlFesi does not take place at any
appreciable rate. This is true, especially for the cooling rates encountered in the current
industrial homogenization practices and is due to the very low diffusivity of Fe in the Al
matrix. A DICTRA calculation in Chapter 3 indicated that there is actually no transformation
of α to β-AlFeSi during cooling of the cast billet. Therefore, the only transformation to be
studied is the precipitation of the strengthening Mg2Si phase. This chapter deals with three
issues:
109
110
The majority of phase transformations, which take place in alloys, are diffusional and
proceed by thermally activated movements of atoms across a concentration gradient. Of
significant importance in relation to thermal industrial processing, are phase
transformations involving precipitation reactions resulting to a second phase in the form of
a particle population. During these transformations three basic physical mechanisms are
taking place, the nucleation of new particles the growth of the nucleated particles and the
coarsening of the particles resulting to a particle population which can be described by a
particle size distribution (PSD). The material final mechanical properties depend on the
particles volume fraction and radius and consequently on the shape of the PSD. Since the
original works of Lifshitz & Slyozov [4] and Wagner [5] only a few of the published
precipitation models are comprehensive to deal with non-isothermal transformations
treating nucleation, growth and coarsening taking place concurrently. The two main works
in this area were performed by Langer & Schwartz [6] who studied nucleation and growth
of droplets in near-critical fluids and Kampmann et al. [7,8] who proposed a numerical
procedure (KWN model) for the calculation of the particle size distribution in solid-state
transformations treating the concurrent nucleation, growth and coarsening of precipitates.
As mentioned above the KWN model was modified to handle precipitation of Mg2Si during
cooling from the homogenization temperature. The results of the model were introduced to
the strength model for the calculation of the final strength of the cooled billet, after
homogenization. The strengthening mechanisms considered we lattice resistance, solid
solution strengthening and precipitation strengthening. The details of the models are
discussed in [1-3]. Results will be presented for alloy 6063, where the experimental
validation was performed.
Three homogenization temperatures were considered 540, 560 and 580oC.
Homogenization time is not a parameter in these simulations, since it is considered that the
time is enough for complete dissolution of the Mg2Si phase during holding at the
homogenization temperature. Therefore all Mg and Si are in solid solution at the end of the
homogenization holding period. A critical parameter is the cooling rate. Two cooling
111
Figure 5.1: Measured cooling profiles used in simulations and in the experimental
validation, (a) natural cooling, (b) forced cooling
The simulation results from the KWN model regarding the evolution of the volume
fraction and particle size of the Mg2Si phase as a function of time during cooling from the
homogenization temperature are shown in Figure 5.2. Regarding the volume fraction, it
appears that the homogenization temperature influences the initial stages of the
precipitation. Towards the end of cooling the curves corresponding to the three
homogenization temperatures coincide. The amount of Mg2Si formed under forced cooling
is slightly lower than the respective amount formed under natural cooling conditions, 0.79
and 0.82% respectively. However the final particle size at the end of cooling is lower in the
forced cooling condition, about 6.9nm, compared to 7.8nm for the natural cooling condition.
112
Regarding the shape of the curves describing the size evolution during cooling, is
evident from Figure 5.2b that there is a competition between nucleation and growth
processes, which as mentioned in the Introduction of this chapter, take place
simultaneously. Three stages can be described. Following initial nucleation the mean size
drops due to the increase of the nucleation driving force with the decrease in temperature.
The rise of the mean size is followed by growth of the already nucleated particles. However
the further decrease of temperature during cooling induces a substantially higher driving
force for nucleation and the mean size is again reduced to its final value at the end of
cooling.
The respective particle size distributions (PSD) are given in Figure 5.3 for the 580oC
homogenization temperature and the two cooling profiles. The KWN simulation predicts
similar peak sizes for the two cooling profiles, but it also predicts a second peak and larger
particles for the natural cooling profile. This is due to the longer time available for the
precipitation process in the case of natural cooling.
113
The volume fraction of Mg2Si and the size evolution are used as input to the strength
model. The resulting hardness evolution is shown in Figure 5.4 for the two cooling profiles
and the three homogenization temperatures. The results indicate that forced cooling results
in slightly higher hardness 69 relative to 64.9 HV0.3 for natural cooling. This hardness
difference is attributed to the slightly finer dispersion of Mg2Si formed after forced cooling.
Figure 5.4: Evolution of hardness during cooling of the 6063 alloy for natural and forced
cooling and for three homogenization temperatures 540, 560 and 580oC.
114
The hardness resulting from forced cooling is slightly higher than the one resulting
from natural cooling as predicted by the simulation. Regarding absolute values, the
simulation overestimates the hardness by 3-4 Vickers units in both cases, as shown in
Figure 5.4. Taking into account the approximations used in the KWN precipitation model as
well as the strength model, the agreement is considered satisfactory.
