You are on page 1of 13

Available online at www.sciencedirect.

com

Journal of Magnesium and Alloys 7 (2019) 725–737


www.elsevier.com/locate/jma

Full Length Article


A systematic investigation of secondary phase dissolution in Mg–Sn
alloys
Arushi Dev a, Niladri Naskar b, Nishant Kumar b, Ashutosh Jena a, Manas Paliwal a,∗
a Department of Materials Science and Engineering, Indian Institute of Technology, Gandhinagar, Palaj Campus, Gandhinagar, Gujarat 382355, India
b Seoul Nation University, Seoul 08826, Republic of Korea

Received 16 September 2019; accepted 1 November 2019


Available online 18 November 2019

Abstract
As-cast Mg–Sn alloys (3, 6, and 9 wt% Sn) were solution treated at 653, 703 and 753 K (380, 430 and 480 °C) for 1, 4, 8, 12 and 24 h
to determine the variation of secondary phase with respect to Sn content, temperature and time. Mg-3 wt% Sn exhibits Mg2 Sn dissolution at
all solution treatment temperatures whereas Mg-6 and 9 wt% Sn alloy displays Mg2 Sn reprecipitation and dissolution depending on the heat
treatment temperature. In addition, a combined mathematical model that predicts the secondary phase dissolution and solute redistribution
as a function of temperature and time is presented in this work. This model is a significant improvement compared to the previous studies
where the dissolution and homogenization processes are considered independently. The effect of grain size and solute mobility upon the
dissolution and homogenization kinetics is discussed as well.
© 2019 Published by Elsevier B.V. on behalf of Chongqing University.
This is an open access article under the CC BY-NC-ND license. (http://creativecommons.org/licenses/by-nc-nd/4.0/)
Peer review under responsibility of Chongqing University

Keywords: Solution treatment; Homogenization; Dissolution; Mg–Sn alloys.

1. Introduction Homogenization or solution treatment is an important


downstream process in the wrought alloy production. Reduc-
Improving high temperature strength, creep resistance and ing microsegregation and dissolution of secondary phase is an
room temperature formability are the key properties desired important objective of this step as a homogenous single phase
in application of Mg and its alloys in transportation and is ideal for further deformation processes such as rolling or
aerospace sectors. The most important Mg alloys in com- forging. The homogenization parameters such as time and
mercial use are AZ (Al and Zn), AM (Al and Mn) and AS temperature are dependent on the as-cast microstructural fea-
(Al and Si) as they are used in both as-cast and wrought tures. These parameters are very well established for AZ al-
conditions. However, these alloys are unsuitable for compo- loys and numerous studies exist in literature that present de-
nents working at high temperatures (120 °C and above) be- tailed microstructural evolution with respect to homogeniza-
cause their microstructure becomes thermally unstable (rel- tion temperature and time. Mostly, a hit and trial approach
atively low eutectic temperature of Mg–Al system, about is followed to predict the homogenization process for an un-
410 °C, Mg + Mg17 Al12  Liquid) and discontinuous precip- known alloy but this approach is time consuming and costly.
itation occurs at grain boundaries [1], leading to lower creep Hence, a suitable homogenization model that predicts the so-
resistance and decreased strength. Accordingly, new Mg al- lute distribution and dissolution kinetics of a secondary phase
loys which include Mg-Sn series are being actively investi- with time and temperature is essential for development of new
gated [2–7]. Mg alloys. In addition, few key experiments are required to
validate the modelling results.
∗ Corresponding author.
Mg–Sn based alloys are emerging as promising alloys and
E-mail address: manas.paliwal@iitgn.ac.in (M. Paliwal). offer good combination of mechanical properties [7,8]. The
https://doi.org/10.1016/j.jma.2019.11.002
2213-9567/© 2019 Published by Elsevier B.V. on behalf of Chongqing University. This is an open access article under the CC BY-NC-ND license.
(http://creativecommons.org/licenses/by-nc-nd/4.0/) Peer review under responsibility of Chongqing University
726 A. Dev, N. Naskar and N. Kumar et al. / Journal of Magnesium and Alloys 7 (2019) 725–737

influence of alloying elements upon the final mechanical prop-


erties of Mg–Sn wrought alloys is extensively investigated.
However, there is no systematic study that reports the disso-
lution and homogenization kinetics during the solution treat-
ment process of Mg–Sn alloys. The microstructural control
during the deformation and final ageing treatment of wrought
alloys is only achieved if the homogenized microstructure is
properly optimized at suitable time and temperature.
Recently, present authors performed a combination of ex-
perimental and solidification simulations to evaluate the re-
lationship between solidification parameters such as thermal
gradient, solidification velocity and cooling rate upon the as-
cast microstructural features (secondary dendrite arm spacing,
secondary phase fraction and microsegregation) [9]. Wedge
casting experiments were performed to obtain solidified mi-
crostructures in the cooling rate range of 5–150 K/s for Mg-
3, 6 and 9 wt% Sn alloys. Solidification calculations that in-
corporate solute back diffusion, secondary arm coarsening,
dendrite tip undercooling, dynamically linked with accurate
thermodynamic databases, were performed for accurate anal-
ysis of the experimental results. Shi and Luo [10] investigated
the microstructural evolution during the ageing of Mg–Al–Sn
alloys using a combination of experiments and simulations. Fig. 1. Schematic illustration of the sample from water-cooled Cu plate
shaped mould.
Sazol et.al. [11] performed a systematic investigation on the
dissolution and homogenization kinetics of binary Mg–Al,
Mg–Zn and ternary Mg–Al–Zn alloy. They performed homog- (OES) and Inductive Coupled Plasma (ICP) analysis. Bars of
enization of as-cast Mg–Al, Mg–Zn and Mg–Al–Zn alloys 1-inch thickness (as shown in Fig. 1) were cut from the centre
at 603 and 673 K for 1, 2, 4 and 8 h to determine the sec- of the plate in order to avoid inhomogeneous microstructure.
ondary phase dissolution and solute profile in hcp Mg using Next, small cubed specimens (1 × 1 × 1 cm) were sectioned
SEM-Image analysis and EPMA-WDS respectively. In addi- from large 1-inch bars that were cut earlier. Sharp edges of
tion, they correlated the experimental data with a two-step the specimens were removed by polishing the on 400 grit
explicit diffusion model where two separate dissolution and size SiC sand papers prior to the homogenization experiments.
solute distribution calculation routines were adopted. Samples from parts near the surface were discarded to avoid
The purpose of the present study is to systematically in- heterogeneous chilled microstructure. The homogenization ex-
vestigate the microstructural change of Mg–Sn (3, 6, 9 wt% periments were carried out at 653, 703 and 753 K (380, 430,
Sn) alloys during the solution treatment using a combination and 480 °C) using a box furnace (Nabertherm GmbH 1100 °C,
of experiments and simulations. An implicit diffusion-based 230 V, 2.4 kW furnace). The calibration of the box furnace
solution treatment model is presented in this study that com- was performed at 520 °C for 8 h using a K-type sheathed
bines the dissolution of secondary phase and redistribution of thermocouple. The temperature sensitivity was reported to
solute (Sn) in hcp Mg using a robust numerical technique. The be +3 °C. The specimens were placed exactly at the cali-
effect of grain size and solute diffusivity upon the dissolution bration positions and the duration of the solution treatment
and homogenization kinetics is discussed and a strategy to varied between 1 and 24 h (1, 4, 8, 12 and 24 h). Heating rate
decide the optimum solution treatment is outlined as well. In was kept constant as 10 °C/min followed by isothermal hold-
addition, the as-cast microstructural features (secondary phase ing at various temperatures and time. After the isothermal
fraction and microsegregation) that are used as input to the holding, the samples were quickly taken out from the fur-
solution treatment model are calculated using our in-house nace and quenched in cold water. Next, the specimens were
solidification model. subjected to standard sample preparation method for further
characterization using SEM JEOL JSM-7600F (Field Emis-
2. Experimental sion Scanning Electron Microscope) equipped with Electron
Dispersive Spectroscopy (EDS) to acquire backscattered elec-
As-cast samples of Mg–Sn alloys (3, 6, 9 wt% Sn) were tron (BSE) images and identify the second phase (Mg2 Sn)
prepared using rectangular plate shape water cooled copper in the samples. Multiple images were taken in order to re-
mould. The geometry of the plate is indicated in Fig. 1. The duce the statistical error. The second phase fraction was calcu-
cooling rate during the solidification was measured by insert- lated with the help of image analysis technique using ImageJ
ing a K-type couple at the centre of the mould. The cooling software. The two-dimensional area fraction of second phase
rate was measured to be 30 K/s and the melt chemical com- was considered equal to the secondary phase fraction in the
positions were confirmed by Optical Emission Spectroscopy alloy.
A. Dev, N. Naskar and N. Kumar et al. / Journal of Magnesium and Alloys 7 (2019) 725–737 727

Fig. 2. Mg rich side of the Mg–Sn phase diagram. Alloy compositions along with heat treatment temperatures are indicated in the phase diagram.