115
Three alloys were investigated, 6082, 6005 and 6063, the chemical composition of
which is given in Table 2.2. The billets were sectioned and specimens were prepared and
homogenized according to the procedures described in Chapter 3. Three homogenization
temperatures 540, 560 and 580oC as well as four homogenization times 0.5, 4, 8 and 32
hours were selected. Similar homogenization parameters have been studied by other
researchers [10-13]. Three conditions regarding the cooling rate were studied: Water
Quenching (Q), air cooling (AC) and forced cooling (FC). A K-type thermocouple was
attached at the specimens in order to monitor the cooling profile. The cooling profiles are
depicted in Figure 5.5. The AC and FC specimens reach 150oC in 6 and 2.5 min respectively.
116
Brinnel hardness testing was performed in all specimens with a 5mm ball indenter
diameter and a load of 250Kg for 15 sec. The indentation diameter was measured by
microscopy. In order to determine the hardening potential, the hardness measurements
were performed immediately after cooling and also after 5 days of natural aging at room
temperature. The 5 day period has been also suggested in [12,14].
5.3.3. Results
117
118
119
In all three alloys, the 8h homogenization treatment appears softer in almost all
homogenization temperatures and cooling rates. Moreover with increasing the Mg and Si
alloying elements the hardness increases for each specific cooling rate. This effect can be
seen more easily in Figures 5.9, 5.10 and 5.11, which depict the comparison between the
three alloys for a 4h homogenization time and for the 540, 560 and 580oC homogenization
temperature respectively.
120
Figure 5.10: Brinnel hardness immediately after cooling from 560oC homogenization
temperature
Figure 5.11: Brinnel hardness immediately after cooling from 580oC homogenization
temperature
121
122
Figure 5.13: Brinnel hardness after homogenization and natural aging for the 6005 alloy
Figure 5.14: Brinnel hardness after homogenization and natural aging for the 6063 alloy
123
H NA H AH
Hp 100
H AH
where, HAA is the hardness after 5 days of natural aging and HAH is the hardness immediately
after homogenization. The results are shown in Figures 5.15, 5.16 and 5.17 for the 6082,
6005 and 6063 alloys.
Figure 5.15: Hardening potential Hp for the 6082 alloy at (a) 540, (b) 560 and (c) 580oC
homogenization temperatures
124
125
For the 6082 alloy, the Q state exhibits the highest hardening potential followed by
the FC and AC states, this is due to the higher amount of Mg and Si retained in solid solution
after quenching from the homogenization temperature. The same effect is observed for the
6005 and 6063 alloys. Regarding the effect of homogenization time, the hardening potential
for the Q, FC and AC states increases with homogenization time, reaches a maximum at 8h
end then decreases. This is in line with the results presented above, indicating that the
softer state following homogenization is the one corresponding to the 8 h homogenization
time. It appears from these results that transformations other than the Mg2Si dissolution
during homogenization influence the Mg and Si in the matrix. The most probable influence
is from the β-AlFeSi→α-AlFeSi transformation, studied in Chapter 3, which liberates Si in
126
5.4.2. Methodology
128
The first peak during cooling is attributed to β-Mg2Si, while the second peak at lower
temperatures is attributed to β΄-Mg2Si precipitation. It is clear that the application of a
higher cooling rate suppresses the precipitation of the β-Mg2Si, leaving only the
precipitation of the β΄-Mg2Si. Literature studies regarding the effect of cooling rate on the
precipitation of Mg2Si are limited. Only one publication regarding the development of
Continuous Cooling Precipitation (CCP) diagram was found [18] and one publication
regarding the isothermal TTP diagram of 6082 [19]. The work presented in this section is
based on the work presented in [18] where the CCP diagrams for five alloys 6060, 6063,
6005, 6082(L) and 6082(H) were determined. The composition of the 6060 alloy studied in
the above work is 0.43 Mg-0.40Si. The aim is to quantify the shift of the CCP with the new
alloy composition of 6060_3. It is evident from the above work that the position of CCP is
influenced by mainly the Mg content and secondly by the excess Si in the alloy. Mg
influences directly the amount of Mg2Si while the excess Si influences the heterogeneous
nucleation sites of the β΄-Mg2Si. The methodology followed is the following:
129
5.4.3. Results
The CCP diagrams for the alloys investigated in [18] are depicted in Figure 5.19.