The Mg–Sn binary alloys along with solutions treatment plicit finite difference method was implemented to solve the
temperatures are superimposed in the Mg rich side of the diffusion equations while adopting the appropriate stability
binary phase diagram in Fig. 2. The phase diagram is calcu- criteria. At the end of dissolution simulation, the amount of
lated using FactSage 7.2 along with FTlite database [12]. As solute re-dissolved in hcp Mg matrix was known and used
observed in Fig. 2 the final equilibrium microstructure of Mg- as an input to the next homogenization routine. In this step,
3 wt% Sn alloy only constitutes hcp Mg after solution treat- solute distribution profile was calculated as a function of tem-
ment at 380, 430, and 480 °C (653, 703 and 753 K). In case of perature and time by incorporating the correct boundary con-
Mg-6 wt% Sn, solution treated microstructure at 480 °C com- ditions. In the current study, we developed an implicit control
prises of hcp Mg whereas, at 430 and 480 °C the phase as- volume finite difference model that combines the dissolution
semblage constitutes of hcp Mg and Mg2 Sn phase. Mg-9 wt% and homogenization kinetics in a single calculation routine.
Sn alloy lays in the two-phase region (hcp Mg + Mg2 Sn) for Therefore, dissolution and homogenization results are gener-
all the three solution treatment temperatures. ated using a single calculation routine whereas Sazol et al.
[11] implemented these steps separately. The present model
has following steps which are schematically shown in Fig.
3. Model 3(a)–(d).
Step 1: The following flux balance equation is solved to
3.1. Model explanation calculate the moving boundary velocity for the current time
step:
The dissolution of second phase particles (Mg2 Sn) and
evolution of solute profile (Sn) with solution treatment time  
dx JkR/L − JkL/R
and temperature is explained using a one dimensional (1D) ν= = (1)
implicit diffusion model. Recently, Sazol et al. [11] proposed dt CkR/L − CkL/R
a similar model to calculate the homogenization and dissolu-
tion kinetics in Mg–Al binary alloys. Their model comprised where x is the location of the interface, JkR/L ,JkL/R , CkR/L , and
of two separate calculation routines to capture the complete CkL/R are the fluxes and concentrations of component k at the
solution treatment kinetics. In the first routine, the dissolu- interphase for each phase, respectively.
tion kinetics was calculated by solving a moving boundary The flux Jk for any component k at the interphase is given
problem between hcp Mg and secondary phase. The flux bal- using Fick’s law:
ance equation was adopted to calculate the moving boundary
velocity (dissolution) and the interface between hcp Mg and ∂ Ck
secondary phase was assumed to be under equilibrium. Ex- Jk = −Dkj
n
(2)
∂x
728 A. Dev, N. Naskar and N. Kumar et al. / Journal of Magnesium and Alloys 7 (2019) 725–737

Fig. 3. Illustration of the schematics of the diffusion model (a) model setup with hcp Mg phase and Mg2 Sn secondary phase (b) dissolution of Mg2 Sn phase
(c) redistribution of solute in the hcp Mg matrix (d) Flow chart of the model.

The dissolution length for a given time step is given as Table 1


ν ∗ dt. This new length is added to hcp Mg and subtracted The mobility functions used in the present study.
from the secondary phase. The Sn concentration profile for Interdiffusion coefficient All temperatures Reference
the new incremented length of hcp Mg is taken constant and DM g2 Sn 3.24 ×10−4 ex p ( −109 ,617
) [11]
R∗T
its value is equal to the solubility limit of Sn in hcp Mg at a Mobility Mg HCP solid solution Reference
0.258 ×10−4
given temperature. HCP−Mg
Mg R∗T ex p ( −137
R∗T
,979
) [14,15]
4.869 ×10−4
Step 2: The next step is to calculate the diffusion coeffi- HCP−Mg
Sn R∗T ex p ( −155R∗T
,225.63
) [14,15]
cients as a function of solute concentration. The interdiffusion
coefficient Dkn j in hcp Mg is expressed as [13]:
which is expressed as follows [13]:

n  
∂ μi ∂ μi ∂ Ck  
Dkn j = (δik − ck )ci Mi − (3) = ∇ . Dkn j ∇Ck (6)
i=1
∂cj ∂ cn ∂t
where, Ck is the concentration of solute k, Dkn j is the diffusion
where δ ik is the Kronecker delta, μi is the chemical potential
coefficient and t is time in seconds. The dissolution of inter-
of specie i, ck is the mole fraction of solute k, Mi is the
metallic phase is evaluated by considering a moving boundary
atomic mobility of species which is expressed as [13]:
(also known as Stefan problem) interface between Mg-hcp
 
◦ −Q 1 phase and the secondary phase (Mg2 Sn). Various methodolo-
Mi = Mi exp (4) gies have been to solve the moving boundary problem. Sazol
RT RT
et al. [11] used “Murray and Landis Transformation” to solve
where Q is the activation energy, Mi◦ is a function of jump the moving boundary problem, that uses a variable space grid
frequency and distance, T is the temperature in (K), and R is with constant number of space intervals between fixed and
the universal gas constant. moving boundary [16]. In the present study “Landau Trans-
n
The interdiffusivity Dkj , for the binary Mg–Sn alloy is formation” is used implicitly to solve the moving boundary
given as: problem. In this approach, the length scale of the diffusion
  calculation, that changes as a function of time, is fixed by
∂ μSn ∂ μSn
DSn = (1 − cSn )cSn MSn − coordinate transformation [17,18]. Accordingly, Eq. (1) is re-
∂ CSn ∂ CMg
  arranged in the following manner. The Landau coordinate η
∂ μMg ∂ μMg (x, t) defined as:
−(1 − cSn )cSn MMg − (5)
∂ CSn ∂ CMg x
η= (7)
x (t )
The parameters used to evaluate Eq. (5) are shown in
Table 1 [14,15]. where x(t) is the length of the phase at time t. Substituting
Step 3: The third step involves the calculation of solute Eq. (7) in 6 gives the following transformed equation:
redistribution in hcp Mg phase. The solute diffusion in pri- ∂ Ck 1 ∂ Ck 1
+ ν= ∇ .(D∇Ck ) (8)
mary or the secondary phase is given by Fick’s second law, ∂t x (t ) ∂η x (t )2
A. Dev, N. Naskar and N. Kumar et al. / Journal of Magnesium and Alloys 7 (2019) 725–737 729

Fig. 3. Continued

where ν is the rate of change of element length. The dis- and is uniform throughout the grid. The diffusion equation at
cretization of the transformed diffusion equation in landau interior nodes (η1 ,η2 ,…........, ηN − 1 , ηN ) is discretized in time
coordinates is conducted by introducing spatial discretization using an implicit Euler scheme and in space using control vol-
involving N + 2 nodes (η◦ = 0, η1 ,…, ηN ,ηN + 1 ). In the above ume finite difference method. The final discretized equation
scheme, node 0 is situated on the phase interface and node at any node i is given as:
N + 1 is situated at the phase centre. The grid spacing is given
as:
j+1
Cij+1 − Cij νj C − Cij+1
ηi = ηi+1 − ηi (9) + i i+1
t x (t ) η
730 A. Dev, N. Naskar and N. Kumar et al. / Journal of Magnesium and Alloys 7 (2019) 725–737
 j+1

1 1 Ci+1 − Cij+1 j Ci
j+1 j+1
− Ci−1 the calculated Mg2 Sn phase, and the second phase fraction
= D i
− D (10) are 18.15 μm, 0.61 μm and 3.39% respectively.
x (t )2 η η η
i+1 i