Mg=0.43 Mg=0.47
Mg=0.57
Mg=0.62 Mg=1.05
Figure 5.19: CCP diagram for 6060, 6063, 6005 and 6082 alloys [18]
130
Table 5.2 Position of β and β noses from the CCP diagrams of Fig.2
Time (sec) Time (sec) Logt Logt
Mg(wt%)
(β-nose) (β-nose) (β-nose) (β-nose)
0.43 1200 450 3.07 2.65
0.47 180 100 2.25 2.00
0.57 170 25 2.23 1.39
0.61 40 10 1.60 1.00
1.05 15 1.8 1.17 0.255
Nonlinear curve fitting has applied to the data of Table 5.3 in order to get an expression for
the position of the noses as a function of Mg content. An Arrhenius expression was selected
for the fitting. The fitting results are displayed in Figure 5.20
(a) (b)
Figure 5.20: Arrhenius fitting of the data of Table 2 a) Fitting the β-nose, b) fitting β’ - nose
For Mg=0.38 for 6060_3 alloy the position of the noses are
β-nose at t=4750 sec
β΄-nose at t=3750 sec
Thus the β and β΄ noses are displaced by 1000 sec based only on the Mg content of the alloy.
The next step is to find the displacement of the β΄-nose due to the excess Si in the alloy. The
excess Si is the Si remaining in the alloy after subtracting the Si necessary for the formation
of Mg2Si and the Si present in the α-AlFeMnSi phase after the β-AlFeSi→α-AlFeMnSi
transformation during homogenization.
The relevant equations are [20]:
[Si]xs = [Si]alloy – [Si]α-AlFeMnSi – [Si]Mg2Si (3)
132
The shift in the CCP diagram of the 6060_3 alloy is shown in Figure 5.22. The dotted β΄
line is shifted by taking into account only the Mg effect as per equation (2). The full line
corresponds to the β΄-nose, taking into account the effect of excess Si in the alloy.
133
The resulting CCP diagram can be used for the design of a homogenization cooling program,
in order to achieve precipitation of the metastable phase β΄-Mg2Si.
The aim of the present section is to formulate design rules regarding the
homogenization practice in order to obtain high extrudability. From the literature and from
the results of this thesis, regarding the simulation of solidification and homogenization of
6xxx alloys, the following preliminary remarks can be made:
Extrudability is a complex property, not measurable directly. Extrudability is usually
determined at the extrusion shop with several ways, like the maximum extrusion speed
accomplished without surface defects or the maximum pressure, or even the contact time in
the press. However, since extrudability is a property, it depends on the microstructure,
which in turn depends on the composition and the process chain from melt treatment and
134
135
136
From the results presented in this chapter the following conclusions can be drawn:
Precipitation during cooling from the homogenization temperature was simulated with the
KWN precipitation model. The relevant particle size distribution (PSD), evolution of
volume fraction and particle size during cooling were calculated. The results were
introduced in a suitable strength model to calculate the resulting hardness. The simulation
results are in satisfactory agreement with experimental data. Forced air cooling results in
higher hardness in relation to natural cooling for the alloy 6063 investigated.
The hardening of the homogenized billets was investigated experimentally. Regarding the
effect of the homogenization cooling rate, water quenching results in higher hardness
relative to air cooling and forced air cooling. The hardness difference increases with
alloying in the order 6063→6005→6082. The effect is attributed to the higher amounts of
Mg and Si retained in solid solution during cooling.
Regarding the effect of alloy composition, homogenization temperature, homogenization
time and homogenization cooling rate on the hardening potential of the homogenized
billets, the alloys water that were quenched from the homogenization temperature
exhibited the higher hardening potential relative to the air and forced cooling. In addition,
the maximum hardening potential is exhibited for the 8h homogenization time,
irrespective of the homogenization temperature.
For the lower alloyed 6063 and 6005 billets, the hardening potential at shorter
homogenization time increase with the homogenization temperature.
The effect of excess Si on the precipitation of β΄-Mg2Si was determined using the KWN
precipitation model. The resulting shift in the continuous cooling precipitation (CCP)
diagram was calculated. This allows the design of a suitable cooling program in order to
avoid the precipitation of β-Mg2Si and promote the precipitation of β΄-Mg2Si. This allows
the use of a lower preheating temperature prior to extrusion and leads to the use of higher
press loads in order to obtain higher extrusion speeds.
Specific design rules were formulated, regarding the as-cast microstructure, the
homogenization temperature and time as well as the cooling from the homogenization
temperature in order to obtain high extrudability.
137
138
The Dual grain model (DGM) has been developed to treat the homogenization
process in multicomponent and multiphase Al-alloys exhibiting a large variability of
the as-cast grain size. With the model it is possible to simulate the temporal and
spatial evolution of phase fractions and element concentration during
homogenization.
139
141
Specific design rules were formulated, regarding the as-cast microstructure, the
homogenization temperature and time as well as the cooling from the
homogenization temperature in order to obtain high extrudability.
The microstructure developed during DC casting is affected by the solidification
conditions. Grain size and dendrite arm spacing (DAS) are the most important
factors. They can be controlled by the casting speed (solidification rate) and seeding
for grain refining. A smaller grain size and DAS improves the strength and also
shortens the diffusion distances, leading to a more efficient homogenization for a
given homogenization cycle.
Phase fraction mapping, developed in Chapter 2, can be used for the selection of
alloy compositions to minimize the undesirable β-AlFeSi intermetallic in the as-cast
microstructure, in order to achieve a good starting point for the homogenization to
follow. In addition, it has been shown that Mn reduces the amount of β-AlFeSi for a
given Mg-Si-Fe combination.
With the DGM model presented in Chapter 3, it is possible to determine the required
time for Mg2Si dissolution and the transformation β→α-AlFeSi as a function of
homogenization temperature and alloy composition. It has also been shown that Mn
has a beneficial effect on the kinetics of the β to α transformation. The DGM model
also indicated that a small as-cast grain size accelerates the kinetics of
homogenization, a point that was indicated above.
Microstructural indices, such as the aspect ratio and the circularity, were
determined in Chapter 4 and can be used to characterize the homogenization stage
quantitatively. A fully homogenized billet, with the potential for high extrudability
should have all β-AlFeSi transformed to α-AlFeSi, with necklace morphology and
with aspect ratio and circularity approaching unity. In addition all Mg2Si should be
dissolved and Mg and Si should be distributed as uniformly as possible in the grain
interiors.
142
143
Model integration
While the simulation of microsegregation has been linked to the DGM, the
precipitation during homogenization cooling is an independent simulation based on the
KWN model. It is proposed to link this model to the DGM in the sense that the results of the
DGM can be used as a direct input to the KWN model. A Matlab code could be written to
perform this task. In this way it would be possible to link also the preheating prior to
extrusion; so that all parts of the process chain prior to extrusion can be integrated. This
will increase the level of model integration and provide an easier-to-handle computational
tool for design purposes.
Homogenization cooling
In this thesis the KWN model was used to simulate the precipitation during
homogenization cooling. The current version of the model handles only one set of
precipitates in binary systems (matrix-precipitate) as in the Al-Mg2Si system. Furthermore
only equilibrium phases are considered. It is proposed to investigate the application of the
new software TC-Prisma, which is based on the KWN model, but additionally relies on the
thermodynamic and kinetic databases of Thermo-Calc and Dictra. The new versions of
Thermo-Calc provide values for the interfacial energies and contain data for metastable
phases, making it possible in the near future to simulate precipitation of metastable or
transition phases, such as the β΄-Mg2Si.
145
The DGM for homogenization can be further refined through parametric simulation
studies such as (a) looking at different geometries (spherical or cylindrical), (b) simulating
different grain sizes, (c) taking into account increased diffusivities due to high dislocation
density. In addition the effect of composition such as the effect of Mn or other alloying
elements on the α-AlFeSi→β-AlFeSi transformation could be investigated.
A next step to process integration is the link of the homogenization model with
available or to-be-developed models for hot extrusion based on the finite element method.
Extrusion is a deformation process carried out under high deformation rates at high
temperatures. Several metallurgical phenomena, such as recovery, recrystallization, phase
146
147
1. P.I. Sarafoglou and G.N. Haidemenopoulos, Phase Fraction Mapping in the As-Cast
Microstructure of Extrudable 6xxx Aluminum Alloys, Int. Journal of Materials Research
105, No.12, pp. 1202-1208 (2014).
2. G.N. Haidemenopoulos and P.I. Sarafoglou, Extrudable Al-alloys: microsegregation
and homogenization, to appear in the Encyclopedia of Aluminum Alloys, Taylor and
Francis, NY, (invited paper, under preparation)
3. P.I. Sarafoglou, I. Aristeidakis, M. Tzini and G,N. Haidemenopoulos, Index-based
quantification of the homogenization state in extrudable Al-alloys, To be submitted in
Metals, 2016 (under preparation)
4. P.I. Sarafoglou and G.N. Haidemenopoulos, Simulation of the microsegregation and
homogenization, including hardening during homogenization cooling, of the 6082 Al-
alloy, to be submitted in Materials Science and Engineering A, 2016 (under
preparation)
5. P.I Sarafoglou and G.N. Haidemenopoulos, On the computational-based design of 6xxx
series extrudable Al-alloys, MSE Conference, Darmstadt, Germany, 23-25 September
2014.
6. P.I. Sarafoglou and G.N. Haidemenopoulos, Design of 6xxx alloys: Control of
intermetallic phases in the as-cast microstructure, 5th Hellenic Conference on
Metallic Materials, Volos, 2013
7. Maria-Ioanna Tzini, Panagiota Sarafoglou, Andreas Stieben, Gregory
Haidemenopoulos, Wolfgang Bleck, Austenite evolution and solute
partitioning during thermal cycling in the intercritical range of a medium-Mn steel,
Junior Euromat, 10-16 July, 2016, Lausane, Switzerland
In addition to the publications based on the thesis the following publications were
concluded:
149
150