In the next step, the proper boundary conditions are applied 4. Experimental results
at the interface between hcp Mg/intermetallic phase and at
the centre of the Mg hcp grain (node N + 1). The symmetry The SEM microstructures of as-cast and annealed Mg-Sn
boundary condition is given as: (3, 6 and 9 wt% Sn) alloy are shown in Fig. 4. As-cast mi-
∂C crostructures of the Mg–Sn alloys show equiaxed dendrite
j
DN+1 =0 (11) structure. SEM-BSE images of as-cast samples clearly show
∂η
microsegregation in the grain boundary. The second phase
The easiest method to implement this condition is to equate particles were confirmed to be Mg2 Sn from EDS analysis.
the values at node N and N + 1 but this creates numerical in- The variation of second phase fraction in Mg-Sn alloys
stability during the solution. In the current numerical scheme, with time for the three solution treatment temperatures: 653,
a zone of length 0.5∗ η between node N and N + 1 is consid- 703 and 753 K (380, 430, and 480 °C) is shown in Fig. 5. As
ered. If we consider the input flux through the left boundary seen in Figs. 4(a) and 5(a), one hour of solution treatment
j j+1
as DN+1 (CN+1 − CNj+1 )/η, and the flux through right bound- at 753 K is sufficient to dissolve the secondary phase in Mg-
ary as zero, the final discretized diffusion equation at node 3 wt% Sn alloy. The subsequent treatment times has no effect
N + 1 is given as: on the microstructure. At 703 and 653 K the dissolution of
j+1
CN+1 − CN+1j j
νN+1 j+1
CN+1 j+1
− CN−1 Mg2 Sn phase is completed at 12 and 24 h, respectively. As
+ expected the experimental data for Mg-3 wt% Sn alloy indi-
t x (t ) η
1
cates faster dissolution kinetics at higher solution treatment
1 j j+1 j+1
=
−D N+1 C N+1 − CN (12) temperatures. In Mg-6 wt% Sn alloy (Fig. 4(b) and 5(b)) the
x (t )2 η 2 microstructural evolution and dissolution kinetics at 653 and
2
703 K are different compared to Mg-3 wt% Sn. However, at
Similarly, the boundary condition at node 0 is given as: 753 K, secondary phase is dissolved within one hour of an-
∂C nealing in Mg-6 wt% Sn showing similar kinetics compared
D0j =0 (13) to Mg-3 wt% Sn. In case of solution treatment temperature at
∂η
703 K which lies close to solvus temperature, Mg2 Sn phase is
The concentration at node 0 is always updated before dif-
present in the microstructure even after 24 h of holding time.
fusion and following a similar mathematical treatment to node
At 653 K, as seen in Fig. 4(b) there is a significant change
N + 1, the final discretized Equation at node 0 is given as:
in the microstructure evolution as the holding time increases
C0j+1 − C0j ν j C j+1 − C0j+1 from 1 to 24 h. Actually at 653 K, second phase fraction in-
+ 0 1 creases with holding time and is even higher than the frac-
t x (t ) η
1 1
tion observed in as-cast microstructure. It is worth mention-
j j+1 j+1
= 2
2 D1 C1 − C0 (14) ing that at 653 K, this alloy lies in the two-phase region (hcp
x (t ) η
2 Mg + Mg2 Sn) and it appears that Mg2 Sn phase reprecipitates
in the microstructure. The homogenization kinetics are very
3.2. Model set up interesting in case of Mg-9 wt% Sn at all the three solution
treatment temperatures. As seen in Figs. 4(c) and 5(c), at
The inputs to the diffusion simulation are the microstruc- 753 K as holding time increases between 1 and 24 h, Mg2 Sn
tural features of the as-cast Mg–Sn alloys. These inputs are secondary phase, a eutectic constituent dissolves in Mg ma-
Sn solute concentration profile, secondary phase fraction and trix and Sn microsegregation along the grain boundary is re-
grain size in the final solidified microstructure. The solute duced as well. After 24 h, only globular shaped Mg2 Sn pre-
profile is calculated for Mg–Sn binary alloys using the solid- cipitates are observed in the microstructure. At 703 K, the sec-
ification model for a cooling rate of 30 K/s. For Mg-3 wt% ond phase fraction remains almost constant with time. In this
Sn alloy, the initial total length of hcp and Mg2 Sn phase case, as the holding time increases, Mg2 Sn phase is dissolv-
together is 42.8 μm (this is the average core to grain bound- ing and re-precipitating simultaneously in the matrix. These
ary distance measured from as-cast binary alloys, from the two-opposing phenomenon results in constant Mg2 Sn phase
SEM micrographs). The input Mg2 Sn phase length is fixed evolution with holding time, as observed in Fig. 4(c). How-
at 0.22 μm. In this way, the second phase fraction of Mg2 Sn ever, at 653 K, the dissolution kinetics of secondary phase is
phase is initialized to be 0.51%, which is equal to the area sluggish, and the eutectic microstructure is preserved even af-
fraction determined from BSE image analysis of as-cast sam- ter 24 h of holding time. On the other hand, re-precipitation
ples. Similarly, for Mg-6 wt%Sn alloy, the initial total length of Mg2 Sn phase is dominant at this temperature as globular
(hcp and Mg2 Sn phase) was taken as 23.92 μm and the length Mg2 Sn particles can be observed in the microstructure at both
of Mg2 Sn phase was determined to be 0.41 μm, the second the grain boundaries and inside the hcp Mg phase (Fig. 4(c)).
phase fraction here being 1.76%, determined from BSE im- Overall, at 653 K, the total secondary phase fraction increases
age analysis. In Mg-9 wt%Sn alloy, the initial total length, with holding time and is higher than the as-cast fraction for
A. Dev, N. Naskar and N. Kumar et al. / Journal of Magnesium and Alloys 7 (2019) 725–737 731

Fig. 4. SEM-BSE images of as-cast and solution treated microstructure of Mg–Sn alloys (a) 3 wt% Sn (b) 6 wt% Sn (c) 9 wt% Sn.

all the holding times. At 653 K, the solution treatment ki- To further confirm the simultaneous dissolution and pre-
netics are dominated by precipitation phenomena as there is cipitation phenomena during the solution treatment of Mg–
more thermodynamic propensity for nucleation and growth 9 wt% Sn alloy, a systematic SEM-BSE image mapping of
of Mg2 Sn. With increase in holding time, the precipitation as-cast and solution treated samples was performed. Exactly
kinetics are accelerated and the Mg2 Sn phase fraction be- similar areas in as-cast and solution treated samples of Mg–
comes equal to the equilibrium fraction after 24 h. Of course, 9 wt% Sn alloy (4 h at 703 K) were imaged and the micro-
at 653 K the dissolution kinetics are slower, and the solution graphs are shown in Fig. 6. A significant amount of Mg2 Sn
treatment kinetics is dominated by precipitation phenomena. phase present in the eutectic colonies, has clearly dissolved in
732 A. Dev, N. Naskar and N. Kumar et al. / Journal of Magnesium and Alloys 7 (2019) 725–737

Fig. 4. Continued

the Mg matrix during the heat treatment process. Also, coars- The simulation data is compared with the experimental re-
ening of the Mg2 Sn phase is observed as well. In addition, sults of second phase evolution in Mg-6 wt% Sn at 603, 653
as indicated in Fig. 6, the solution treated microstructure also and 703 K in Fig. 9(a)–(c). Since the simulation only consid-
exhibits newly precipitated globular Mg2 Sn phase in the hcp ers the dissolution and redistribution kinetics, there is mis-
Mg matrix. match between the experiments and modelling data at 653 K.
The homogenization treatment at 653 K, induces precipitation
of Mg2 Sn phase in the hcp Mg matrix which increases the
overall second phase fraction. Of course, the dissolution of
5. Modelling results
Mg2 Sn also occurs concurrently but the overall solution treat-
ment is dominated by precipitation phenomena. The simula-
The comparison between the experimental evolution of
tion data and the experimental results show the same decreas-
Mg2 Sn phase fraction and simulation data for Mg-3 wt% Sn
ing trend at 703 K, however, there is difference in the absolute
at 653, 703 and 753 K (380, 430, and 480 °C) is shown in
values as observed in Fig. 9(b). The model only captures the
Fig. 7(a)–(c). The experimental second phase evolution is very
dissolution phenomena and underestimates the experimental
well captured by the dissolution model at 653 K as seen in
second phase fraction. At 703 K this is natural as precipita-
Fig. 7(a). The modelling results in Fig. 7(b) suggest that at
tion phenomenon increases the second phase fraction, but the
703 K, Mg2 Sn phase completely dissolves in the matrix af-
dissolution kinetics dominate the solution treatment process
ter 4 h of annealing whereas the experimental data indicates
that results in overall decrease of second phase fraction with
that higher holding times are required for complete secondary
time. As seen in Fig. 9(c), the solution treatment at 753 K is
phase dissolution. At 753 K, as seen in Fig. 7(c) the dissolu-
entirely dominated by dissolution and the experimental data
tion kinetics of Mg2 Sn phase are very well reproduced by the
of Mg2 Sn dissolution matches very well with the modelling
model. The calculated solute profile hcp grain for Mg-3 wt%
results. The calculation of Sn redistribution profile for Mg-
Sn at 653, 703 and 753 K is shown in Fig. 8(a)–(c). As ex-
6 wt% Sn for different solution treatment times at 653, 703
pected the redistribution of Sn within hcp grain is fastest at
and 753 K is shown in Fig. 10(a)–(c). It is noted that for 653
753 K. The solute profile calculations at 704 and 653 K in-
and 703 K (Fig. 10(a)–(b)), the current redistributions calcu-
dicate that 24 h of solution treatment is insufficient for com-
lations are inadequate to assess the proper homogenization
plete redistribution of Sn in the hcp Mg matrix. On the other
time. Only the dissolution is accounted in the calculations
hand, dissolution results shown in Figs. 4(a) and 5(a) indi-
whereas the experimental data suggests that both dissolution
cate that Mg2 Sn phase dissolves in hcp matrix well before
and precipitation of Mg2 Sn phase should be accounted in the
24 h. Henceforth, to achieve a homogenized microstructure
Sn redistribution calculations. The calculations at 753 K sug-
with complete dissolution of secondary phase and redistri-
gests that 4–8 h are sufficient for complete distribution of Sn
bution of solute within the matrix, both the dissolution and
in the hcp Mg. No dissolution or redistribution calculations
redistribution kinetics are important to consider.
A. Dev, N. Naskar and N. Kumar et al. / Journal of Magnesium and Alloys 7 (2019) 725–737 733

Fig. 6. SEM-BSE image of the same area in as-cast and solution treated
(703 K for 4 h) Mg-9 wt% Sn alloy.

are shown for Mg-9 wt% Sn as precipitation phenomena needs


to be included in the current model to account for proper so-
lution treatment kinetics.

6. Effect of grain size and solute diffusivity

The solution treatment results for Mg 3 wt% Sn indicate


significant variation in dissolution and redistribution kinetics
especially for treatments at low temperature. For instance, at
653 K, the dissolution of Mg2 Sn phase is completed within
12 h whereas complete homogenization of Sn in hcp Mg is not
achieved even after 24 h. The difference between dissolution
and redistribution kinetics is also noticed at 703 and 753 K.
In case of Mg-3 wt% Al alloy, Sazol et al. [11] confirmed
that at 673 K both dissolution and homogenization processes
are complete with 8 h of annealing. In other words, for Mg-
3 wt% Al alloy, 8 h of solution treatment at 673 K is sufficient
to attain a complete homogenized microstructure, devoid of
secondary phases. However, for Mg-3 wt% Sn the solution
treatment kinetics is controlled by solute redistribution phe-
nomena and more than 24 h of holding times are required to
achieve a homogenized microstructure at 653 K.
To investigate the difference in the dissolution and redis-
tribution kinetics between Mg-3 wt% Sn and Al alloys, addi-
tional simulations were performed. The objective of the cal-
culations was to ascertain the effect of solute mobility and
grain size upon the solute treatment kinetics. The dissolution
of Mg2 Sn phase at 653 K for different grain sizes (20, 40 and
60 μm) and Sn mobilities is presented in Fig. 11(a)–(c). As
seen in Fig. 11(a), with Sn mobility as 3.46 × 10−20 m2 /s in
hcp Mg (original Sn mobility as input) and 20 μm as grain
size, the dissolution of Mg2 Sn phase is completed within 4 h.
Fig. 5. Variation of second phase fraction with temperature and time in Mg–
Sn alloys solution treated at 653, 703 and 753 K (a) 3 wt% Sn (b) 6 wt% Sn
In addition, if Sn mobility is increased 10 or 100 times the
(c) 9 wt% Sn alloys. dissolution is completed within one hour for 20 μm grain size.
Exactly similar trend is seen in simulations for 40 and 60 μm
734 A. Dev, N. Naskar and N. Kumar et al. / Journal of Magnesium and Alloys 7 (2019) 725–737

Fig. 7. The comparison between modelling results and the experimental data Fig. 8. The calculated Sn redistribution profiles for Mg-3 wt% Sn alloy at
for Mg-3 wt.% Sn alloy at different temperatures (a) 653 K (b) 703 K (c) different temperatures (a) 653 K (b) 703 K (c) 753 K.
753 K.
A. Dev, N. Naskar and N. Kumar et al. / Journal of Magnesium and Alloys 7 (2019) 725–737 735

Fig. 9. The comparison between modelling results and the experimental data
for Mg-6 wt% Sn alloy at different temperatures (a) 653 K (b) 703 K (c) Fig. 10. The calculated Sn redistribution profiles for Mg–6 wt% Sn alloy at
753 K. different temperatures (a) 653 K (b) 703 K (c) 753 K.
736 A. Dev, N. Naskar and N. Kumar et al. / Journal of Magnesium and Alloys 7 (2019) 725–737

Fig. 11. The modelling results showing the dissolution of Mg2 Sn phase for Fig. 12. The modelling results showing the Sn redistribution in the hcp Mg
Mg-3 wt% Sn as a function of initial grain size and Sn mobility (a) 20 μm phase for Mg-3 wt% Sn as a function of initial grain size and Sn mobility
(b) 40 μm (c) 60 μm. (a) 20 μm (b) 40 μm (c) 60 μm.
A. Dev, N. Naskar and N. Kumar et al. / Journal of Magnesium and Alloys 7 (2019) 725–737 737

grain size (Fig. 11(b) and (c)) with dissolution time decreas- The current modelling results suggest that dissolution and
ing with reduction in grain size and increase in Sn mobility. redistribution kinetics should be considered simultaneously to
The Sn redistribution profile with varying grain sizes (20, determine the optimum solution treatment temperature and
40 and 60 μm) and Sn mobilities is presented in Fig. 12(a)- time. Both these processes are dependent on as-cast mi-
(c). As seen in Fig. 12(a) for 20 μm grain and Sn mobility crostructural features (second phase fraction, microsegrega-
as 3.46 × 10−20 m2 /s in hcp Mg, Sn has not completely re- tion and grain size) and solute mobility. In case of Mg-3 wt%
distributed in Mg matrix even after 24 h. However, when Sn Sn alloys, the dissolution kinetics are faster than redistribution
mobility is increased 100 times, Sn redistribution is achieved kinetics, therefore the correct solution treatment temperature
within one hour of holding time. In this case, Mg2 Sn dis- should be fixed based on Sn redistribution time at a given
solution is also complete within one hour. The results with temperature.
40 μm grain size (Fig. 12(b)) indicate a very similar trend. If
the original Sn mobility: 3.46 × 10−20 m2 /s is used as input, Declaration of Competing Interest
complete Sn homogenization is not achieved even after 24 h
of holding whereas if Sn mobility is increased by one and None.
two orders, complete Sn redistribution is attained in 8 and
1 h, respectively. With 60 μm grain size as input, the redistri- References
bution kinetics are extremely slow, and the results presented
[1] M. Dargusch, The Role of Microstructure in the Creep of Die Cast
in Fig. 12(c) indicate that more than 24 h of holding times
Magnesium Alloy, Univ. Queensland, Brisbane, Australia, 1998.
are required for both original and one order increase in Sn [2] M.S. Yoo, Y.C. Kim, S. Ahn, N.J. Kim, Mater. Sci. Forum 419 (2003)
mobility. 419–422, doi:10.4028/ www.scientific.net/ MSF.419-422.419.
The above calculations clearly elucidate the effect of grain [3] D.H. Kang, M.S. Yoo, S.S. Park, N.J. Kim, Mater. Sci. Forum 475
size and solute mobility upon the dissolution and redistri- (2005) 521–522, doi:10.4028/ www.scientific.net/ MSF.475-479.521.
bution kinetics. Typically, the second phase dissolution pro- [4] D.H. Kang, M.S. Yoo, S.S. Park, N.J. Kim, Mater. Sci. Forum 488
(2005) 759–762, doi:10.4028/ www.scientific.net/ MSF.488-489.759.
cess is faster whereas the redistribution process is relatively [5] D.H. Kang, S.S. Park, N.J. Kim, Mater. Sci. Eng. 413A (2005) 555–560,
slower for a given temperature. The final solution treatment doi:10.1016/j.msea.2005.09.022.
time should be decided based on the redistribution kinetics. In [6] I.-.H. Jung, D.H. Kang, W.-.J. Park, N.J. Kim, S.H. Ahn, Int. J. Mater.
some circumstances both the dissolution and solute homog- Res. 98 (9) (2007) 807–815, doi:10.3139/146.101542.
[7] H.M. Liu, Y.G. Chen, Y.B. Tang, D.M. Huang, G. Niu, Sci. Eng. A 437
enization time can be similar, but it primarily depends on
(2006) 348–355, doi:10.1016/j.msea.2006.07.149.
the grain size, solute mobility and the as-cast second phase [8] H.M. Liu, Y.G. Chen, Y.B. Tang, S.H. Wei, G. Niu, Mater. Sci. Eng. A
fraction. 464 (2007) 124–128, doi:10.1016/j.msea.2007.02.061.
[9] A. Dev, M. Paliwal, J. Cryst. Growth 503 (2018) 28–35, doi:10.1016/j.
7. Summary jcrysgro.2018.09.032.
[10] R. Shi, A.A. Luo, CALPHAD 62 (2018) 1–17, doi:10.1016/j.calphad.
2018.04.009.
A systematic investigation of Mg–Sn (3, 6, 9 wt% Sn) sam- [11] S.K. Das, D.H. Kang, I.H. Jung, Metall. Mater. Trans. A 45 (2014)
ples were performed to determine the variation of Mg2 Sn 5212–5225, doi:10.1007/s11661- 014- 2443- 6.
phase as a function of temperature and time. The experiments [12] C.W. Bale, E. Be´lisle, P. Chartrand, S.A. Degterov, G. Eriksson,
were carried out at 653, 703 and 753 K (380, 430 and 480 °C) K. Hack, I.-.H. Jung, Y.B. Kang, R.B. Mahfoud, J. Melanc¸on, A.D. Pel-
ton, C. Robelin, S. Petersen, CALPHAD 33 (2009) 295–311, doi:10.
for 1, 4, 8, 12 and 24 h. In case of Mg-3 wt% Sn, the solution
1016/j.calphad.2016.05.002.
treatment at all the temperatures is characterized by secondary [13] J.-.O. Andersson, J. Agren, Models for numerical, Appl. Phys. 172
phase dissolution and solute redistribution phenomena which (1992) 1350–1355, doi:10.1063/1.351745.
is well explained by the modelling results. However, for Mg- [14] Z.L. Bryan, P. Alieninov, I.S. Berglund, M.V. Manuel, CALPHAD 48
6 and 9 wt% Sn, re-precipitation of Mg2 Sn phase is also ob- (2015) 123–130, doi:10.1016/j.calphad.2014.12.001.
[15] W. Zhong, J.C. Zhao, Metall. Mater. Trans. A 48 (2017) 5778–5782,
served during the solution treatment process. All the three
doi:10.1007/s11661- 017- 4378- 1.
processes: secondary phase dissolution, precipitation and so- [16] M. Błasik, Sci. Res. Inst. Math. Comput. Sci. 11 (2012) 9–14, doi:10.
lute redistribution are concurrent in these alloys. In Mg-6 and 17512/jamcm.2012.3.02.
9 wt% Sn alloys, low temperature solution treatment is dom- [17] L.N. Tao, SIAM J. Appl. Math. 46 (1986) 254–264, doi:10.1137/
inated by Mg2 Sn precipitation which ultimately controls the 0146018.
overall variation of phase fraction with time, whereas, disso- [18] H.G. Landau, Q. Appl. Math. 8 (1950) 81–94.
lution process is dominant at higher treatment temperatures.

You might also like